Permanent Magnet and Electromechanical Devices
Academic Press Series in Electromagnetism Edited by ISAAK MAYERGOYZ, U...
227 downloads
1763 Views
3MB Size
Report
This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
Report copyright / DMCA form
Permanent Magnet and Electromechanical Devices
Academic Press Series in Electromagnetism Edited by ISAAK MAYERGOYZ, UNIVERSITY OF MARYLAND, COLLEGE PARK, MARYLAND Electromagnetism is a classical area of physics and engineering that still plays a very important role in the development of new technology. Electromagnetism often serves as a link between electrical engineers, material scientists, and applied physicists. This series presents volumes on those aspects of applied and theoretical electromagnetism that are becoming increasingly important in modern and rapidly developing technology. Its objective is to meet the needs of researchers, students, and practicing engineers.
Books Published in the Series Giorgio Bertotti, Hysteresis in Magnetism: For Physicists, Material Scientists, and Engineers Scipione Bobbio, Electrodynamics of Materials: Forces, Stresses, and Energies in Solids and Fluids Alain Bossavit, Computational Electromagnetism: Variational Formulations, Complementarity, Edge Elements M. V. K. Chari and S. J. Salon, Numerical Methods in Electromagnetism Göran Engdahl, Handbook of Giant Magnetostrictive Materials Vadim Kuperman, Magnetic Resonance Imaging: Physical Principles and Applications John C. Mallinson, Magneto-Resistive Heads: Fundamentals and Applications Isaak Mayergoyz, Nonlinear Diffusion of Electromagnetic Fields Giovanni Miano and Antonio Maffucci, Transmission Lines and Lumped Circuits Shan X. Wang and Alexander M. Taratorin, Magnetic Information Storage Technology
Related Books John C. Mallinson, The Foundations of Magnetic Recording, Second Edition Reinaldo Perez, Handbook of Electromagnetic Compatibility
Permanent Magnet and Electromechanical Devices Materials, Analysis, and Applications
Edward P. Furlani Research Laboratories Eastman Kodak Company Rochester, New York
San Diego San Francisco New York Boston London Sydney Tokyo
This book is printed on acid-free paper. Copyright 2001 by Academic Press All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the publisher. Requests for permission to make copies of any part of the work should be mailed to: Permissions Department, Harcourt, Inc., 6277 Sea Harbor Drive, Orlando, Florida 32887-6777. Explicit permission from Academic Press is not required to reproduce a maximum of two figures or tables from an Academic Press chapter in another scientific or research publication provided that the material has not been credited to another source and that full credit to the Academic Press chapter is given. Academic Press A Harcourt Science and Technology Company 525 B Street, Suite 1900, San Diego, CA 92101-4495, USA http://www.academicpress.com Academic Press Harcourt Place, 32 Jamestown Road, London NW1 7BY, UK http://www.academicpress.com Library of Congress Catalog Card Number: 2001089410 International Standard Book Number: 0-12-269951-3 Printed in the United States of America 01 02 03 04 05 06 EB 9 8 7 6 5 4 3 2 1
This book is dedicated to my parents Edward and Helen for their love and sacrifice, my wife Karen for her patience and encouragement, and my two children Edward and Amanda for the joy they give me.
a
This Page Intentionally Left Blank
Contents
Foreword Preface 1. Materials 1.1 1.2 1.3 1.4
1.5 1.6 1.7 1.8 1.9
1.10 1.11 1.12 1.13 1.14 1.15 1.16 1.17 1.18 1.19 1.20
Introduction Units Classification of Materials Atomic Magnetic Moments 1.4.1 Single Electron Atoms 1.4.2 Multielectron Atoms Paramagnetism Ferromagnetism Magnetostatic Energy Demagnetization Field Anisotropy 1.9.1 Magnetocrystalline Anisotropy 1.9.2 Shape Anisotropy Domains Hysteresis Soft Magnetic Materials Hard Magnetic Materials Ferrites Alnico Samarium-Cobalt Neodymium-Iron-Boron Bonded Magnets Magnetization Stability
2. Review of Maxwell’s Equations 2.1 Introduction
xi xiii 1 1 2 6 7 9 14 17 19 23 23 27 27 31 33 36 39 44 46 48 50 51 53 55 66
73 73 vii
viii 2.2 Maxwell’s Equations 2.2.1 Constitutive Relations 2.2.2 Integral Equations 2.2.3 Boundary Conditions 2.2.4 Force and Torque 2.3 Potentials 2.4 Quasi-static Theory 2.5 Static Theory 2.5.1 Magnetostatic Theory 2.5.2 Electrostatic Theory 2.6 Summary
3. Field Analysis
Contents 74 75 77 79 82 82 85 87 87 89 91
97
3.1 Introduction 3.2 Magnetostatic Analysis 3.2.1 Vector Potential 3.2.2 Force and Torque 3.2.3 Maxwell Stress Tensor 3.2.4 Energy 3.2.5 Inductance 3.3 The Current Model 3.4 The Charge Model 3.4.1 Force 3.4.2 Torque 3.5 Magnetic Circuit Analysis 3.5.1 Current Sources 3.5.2 Magnet Sources 3.6 Boundary-Value Problems 3.6.1 Cartesian Coordinates 3.6.2 Cylindrical Coordinates 3.6.3 Spherical Coordinates 3.7 Method of Images 3.8 Finite Element Analysis 3.9 Finite Difference Method
97 97 102 110 112 116 118 126 131 135 141 144 144 153 161 162 175 180 185 190 200
4. Permanent Magnet Applications
207
4.1 Introduction 4.2 Magnet Structures 4.2.1 Rectangular Structures 4.2.2 Cylindrical Structures 4.3 High Field Structures 4.4 Magnetic Latching 4.5 Magnetic Suspension
207 208 208 218 265 268 273
Contents
ix
4.6 4.7 4.8 4.9 4.10 4.11
282 286 306 312 318 325
Magnetic Gears Magnetic Couplings Magnetic Resonance Imaging Electrophotography Magneto-Optical Recording Free-Electron Lasers
5. Electromechanical Devices 5.1 Introduction 5.2 Device Basics 5.3 Quasi-static Field Theory 5.3.1 Stationary Reference Frames 5.3.2 Moving Reference Frames 5.4 Electrical Equations 5.4.1 Stationary Circuits 5.4.2 Moving Coils 5.5 Mechanical Equations 5.6 Electromechanical Equations 5.6.1 Stationary Circuits 5.6.2 Moving Coils 5.7 Energy Analysis 5.8 Magnetic Circuit Actuators 5.9 Axial-Field Actuators 5.10 Resonant Actuators 5.11 Magneto-Optical Bias Field Actuator 5.12 Linear Actuators 5.13 Axial-Field Motors 5.14 Stepper Motors 5.15 Hybrid Analytical-FEM Analysis 5.16 Magnetic MEMS
A. Vector Analysis A.1 A.2 A.3 A.4 A.5 A.6
Cartesian Coordinates Cylindrical Coordinates Spherical Coordinates Integrals of Vector Functions Theorems and Identities Coordinate Transformations
335 335 335 337 338 341 354 354 358 361 363 363 366 376 383 393 402 406 413 421 437 446 455
469 469 473 476 479 485 491
B. Green’s Functions
495
C. Systems of Equations
497
C.1 Euler’s Method
498
x
Contents C.2 Improved Euler Method C.3 Runge-Kutta Methods
502 503
D. Units
509
Index
513
Foreword
This volume in the Academic Press Series in Electromagnetism presents an in-depth, self-contained, and up-to-date treatment of permanent magnets and related electromechanical devices. The emergence of modern rare-earth permanent magnets has led to the tremendous proliferation of devices that utilize these magnets. Currently, there does not exist any book that presents comprehensive treatment of permanent magnet technology and covers material aspects, field analysis, and wide-ranging device applications of permanent magnets. This book, written by Dr. Edward P. Furlani, represents the first and successful attempt to give the exposition of all these issues within one volume. The book reflects the author’s extensive experience as well as his intimate and firsthand knowledge of the area of permanent magnet technology. The unique feature of this book is its broad scope and the proper and delicate balance between theoretical and applied aspects of permanent magnet devices. The book is intended for readers without extensive experience or knowledge in the area of permanent magnets. It reviews the physics of magnetic materials and the basics of electromagnetic field theory. Then, it presents in detail the analysis and design of the wide range of permanent magnet devices such as magnetic gears, couplings, bearings, magnetic resonance imaging devices, free-electron laser magnets, magnetooptical recording systems, permanent magnet motors and stepper motors, and, finally, magnetic microactuators (MEMS). The book contains a large number of practical design examples that are completely worked out. This feature will help readers to follow all the computational and design steps involved in the development of new devices and will be very beneficial for graduate students and practicing engineers. I believe that this book will be a valuable reference for both experts and beginners in the field. Electrical and mechanical enginers, applied physicists, material scientists, and graduate students will find this book very informative. Isaak Mayergoyz Series Editor xi
a
This Page Intentionally Left Blank
Preface
Permanent magnet and electromechanical devices can be found in a wide variety of products that have an impact on our daily lives. Examples of these include audio and video players, telecommunications equipment, personal computers, printers, copiers, automobiles, appliances, power tools, data storage devices, and biomedical apparatus. Furthermore, the use of modern rare-earth magnets is proliferating at a remarkable rate, which is giving rise to new products and markets and stimulating research into a variety of new devices. However, despite the substantial theoretical and practical interest in this technology, there is no single comprehensive text that adequately treats the subject matter required for the research and development of such devices. Instead, one needs to extract crucial information from a variety of texts, each with its own notation and style and emphasis of presentation. It is hoped that this book will eliminate the need for this cumbersome approach. This book is a comprehensive and self-contained exposition of the theory and methods used in the analysis and design of permanent magnet and electromechanical devices. It is intended to be used as both a text and a reference for researchers, professors, students, and engineers who are engaged in the research and development of new and conventional permanent magnet and electromechanical devices. It is written to provide the reader with a first-principles understanding of the underlying theory of such devices, and to enable him or her to analyze, design, and simulate practical devices. The salient features of the book are as follows:
· Extensive discussion of magnetic materials including the physical
·
principles of magnetism, domains, hysteresis, soft and hard magnetic materials and their properties, and the magnetization and stability of permanent magnets. Comprehensive presentation of analytical and numerical methods for the analysis of steady currents, permanent magnets, and xiii
xiv
Preface
· · · · · · · ·
magnetic circuits. Topics include magnetostatic analysis, the Maxwell stress tensor, the current model, the charge model, magnetic circuit analysis, boundary-value theory, finite element analysis, and finite difference analysis. Over 60 examples of practical permanent magnet and electromechanical applications with detailed solutions. Analytical analysis and design formulas for the field distributions of the most common rare-earth permanent magnet structures. Analytical analysis and design formulas for the performance of rare-earth magnetic couplings, gears, and bearings. Comprehensive presentation of the theory of electromechanical devices with numerous solved examples. Analytical analysis and design formulas for linear and rotational actuators. Analytical analysis and design formulas for permanent magnet brushless dc motors and stepper motors. Presentation of a hybrid analytical FEM approach for the analysis of electromechanical devices. Introduction to magnetic microactuators.
This book consists of five chapters and four appendices. The organization of the material is illustrated in the following diagram. Materials Review of Magnetic Field Analysis Theory $ Maxwell’s $ Analytical and Properties Equations Numerical Methods Appendices Permanent Vector Analysis Electromechanical $ Magnet $ $ Green’s Functions Devices Devices ODE Systems Units Book Organization
The first three chapters provide a theoretical background, while the final two chapters emphasize the analysis of practical devices. The first chapter covers magnetic materials. It begins with a brief review of the different units of magnetism, followed by a summary of the major classifications of materials. This is followed by a discussion of the basic mechanisms of magnetism starting at the atomic level with an analysis of an isolated single electron atom. The atomic magnetic mo-
Preface
xv
ment is principally due to the orbital and spin motion of the electron. The discussion is extended to multielectron atoms where the spin and orbital moments of the constituent electrons couple to give a net atomic moment. The theory is then generalized to the case of an ensemble of atoms. If the atomic moments are noninteracting, the ensemble exhibits a bulk paramagnetic behavior. If the atomic moments are interacting, there is a cooperative alignment of the moments (ferromagnetism) below the Curie temperature. This ferromagnetic behavior is combined with the concept of magnetic domains to explain the B-H (hysteresis) curves of bulk magnetization. A survey is given of commercially available soft and hard magnet materials and their properties. The chapter concludes with a discussion of the magnetization and stability of permanent magnets. The second chapter contains a brief but thorough review of Maxwell’s equations. The field equations are presented in both differential and integral form along with the constitutive relations and boundary conditions. The scalar and vector potentials are introduced and shown to provide an alternate and often more tractable formulation of field theory. Quasi-static field theory is discussed. Here, there is a partial uncoupling of the field equations that simplifies their solution. Static field theory is also discussed, and it is shown that the magnetic and electric fields uncouple into separate magnetostatic and electrostatic field equations. The third chapter builds upon the second. It consists of a comprehensive and systematic treatment of the various analytical and numerical methods used for the analysis of steady currents, permanent magnets, and magnetic circuits. Magnetostatic field theory is presented in some detail, including a discussion of force, torque, energy, and inductance. The current and charge models for magnetic materials are introduced. These are used to reduce a permanent magnet to an equivalent source term that can be analyzed using the field equations. Magnetic circuits are discussed, and the concept of reluctance is introduced and used to transform a physical circuit into an equivalent lumped-parameter circuit. Various analysis methods are also discussed, including boundary-value theory, the method of images, finite element analysis (FEA), and the finite difference method. These methods are demonstrated via solved examples. The final two chapters cover permanent magnet applications and electromechanical devices, respectively. These chapters follow a problem solving approach in which numerous practical examples are worked out in sufficient detail to be of immediate use to researchers and practitioners in the field. Many of the examples contain design formulas that cannot be found in any other text.
xvi
Preface
The fourth chapter deals with permanent magnet applications. These include bias magnet structures, high field structures, latching magnets, magnetic suspensions, and magnetic gears and couplings. Analytical methods are used to derive design formulas for many of these applications. Specifically, analytical formulas are derived for the most common rare-earth permanent magnet structures, and the performance of rareearth gears, couplings, and bearings In addition, various miscellaneous applications that utilize magnets are discussed, including magnetic resonance imaging (MRI), electrophotography, magneto-optical recording, and free-electron lasers. The fifth chapter covers electromechanical devices, with an emphasis on magnetically linear, singly excited devices with a single degree of mechanical freedom. Quasi-static field theory is used to derive Kirchhoff’s circuit laws for actuators with stationary circuits, and circuits with coils that move through an external magnetic field. The circuit laws are combined with the laws of rigid body dynamics to obtain the coupled electromechanical equations that govern linear and rotational actuators. The theory is applied to numerous practical devices, including magnetic circuit actuators, linear and rotary actuators, permanent magnet direct current (dc) motors, and stepper motors. The chapter concludes with a brief introduction to magnetic microactuators. In addition to the five chapters, there are four appendices. Appendix A contains a review of vector analysis. The topics covered include vector calculus in the Cartesian, cylindrical, and spherical coordinate systems, line and surface integrals, Stokes’ theorem, the Divergence theorem, various vector identities, and coordinate transformations. In Appendix B the free-space Green’s functions for Poisson’s equation are introduced. These are used to obtain the field distributions of free-standing permanent magnet structures. Appendix C contains a discussion of various numerical methods that are used for the solution of first-order initialvalue problems. Specifically, the Euler, modified Euler, and Runge-Kutta methods are discussed. These are used for the analysis and design of electromechanical devices. Last, Appendix D contains a summary of the different systems of units and conversion factors that are used in the analysis of permanent magnet and electromechanical devices.
Acknowledgments Finally, I would like to acknowledge and thank the individuals who assisted me in the preparation of this book. First, I would like to thank
Preface
xvii
Svetlana Reznik for many insightful discussions on soft and hard magnetic materials. I am also grateful for the patient and diligent work of Christopher Devries in the preparation of numerous drawings and figures. Special thanks go to Dr. Isaak Mayergoyz for his encouragement and support during this project. I would also like to thank Marsha Filion, Julie Bolduc, and the rest of the editorial and production staff at Academic Press for their help and guidance with the preparation of the manuscript. Last, I would like to thank the research managers at the Eastman Kodak Company for their support. Edward P. Furlani
a
This Page Intentionally Left Blank
CHAPTER
1
Materials
1.1 INTRODUCTION In this chapter we study the principles and properties of magnetic materials. Our study begins with a brief review of the basic units of magnetism. This is followed by a summary of the major classifications of magnetic materials. Next, we consider the basic mechanisms of magnetism. We start at the atomic level with an analysis of an isolated single electron atom. Our analysis shows that the electron’s magnetic moment is due to both its orbital motion and spin. The same analysis is then applied to an isolated multielectron atom where the spin and orbital moments of the constituent electrons couple to give a net atomic moment. After discussing single atoms we consider a collection of atoms. Specifically, we apply classical statistical mechanics to determine the bulk magnetization of an ensemble on noninteracting atomic moments. The ensemble exhibits a bulk paramagnetic behavior. A similar analysis is applied to an ensemble of interacting atomic moments. In this case, there is a cooperative alignment of the moments (ferromagnetism) below the Curie temperature. Following this, we introduce the concepts of magnetostatic energy, magnetic anisotropy and domains, and then use these to explain the B-H (hysteresis) curves of bulk magnetization. After describing bulk magnetization, we discuss soft and hard magnetic materials. For soft materials, we give a brief description of the most common materials and list their key physical properties. This is followed by a survey of commercially available hard materials (permanent magnets). The chapter concludes with a discussion of the magnetization and stability of permanent magnets. The topics covered in this chapter are summarized in Fig. 1.1.
1
2
CHAPTER 1 Materials
FIGURE 1.1
Organization of chapter topics.
1.2 UNITS There are three systems of units that are commonly used in the magnetism. These are the CGS or Gaussian system, and two MKS or SI systems that are referred to as the Kennelly and Sommerfeld conventions, respectively. Throughout this book we use SI units in the Sommerfeld convention [1, 2]. In the Sommerfeld convention the magnetic flux is in webers (Wb), the flux density B is in teslas (T) or Wb/m, and both the field strength H and magnetization M are in A/m. The units for the SI (Sommerfeld) and CGS systems are as follows: Symbol
Description
SI
CGS
H
Magnetic Field Strength
A/m
Oe
B
Flux Density
Tesla
Gauss
M
Magnetization
A/m
emu/cm
Flux
Webers
Maxwells
(1.1)
3
1.2 UNITS
The conversion factors for these systems are 1 Oe : 1000/4 A/m 1 Gauss : 10\ T 1 emu/cm : 1000 A/m
(1.2)
1 Maxwell : 10\ Webers Additional conversion factors are discussed in Appendix D. We can obtain an intuitive feeling for the magnitude of the fields (units) described in the foregoing text by considering the following physical examples: 1. Field strength H: Consider a straight, infinitely long wire carrying a current i : 2A. The conductor generates a tangential field strength H : 1 A/m at a radial distance r : 1 m from its center. As another example, consider a long solenoid that has n turns per meter and carries a current of 1/n A. It generates a field strength H : 1 A/m along its axis. 2. Flux density B: Consider an infinitely long conductor that carries a current i : 1 A perpendicular to an external B-field. A force of 1 newton will be imparted to each meter of the conductor when B : 1 T. 3. Flux : Consider a single turn coil with 1 Wb of magnetic flux passing through it. One volt will be induced in the coil when the flux is uniformly reduced to zero in one second. The fundamental element in magnetism is the magnetic dipole. This can be thought of as a pair of closely spaced magnetic poles (Section 3.4.1), or equivalently as a small current loop (Section 1.4) (Fig. 1.2). A magnetic dipole has a magnetic dipole moment m. In the Sommerfeld convention m is measured in A · m. In the Kennelly and CGS systems it is measured in Wb · m and emu, respectively (1 emu : 4 ; 10\ Wb · m). Magnetization M is a measure of the net magnetic dipole moment per unit volume. Specifically, it is given by m M : lim G G , V 4 where m is a vector sum of the dipole moments contained in the G G elemental volume V.
4
CHAPTER 1 Materials
FIGURE 1.2 Magnetic dipole: (a) magnetic charge model and H-field; and (b) current loop and B-field.
If a magnetic dipole is subjected to an external B-field, it acquires an energy E : 9m · B (1.3) and experiences a torque T : m ; B.
(1.4)
In both the Kennelly and CGS systems these are given by E : 9m · H and T : m ; H. In the Sommerfeld convention the fields are related by the constitutive relation B : (H ; M)
(Sommerfeld convention),
(1.5)
5
1.2 UNITS
where : 4 ; 10\ T m/A is the permeability of free space. In the Kennelly convention, the constitutive relation is B: H;J (Kennelly convention), where J is called the magnetic polarization and is measured in tesla. Notice that J : M. The constitutive relation (1.5) simplifies for linear, homogeneous and isotropic media. For such materials, both B and M are proportional to H. Specifically, B : H,
(1.6)
and M : H, (1.7) K where and are the permeability and susceptibility of the material, K respectively. These coefficients are related to one another. From Eqs. (1.5), (1.6), and (1.7) we find that or
: ( ; 1), K
(1.8)
: 9 1. (1.9) K The constitutive relations (1.6) and (1.7) need to be modified for nonlinear, inhomogeneous or anisotropic materials. A material is magnetically nonlinear if depends on H, otherwise it is linear. For nonlinear materials, Eqs. (1.6) and (1.7) become B : (H)H, and M : (H)H, K respectively. A material is inhomogeneous if is a function of position, otherwise it is homogeneous. In inhomogeneous materials the permeability and susceptibility are functions of the coordinate variables, : (x, y, z), : (x, y, z). Last, a material is said to be anisotropic if K K depends on direction, otherwise it is isotropic. In anisotropic materials Eq. (1.6) generalizes to B : H ; H ; H , V V W X B : H ; H ; H , W V W X B : H ; H ; H . X V W X A similar set of equations hold for Eq. (1.7).
6
CHAPTER 1 Materials
1.3 CLASSIFICATION OF MATERIALS Magnetic materials fall into one of the following categories: diamagnetic, paramagnetic, ferromagnetic, antiferromagnetic, and ferrimagnetic (Fig. 1.3). Diamagnetic materials have no net atomic or molecular magnetic moment. When these materials are subjected to an applied field, atomic currents are generated that give rise to a bulk magnetization that opposes the field. Bismuth (Bi) is an example of a diamagnetic material. Paramagnetic materials have a net magnetic moment at the atomic level, but the coupling between neighboring moments is weak. These moments tend to align with an applied field, but the degree of alignment decreases at higher temperatures due to the randomizing effects of thermal agitation. Ferromagnetic materials have a net magnetic moment at the atomic level, but unlike paramagnetic materials there is a strong coupling between neighboring moments. This coupling gives rise to a spontaneous alignment of the moments over macroscopic regions called domains. The domains undergo further alignment when the material is subjected to an applied field. Finally, antiferromagnetic and ferrimagnetic materials have oriented atomic moments with neighboring moments antiparallel to one another.
FIGURE 1.3
Classifications of magnetic materials [16].
1.4 ATOMIC MAGNETIC MOMENTS
7
In antiferromagnetic materials the neighboring moments are equal, and there is no net magnetic moment. In ferrimagnetic materials the neighboring moments are unequal, and there is a net magnetic moment. Rigorous treatments of these materials require a full quantum mechanical analysis that is beyond the scope of this text. We are primarily interested in ferromagnetic materials, and for our purposes it is sufficient to review some of the key principles and results. The interested reader can find more detailed presentations in numerous texts [1—7].
1.4 ATOMIC MAGNETIC MOMENTS In this section we study the magnetic moment of an isolated single electron atom. We start with a classical analysis. Consider an isolated atom with an electron of mass m and charge e moving in a circular orbit C of radius r with a linear velocity u (angular velocity : u/r) as shown in Fig. 1.4. The electron completes an orbit in : 2/ s and has an orbital angular momentum L given by L:m r;u C : m r. (1.10) C The orientation of L is given by a right-hand rule in which the fingers of the right hand curl in the direction of the particle motion, and the thumb points in the direction of L (Fig. 1.4). The orbiting electron defines a closed loop of current i where e i: .
FIGURE 1.4
Atomic system.
8
CHAPTER 1 Materials
The current i defines the flow of positive charge and therefore circulates in a direction opposite to the motion of the electron (Fig. 1.4). The current loop gives rise to a magnetic dipole moment (Fig. 1.2b). In general, if a current i circulates around an enclosed area ds, it gives rise to a magnetic dipole moment m, m : ids
(magnetic moment).
(1.11)
The vector ds defines the orientation of m relative to the circulation of i. This is given by the right-hand rule, which states that if the fingers of the right hand follow the direction of current, then the thumb points in the direction of ds (Fig. A7). The magnitude of the moment is m : ids,
(1.12)
and is measured in units of A · m. Let m denote the magnetic moment of the electron due to its orbital * motion. The magnitude of this moment is given by Eq. (1.12) m : ir * e : r e : r. 2 This can be expressed in terms of L. Specifically, from Eq. (1.10), and the rules for determining the orientations of m and L, we find that * e m :9 L. (1.13) * 2m C Notice that m is antiparallel to L as shown in Fig. 1.4. Amazingly, Eq. * (1.13) holds even at the atomic level when L is replaced by its quantized expression. If the orbiting electron is subjected to an external B-field, its moment m acquires an energy * E : 9m · B * e : L · B, 2m C and experiences a torque T:m ;B * e :9 L ; B. 2m C
9
1.4 ATOMIC MAGNETIC MOMENTS
FIGURE 1.5
Precession of a magnetic moment about an applied field.
As shown in Fig. 1.5, the torque causes L to precess around B. In classical mechanics, L can take on a continuum of orientations relative to B. However, when quantum effects are taken into account, the orientation of L is restricted to a discrete set of values as we shall see.
1.4.1 Single electron atoms Atomic systems are governed by quantum theory with electronic states specified in terms of wavefunctions [8, 9]. The wavefunctions are obtained by solving Schrödinger’s equation
9
; V : E , 2m
where m is the mass of the orbiting particle, V is the potential energy, and E is the total energy. The wavefunctions are indexed by a discrete set of quantum numbers (n, l, m , m ) that distinguish the allowable J Q energy levels, orbital configurations, etc. The principle quantum number is n and this specifies the energy of a given orbit or shell. Observables such as energy and angular momentum are represented by operators that act on . The observed value L of an observable L for a given state
10
CHAPTER 1 Materials
(n, l, m , m ) is obtained by computing the expectation value of its J Q operator L , L
:
*(n, l, m , m )L (n, l, m , m ) dv, J Q J Q
(1.14) where * is the complex conjugate of . It follows from Eq. (1.14) that observables are also quantized and indexed. We now discuss the angular momentum of an atomic electron. Consider a hydrogenlike single electron atom. There are two contributions to the angular momentum: one due to orbital motion and the other due to spin. From quantum mechanics we know that for a given value of n the electron’s orbital angular momentum L has a magnitude
L : [l(l ; 1)] .
(1.15)
In Eq. (1.15), l is the orbital angular momentum quantum number with allowable values l : 0, 1, 2, . . . , (n 9 1), and : h/2 where h is Planck’s constant (h : 6.6260755 ; 10\ J s) [3]. Electrons with l : 0, 1, 2, 3, . . . are referred to as s, p, d, f, . . . electrons, respectively. For example, the n : 1 shell has an s electron, the n : 2 shell has s and p electrons, etc. Notice that the magnitude of L is restricted to a discrete set of values. This is in contrast to classical theory, which allows for a continuum of values. The orientation of L is also restricted. Specifically, its projection onto a given axis (say, the z-axis) is given by L : m , (1.16) X J where m is called the magnetic quantum number, and is restricted to the J values m : l, (l 9 1), . . . , 0, . . . , 9 (l 9 1), 9l. Thus, for a given value of J l there are 2l ; 1 possible orientations of L relative to the axis (Fig. 1.6a). An electron also has a spin angular momentum. The concept of spin was originally proposed to explain the multiplet structure of atomic spectra. Initially, it was thought that spin could be explained in terms of the electron spinning about an internal axis. However, predictions based on this model do not yield the correct magnetic moment. Therefore, spin is viewed as a purely quantum mechanical phenomenon. The spin angular momentum S is quantized and has a magnitude S : [s(s ; 1)] , where s : 1/2. The projection of S along a given axis is also quantized with S : m , (1.17) X Q where m : <1/2 (Fig. 1.6b). Q
1.4 ATOMIC MAGNETIC MOMENTS
11
Spatial quantization of angular momentum: (a) orbital angular momentum (l : 2); and (b) spin angular momentum [4].
FIGURE 1.6
The total angular momentum of the electron, which is denoted by J, is the sum of its orbital and spin angular momenta J : L ; S. It has a magnitude
J : [ j( j ; 1)] , where j : l < 1/2. The total angular momentum depends on the magnitude and relative orientations of the orbital and spin angular momenta. We now consider the electron’s magnetic moment. There are two contributions to the moment: one due to the orbital motion and the other due to spin. Recall from the classical result of Eq. (1.13) that the orbital moment m is proportional to L. This same relation holds at the atomic * level, e m :9 L, * 2m C but now the magnitude of m is quantized, * e m :9 [l(l ; 1)]
* 2m C : 9 [l(l ; 1)],
(1.18)
where : e /2m is the Bohr magneton. Notice that m is restricted by C *
12
CHAPTER 1 Materials
the allowable values of l. If the electron is subjected to an external H-field, its orbital magnetic moment couples to H, and the electron experiences a torque that causes m to precess around H. However, the * orientation of m relative to H is restricted to a discrete set of values as * described in the preceding. Specifically, the projection of m along H * (denoted m ) is *& m : 9 m , *& J where m is as specified in Eq. (1.16). Thus, for a given value of l there J are 2l ; 1 possible orientations of m relative to H (Fig. 1.6a). * The magnetic moment m due to spin is given by 1 e m : 9 S. 1 m C It has a magnitude m : 92 [s(s ; 1)]. (1.19) 1 The spin also couples to an external field, and has a projection : 92 m . 1& Q The electron’s total magnetic dipole moment m is the sum of its orbital and spin moments (Fig. 1.7): m
m :m ;m * 1 e :9 (L ; 2S) 2m C As J : L ; S, this can also be written as
Notice that m
e m :9 (J ; S). 2m C is not antiparallel to J.
FIGURE 1.7
Addition of angular momenta.
(1.20)
(1.21)
13
1.4 ATOMIC MAGNETIC MOMENTS
From Eq. (1.20) we see that m depends on both the magnitude and relative orientations of L and S. However, L and S are neither independent of one another nor constant in time. Rather, they are coupled via a spin-orbit interaction. This interaction can be understood by considering a reference frame attached to the electron. In this frame, the nucleus appears to revolve around the electron. Therefore, the electron is subjected to a magnetic field B due to the orbiting charged nucleus. L Moreover, B couples to the spin magnetic moment m with an energy L 1 E . m · B . As B is parallel to L and m is antiparallel to S, we have 1 L L 1 E : L · S, where is the spin orbit parameter. Notice that E depends on the angle between S and L. Consequently, a torque is imparted perpendicular to both L and S causing them to precess. We may visualize this as a precession of L and S about the total angular momentum J, which is constant (absent any external influences). It follows from Eq. (1.20) that m also precesses around J. Therefore, it is time dependent and cannot be specified as a fixed vector. However, we can define a constant average-value vector m . Specifically, as J is constant, we can specify m by taking the time-averaged value of m in the direction of J (Fig. 1.8). This is given by m
J J : m · . J J
(1.22)
Substituting Eq. (1.21) into Eq. (1.22), and making use of the fact that J : L ; S $ S · J : 1/2(J · J ; S · S 9 L · L), we get
(1.23)
j( j ; 1) ; s(s ; 1) 9 l(l ; 1) . 2j( j ; 1)
(1.24)
m
:9
e gJ 2m C
where g:1;
FIGURE 1.8
Spin-orbit coupling and the orientation of m
[9].
14
CHAPTER 1 Materials
This is called the Landé g factor. From Eq. (1.23) we see that the magnetic moment of an electron can be determined given its angular momentum. This generalizes to the case of multielectron atoms, which we study next.
1.4.2 Multielectron atoms In this section we consider the magnetic moment of an isolated multielectron atom. When multiple electrons are present, each must occupy a different orbit as defined by the four quantum numbers n, l, m , and m . This is the Pauli exclusion principle which states that no two J Q electrons may have the same set of quantum numbers. The magnetic moment of a multielectron atom can be determined from the total angular momentum J using a relation similar to Eq. (1.23). Since the atom is isolated, J is a constant with a magnitude
J : [J(J ; 1)] . The projection of J onto a given axis is J : M , X ( where M : J, J 9 1, J 9 2, . . . , 9(J 9 2), 9(J 9 1), 9J. To determine J ( or J we need to know the allowed values of J. These depend on the X orbital and spin configurations of the constituent electrons, and the way in which their respective moments couple. There are two different modes of coupling. These are known as Russell-Saunders and spin-orbit coupling,
L ;S (Russell-Saunders) J : or (1.25) (L ; S ) (spin-orbit). G G G The rule for evaluating Eq. (1.25) is that vectors with the strongest coupling are summed first. Russell-Saunders coupling applies when the orbital-orbital (L -L ) and spin-spin (S -S ) couplings between different G H G H electrons are stronger than the spin-orbit (S -L ) coupling of each separG G ate electron. This applies to all but the heaviest atoms. On the other hand, spin-orbit coupling is used when the spin-orbit coupling of each separate electron dominates the couplings between different electrons. Russell-Saunders coupling applies to most magnetic atoms. In this coupling the total orbital angular momentum is the vector sum of orbital momenta of the individual electrons L :L . G G
1.4 ATOMIC MAGNETIC MOMENTS
15
It has a magnitude
L : [L (L ; 1)] , and its projection along a given direction (say, the z-axis) is L : M , X * where M : L, (L 9 1), . . . , 0, . . . , 9 (L 9 1), 9L, and M : m . * * G JG Similarly, the total spin angular momentum is given by
This has a magnitude and a projection
S :S . G G
S : [S(S ; 1)] ,
S : M , X 1 where M : S, (S 9 1), . . . , 0, . . . , 9(S 9 1), 9S, and M : m . The net 1 1 G QG angular momentum of the electrons in a closed shell is zero,
and
L : 0, G
S : 0. G Therefore, only the partially filled shells contribute to J . An example of Russell-Saunders coupling is shown in Fig. 1.9. Recall that the total angular momentum is J : L ; S . This has a magnitude
J : [J(J ; 1)] . In Russell-Saunders coupling J is restricted to the following values: J : L ; S, L ; S 9 1, . . . , L 9 S . To evaluate J we need to know the ground state orbital and spin configurations of the electrons. These are determined using Hund’s rules, which state that: (i) the spins S combine to give the maximum G value of S consistent with the Pauli exclusion principle; (ii) the orbital vectors L combine to give the maximum value of L compatible with the G Pauli principle and condition (i); and (iii) the resultant L and S combine to form J , with J : L 9 S if the shell is less than half filled, and J : L ; S if it is more than half filled. When the shell is exactly half full, L : 0 and J : S.
16
CHAPTER 1 Materials
FIGURE 1.9
Russell-Saunders coupling [9].
Finally, for an isolated atom J is constant. We determine an average magnetic moment of the atom m by taking the projection of its magnetic moment m along the direction of J . Following an analysis similar to that for Eq. (1.22) we obtain
m : g (J(J ; 1), where g:1;
J(J ; 1) ; S(S ; 1) 9 L(L ; 1) . 2J(J ; 1)
Notice that if the net orbital angular momentum is zero (L : 0), then J : S and g : 2. On the other hand, if the net spin angular momentum is zero (S : 0), then J : L and g : 1. The projection of m along the direction of an applied H-field is denoted by m &
m : g M , (1.26) & ( where M : J, J 9 1, J 9 2, . . . , 9(J 9 2), 9(J 9 1), 9J. Thus, m is ( & constrained to a discrete set of orientations relative to H. This completes our study of isolated atoms. In summary, we have found that the magnetic moment of the atom is due principally to the magnetic moments of its constituent electrons. These, in turn, are due to both orbital motion and spin. The magnitude of the atomic moment can
17
1.5 PARAMAGNETISM
be determined from the total angular momentum, and this is obtained by summing the contributions of the individual electrons using RussellSaunders coupling, subject to Hund’s rules.
1.5 PARAMAGNETISM Recall that paramagnetic materials consist of atoms that have a net magnetic moment, but negligible coupling between neighboring moments. In this section we study the response of such materials to an applied field. This serves as a prerequisite to our subsequent discussion of` ferromagnetism. Consider a paramagnetic specimen with N noninteracting atomic magnetic moments per unit volume. We analyze its response to an applied field H . Initially, we treat the moments classically assuming ? that they can take on a continuum of orientations. Later, we impose quantum constraints. Let m denote the magnetic moment of a single atom. We seek an expression for the magnetization M, which is the total magnetic moment per unit volume in the direction of H . Each moment m couples to H ? ? and acquires a magnetostatic energy E : 9 m · H ? : 9 mH cos(), (1.27) ? where is the angle between m and H . We model the material as an ? ensemble of noninteracting atomic moments [10]. Specifically, we use Boltzmann statistics to describe the probability p(E) of a moment having an energy E at a temperature T p(E) : exp : exp
9E k T
mH cos() ? , k T
(1.28)
where k is Boltzmann’s constant. Physically, the alignment of m is opposed by thermal agitation and Eq. (1.28) gives the probability of m obtaining an angular orientation relative to H at temperature T. The ? magnetization is given by M:
L
m cos()n() d,
(1.29)
18
CHAPTER 1 Materials
FIGURE 1.10
Langevin function.
where m cos() is the projection of m (with orientation relative to H ) ? onto the field direction and n() d is the number of atomic moments per unit volume with angular orientations between and ; d. Specifically, n() d : 2C exp
mH cos() ? sin() d, k T
(1.30)
where C is a normalization constant determined from L n() d : N. An evaluation of Eq. (1.29) yields M : NmL
mH ? , k T
(1.31)
where L() is the Langevin function 1 L() : coth() 9 , (see Fig. 1.10). Notice that 91 L() 1 and that lim L() : 1. (1.32) A From Eqs. (1.31) and (1.32) we find that for low T and high H , M ? approaches its maximum value Nm, which means that the atomic moments tend towards perfect alignment in this limit.
19
1.6 FERROMAGNETISM
In most cases 1, and L() can be represented as an infinite power series of the form, L() : 9 ; · · · . 3 45 For small we ignore all but the first term and obtain NmH ?. M: 3k T This gives a parametric susceptibility of the form :
M Nm : . 3k T H ?
This is known as Curie’s law. The expression (1.31) follows from a classical analysis that allows for arbitrary orientations of the moments. However, from Eq. (1.26) we know that m is constrained to a discrete set of orientations relative to H . ? We modify Eq. (1.29) taking this into account and obtain ( m g exp( m g H /k T) ( ( ? M : N K(\( ( exp( m g H /k T) ( ? K(\( This can be written as M:M B (
gJ H ? , k T
(1.33)
where M : NgJ is the maximum possible moment in the direction of H , and ? 2J ; 1 (2J ; 1) 1 B () : coth 9 coth (1.34) ( 2J 2J 2J 2J
is the Brillouin function. In the limiting case when J ; - the allowed values and orientations of the atomic moments approach a continuum, and we recover the classical Langevin result of Eq. (1.31).
1.6 FERROMAGNETISM Ferromagnetic materials consist of atoms with a net magnetic moment, and there is substantial coupling between neighboring moments. These
20
CHAPTER 1 Materials
materials exhibit a magnetic ordering that extends over a substantial volume and entails the cooperative orientation of a multitude of atomic moments. An early attempt to explain this phenomenon was made by Weiss, who postulated the existence of a ‘‘molecular field’’ H that is K proportional to the magnetization H : M. (1.35) K In Eq. (1.35) is the mean field constant [11]. Consider a material characterized by Eq. (1.35). We seek an expression for M and follow the same analysis as in the previous section, but now H : H . Specifically, the magnetostatic energy is given by ? K E : 9 m · H K : 9 mM cos(), (1.36) where is the angle between the local moment m and the bulk magnetization M. Substitute H : M into Eq. (1.31) and obtain ? mM M : NmL . (1.37) k T
Notice that Eq. (1.37) does not give M explicitly in terms of T. Instead, graphical or self-consistent methods are used to determine M for a given T. A careful analysis of Eq. (1.37) shows that spontaneous magnetization occurs at low temperatures, and the material becomes saturated with M : M : Nm. However, as T increases, the magnetization gradually decreases until T approaches the Curie temperature, Nm T : A 3k
(Curie temperature).
Near the Curie temperature, the magnetization drops dramatically to zero, and as shown in Fig. 1.11 the material becomes paramagnetic. Below the Curie temperature, the material divides into numerous distinct localized regions called domains. Each domain is uniformly magnetized to its saturation value. However, the bulk magnetization can be zero if the domains are randomly oriented. Domains are discussed in more detail in Section 1.10. If we subject a ferromagnetic material to an applied field H , the ? same analysis as in the preceding applies, but now H ; H ; H . The K ? K
21
1.6 FERROMAGNETISM
FIGURE 1.11
Spontaneous magnetization vs temperature.
magnetostatic energy is given by E : 9 m · (H ; H ) ? K : 9 m(H ; M) cos() ? and we obtain M : NmL
m(H ; M) ? . k T
(1.38)
In ferromagnetic materials M H and, therefore, the applied field ? does not alter the local magnetization appreciably. That is, H does ? not alter the magnetization within a domain. However, an applied field does alter the bulk magnetization of a specimen by altering the orientation and structure of individual domains. This is discussed in Section 1.11. Expression (1.38) is based on classical theory in which the magnetic moments can take on a full range of orientations. If we apply quantum
22
CHAPTER 1 Materials
theory, the orientations of the moments are restricted, and we obtain an expression similar to Eq. (1.33) with H ; H ; M , ? ? Q M :M B Q (
gJ (H ; M ) ? Q , k T
(1.39)
where M : NgJ , and B () is the Brillouin function (1.34). Because ( M H , Eq. (1.39) reduces to ? M :M B Q (
gJ M Q . k T
(1.40)
As J ; - the allowed values and orientations of the atomic moments approach a continuum, and Eq. (1.40) reduces to the classical result seen in Eq. (1.37). Although the Weiss mean field theory explained some key features of ferromagnetism such as spontaneous magnetization below the Curie temperature, it nevertheless lacked a rigorous foundation. This would come later with the discovery of quantum mechanics [12]. Specifically, Heisenberg postulated the existence of an exchange interaction between spin angular momenta. This is characterized by a potential energy (exchange energy) of the form, w : 92J S · S GH G H : 92J S S cos() G H
(exchange energy),
(1.41)
where J is called the exchange integral, S and S are the coupled spin G H angular momenta, and is the angle between S and S . Notice G H that if J 0, the exchange energy is minimum when the spins are parallel ( : 0). However, if J 0, the exchange energy is a minimum when the spins are antiparallel ( : ). Therefore J 0 gives rise to ferromagnetic behavior, and J 0 gives rise to antiferromagnetic be havior. The exchange force falls off rapidly and is rarely effective beyond the nearest neighbors. The forementioned analysis describes the local ordering (spontaneous alignment) of atomic moments and the temperature dependence of this ordering. However, it does not describe the hysteretic behavior of bulk materials. For this, we need to understand magnetic domains and their behavior in an external field. In the next few sections we introduce the concepts of magnetostatic energy, demagnetizing field, and magnetic anisotropy. These are key to understanding domains. Following this, we study domains specifically and then use them to describe hysteresis.
23
1.8 DEMAGNETIZATION FIELD
1.7 MAGNETOSTATIC ENERGY In this section we discuss the magnetostatic self-energy of a magnetized specimen, and the energy it acquires in an external field. Consider a specimen of volume V with a fixed magnetization M. It possesses a magnetostatic self-energy that is given by W :9 Q 2
4
M · H dv +
(self-energy),
(1.42)
where H is the field in the specimen due to M (i.e., due to the + continuum of dipole moments of which the specimen is composed [3]). The self-energy equation (1.42) can be derived by considering the energy required to assemble a continuum of dipole moments in the absence of an applied field [6]. Notice that the energy density associated with W is Q w :9 M·H . (1.43) Q + 2 The self-energy equation (1.42) plays an important role in the formation of magnetic domains. We discuss this more in Section 1.10. If a magnetized specimen is subjected to an applied field H , it ? acquires a potential energy W : 9 ?
4
M · H dv. ?
(1.44)
This can be viewed as the work required to move the specimen from an environment with zero field to a region permeated by H . The express? ion (1.44) can be used to determine the force or torque imparted to a magnetized body by an external field. From Eq. (1.44) we see that the energy density due to the coupling of M to H is given by ? w : 9 M · H ? ? : 9 MH cos(), (1.45) ? where is the angle between M and H . ?
1.8 DEMAGNETIZATION FIELD When a specimen is magnetized, a self-field develops within it that opposes the magnetizing field. This is called the demagnetization field.
24
CHAPTER 1 Materials
We study this in some detail because it plays an important role in the magnetization process. Consider a uniformly magnetized specimen with a volume V and a surface S. Its magnetization M gives rise to surface poles (Section 3.4). These, in turn, give rise to a demagnetization field H within the B specimen. The field H is proportional to M, but in the opposite B direction. For example, along the x-axis we have H : 9N M , (1.46) BV V V where H and M are the x-components of H and M, respectively, and B V B N is theV demagnetizing factor along the axis. If the specimen is V subjected to an applied field H , then H inside the specimen (denoted ? H ) is the vector sum of H and H , ? B H :H ;H . (1.47) ? B For example, consider the bar magnet shown in Fig. 1.12. Outside the magnet H and B are in the same direction (Fig. 1.12a,b). However, inside the magnet H : H , which opposes both M and B (Fig. 1.12c). B Demagnetizing factors depend on the permeability and shape of a specimen and are difficult to obtain in closed-form. However, analytical expressions have been derived for ellipsoidal shapes. Ellipsoids have
Field distribution of a bar magnet: (a) H-field; (b) B-field; and (c) B, H, and M inside the magnet.
FIGURE 1.12
25
1.8 DEMAGNETIZATION FIELD
FIGURE 1.13
Ellipsoidal shapes: (a) general ellipsoid; and (b) prolate ellipsoid.
three demagnetizing factors N , N , and N corresponding to the three ? @ A orthogonal axes with lengths a, b, and c, respectively (Fig. 1.13a). These factors satisfy the following relation, N ;N ;N :1 (ellipsoid). (1.48) ? @ A For the special case of a sphere where a : b : c we have N : N : ? @ N : 1/3 and A H : 9M (sphere). (1.49) B Another shape of interest is the prolate spheroid with a semimajor axis c and semiminor axis a (i.e., a : b c) (Fig. 1.13b). The demagnetization factor along the c-axis is
1 k N : ln(k ; (k 9 1) 9 1 , A k 9 1 (k 9 1 where k : c/a [2]. For a long cylinder that is magnetized along its axis, the demagnetizing factors can be approximated by N : 0, and A N : N : 1/2, where the c-direction is along the axis of the cylinder. ? @ Finally, consider a magnetized specimen with H : 9NM. Its selfB energy is given by Eq. (1.42) with H : H . Specifically, we have + B W : 9 M · H dv Q B 2 4 N : M dv. (1.50) 2 4 If M is constant throughout V, then Eq. (1.50) reduces to
NMV W : . Q 2 We demonstrate some of these principles in the following example.
26
CHAPTER 1 Materials
FIGURE 1.14
Sphere with linear permeability in an external field.
EXAMPLE 1.8.1 Determine the magnetization and dipole moment of a sphere of linear isotropic material in free space that is subjected in an external field H (Fig. 1.14). SOLUTION 1.8.1 Let and denote the permeability and susceptibility of the K sphere, and let R denote its radius. The field inside the sphere is H :H ;H B : H 9 M. Notice that we have used Eq. (1.49). Recall from Eq. (1.7) that M: H . K
(1.51)
Therefore,
As : / 9 1 we have K
H H : . 1 ; /3 K
3H . H : 2 ; / The magnetization inside the sphere is obtained from Eq. (1.51), M:
3( 9 ) H . ( ; 2 )
The dipole moment m is m : Volume ; M : RM.
(1.52)
(1.53)
27
1.9 ANISOTROPY
Therefore, we find that m:
4R( 9 ) H . ( ; 2 )
(1.54) )
1.9 ANISOTROPY A material with magnetic anisotropy is one in which the magnetic properties are different in different directions. The principal classifications of anisotropy are as follows: 1. 2. 3. 4.
magnetocrystalline anisotropy; shape anisotropy; stress anisotropy; and exchange anisotropy.
A thorough discussion of these can be found in Cullity [5]. In this section we review magnetocrystalline and shape anisotropy. These play an important role in the magnetization process.
1.9.1 Magnetocrystalline anisotropy Magnetocrystalline anisotropy is an intrinsic property of a material in which the magnetization favors preferred directions (easy axes). For example, iron (Fe) has a body-centered cubic lattice structure, and there are six easy, equally preferred directions of magnetization: [0,0,1]; [0,1,0]; [1,0,0]; [0,0,91]; [0,91,0]; and [91,0,0] (Miller indices) (Fig. 1.15). The preferred alignment within a material constitutes a lower energy state and is primarily due to spin-orbit coupling. Specifically, atomic electrons couple to the crystal lattice in such a way that their orbital moments tend to align along the crystallographic axes, and because of spin-orbit coupling the spin moments tend to align as well. The most elementary form of magnetocrystalline anisotropy is uniaxial anisotropy. If a specimen with uniaxial anisotropy is polarized at an angle relative to its easy axis, the anisotropy energy density is w K sin() (J/m). (1.55)
Notice that energy minima occur when : 0° and 180°. Consider the response of a single domain specimen with uniaxial anisotropy when it is subjected to an applied field H . Assume that H ? ?
28
FIGURE 1.15
CHAPTER 1 Materials
Crystal structure of iron and magnetization curves along various
axes.
is directed at an angle with respect to the easy axis (Fig. 1.16). We analyze the behavior of this specimen following the presentation of Stoner and Wohlfarth [13]. Assume that the specimen is magnetized to saturation with M at an angle with respect to the easy axis. Further assume that demagnetization field H is negligible. Therefore, inside the B specimen H : H in accordance with Eq. (1.47). As the specimen ? constitutes a single domain, its energy density w() consists of two parts: the magnetostatic energy density due to the coupling of M to H as given ? by Eq. (1.45) and the magnetocrystalline anisotropy energy density w
as given by Eq. (1.55). The total energy density is w() : K sin() 9 M H cos( 9 ) Q ? : K sin() 9 M [H cos() ; H cos()], (1.56) Q V W where H : H cos() and H : H sin(). We identify the values of V ? W ?
29
1.9 ANISOTROPY
FIGURE 1.16
Specimen with uniaxial anisotropy.
that represent energy minima using the second derivative test where w()/ : 0 identifies the critical points, and w()/ 0 confirms a minimum. We consider the case where the field is applied along the easy axis (H : 0) and obtain W w() : 2K sin() cos() ; M H sin() : 0, (1.57) Q V and w() : 2K cos(2) ; M H cos() 0. Q V
(1.58)
Notice that minima occur when : 0 or . When : 0 a minimum occurs when 2K H , (1.59) V M Q and we have M : M cos() : M . When : a minimum occurs when V Q Q 2K H (1.60) V M Q and we have M : 9M . The conditions (1.59) and (1.60) identify the V Q intrinsic coercivity H (or switching field) of the specimen. Specifically, H is given by 2K H Y . M Q The intrinsic coercivity is an important property of permanent magnet materials. It gives a measure of the field required to magnetize and demagnetize a specimen (Section 1.19). 9
30
FIGURE 1.17
CHAPTER 1 Materials
Response of a specimen with uniaxial anisotropy: (a) M vs H; and
(b) B vs H.
We apply these results to the behavior of a single domain specimen with uniaxial anisotropy. Assume that the specimen is initially magnetized along its easy axis with : 0 (M : M x ) (Fig. 1.17a). We apply a Q reversal field in the direction : (H negative). As we gradually V increase the field (in a negative sense), the specimen remains polarized along its easy axis (at an energy minimum) as long as Eq. (1.59) is satisfied (i.e., as long as 9H H 0). When the field increases to the V
1.9 ANISOTROPY
31
point where H : 9H then Eq. (1.59) is violated and M flips orientation V to : (M : 9M x ). When this happens Eq. (1.60) is satisfied and the Q specimen is once again in a state of minimum energy. The specimen stays in this state as H becomes more negative. V Next, we gradually decrease H (making it less negative). The V specimen stays with M : 9M x for all negative field values. Eventually, Q H reaches zero. We then gradually increase H from zero until we reach V V H : H . At this point, Eq. (1.60) is violated and M flips orientation to V : 0 (M : M x ) in order to satisfy Eq. (1.59). Once this occurs the Q specimen is again at an energy minimum. The specimen remains in this state for any further increase H . V This entire sequence of events can be summarized in graphical form by plotting M vs H (Fig. 1.17a). Notice that this plot has the form of a square loop. In particular, the specimen starts out with M : M . When Q H is reversed and reaches H : 9H , the magnetization abruptly re verses direction to M : 9M . When H is reversed again and increases Q to H : H , the magnetization abruptly reverses back to its original value M : M . Thus, except for the two points H :
1.9.2 Shape anisotropy Shape anisotropy is not an intrinsic property of a material. Rather, it is due to a geometrically induced directional dependency of the demagnet-
32
CHAPTER 1 Materials
FIGURE 1.18
Specimen with shape anisotropy.
ization field H within the specimen. For example, consider a specimen in B the shape of` a prolate spheroid with a semimajor axis c and semiminor axis a (Fig. 1.18). Assume that the specimen is subjected to an applied field H ? at an angle with respect to the c-axis. Further assume that the specimen is uniformly polarized to a level M with M at an angle with respect to the Q c-axis. The field inside the specimen is H : H ; H . The energy density ? B w() consists of the self-energy density w as given by Eq. (1.43) with Q H : H , and the magnetostatic energy density due to the coupling of M + B to H as given by Eq. (1.45). Thus, the total energy density is ? 1 w() : 9 M · H 9 M · H B ? 2 : M[N sin() ; N cos()] 9 M H cos( 9 ) A Q ? 2 Q ? : MN ; K sin() 9 M [H cos() ; H cos()], (1.62) Q Q V W 2 Q A where M K : Q (N 9 N ). Q ? A 2
(1.63)
Notice that except for the leading constant term, Eq. (1.62) is formally similar to Eq. (1.56) of Section 1.9.1. We follow the same energy minimization analysis presented there except that here the c-axis plays the role of the easy axis. We find that the intrinsic coercivity (switching field) is given by 2K Q H Y M Q : M (N 9 N ). (1.64) Q ? A
33
1.10 DOMAINS
FIGURE 1.19
The B-H loop for a specimen with shape anisotropy.
From Eq. (1.48), and the fact that N : N for a prolate spheroid, we have ? @ 2N ; N : 1. Therefore, ? A
H M (shape anisotropy). (1.65) Q From the discussion following Eq. (1.61) we know that in this case, H : H , and the demagnetization curve is linear over only a portion of` A the second quadrant as shown in Fig. 1.19.
1.10 DOMAINS We have seen that the coupling of atomic moments in ferromagnetic materials gives rise to a spontaneous alignment of the moments below the Curie temperature (Section 1.6). Thus, one would expect ferromagnetic materials to be magnetically saturated at room temperature. However, we observe that these materials are often unmagnetized at a macroscopic level. This can be explained by the concept of magnetic domains [1, 11]. Domains typically contain 10—10 atoms. The atomic moments within a domain are subject to a magnetocrystalline anisotropy that favors spontaneous alignment along certain preferred crystallographic axes (Section 1.9). Thus, the moments within a domain align parallel to
34
CHAPTER 1 Materials
one another and the domain represents a local region of magnetic saturation. A macroscopic or bulk sample of material consists of a multitude of domains that vary in size, shape, and orientation. The magnetization of the sample is defined by the collective structure and orientation of all its domains. For example, it is unmagnetized if the domains are randomly oriented, and saturated if the domains are uniformly aligned. This explains the apparent contradiction between spontaneous ferromagnetic alignment of atomic moments and the existence of macroscopically unmagnetized ferromagnetic materials. In homogeneous materials, domains form in such a way as to minimize the total energy of the specimen. For example, consider the block of material shown in Fig. 1.20. The magnetostatic energy of this block is highest in configuration a, and lowest in configuration c. In inhomogeneous materials the domain structure is affected by inhomogeneities such as inclusions, dislocations, crystal boundaries, internal stress, etc. Inclusions are regions with different magnetic properties than the surrounding material, and dislocations are line imperfections in the crystal lattice. Neighboring domains are separated by a transition layer called a domain wall (Fig. 1.21a). These are also called Bloch walls since Bloch
Magnetized material: (a) high energy state; (b) intermediate energy state; and (c) low energy state.
FIGURE 1.20
1.10 DOMAINS
35
Domain structure: (a) domain wall; and (b) domain wall movement in an applied field.
FIGURE 1.21
was the first to study their properties [14]. A domain wall has a finite thickness and an energy associated with it. The wall thickness and energy follow from an analysis of two competing effects: magnetocrystalline anisotropy that favors spin alignment along a preferred axis (thinner wall) and exchange interactions that favor parallel alignment of neighboring spins (thicker wall). The orientation of the spins within a wall varies gradually across its width so that the spins at either side of the wall are aligned with the respective adjacent domains. For example, if two neighboring domains are polarized in opposite directions, the spins in the wall between them rotate 180° across the wall from one side to the other (Fig. 1.21a). Domain walls can move in response to an applied field. For example, if an external field is applied parallel to the magnetization in a domain, the spins within the wall experience a torque that tends to rotate them toward the field direction. As the spins rotate, the wall effectively moves with the aligned domain expanding and antialigned domain contracting (Fig. 1.21b).
36
CHAPTER 1 Materials
The wall motion can be impeded by the forementioned listed inhomogeneities. For example, an inclusion lowers the energy of a domain wall when they intersect. Thus, a domain wall will tend to ‘‘stick’’ to an inclusion when it contacts it. Similarly, localized regions of inhomogeneous strain (which can be caused by defects and dislocations) give rise to local energy barriers that impede the motion of a domain wall. The energy barriers are caused by the interaction of the local magnetostriction with the magnetic moments of the wall. The various impediments to the movement of a domain wall tend to make its motion jerky and discontinuous. This effect is named after Barkhausen who was the first to observe it. Domain wall motion plays an important role in the magnetization process. We discuss this next.
1.11 HYSTERESIS For the analysis and design of permanent magnet applications we need the macroscopic or bulk magnetic properties of a material. These are expressed in graphical form by plotting the magnetic induction B or the magnetization M as a function of H. Either of these plots can be obtained from the other using the constitutive relation B : (H ; M). A typical B vs H plot is shown in Fig. 1.22. This plot is nonlinear and multivalued, reflecting the fact that the response of a material depends
FIGURE 1.22
Hysteresis loop.
1.11 HYSTERESIS
37
Distribution of domain orientations at various points along a hysteresis loop [4].
FIGURE 1.23
on its prior state of magnetization. The B-H plot is called a hysteresis loop. The word hysteresis is derived from the Greek word meaning ‘‘to lag.’’ The energy expended in traversing a hysteresis loop is equal to the area of the loop. Detailed discussions of this phenomenon have been given by Jiles [1] and Mayergoyz [15]. For our purposes, it suffices to review the magnetization cycle plotted in Fig. 1.22. Consider a specimen with an isotropic distribution of preferred axes orientations (magnetocrystalline anisotropy). Initially, it is unmagnetized with its domains isotropically oriented as shown at point O in Fig. 1.23. We apply an H-field, and as H increases from zero in the plus direction (weak field), the domains aligned with H grow, and the antialigned domains shrink due to domain wall motion. As H increases further (moderate field), the magnetic moments in the remaining unaligned domains rotate (or flip) into alignment along preferred axes that are in the general direction of H. For sufficiently high H, the magnetic moments along the preferred axes in domains that are not strictly aligned with H rotate away from these axes and align with H (point S) (Figs. 1.22 and 1.23). At this point the material is saturated, and has achieved its saturation induction B . The segment OS is called the initial magnetizQ ation curve. As H decreases from its saturation value, the moments along preferred axes strictly aligned with H retain their orientation. However, the moments in domains with easy axes offset from H, rotate back towards the nearest easy axis, away from H. This results in a distribution of domain alignments about the plus direction. Even when H is reduced to
38
CHAPTER 1 Materials
zero, there is still a net alignment in the plus direction, and the specimen exhibits a remanent (residual) induction B (Fig. 1.23). P Now, as H is reversed to the negative direction, the moments in domains oriented along the plus direction are first to rotate (or flip) into alignment along the negative direction. As H increases further (in the negative direction), a sufficient number of domains obtain a predominant negative orientation, thereby offsetting the magnetization of the remaining positively oriented domains so that the net induction in the specimen is zero, B : 0. The value H at which this occurs is called the coercivity and is denoted H , H : 9H . Notice that although B : 0 at A A this point, the distribution of domain orientations is not isotropic as it was in the initial demagnetized state (Fig. 1.23). As H increases beyond 9H (in the negative direction), a progressA ively larger fraction of the domains obtain a predominant negative orientation until saturation is reached (point S). From this point, H is decreased to zero and a fraction of the domains reverse their orientation, resulting in a remanent induction 9B . As H increases from zero in a P positive sense, a sufficient number of domains obtain a predominant positive orientation, thereby offsetting the magnetization of the remaining negatively oriented domains so that the net induction in the specimen is zero. This occurs when H : H . As H increases further, the A domains once again align themselves with H until saturation is achieved. Figure 1.22 is called the major hysteresis loop. There are also minor hysteresis loops as shown in Fig. 1.24. These are formed by cycling H within the range 9H H H . A A
FIGURE 1.24
Minor hysteresis loops.
1.12 SOFT MAGNETIC MATERIALS
39
This section completes our study of the principles of magnetism. In the remaining sections we discuss various magnetic materials.
1.12 SOFT MAGNETIC MATERIALS Ferromagnetic materials are classified as either soft or hard depending on their coercivity H . Soft materials are characterized by a high permeaA bility and a low coercivity (H 1000 A/m), which makes them easy to A magnetize and demagnetize. Hard materials have a relatively low permeability and a high coercivity (H 10,000 A/m), which makes A them more difficult to magnetize and demagnetize. The difference between these two materials is best illustrated by comparing their hysteresis loops. These are shown in Fig. 1.25. In this section we give a brief survey of common soft magnetic materials. More detailed treatments can be found in the texts by Jiles and Bozorth [1, 7]. Soft magnetic materials are used as flux conduits to confine and direct flux, as flux amplifiers to amplify the flux density across a region, and as magnetic shields to shield a region from an external field. The most commonly used soft magnet materials are soft iron, alloys of iron-silicon, nickel-iron, and soft ferrites. They can be found in numerous devices including transformers, relays, motors, inductors, and electromagnets. When selecting a soft material, the properties that are weighed are its permeability, saturation magnetization, resistance, and coercivity. A higher permeability and saturation magnetization are desired for flux confinement and focusing. The resistance and coercivity are important for high-frequency applications. A higher resistance
FIGURE 1.25
The B-H loops for soft and hard magnetic materials.
40
CHAPTER 1 Materials
reduces eddy currents, whereas a lower coercivity reduces hysteresis loss. Soft materials are magnetically linear over limited portions of their B-H curve where the permeability is constant (independent of H). However, they are nonlinear over other portions where is a function of H, : (H). The B-H curves for various soft materials are shown in Figs. 1.26 and 1.27.
The B-H curves for various soft magnetic materials: (a) H 400 A/m; and (b) H 400 A/m ([3] of Appendix A).
FIGURE 1.26
41
1.12 SOFT MAGNETIC MATERIALS
The B-H curves of various soft magnetic materials.
FIGURE 1.27
Various measures of permeability are used. Specifically, there is the absolute permeability :
B H
(absolute permeability),
that changes along the B-H curve. There is also the relative permeability (relative permeability), : P that is normalized to the free space value : 4 ; 10\ Wb/A · m. A plot of vs H for silicon steel is shown in Fig. 1.28. This plot shows the P nonlinear behavior of the material. There is also the differential permeability
:
dB dH
(differential permeability),
which is the slope of the tangent line to the B-H curve. The initial permeability is evaluated at the origin of the initial magnetization G curve (segment OS in Fig. 1.22), dB : G dH
. & Other measures of permeability include the initial and maximum relative permeability and , respectively. GP
P
42
CHAPTER 1 Materials
FIGURE 1.28
Relative permeability for silicon steel vs H. P
Similar definitions apply to the susceptibility of soft materials. Specifically, there are the absolute, differential and initial susceptibilities that are given by M : K H dM : K dH
(susceptibility), (differential susceptibility),
and dM : KG dH
, +&
respectively. Soft iron was the original material of choice for electromagnetic applications, and is still widely used as a core material for dc electromagnets. However, for ac applications it has been supplanted by materials with a higher resistivity because these have lower eddy current losses. Commercially available soft iron typically contains low levels of impurities such as carbon (C) (0.02%), manganese (Mn) (0.035%), sulfur (S) (0.015%), phosphor (P) (0.002%), and silicon (Si). Iron with such impurities has a coercivity on the order of 80 A/m (1 Oe), a saturation magnetization of 1.7 ; 10 A/m, and a maximum relative permeability
1.12 SOFT MAGNETIC MATERIALS
43
of 10,000. However, these properties can be improved by annealing the iron in hydrogen, which removes the impurities. Such treatment can lower the coercivity to 4 A/m (0.05 Oe), and increase the maximum relative permeability to 100,000. The properties of soft iron also degrade when it is subjected to mechanical stress, such as when it is flexed or bent. However, these properties can also be restored by annealing, which relieves the internal stress. Iron-silicon alloys typically contain 2—4% silicon. The addition of silicon has the beneficial effects of increasing the resistivity and reducing both magnetostriction and magnetic anisotropy. For example, an alloy with 3% silicon has a resistivity 14 cm, which is approximately four times the resistivity of pure iron. Consequently, these alloys are superior to soft iron for low-frequency applications such as transformer cores. Their increased resistivity reduces eddy current loss. Iron-silicon alloys have an additional advantage in that their reduced magnetostriction reduces acoustic noise (transformer hum). Furthermore, their reduced anisotropy gives rise to increased permeability in nonoriented alloys. Alloys with 1—3% Si are used in rotating machines, whereas alloys with 3—4% Si are used in transformers. Iron-silicon alloys are inferior to soft iron in two respects. Specifically, they have a reduced saturation induction B , and they are more brittle. This latter property limits the level of Q silicon to 4% for most applications. Iron-silicon alloys have coercivities in the range of 5—120 A/m, and a maximum relative permeability in the range of 4000 to 20,000, depending on the percentage content of the silicon [1]. Nickel-iron alloys are used for a variety of applications such as inductors, magnetic amplifiers, and audio frequency transformer cores. Commercially available Ni-Fe alloys contain 50—80% Ni, and are characterized by a very high permeability. These include Permalloy (78% Ni, 22% Fe) with : 8000 and : 100,000; Supermalloy (79% Ni, 16% GP
P Fe, 5% Mo) with : 100,000 and : 1,000,000; and Mumetal (77% GP
P Ni, 16% Fe, 5% Cu, 2% Cr) with : 20,000 and : 100,000 ( and GP
P GP are the initial and maximum relative permeability, respectively).
P The resistivity of Ni-Fe alloys decreases from a peak value of 72 cm for 30% Ni to 15 cm for 80% Ni [1]. Permalloy has a coercivity H : A 4.0 A/m (0.05 Oe), and a saturation magnetization M :0.86;10 A/m. Q The corresponding values for Supermalloy are H : 0.16 A/m and A M : 0.63;10 A/m. Q Finally, there are the soft ferrites. These have a cubic crystalline structure with a chemical formula of the form MO · Fe O where M is a transition metal such as iron, nickel (Ni), magnesium (Mg), manganese, or zinc (Zn). Most soft ferrites have a relatively high resistivity (up to
44
CHAPTER 1 Materials
10 m), a relatively low permeability (initial relative permeability in the range of 10 to 10,000), and a relatively low saturation magnetization (M : 4 ; 10 A/m) as compared to the soft magnetic alloys. Their low Q permeability and saturation magnetization render these materials inferior to soft magnetic alloys for dc and low-frequency applications. However, their high resistivity makes them the material of choice for high-frequency applications. Soft ferrites are sometimes categorized as either nonmicrowave ferrites (for use with audio frequencies to 500 MHz), or microwave ferrites (for use from 100 MHz to 500 GHz). Many soft ferrites have a permeability that is relatively constant up to frequencies of between 10—100 MHz. This makes them useful for broadband telecommunications applications such as specialized transformers and inductors [1, 4].
1.13 HARD MAGNETIC MATERIALS Hard magnetic materials are characterized by a low permeability and high coercivity, typically 10,000 A/m. This latter property makes them difficult to magnetize and demagnetize. Such materials are referred to as permanent magnets because once magnetized they tend to remain magnetized. Permanent magnets are used as field source components in a wide range of products including consumer electronic equipment, computers, data storage devices, electromechanical devices, telecommunications equipment, and biomedical apparatus [2, 16—18]. The properties of primary importance in the selection of a magnet are those that define the magnitude and stability of the field that it can provide. These include the coercivity H , saturation magnetization M , A Q and remanence B , as well as the behavior of the hysteresis loop in the P second quadrant (Fig. 1.29a). This portion of the hysteresis loop is called the demagnetization curve. The points (B, H) on the demagnetization curve define an energy product BH that obtains a maximum (BH) for some point in the
interval 9H H 0 as shown in Fig. 1.29b. When a magnet is used as A a field source it becomes biased at an operating point (B ,H ) on its K K demagnetization curve. The operating point depends on the circuit in which it is used. It can be determined from the load line of the circuit. This intersects the demagnetization curve at the operating point (B , H ) K K as shown in Fig. 1.30. The load line is discussed in Section 3.5. We show there that it is desirable to bias a magnet at the point of maximum energy (BH) . This
minimizes the volume of the magnet and reduces its cost. The energy
1.13 HARD MAGNETIC MATERIALS
FIGURE 1.29
45
Second quadrant B-H curve: (a) demagnetization curve; and (b)
B · H vs H.
FIGURE 1.30
Demagnetization curve and load line.
46
CHAPTER 1 Materials
FIGURE 1.31
Second quadrant constant energy contours.
product for a given material and circuit can be determined from second quadrant constant energy contours (Fig. 1.31). These are usually plotted on the demagnetization curves supplied by permanent magnet vendors. In the following sections we give a brief survey of various commercially available hard materials. These include ferrites, alnico, samariumcobalt (Sm-Co), neodymium-iron-boron (Nd-Fe-B) and bonded magnets. A more extensive treatment of hard materials is given by McCaig and Clegg, Parker, and Campbell [2, 16, 17].
1.14 FERRITES Hard ferrites are the least expensive and most commonly used permanent magnet material. The development of ferrite magnets dates back to the 1950s [19]. These magnets are made from fine particle powders with compositions of the form XO · 6(Fe O ) where X is either barium (Ba), strontium (Sr), or lead (Pb). They are fabricated using powder metallurgical methods and are commonly referred to as ceramic magnets. The fabrication cycle starts by either wet or dry mixing the correct proportions of iron oxide (Fe O ) with the carbonate of either Ba, Sr, or Pb. The mixture is then calcined at a temperature of between 1000 and 1350 °C. The calcined material is then crushed and milled into a fine powder. Isotropic magnets are formed by drying and pressing the powder into a desired shape, with subsequent sintering at a temperature in the range of 1100—1300 °C.
47
1.14 FERRITES TABLE 1.1
Properties of ceramic magnets
Ceramic $ B P
(T) (G)
H A (kAm\) (Oe) H (kAm\) (Oe) (BH)
(kJm\) (MG Oe) P Density (kg/m) Curie temp (°C)
1
5
7
8
10
0.23 2300
0.38 3800
0.34 3400
0.385 3850
0.41 4100
147 1850
191 2400
259 3250
235 2950
223 2800
259 3250
203 2550
318 4000
243 3050
231 2900
8.36 1.05 1.1 4982 450
27.0 3.40 1.1 4706 450
21.9 2.75 1.1 4706 450
27.8 3.50 1.1 4706 450
31.8 4.00 1.1 4706 450
Anisotropic ferrite magnets are fabricated using a fine grain powder with the grain size on the order of a single domain (:1 m). The powder is mixed with water to form a slurry, and the slurry is pressed and then sintered. An alignment field is applied during the compression of the powder, and the water acts as a lubricant to facilitate the alignment process. A shrinkage of 15% is common during sintering. Finished magnets are produced by grinding down the sintered material. Ferrite magnets are characterized by a high coercivity and a nearly linear second quadrant B vs H curve. This is due to the dominant magnetocrystalline anisotropy of the constituent particles. Ceramic magnets are produced with a wide range of properties. These depend on various process variables such as grain size, milling sequence, and heating schedule in both the calcination and sintering phases of production. Demagnetization curves for different grades of ceramic magnets are given in Section 1.20. Various magnetic properties are listed in Table 1.1. Iron oxides are also used for various applications. The two most commonly used oxides are ferric oxide (Fe O ), also known as magnetite or lodestone, and gamma ferric oxide (Fe O ). Ferric oxide has the following properties: H : 7.16 ; 10 A/m (90 Oe); M : 4.8 ; 10 A/m; A Q density : 5000 kg/m; and T : 585°C. Some of the corresponding A properties for Fe O are H : 21 ; 10 9 32 ; 10 A/m (270—400 Oe), A : 4700 kg/m, and T : 590 °C. A
48
CHAPTER 1 Materials
1.15 ALNICO Alnico alloys date back to the early 1930s. Their principal components are iron, cobalt, nickel, and aluminum (Al) along with lesser amounts of other metals such as copper (Cu). Alnicos are hard as well as too brittle for cold working. Production methods are limited to casting the liquid alloy, or pressing and sintering the metal powders [2]. Once formed, the alloy is subjected to controlled heat treatment so as to precipitate a dispersion of fine magnetic particles throughout the Al-Ni-Fe-Co matrix. These particles are elongated and rod shaped with a dominant shape anisotropy that is manifest in the relatively high coercivity of the finished magnet. Alnico magnets can be either isotropic or anisotropic, depending on whether or not the magnetic particles are oriented during their formation. These magnets exhibit a wide range of properties depending on the nature of the heat treatment, and whether the original alloy was prepared by casting or sintering. Demagnetization curves for various grades of Alnico are shown in Fig. 1.32a,b. The properties of Alnico magnets are listed in Tables 1.2 and 1.3.
TABLE 1.2
Properties of sintered Alnico magnets
Alnico (sintered) $ B P
(T) (G)
H A (kAm\) (Oe) H (kAm\) (Oe) (BH)
(kJm\) (MG Oe) P Density (kg/m) Curie temp. (°C)
2
5
6
8
0.71 7100
1.05 10,500
0.94 9400
0.76 7600
43.8 550
47.8 600
62.9 790
119 1500
45.3 570
49.3 620
65.2 820
134 1690
11.9 1.5 6.4 6837 810
24 3.0 4.0 7000 900
23 2.9 4.5 6892 860
36 4.5 2.1 6975 860
1.15 ALNICO
49
Demagnetization curves for Alnico magnets: (a) Alnico 1, 2, 3, and 4; (b) Alnico 6, 7, and 8 [1].
FIGURE 1.32
50 TABLE 1.3
CHAPTER 1 Materials Properties of cast Alnico magnets
Alnico (cast) $ B P
(T) (G)
H A (kAm\) (Oe) H (kAm\) (Oe) (BH)
(kJm\) (MG Oe) P Density (kg/m) Curie temp. (°C)
2
5
6
8
0.75 7500
1.24 12,400
1.5 10,500
0.82 8200
44.6 560
50.9 640
62.1 780
131.3 1650
46.1 580
54.1 680
63.7 800
148 1860
13.5 1.7 6.4 7086 810
43.8 5.5 4.3 7308 900
31.0 3.9 5.3 7418 860
42.2 5.3 2.0 7252 860
From these tables we see that Alnico magnets have impressive magnetic properties. Unfortunately, they also have poor physical properties. As already noted, they are extremely hard and brittle. Thus, production of finished magnets with tight tolerances requires tedious and costly machining.
1.16 SAMARIUM-COBALT The development of samarium-cobalt (SmCo) magnets began in the 1960s [20]. Samarium is one of the rare-earth elements that form a transition group with atomic numbers ranging from 58 (Ce) to 71 (Lu). The development of SmCo resulted from research that was directed toward the formation of alloys of the rare-earth elements with the transition series ferromagnets: iron, cobalt, and nickel. The two most common SmCo materials are SmCo and Sm Co . The first samarium-cobalt magnets were made by bonding SmCo powder in a resin. Subsequent production entailed the formation of the base alloy using either a reduction/melt or reduction/diffusion process. In the reduction/melt process Sm and Co are mixed and induction melted to form the alloy. The cast alloy is brittle, and readily ground into a fine grain powder.
1.17 NEODYMIUM-IRON-BORON
51
In the reduction/diffusion process samarium oxide (Sm O ) and cobalt powder are reacted with calcium (Ca) at approximately 1150 °C to form a compound: 10Co ; Sm O ; 3Ca ; 2SmCo ; 3CaO. The CaO is separated from the compound via a sequence of steps starting with a reaction with water, followed by a gravimetric separation of the hydroxide, followed by an acid rinse, and then drying. Once in powder form, powder metallurgical methods are used to form the desired magnet shape. Samarium-cobalt magnets are characterized by high coercivities and nearly linear second quadrant demagnetization curves. This suggests that the dominant magnetic mechanism for this material is magnetocrystalline anisotropy (Section 1.9). However, the grains in SmCo powders are typically 5—10 m in diameter, which is approximately an order of magnitude larger than single domain size. This relatively large grain size makes it energetically favorable for the formation of domain walls within each grain. Thus, each grain may comprise multiple domains. In SmCo , the domain walls move with relative ease within each grain. The high coercivity of these magnets can be explained by the limited nucleation sites for domain walls, and the pinning of domain walls at the grain boundaries that hinders the growth of domains from grain to grain. In Sm Co , the high coercivity is attributable to the pinning of domain walls. The magnetic properties of SmCo and Sm Co are given in Table 1.4. Demagnetization curves for these materials can be found in Section 1.20.
1.17 NEODYMIUM-IRON-BORON The development of neodymium-iron-boron (NdFeB) followed that of SmCo and dates back to the early 1980s [21, 22]. The initial motivation for its development was a desire for a cost effective alternative to SmCo. This was based on concern over the cost and availability of cobalt. Also, neodymium is much more abundant than samarium. The NdFeB powders are produced using various methods including a reduction/diffusion process, and a rapid quenching (melt spinning) method. The finished magnets can be either isotropic or anisotropic [17]. The reduction/diffusion process is similar to that of SmCo (Section 1.16). The finished magnets are produced from the powder via the steps of pressing and sintering, followed by heat treatment, and then grinding
52
CHAPTER 1 Materials TABLE 1.4
B P
Properties of samarium-cobalt magnets
(T) (G)
H A (kAm\) (Oe) H (kAm\) (Oe) (BH)
(kJm\) (MG Oe) P Density (kg/m) Curie temp. (°C)
SmCo5
Sm2Co17
0.83 8300
1.0 10,000
600 7500
480 6000
1440 18,100
558 7000
128 16 1.05—1.1 8200 700
192 24 1.05 8100 750
into the desired shape. The powder can be aligned with a bias field as it is pressed. The finished magnets have a high coercivity that is attributed to domain wall pinning at the grain boundaries. The rapid quenching or melt spinning method is commonly referred to as Magnequench [23, 24]. In this process the NdFeB alloy is melted and forced under argon (Ar) pressure through a small orifice onto the surface of a water-cooled revolving metal wheel. This yields a rapidly quenched thin ribbon of the alloy 35 m thick and 1—3 mm wide. This material has a very fine microcrystalline structure consisting of randomly oriented single domain NdFeB grains with a diameter on the order of 0.030 m. The ribbon is milled into thin platelets 200 m across and 35 m thick. There are three grades of Magnequench known as MQ I, MQ II, and MQ III. In the MQ I process the milled platelets are annealed and then blended with an epoxy resin. The compound is pressed into a desired shape and then oven cured. The finished magnets are isotropic due to the random orientation of the NdFeB grains within the platelets. In the MQ II process the magnets are formed by hot-pressing the platelets at 700 °C. This results in a nearly fully densified magnet that is also isotropic. In the MQ III process magnets are formed in a two-step sequence. First, the milled platelets are hot-pressed as in the MQ II process. Next,
53
1.18 BONDED MAGNETS TABLE 1.5
B P
Properties of Magnequench
(T) (G)
H A (kAm\) (kOe) H (kAm\) (kOe) (BH)
(kJm\) (MG Oe) Density (kg/m) Curie temp. (°C) Maximum operat. temp. (°C) P
MQ I
MQ II
MQ III
0.61 6100
0.80 8000
1.18 11,800
424 5.3
520 6.5
840 10.5
1200 15.0
1280 16.0
1040 13.0
64 8 6000 312 125 1.15
104 13 7500 312 150 1.15
256 32 7500 312 150 1.05
the hot-pressed material is subjected to a second hot-pressing in a die with a larger cross section. This process compresses the height by close to 50%, and is known as die-upsetting [25]. As the material thins, internal shear stresses develop at the granular level, and these tend to align the preferred magnetic axes parallel to the pressing direction. Thus, the finished magnet is anisotropic. The properties of the Magnequench materials are given in Table 1.5.
1.18 BONDED MAGNETS Permanent magnets made from fully dense materials tend to be hard and brittle. Consequently, substantial machining is required to form these materials into finished magnets with precession tolerances. A cost-effective solution to this problem is the use of bonded magnets. These magnets are fabricated by blending a magnetic powder with a binder material to form a compound, and then molding or extruding the compound into a desired shape. Typical binder materials include rubber, resins, and plastics. Bonded magnets are formed using compressionmolding, injection-molding, and extrusion. Compression-molded magnets are made using a thermosetting binder such as epoxy resin, whereas
54
CHAPTER 1 Materials
injection-molded magnets are formed using a thermoplastic binder such as nylon. Extruded magnets are produced using an elastomer such as rubber. Bonded magnets can be made rigid or flexible depending on the volume fraction of powder used. Typical volume fractions are 80—85% for compression-molding, 60—65% for injection-molding, and 55—60% for extrusion [17]. The most common bonded magnets are ferrite based. These are usually fabricated using a fine powder (1-m particle diameter) made from sintered BaFe O . The particles in this powder form as thin platelets with their planes perpendicular to the magnetic easy axis. Isotropic grades of these magnets can be produced by injection-molding the powder/binder compound. Typical energy products for these magnets are on the order of (BH) : 4 kJ/m (0.5 MG Oe).
Anisotropic magnets can be produced using an injection-molding or extrusion process with an alignment field applied during the formation of the magnet. In the injection molding process, the alignment field is applied across the mold cavity while the powder/binder compound is still fluid. The alignment field can be provided by coils embedded in the mold itself. The energy product of anisotropic injection-molded ferrite magnets is typically (BH) : 14 kJ/m (1.76 MG Oe).
A lesser degree of alignment can be achieved by applying mechanical pressure to the powder/binder compound during the formation of the magnet. The pressure tends to orient the platelets so that they face the pressing direction. Such mechanical alignment is sometimes implemented in the extrusion process. Magnetic properties of various grades of bonded ferrite magnets are listed in Table 1.6 [16]. Bonded magnets are also made using samarium-cobalt materials SmCo and Sm Co . Such magnets have reduced performance relative to their sintered counterparts, but they are far less brittle and more easily machined. SmCo powder lends itself to the injection-molding and extrusion processes where gate sizes limit the particle sizes to 5—10 m. Energy products on the order of (BH) : 60 kJ/m (7.5 MG Oe) are
achieved with these magnets. The Sm Co particles suffer a reduction in magnetic properties if their diameter is reduced much below 40 m [26]. This material is best suited for compression molding where energy products on the order of (BH) : 130 kJ/m (16 MG Oe) are achieved.
Additional properties of bonded SmCo magnets are given in Table 1.7 [2]. Bonded magnets can also be made using NdFeB powder. The Magnequench MQ I material is one example of this (Section 1.17). This section completes our survey of permanent magnet materials. In the remaining sections we discuss the magnetization and stability of magnets.
55
1.19 MAGNETIZATION
Properties of bonded ferrite magnets (isotropic (i), anisotropic (a))
TABLE 1.6
Bonded ferrite $ B P
(T) (G)
H A (kAm\) (Oe) H (kAm\) (Oe) (BH)
(kJm\) (MG Oe) P
TABLE 1.7
Flexible (i)
Flexible (a)
Rigid (i)
Rigid (a)
0.17 1700
0.25 2500
0.14 1400
0.30 3000
127 1600
175 2200
84 1050
191 2400
239 3000
239 3000
— —
223 2800
5.57 0.7 1.05
11.9 1.5 1.05
3.18 0.4 1.05
15.9 2.0 1.05
Properties of bonded samarium-cobalt magnets
Bonded SmCo $ B P
(T) (G)
H A (kAm\) (Oe) H (kAm\) (Oe) (BH)
(kJm\) (MG Oe) P
SmCo5
Sm2Co17
0.65 6500
0.86 8600
460 5750
497 6250
620 7790
800 10,050
80 10 1.1
130 16.2 1.1
1.19 MAGNETIZATION Permanent magnet materials are of little use until they are magnetized. In this section we discuss the magnetization process. It is widely held that the field H required to saturate a magnet is 3—5 times its intrinsic Q
56
CHAPTER 1 Materials Magnetizing fields of various permanent magnet materials
TABLE 1.8
Hs Magnetizing field $ Ceramic ferrite Alnico (2,5,6) Alnico (8,9) SmCo Sm Co NdFeB MQ 1—MQ II MQ III
kA/m
Oe
796 199—318 636 1193 1989
10,000 2500—4000 8000 15,000 25,000
2784 1989
35,000 25,000
coercivity H . Here, H is the field inside the material. The applied field Q H must be greater than H to compensate for the internal demagnetiz? Q ation field H (Section 1.8). Specifically, B H :H ;H , Q ? B or H :H 9H . ? Q B Recall that H : 9NM where N is the (geometry dependent) demagnetB ization factor and M is the magnetization. For a saturated magnet H : 9NM . Therefore, the (external) field required to saturate the B Q magnet is H : H ; NM . ? Q Q The values of H for various materials are listed in Table 1.8. Q Conventional materials such as Alnico and hard ferrites have lower values of H and can be magnetized with fields H 8.0 ; 10 A/m. Q Such fields can be obtained by passing direct currents through conventional solenoids (Fig. 1.33). Magnetizing fixtures are also used to magnetize conventional materials. A typical fixture consists of a coil wrapped around a soft magnetic core that has a gap in which the magnet is placed (Fig. 1.34). To polarize the magnet, a direct current is applied to the coil and this generates a field through the core. The core guides and focuses the field across the magnet to produce the desired polarization pattern. High-grade core materials are available that are capable of carrying flux densities of up to 2.0 T, which is sufficient for the
57
1.19 MAGNETIZATION
FIGURE 1.33
Magnetization of a bar magnet using a solenoid.
magnetization of low coercivity materials (i.e., B : 2 T ; H : B/ : 16 ; 10 A/m H ). Q The design of such fixtures entails the following steps: (a) estimate the demagnetization field H in the magnet assuming it is saturated; (b) B compute the applied field H : H 9 H required to saturate the magnet; ? Q B and (c) determine the current required to generate H assuming the ? magnet is an air gap. We demonstrate this procedure using the fixture shown in Fig. 1.34. This circuit (without the coil) is analyzed in Example 3.5.3. If the circuit is ideal ( -), and there is no leakage, then the demagnetization field in the magnet is essentially zero, H : 0. ThereB fore, to saturate the magnet the applied field must equal the saturation field H : H . To determine H we assume that the permeability of the ? Q ? magnet is . Then, the magnet itself represents an air gap, and the field through it is given by ni B: , l K
or
H:
FIGURE 1.34
ni , l K
Conventional magnetizing fixture.
(1.66)
58
CHAPTER 1 Materials
FIGURE 1.35
Components of an impulse magnetizing system.
where n is the number of turns, i is the current, and l is the length of K the magnet (height of the gap) (Example 3.5.1). Thus, to saturate the magnet we choose ni : l H so that H : H . K Q Q While conventional magnetization methods are appropriate for low energy magnets, they are impractical for rare-earth materials. These require magnetizing fields on the order of H : 2 ; 10 A/m (25,000 Oe), Q and the current required to generate such fields cannot be sustained without overheating and damaging the conductors. The nominal limit of current density for continual operation is generally taken to be 1.5 to 2 A/mm Fortunately, the magnetizing field needs to be applied only momentarily. Therefore, rare-earth magnets can be magnetized using impulse magnetizers that provide a high current pulse of short duration. Such magnetizers are relatively inexpensive and can operate from standard power lines. A typical impulse magnetizing system consists of a charging circuit, a capacitor bank for energy storage, a thyristor or ignitron tube for switching the current, a step-down transformer, and a magnetizing fixture (Fig. 1.35). An equivalent electrical circuit for a basic magnetizing system (without the transformer) is shown in Fig. 1.36. The charging circuit consists of a voltage source V and a resistor R . The ignitron tube BA is represented by a toggle switch, and the capacitor bank is represented by capacitance C. The magnetizing fixture is represented as an RL load.
FIGURE 1.36
Equivalent circuit of an impulse magnetizing system.
59
1.19 MAGNETIZATION
The magnetization process entails two distinct steps. In the first step, the capacitor bank is slowly charged to a predefined voltage. This corresponds to the switch being set to position (1) in Fig. 1.36. The switch is held in this position until the capacitor bank C is charged to the voltage V . The duration of the charging phase is determined by the resistance R . In the second step, the ignitron tube is activated, which in effect toggles the switch to position (2). The current from the capacitor is discharged into the fixture in the form of a single high-current pulse. This produces a peak field of short duration across the magnet that is sufficient to magnetize it. Impulse magnetizing fixtures typically consist of a coil wound on a nonmagnetic support structure that houses the magnet. There is usually little benefit in using a soft magnetic core because it would saturate at a level below the required magnetizing field. Moreover, such cores are conductive and can therefore generate eddy currents that oppose the magnetizing field. Without such a core, a fixture is magnetically linear and the magnetizing field that it generates scales linearly with the applied current. It is instructive to study the electrical response of an impulse magnetizing system. Consider the circuit of Fig. 1.36. After the capacitor is charged to the voltage V , the switch is moved to position (2), which sends current to the fixture. The current is governed by the equation di R di i ; ; : 0, dt L dt LC
(1.67)
with initial conditions i(0) : 0 and di/dt : V /L. The solution to Eq. R (1.67) is given by V i(t) : 9 [e\RO 9 e\RO] 2L
(1.68)
1 : , R/2L ;
(1.69)
1 : , R/2L 9
(1.70)
where
and : ((R/2L) 9 1/LC. Notice that . Initially, the current surges through the fixture and rises with a time constant . After reaching a peak value, the current decays with a time constant . The system is critically damped when : 0, and overdamped when 0.
60
CHAPTER 1 Materials
Thus, the system is overdamped when R2
L C
and the current through the fixture will be unidirectional without oscillations. This is desired for magnetization. A fixture needs to be designed with appropriate values of R and L so that sufficient energy with be transferred from the capacitor bank to the magnetizing field. To saturate a magnet, the fixture must provide a field H throughout the magnet. Let E denote the magnetostatic Q
energy required to saturate the magnet. We estimate E using Eq.
(3.66), E
: H ; (volume of the magnet), ?
where H 3H , and is the permeability of the unmagnetized material ? ( for rare-earth materials). The energy stored in an impulse mag netizer is E : 1/2CV when its capacitor bank is fully charged. The energy dissipated through the resistor R is E : 1/2R i dt, where i is given by Eq. (1.68). Thus V , C, R, and L are chosen so that E E
; E . Moreover, the fixture needs to be designed to render the desired magnetization pattern throughout the magnet. The design of an impulse magnetizing fixture is straightforward [27]. First, a coil structure is chosen that will render the desired magnetization pattern within the magnet. Next, an initial wire gauge is selected for the coil. Third, an analysis is performed to determine the resistance R and inductance L of the coil as well as the average B-field that it generates across the magnet per ampere of applied current. This latter value is referred to as the tesla per amp ratio, or TPA. This ratio is used in a subsequent analysis to determine the level of current required to saturate the magnet, B : H . Q The coil resistance R is easy to determine given the length and gauge of the wire. However, the inductance and TPA calculations are more complicated. The topology of the coil often precludes a simple analytical evaluation of these values. However, they can be determined using numerical methods such as finite-element analysis (FEA) (Section 3.8). The FEA for fixture design is essentially a steady-state magnetostatic analysis. Briefly, the finite-element method amounts to dividing the fixture geometry and surrounding region into small elements (collectively known as a mesh) and assigning to each element a polynomial that will approximate the behavior of the magnetic potential in that element.
61
1.19 MAGNETIZATION
The coefficients of these polynomials are arranged in a matrix, and their values are obtained via the solution of a matrix equation that contains all the information of the problem, including the location of each element relative to its neighbors, the underlying field equations, and the boundary conditions. This is discussed in detail in Section 3.8. Once R, L, and TPA are known, a circuit analysis is performed to determine the current through the fixture and the field that develops across the magnet. A thermal analysis is also performed to determine the temperature rise in the fixture. The temperature T is predicted using the formula T(t) :
1 R J dt, c
where , c , and are the density, constant-volume specific heat, and conductivity of the conductor, and J is the current density through it. For copper at room temperature, : 8950 kg/m, c : 383 J/(kg °C), and : 5.8 ; 10 mho/m. The calculations are performed for a series of wire gauges and an optimum wire gauge is selected based on achieving the appropriate field strength, the sharpest pole transitions, and a tolerable fixture temperature (220 °C). We demonstrate the design procedure via a simple example. Consider the bipolar disk magnet shown in Fig. 1.37. Such magnets are typically used for clamping or holding applications. Assume that the magnet is made from NdFeB MQ I material that requires a magnetizing field H : 2785 k Am\. This corresponds to a saturation flux density Q B : 3.5 T (Table 1.8). The design starts with a wire configuration design such as the one shown in Fig. 1.38. The wires are arranged so as to induce the desired bipolar magnetization across the magnet. They are connected at their terminals to form a continuous circuit. Several magnets are to be magnetized at once and therefore the length of the fixture is much greater than its width. We perform a two-dimensional (2D) FEA of the cross-sectional geometry of the wires. This cross section is shown
FIGURE 1.37
Bipolar disk magnet.
62
CHAPTER 1 Materials
FIGURE 1.38
Wire configuration for magnetizing a bipolar disk magnet.
in Fig. 1.39a. The first step is the generation of the mesh (Fig. 1.39b). Once the geometry is meshed, a nominal value of current is defined for the wire regions and the appropriate boundary conditions are imposed. In this case the field is set to zero far from the fixture (Dirichlet condition). The field problem is then solved, and plots of the flux lines and the B-field are obtained (Fig. 1.40a,b). A flux line is a line that indicates the direction of the B-field at any point along its length (Section 3.2.1). The inductance L and TPA are obtained from this analysis. The TPA is computed by dividing the field generated at the center of a pole by the number of amps of applied current. At this point V , R , R, L, and TPA are known. A circuit analysis can now be performed to determine the optimum wire gauge. A sample output of an analysis is shown in Fig. 1.41. The vertical axis has been divided into 10 parts, and the top of the scale corresponds to the values B : 5.0 T, I : 50,000 A, and TEMP : 200 °C. The negative swing
in current is prevented in practice and therefore only the positive half of the cycle is shown. The analysis shows that a peak field of 4.5 T is achieved with a maximum current of 25,000 A. The operating temperature of the coil reaches 140 °C, which is acceptable since the maximum working temperature for common grades of insulation is typically 200 °C.
1.19 MAGNETIZATION
FIGURE 1.39
FEA mesh.
63
Fixture analysis: (a) cross section of the wire configuration; and (b)
64
CHAPTER 1 Materials
FIGURE 1.40
Fixture analysis: (a) flux lines; and (b) B-field.
65
1.19 MAGNETIZATION
FIGURE 1.41
Analysis of the magnetization process.
As a further example, consider the multipole disk magnet shown in Fig. 1.42. Eight north-south alternating magnetic poles are arranged in pie-shaped zones with vertical magnetization over each zone. This type of magnet is used in axial-field permanent magnet motors (Section 5.13). A fixture for this magnet is shown in Fig. 1.43. The coil is supported in a nonmagnetic structure and is wound to provide the desired magnetization. The resistance of the coil is easily calculated once an initial wire gauge is specified. The inductance and TPA are calculated using FEA. Once these have been determined, parametric circuit calculations are performed to determine the optimum wire gauge.
FIGURE 1.42
Multipole disk magnet.
66
FIGURE 1.43
CHAPTER 1 Materials
Exploded view of a magnetizing fixture for an eight-pole disk
magnet.
1.20 STABILITY The operating point of a magnet depends on its load line as well as its demagnetization curve. These, in turn, depend on the operating environment. A good design must account for variations in the environment so that the performance of the magnet remains within an acceptable range throughout its working life. This is the issue of stability. A magnet’s load line is a function of the magnetic circuit used. Thus, load line stability needs to be considered on a case-by-case basis. However, the stability of the demagnetization curve depends on factors that are common to all applications. In this section we briefly review and summarize these factors. A thorough discussion of this subject is given by McCaig and Clegg [2].
67
1.20 STABILITY
The magnetic properties and hence demagnetization curve of a magnet depend on many variables including temperature, pressure, and the applied field. The magnetic properties change due to variations in these variables, and such changes can be classified as reversible, irreversible, or structural. Reversible changes occur when the magnetic properties change in response to a variation in a variable, but return to their original state when the variable returns to its initial value. For example, a magnet’s hysteresis loop changes with temperature, and for a limited range of temperatures these changes are reversible and essentially linear. Such changes can be described by reversible temperature coefficients for the remanence B , coercivity H , and intrinsic coercivity H . The coefficients P A are expressed as a percentage change per °C. All magnets suffer a reduction in B as temperature increases. However, the change in H can P A be positive or negative depending on the dominant anisotropic mechanism of the material. Most magnets exhibit a decrease in H with A temperature (a negative coefficient for H ). However, ceramic magnets A have a positive coefficient of 0.02%/°C. The designer must account for these changes to ensure that a magnet’s operating point remains within an acceptable tolerance over the range of operating temperatures. The demagnetization curves for ceramic and samarium cobalt materials at various temperatures are shown in Figs. 1.44 and 1.45, respectively. The temperature coefficients for various materials are listed in Table 1.9 [16]. In an irreversible change in magnetization, the magnetic properties change in response to a variation in a variable, and remain changed even when the variable returns to its initial value. To restore the original properties the magnet must be remagnetized. For example, a ferromagnetic specimen undergoes an irreversible change when its temperature is elevated above the Curie temperature (Section 1.6). Once above the
TABLE 1.9
Temperature coefficients of various permanent magnet materials
Maximum service temp. (°C) Curie temp. (°C) B P Reversible coefficient (%/°C) H Reversible coefficient (%/°C)
Alnico 5
Ferrite
SmCo5
NdFeB
520
400
250
150
720
450
725
310
90.02
90.20
90.04
90.12
90.03
;0.40
90.30
90.60
68
CHAPTER 1 Materials
Demagnetization curves of ceramic magnets (dashed lines : M): (a) ceramic 5; (b) ceramic 7; and (c) ceramic 8 [16].
FIGURE 1.44
1.20 STABILITY
69
Demagnetization curves of sintered SmCo materials (dashed lines : M): (a) SmCo ; and (b) Sm Co [17].
FIGURE 1.45
70
CHAPTER 1 Materials
Curie temperature, the specimen becomes paramagnetic. Its domains retain a random orientation as it cools, which renders it unmagnetized. The specimen can be remagnetized to its original state provided that its metallurgical structure has not been altered. Elevated temperatures can also cause changes in domain structure. In fact, immediately after magnetization a specimen undergoes a relatively slow process in which its unstable domains relax to a lower energy state via a variety of mechanisms including domain boundary movement. Such irreversible effects are collectively known as magnetic viscosity. This accelerates with temperature. Another example of an irreversible change is when an external demagnetizing field H is applied that exceeds the intrinsic coercivity H of the specimen. The magnetization irreversibly flips to the opposite direction. Again, the specimen can be remagnetized to its original state. Structural changes entail a permanent change in the metallurgical state of a magnet. Such changes cannot be undone by remagnetization. Corrosion and oxidation are examples of structural changes. Samarium cobalt, ceramic and Alnico magnets are resistant to oxidation, but rare-earth magnets are not. Oxidation is especially severe in sintered Nd Fe B. Oxygen diffuses into this material at elevated temperatures, causing an oxidized layer to form. The layer thickness d grows according to the relation d : f(T)(t, where f(T) represents a nonlinear temperature dependence. The oxidized layer has a lower intrinsic coercivity than the body of the magnet, and is more easily demagnetized by the magnet’s internal field. This layer degrades performance because: (a) it reduces the working volume of the magnet; and (b) it shunts the field from the magnet’s interior. Oxidation can be controlled by varying the composition of the material, and/or coating the finished magnet. Common coating materials include nickel, zinc (Zn), and aluminum (Al) with thicknesses of 10—20 m, or epoxy with a thickness of 20—30 m. References 1. Jiles, D. (1991). Introduction to Magnetism and Magnetic Materials, London: Chapman and Hall. 2. McCaig, M. and Clegg, A. G. (1987). Permanent Magnets in Theory and Practice, 2nd ed., New York: John Wiley and Sons. 3. Craik, D. (1995). Magnetism: Principles and Applications, New York: John Wiley and Sons.
1.20 STABILITY
71
4. Chikazumi, S. and Charap, S. (1964). Physics of Magnetism, New York: John Wiley and Sons. 5. Cullity, B. D. (1972). Introduction to Magnetic Materials, Reading, MA: Addison-Wesley. 6. Morrish, A. H. (1983). The Physical Principles of Magnetism, Malabar, FL: R. Krieger Publishing Co. 7. Bozorth, R. M. (1993). Ferromagnetism, New York: IEEE Press. 8. Messiah, A. (1961). Quantum Mechanics, Amsterdam: North Holland. 9. Alonzo, M. and Finn, E. J. (1968). Fundamentals of University Physics, vol. III: Quantum and Statistical Physics, Reading, MA: Addison-Wesley. 10. Langevin, P. (1905). Annales de Chem. et Phys. 5: 70. 11. Weiss, P. (1907). L’hypothèse du champ moléculaire et la propriété ferromagnétique, J. de Phys. 6: 661. 12. Heisenberg, W. (1928). Zur theorie des ferromagnetismus, Z. Phys. 49: 619. 13. Stoner, E. C. and Wohlfarth, E. P. (1948). A mechanism of magnetic hysteresis in heterogeneous alloys, Phil. Trans. Roy. Soc. A240: 599. 14. Bloch, F. (1932). Zur theorie des austauschproblems und der remanenzerscheinung der ferromagnetika, Z. Phys. 74: 295. 15. Mayergoyz, I. D. (1991). Mathematical Models of Hysteresis, New York: Springer-Verlag. 16. Parker, R. (1990). Advances in Permanent Magnetism, 2nd ed., New York: John Wiley and Sons. 17. Campbell, P. (1994). Permanent Magnet Materials and their Application, Cambridge: Cambridge Univ. Press. 18. Watson, J. K. (1980). Applications of Magnetism, New York: John Wiley and Sons. 19. Went, J. J., Rathenau, G. W., Gorter, E. W., and Van Oosterhout, G. W. (1952). Ferroxdure, a class of new permanent magnet materials, Philips Tech. Rev. 13: 194. 20. Strnat, K., Hoffer, G., Olson, J., Ostertag, W., and Becker, J. J. (1967). A family of new cobalt based permanent magnetic materials, J. Appl. Phys. 38: 1001. 21. Sagawa, M., Fujimura, S., Togawa, N., Yamamoto, H., and Matsuura, Y. (1984). New material for permanent magnets on a base of Nd and Fe, J. Appl. Phys. 55: 2083. 22. Sagawa, M., Horosawa, S., Yamamoto, H., Matsuura, Y., Fujimura, S., Tokuhara, H., and Hiraga, K. (1986). Magnetic properties of the bcc phase at grain boundaries in the Nd-Fe-B permanent magnet, IEEE Trans. Magn. 22: 910. 23. Croat, J., Herbst, J. F., Lee, R. W., and Pinkerton, F. E. (1984). Pr-Fe and Nd-Fe-based materials: a new class of high-performance permanent magnets, J. Appl. Phys. 55: 2078.
72
CHAPTER 1 Materials
24. Lee, R. W. (1985). Hot-pressed neodymium-iron-boron magnets, Appl. Phys. Lett. 46: 790. 25. Lee, R. W., Brewer, E. G., and Schaffel, N. A. (1985). Processing of neodymium-iron-boron melt spun ribbons to fully dense magnets, IEEE Trans. Magn. 21: 1958. 26. Satoh, K., Oka, K., Ishii, J., and Satoh, T. (1985). Thermoplastic resin-bonded Sm-Co magnet, IEEE Trans. Magn. 21: 1979. 27. Lee, J. K. and Furlani, E. P. (1990). The optimization of multipole magnetizing fixtures for high-energy magnets, J. Appl. Phys. 67(3).
CHAPTER
2
Review of Maxwell’s Equations
2.1 INTRODUCTION Maxwell’s equations govern the interaction of charged matter, and the behavior of electromagnetic fields. They provide a fundamental understanding of a wide range of phenomena including the magnetic interactions that are of interest to us. In this chapter, we review electromagnetic field theory and Maxwell’s equations. This review is not intended to provide a working knowledge of the theory because it is assumed that the reader already possesses such knowledge. Instead, it is written to provide an overview of the theory, with an emphasis on the reduction and simplification of the field equations for the applications that we intend to study. We start with Maxwell’s equations in their most general form. These are presented in both differential and integral form along with constitutive relations and boundary conditions. We then introduce the scalar and vector potentials, and show that they provide an alternate, and often more tractable formulation of electromagnetic field theory. Next, we specialize to the case of quasi-static field theory. Here, there is a partial uncoupling of the field equations that simplifies their solution. Finally, we study static field theory in which all time dependence is negligible. We find that the magnetic and electric fields uncouple into separate magnetostatic and electrostatic field equations. Permanent magnet and electromechanical devices are analyzed using magnetostatic and quasistatic field theory, respectively. We study the methods and techniques for analyzing such devices in Chapters 3 and 5, respectively. Throughout this text we use SI units in the Sommerfeld convention. These and other units are discussed in Appendix D.
73
74
CHAPTER 2 Review of Maxwell’s Equations
2.2 MAXWELL’S EQUATIONS Electromagnetic fields arise from sources of charge and current, and are governed by Maxwell’s equations. We begin with the field equations in differential form, Maxwell’s Equations ;H:J;
D t
·B:0 ;E:9
(2.1) (2.2)
B t
· D : .
(2.3) (2.4)
In these equations J (A/m) and (C/m) are source terms, and H, B, E, and D are the fields. These are defined as follows: Field E D H B
Description Electric field intensity (V/m) Electric flux density (C/m) Magnetic field intensity (A/m) Magnetic flux density (T)
(2.5)
In definition (2.5), V/m stands for volt/meter, C/m stands for Coulomb/meter, A/m stands for ampere/meter, and T stands for tesla. The fields (2.5) are vector-valued functions of space and time, for example, B : B (x, y, z, t)x ; B (x, y, z, t)y ; B (x, y, z, t)z . V W X As each of the four fields has three components, the field equations represent a system of equations for 12 unknown field components. However, while these equations are consistent, they are not all independent. Therefore, they do not provide the 12 independent scalar equations required to determine the 12 unknowns [1, 2]. Specifically, the two divergence equations can be obtained from the two curl equations if one imposes the continuity of charge equation (2.6). Thus, for a complete
75
2.2 MAXWELL’S EQUATIONS
theory the field equations must be augmented by additional independent equations that take the form of constitutive relations. These are discussed in more detail in Section 2.2.1. The source terms in the field equations are the free current density J (A/m), and the free charge density (C/m) [2]. Free charge is charge that has been separated from its atomic or molecular source and is capable of moving through a material. The free current density J is due to the motion of free charge. It accounts for such phenomena as the motion of conduction electrons in conductors (conduction current), the motion of conduction electrons and/or holes in semiconductors (conduction current), the migration of positive or negative ions through liquids (electrolytic current), and the motion of electrons or ions in a vacuum (convection current) [3]. The field equations embody the conservation of free charge as expressed by the continuity equation, ·J;
: 0. t
(2.6)
This is obtained by combining the divergence of Eq. (2.1) with the time derivative of Eq. (2.4). On the other hand, if Eq. (2.6) is viewed as an independent equation, we find that the two divergence equations (2.2) and (2.4) can be obtained from the two curl equations (2.1) and (2.3) as noted in the preceeding text. In addition to free charge there is also bound charge. Bound charge is charge that is displaced from its atomic or molecular source, but does not separate from it. Bound charges arise during the polarization of dielectric materials. The term D/t in Eq. (2.1) is called the displacement current density and accounts for the motion of bound charge.
2.2.1 Constitutive relations As already noted, Maxwell’s equations by themselves do not provide a complete set of equations for the fields. To complete the theory, the field equations must be augmented by additional independent equations that are specified by the following constitutive relations: Constitutive Relations B : (H ; M) D:+ E;P
(2.7) (2.8)
76
CHAPTER 2 Review of Maxwell’s Equations
where M (A/m) is the magnetization, P (C/m) is the polarization, and : 4 ; 10\ T m/A and + : 8.854 ; 10\ F/m are the permeability and permittivity of free space, respectively. The magnetization M was discussed in detail in Chapter 1. It represents the net magnetic dipole moment per unit volume of a material. Specifically, M is given by m M : lim G G , V 4 where m is a vector sum of the magnetic dipole moments contained G G in the elemental volume V. Similarly, polarization P represents the net electric dipole moment per unit volume. It is given by p P : lim G G , V 4 where p is a vector sum of the electric dipole moments contained in G G the elemental volume V. From Eq. (2.8) we find that the displacement current density consists of two components D E P :+ ; . t t t The addition of D/t to the field equations is due to James Clerk Maxwell and is perhaps his most notable scientific contribution. This term, by virtue of its first component + (E/t), provides for the propa gation of electromagnetic fields in free space. As already noted here, its second component P/t is due to the motion of bound charge in polarized materials. In stationary, linear, homogeneous and isotropic media the constitutive relations reduce to B : H
(2.9)
D : +E,
(2.10)
and where and + are the permeability and permittivity of the media, respectively. There is an additional constitutive relation that relates J to E, J : E, where is the conductivity with units A/V · m.
(2.11)
77
2.2 MAXWELL’S EQUATIONS
The constitutive relations (2.9)—(2.11) need to be modified for nonlinear, inhomogeneous or anisotropic materials. A material is nonlinear if , +, or depend on the fields, otherwise it is linear. For nonlinear materials Eqs. (2.9)—(2.11) are written as B : (H)H, D : +(E)E, or J : (E)E, depending on the nonlinearity. A material is inhomogeneous if , +, or depend on position, otherwise it is homogeneous. In inhomogeneous materials, , +, or are functions of the coordinate variables, for example, : (x, y, z), + : +(x, y, z), or : (x, y, z). Finally, a material is said to be anisotropic if , +, or depend on direction, otherwise it is isotropic. In anisotropic materials Eq. (2.9) generalizes to three equations: B : H ; H ; H , V V W X B : H ; H ; H , W V W X B : H ; H ; H . X V W X Similar relations hold for Eqs. (2.10) and (2.11).
2.2.2 Integral equations So far we have studied the differential form of Maxwell’s equations. However, the field equations can also be written in integral form. The integral equations are usually more useful for applications that manifest a high degree of geometric symmetry. The integral form of field equations is obtained as follows: Take the surface integral of the curl equations (2.1) and (2.3) over an open surface S bounded by a contour C, and then apply Stokes theorem (A.42) to the left-hand side and obtain
H · dl :
!
J;
1
D · ds t
(2.12)
and
!
respectively.
E · dl : 9
1
B · ds, t
(2.13)
78
CHAPTER 2 Review of Maxwell’s Equations
Similarly, integrate the divergence equations (2.2) and (2.4) over a volume V with a closed surface S, and then apply the Divergence theorem (A.45) to the left-hand side. This gives
B · ds : 0
(2.14)
1
and
D · ds : dv, (2.15) 1 4 respectively. Equations (2.12)—(2.15) constitute the integral form of the field equations. The first equation (2.12) is a generalization of Ampere’s circuital law, which states that the circulation of the magnetic field intensity around any closed path is equal to the free current flowing through the surface bounded by the path. The second equation (2.13) embodies Faraday’s law of electromagnetic induction, which states that the electromotive force induced in a stationary closed circuit equals the negative rate of increase of the magnetic flux linking the circuit. In simpler terms, this law states that an E-field is generated by a time-varying magnetic flux. The negative sign in Eq. (2.13) implies that the induced E-field will act to oppose the change in magnetic flux. This is known as Lenz’s law. Faraday’s law provides the basis for the behavior of many important devices including transformers, generators and electromechanical devices. The third equation (2.14) states that the total outward flux of the B-field over a closed surface equals zero. Recall that the flux of a vector field B through a surface S is defined by, :
B · ds.
1
Flux of B through S
Equation (2.14) implies that there are no isolated sources or sinks of magnetic field, that is, no magnetic monopoles. The fourth equation (2.15) is Gauss’ law. It states that the total outward flux of the D-field over a closed surface equals the total free charge enclosed by the surface. The differential and integral forms of Maxwell’s equations are summarized in Table 2.1. In the next section, we use the integral form of the field equations to determine the boundary conditions for the fields at material interfaces.
79
2.2 MAXWELL’S EQUATIONS TABLE 2.1
Differential and integral form of Maxwell’s equations Maxwell’s Equations (with Eqs. 2.16)
Differential Form ;H:J;
D t
Integral Form
H · dl : !
;E:9
B t
·D:
1
·B:0
J;
B · ds : 0 1
E · dl : 9
!
1
D · ds : 1
D · ds t
(2.16)
B · ds t dv
4
2.2.3 Boundary conditions We often need to solve Maxwell’s equations in a region consisting of various different materials. However, to obtain a field solution we need to know the behavior of the fields at the material interfaces. The boundary conditions are determined by evaluating the integral form of the field equations over control volumes at the material interfaces. Specifically, Eqs. (2.12) and (2.13) are evaluated around an infinitesimal Stokesian path, and Eqs. (2.14) and (2.15) are evaluated over the surface of an infinitesimal Gaussian pillbox (Fig. 2.1). We determine the boundary conditions for Eqs. (2.12) and (2.13) first. Apply Eq. (2.12) around the Stokesian path of Fig. 2.1b and ignore the contributions from the infinitesimal end portions. This yields D H l 9 H l : J lh ; L lh, R R L t where H (i : 1, 2) are the components of H tangential to the interface. GR Take the limit as h ; 0 and obtain H 9H :J , (2.17) R R where J : lim J h is the free surface current density at the inter F L face (into the page as shown in Fig. 2.1b). If both regions have finite conductivity (which is true for all applications we study), then J : 0 and the tangential component of H is continuous H :H . R R
(2.18)
80
CHAPTER 2 Review of Maxwell’s Equations
Boundary analysis: (a) Gaussian pillbox and Stokes path at the material interface; and (b) field components.
FIGURE 2.1
Exceptions to this occur at an interface of a perfect conductor or superconductor. We apply a similar analysis to Eq. (2.13) and obtain E :E , (2.19) R R which shows that the tangential component of E is continuous at an interface. Next, we determine the boundary conditions for Eqs. (2.14) and (2.15). Apply Eq. (2.14) over the surface of the Gaussian pillbox of Fig. 2.1b and obtain This implies that
B A 9 B A : 0. L L B :B . L L
(2.20)
81
2.2 MAXWELL’S EQUATIONS
Thus, the normal component of B is continuous at an interface. A similar analysis applies to Eq. (2.15) and yields D 9D : , (2.21) L L Q where : lim h is the free surface charge density at the interface. F Q This shows that the normal component of D is discontinuous when a free surface charge exists at the interface. If both regions are perfect dielectrics (zero conductivity), then : 0 and the normal component of Q D is continuous, D : D . If, on the other hand, region 1 is a perfect L L conductor (infinite conductivity), and region 2 is a perfect dielectric, then D : 0 (E : 0 in a perfect conductor), and D : , where resides on L L Q Q the surface of the perfect conductor. The boundary conditions are summarized as follows: 1. The tangential component of H is discontinuous at an interface by an amount equal to the free surface current density at the interface (H 9 H : J ). R R 2. The tangential component of E is continuous across an interface (E : E ). R R 3. The normal component of B is continuous across an interface (B : B ). L L 4. The normal component of D is discontinuous at an interface by an amount equal to the free surface charge density at the interface (D 9 D : ). L L Q The derivations of the boundary conditions are summarized in the following diagrams:
H · dl :
!
J;
1
1
1
t
B
E · dl : 9 !
D
t
· ds
· ds
B · ds : 0 1
D · ds :
4
$ dv
$
Integrate around Stokesian path across interface
Integrate over surface of Gaussian pillbox at interface
$
$
H 9H : J R R E :E R R
B :B L L D 9D : L L Q
82
CHAPTER 2 Review of Maxwell’s Equations
2.2.4 Force and torque While the fields themselves are of interest, their interaction with charged matter is equally important. In particular, for electromechanical analysis we need to know the electromagnetic force and torque imparted to steady currents. To determine this, we study the behavior of a charge q moving with velocity u through a region permeated by E- and B-fields. It is well known that the moving charge experiences a Lorentz force given by F : q(E ; u ; B)
(Lorentz force).
(2.22)
In the continuum limit, Eq. (2.22) generalizes to f : E ; J ; B
(N/m),
(2.23)
where is the free charge density and J is the free current density. Notice that f is a force density. We are especially interested in the force and torque imparted to steady currents by an external B-field. From Eq. (2.23) we find that the Lorentz force on a current distribution J is F:
J ; B dv,
(2.24)
4 where V is the volume occupied by J. It follows that the torque is given by T:
r ; (J ; B) dv, (2.25) 4 where r is the vector to the point about which the torque is computed.
2.3 POTENTIALS Maxwell’s equations can be solved directly for the fields. However, it is often more convenient to obtain the fields using potential functions. Specifically, the four coupled first-order field equations can be rewritten in terms of a pair of uncoupled second-order equations by making a change of variables to the vector and scalar potentials A and [1]. In this section we discuss the potentials, their governing equations, and their relation to the fields. For our purposes, it suffices to consider Maxwell’s equations in an unbounded, stationary, homogeneous, isotropic and linear medium with
83
2.3 POTENTIALS
constitutive relations B : H and D : + E. The derivation of the equations for the potentials is as follows: Apply Proposition A.5.1 to Eq. (2.2) and introduce the vector potential A, · B : 0 $ B : ; A.
(2.26)
Next, substitute Eq. (2.26) into Eq. (2.3) and obtain
; E;
A A :0$E; : 9 t t
or E:9
A 9 . t
(2.27)
Notice that we have used Proposition A.5.2 to introduce the scalar potential . Substitute Eqs. (2.26) and (2.27) into the remaining Eqs. (2.1) and (2.4) and obtain
(2.28)
A 9 · A ; + : 9J, t t
(2.29)
A 9 ( · A) : 9J ; +
A ; , t t D term t
or A 9 +
and ;
( · A) :9 . t +
(2.30)
Notice that we have applied the identity ;;A : ( · A)9A to obtain Eq. (2.28). Now, recall from Theorems A.5.1 and A.5.2 that A is uniquely specified if both its curl and divergence are specified (Appendix A). The former is specified by Eq. (2.26). We specify the latter by imposing the Lorentz gauge condition · A : 9+
t
(Lorentz gauge).
(2.31)
84
CHAPTER 2 Review of Maxwell’s Equations
Finally, substitute Eq. (2.31) into Eqs. (2.29) and (2.30) and obtain two second-order equations for the potentials A 9 +
A : 9J, t
(2.32)
9 +
:9 . + t
(2.33)
It is well known that Eqs. (2.32) and (2.33) have solutions 4
1 4+
A(x, t) :
J(x, t 9 x 9 x /u) dv
x 9 x
(2.34)
(x, t 9 x 9 x /u) dv,
x 9 x
(2.35)
4
and (x, t) :
4
where A(x, t) and (x, t) are time-retarded potentials, and u : 1/(+ is the speed of light in the medium. In free space, + : + , : , and u : c : 1/( + : 3 ; 10 m/s. It can be shown that A(x, t) and (x, t) satisfy the Lorentz condition (2.31) as long as J and satisfy the continuity condition (2.6). Thus, given J and we can solve Maxwell’s equations by first solving Eqs. (2.34) and (2.35), and then using Eqs. (2.26) and (2.27) to obtain B and E. The fields D and H are obtained from the constitutive relations D : + E and H : B/. For certain time-harmonic problems it is possible to reconstruct the fields from A(x, t) alone. The boundary conditions for and A must be compatible with the boundary conditions for the fields (2.17)—(2.21). Specifically, and A must be continuous at an interface. If there are discontinuities in or + at an interface, then the normal derivatives of and A must be discontinuous in order to ensure that the normal component of D : 9+ ( ; A/t), and the tangential component of H : 1/( ; A) are continuous across the interface [5]. Maxwell’s equations govern a broad range of phenomena that span the entire electromagnetic spectrum. It is useful to categorize such applications as high-frequency, quasi-static, or static. For high-frequency applications, the field equations (2.1)—(2.4) need to be solved in their entirety. However, for quasi-static and static applications, certain terms are negligible and as we shall see the analysis simplifies.
85
2.4 QUASI-STATIC THEORY
2.4 QUASI-STATIC THEORY Quasi-static field theory applies at low frequencies (long wavelengths) when the dimensions of the region of interest are small relative to the wavelength of the electromagnetic field that permeates it. When this is the case, we ignore the field’s finite speed of propagation and assume that any change in the field is felt instantaneously across the region. For example, consider a system that resides in a cubic region of free space that measures (3 m on each side. Assume that the system generates a time-harmonic electromagnetic field with period and wavelength . The field propagates across this region with a velocity c : 3 ; 10 m/s. Therefore, any change in the field is felt throughout the region within 10 ns. The system is said to be quasi-static as long as the field changes much more slowly than the time required for it to propagate across the region, that is, as long as 10 ns, or f : 1/ 100 MHz. Alternatively, because f : c the system is quasi-static as long as 3 m, which relates back to our original definition. Quasi-static theory governs an important range of applications including electrical circuit analysis, electromechanical devices, and eddy current phenomena [6]. We study these in some detail in Chapter 5. Here, we briefly review the quasi-static field equations. From a mathematical perspective, the quasi-static approximation amounts to ignoring the displacement current term D/t in the field equations. When this is done, the field equations simplify as follows:
;H:J
·B:0 D :0$ B t ;E:9 t · D : .
(Quasi-static)
(2.36)
Notice that in the first and second equations in (2.36) there are no time derivatives. Thus, even when the current source J is time dependent, the fields H and B are determined as if the system were static. Also notice that the first equation implies that · J : 0. The third equation in (2.36) is known as Faraday’s law. It plays an important role in the analysis of electromechanical devices (Chapter 5). The fields can be obtained by direct solution of (2.36) as long as the constitutive relations (2.7) and (2.8) are taken into account. However, the fields can also be obtained using the potentials introduced in Section 2.3.
86
CHAPTER 2 Review of Maxwell’s Equations
To see this, follow the same analysis presented there, except that D/t : 0 in Eq. (2.28). Then, Eqs. (2.32) and (2.33) reduce to A 9 ( · A) : 9J(t),
(2.37)
( · A) (t) :9 , t +
(2.38)
and ;
respectively. The potential A needs to be uniquely specified, and we do so by imposing the Coulomb gauge condition, ·A:0
(Coulomb gauge).
(2.39)
Substituting Eq. (2.39) into Eqs. (2.37) and (2.38) gives A : 9J : 9/+ .
(2.40)
In free space (no material boundaries), the solutions to Eq. (2.40) can be written as A(x, t) : 4
4
J(x, t) dv,
x 9 x
(2.41)
and
1 (x, t) dv. (2.42)
x 9 x 4+ 4 Notice that there is no time retardation in Eqs. (2.41) and (2.42) and, therefore, any change in a source is felt instantaneously. The procedure for obtaining the fields is as follows: Given the source terms J(x, t) and (x, t), evaluate Eqs. (2.41) and (2.42). This gives A(x, t) and (x, t). Once the potentials are known, E and B are obtained using (x, t) :
E(x, t) : 9
A(x, t) 9 (x, t), t
and B(x, t) : ; A(x, t). The remaining fields are obtained from the constitutive relations D : + E and H : B/.
87
2.5 STATIC THEORY
2.5 STATIC THEORY In static field theory there is no time variation. Amazingly, this restrictive theory applies to a wide range of important phenomena involving steady currents (magnetostatic theory), and stationary charge distributions (electrostatic theory). We make extensive use of magnetostatic theory in our study of permanent magnet structures and direct current devices. In static field theory the time-dependent terms in Maxwell’s equations are negligible: D B : : 0. t t
(2.43)
When this is the case, the field equations uncouple into magnetostatic and electrostatic equations as shown in the following diagram: D :0 t B :0 t
Maxwell’s Equations $ Uncouple
;H:J (Magnetostatic) ·B:0 ;E:0 (Electrostatic) ·D:
The uncoupled magnetostatic and electrostatic equations along with their constitutive relations can be solved independently for the magnetic and electric fields.
2.5.1 Magnetostatic theory The equations for magnetostatic analysis can be summarized as follows: Magnetostatic Analysis Differential Form ;H:J
Integral Form
!
·B:0
H · dl :
J · ds 1
B · ds : 0 1 Constitutive Relation B : (H ; M)
88
CHAPTER 2 Review of Maxwell’s Equations
The first integral equation is Ampere’s circuital law, which states that the circulation of the magnetic field intensity around any closed path is equal to the free current flowing through the surface bounded by the path. The second integral equation states that the total outward flux of the B-field over a closed surface equals zero. This implies that there are no isolated sources or sinks of magnetic field. The magnetostatic equations can be solved directly for the fields. However, it is often more convenient to obtain the fields using the vector potential A introduced in Section 2.3. The equation for A follows from the analysis presented there, with the additional constraint that D/t : B/t : 0. In this case, Eq. (2.32) reduces to A 9 ( · A) : 9J.
(2.44)
Here A is uniquely specified by imposing the Coulomb gauge condition (2.39), which gives A : 9J.
(2.45)
The solution to Eq. (2.45) can be expressed in the integral form using the free space Green’s function G(x, x):91/4(1/ x9x ) for (Appendix B). Specifically, we have A(x) :
4
J(x) dv.
x 9 x
(2.46)
4 Once A is known, the B-field follows from B(x) : ; A(x) : :
4 4
; 4
J(x) dv
x 9 x
J(x) ; (x 9 x) dv.
x 9 x
(2.47)
4 The H-field is obtained from the constitutive relation H : B/. A magnetostatic field contains energy. In the absence of dissipation mechanisms, the energy stored in a field equals the energy required to create it [7, 8]. It can be shown that the energy supplied to a system of currents and magnetizable materials in creating a field B : ; A is given by W : K
4
J · dA dv
(J)
(2.48)
89
2.5 STATIC THEORY
or W : K
(J). (2.49) H · dB dv 4 If the system is linear in the sense that A is proportional to J, then Eqs. (2.48) and (2.49) reduce to 1 W : K 2 and 1 W : K 2
A · J dv
(J)
(2.50)
4
B · H dv (J). (2.51) 4 Notice that in Eq. (2.50) the integration extends over the region occupied by the source, which is usually finite, whereas in Eq. (2.51) the integration typically extends over all space. From Eq. (2.51) we find that the magnetostatic energy density for a linear system can be written as 1 w : B·H K 2
(J/m).
(2.52)
This completes our review of magnetostatic theory. A more comprehensive presentation is given in Chapter 3.
2.5.2 Electrostatic theory The equations for electrostatic analysis are as follows: Electrostatic Analysis Differential Form
Integral Form
;E:0 · D :
E · dl : 0
!
D · ds :
1
dv 4
Constitutive Relation D:+ E;P The first integral equation provides the basis for Kirchhoff’s voltage law, which states that around any closed path the algebraic sum of the
90
CHAPTER 2 Review of Maxwell’s Equations
voltage drops equals the algebraic sum of the voltage rises. The second integral equation is Gauss’ law, which states that the total outward flux of the D-field over a closed surface equals the total free charge enclosed by the surface. The fields can be obtained by direct solution of the electrostatic equations. However, they can also be obtained using the scalar potential introduced in Section 2.3. The equation for follows from the analysis presented there if we impose the additional constraint D/t : B/t : 0. In this case, Eq. (2.33) reduces to : 9 . +
(2.53)
The solution to Eq. (2.53) can be expressed in integral form using the free space Green’s function for (Appendix B), that is, (x) :
1 4+
4 Once is known, E is obtained from
(x) dv.
x 9 x
(2.54)
E(x) : 9(x) :9
1 4+
1 : 4+
(x) dv
x 9 x
4 (x)(x 9 x) dv.
x 9 x
(2.55) 4 Here D is obtained from D : +E. Notice that E : 9 implies that the E-field points in the direction of decreasing potential. Consider the line integral of E between two points a and b. As E : 9, we have
@
@
· dl ? : 9 , (2.56) ? @ where and are the potentials at a and b, respectively. The electrical ? @ potential energy E of a charge q at points a and b is E : q and ? ? E : q , respectively. Thus, the potential difference 9 is the @ @ @ ? energy expended in moving a unit charge from point a to point b. If no magnetic field is present, the potential difference is independent of the path taken. The electrostatic field contains energy. Specifically, it can be shown that the energy supplied to a system of charges, isolated conductors, and ?
E · dl : 9
91
2.6 SUMMARY
dielectric materials in creating a field E : 9 is given by 1 W : C 2
M
(J), (2.57) d dv 4 where is the free charge density [7]. When the system is linear in the sense that is linearly proportional to , then Eq. (2.57) reduces to 1 W : C 2
dv
(J).
(2.58)
4 We substitute · D : , make use of the identity · (D) : · D ; D · , and obtain 1 W : C 2
D · E dv
(J),
(2.59)
4 which gives the electrostatic energy in terms of the fields. Notice that the integration in Eq. (2.58) extends over a region containing the free charge, which is usually finite, whereas the integration in Eq. (2.59) typically extends over all space. In addition, Eq. (2.59) implies that the electrostatic field has an energy density 1 w : D·E C 2
(J/m).
(2.60)
2.6 SUMMARY We have presented an overview of classical electromagnetic field theory. We summarize the key results for future reference. Electromagnetic fields are produced by sources of charge and current and are governed by Maxwell’s equations: Differential Form ;H : J ;
D t
Integral Form
H · dl :
!
·B:0 ;E : 9
B t
·D:
J;
1
B · ds : 0 1
E · dl : 9 !
1
D · ds : 1
D · ds t
B · ds t dv
4
Ampere’s circuital law Conservation of flux Faraday’s law Gauss’ law
92
CHAPTER 2 Review of Maxwell’s Equations
The field equations are consistent, but not all independent. For a complete theory they must be augmented by the constitutive relations B : (H ; M), and D : + E ; P. (2.61) In stationary, linear, homogeneous and isotropic media these reduce to B : H and D : + E. There is an additional constitutive relation that relates the E-field to the current density J : E. Maxwell’s equations imply the conservation of charge, ·J;
: 0. t
(2.62)
Moreover, if a charge q moves with velocity u through a region permeated by E- and B-fields, it experiences a Lorentz force F : q(E ; u ; B).
(2.63)
In the continuum limit, Eq. (2.63) generalizes to f : E ; J ; B
(N/m),
(2.64)
where is the free charge density and J is the free current density. When solving Maxwell’s equations in regions consisting of different materials, we need to impose boundary conditions at the material interfaces. The boundary conditions for the tangential components of H and E are H 9H :J , R R (2.65) E :E . R R For most practical applications J : 0 and H : H . The boundary R R conditions for the normal components of B and D are B :B , L L (2.66) D 9D : . Q L L The fields can be determined by direct solution of Maxwell’s equations. However, it is often more convenient to make a change of variables to the scalar and vector potentials and A where B:;A
(2.67)
and E:9
A 9 . t
(2.68)
93
2.6 SUMMARY
FIGURE 2.2
Simplification of Maxwell’s equations in terms of the fields.
94
FIGURE 2.3
CHAPTER 2 Review of Maxwell’s Equations
Simplification of Maxwell’s equations in terms of the potentials.
In stationary, linear, homogeneous and isotropic media we impose the Lorentz condition · A : 9+
, t
and find that the potentials satisfy the inhomogeneous wave equations, A 9 +
A : 9J, t
9 + :9 . t +
(2.69)
95
2.6 SUMMARY
These equations have solutions 4
1 4+
A(x, t) :
4
J(x, t 9 x 9 x /u) dv,
x 9 x
(2.70)
and (x, t) :
(x, t 9 x 9 x /u) dv.
x 9 x
(2.71) 4 The B- and E-fields can be obtained from Eqs. (2.70) and (2.71) using Eqs. (2.67) and (2.68). Once B and E are known, H and D are obtained from the constitutive relations. Finally, electromagnetic applications can be broadly categorized as either high-frequency, quasi-static, or static. In Figs. 2.2 and 2.3 we illustrate the reduction and simplification of the field equations for these categories in terms of the fields and the potentials, respectively. Note that these equations are written for a stationary, linear, homogeneous and isotropic medium. References 1. Jackson, J. D. (1962). Classical Electrodynamics, 2nd ed., New York: John Wiley and Sons. 2. Cheng, D. K. (1989). Field and Wave Electromagnetics, 2nd ed., New York: Addison-Wesley. 3. Lorrain, P., Corson, D. P., and Lorrain, F. (1987). Electromagnetic Field and Waves, New York: W. H. Freeman and Co. 4. Reitz, J. R., Milford, F., and Christy, R. (1979). Foundations of Electromagnetic Theory, 3rd ed., New York: Addison-Wesley. 5. Silvester, P. P. and Ferrari, R. L. (1990). Finite Elements for Electrical Engineers, 2nd ed., New York: Cambridge Univ. Press. 6. Mayergoyz, I. (1998). Nonlinear Diffusion of Electromagnetic Fields with Applications to Eddy Currents and Superconductivity, New York: Academic Press. 7. Johnk, C. T. A. (1988). Engineering Electromagnetic Fields and Waves, 2nd ed., New York: John Wiley and Sons. 8. Morrish, A. H. (1983). The Physical Principles of Magnetism, Malabar, FL: R. Krieger Publishing Co.
a
This Page Intentionally Left Blank
CHAPTER
3
Field Analysis
3.1 INTRODUCTION In this chapter we study the theory and methods used in the analysis of steady currents, permanent magnets, and magnetic circuits. We start with a review of magnetostatic field theory, which includes a discussion of force, torque, energy, and inductance. Next, we introduce the current and charge models for magnetic materials. These are used to reduce a permanent magnet to an equivalent source term. Following this, we study magnetic circuits, introduce the concept of reluctance, and show how to transform a physical circuit into an equivalent lumped-parameter circuit. Finally, we discuss various analysis methods including boundary-value theory, the method of images, finite element analysis (FEA), and the finite difference method.
3.2 MAGNETOSTATIC ANALYSIS Although the magnetostatic field equations were presented in both differential and integral form in Section 2.5, we repeat them here for convenience. The field equations in differential form are and
;H : J,
(3.1)
· B : 0,
(3.2)
where J (A/m) is the free current density, and Field
Description (units)
H
Magnetic field intensity (A/m)
B
Magnetic flux density (T)
(3.3)
97
98
CHAPTER 3 Field Analysis
The field equations in integral form are
H · dl : !
J · ds,
(3.4)
1
and
B · ds : 0.
(3.5)
1
The magnetic flux through a surface S is given by :
B · ds;
(3.6)
1
flux is measured in webers (Wb). The magnetostatic fields and their units were discussed in Section 1.2. We repeat the following physical descriptions for convenience. 1. Field strength H: Consider a straight, infinitely long wire carrying a current i : 2 A. The conductor generates a tangential field strength H : 1 A/m at a radial distance r : 1 m from its center. As another example, consider a long solenoid that has n turns per meter and carries a current of 1/n A. It generates a field strength H : 1 A/m along its axis. 2. Flux density B: Consider an infinitely long conductor that carries a current i : 1 A perpendicular to an external B-field. A force of 1 newton will be imparted to each meter of the conductor when B : 1 T. 3. Flux : Consider a single turn coil with 1 Wb of magnetic flux passing through it. A voltage of 1 V will be induced in the coil when the flux is uniformly reduced to zero in one second. Equation (3.4) is Ampere’s circuital law. It states that the line integral of H around any closed path C is equal to the free current flowing through the surface bounded by C. Equation (3.5) states that the total outward flux of B over a closed surface equals zero. This implies that there are no isolated sources or sinks of magnetic field. The integral equations are easily derived from the differential equations and vice versa (Section 2.2.2). However, neither set of equations alone is sufficient to determine the fields. This is because there are six unknown field components (three for both H and B) and at most only four independent equations (three for Eq. (3.1) and one for Eq. (3.2)). Therefore, the field equations must be augmented by an additional
99
3.2 MAGNETOSTATIC ANALYSIS
independent equation that takes the form of the constitutive relation B : (H ; M). (3.7) For linear, homogeneous and isotropic materials, B and M are proportional to H. Specifically, B : H,
(3.8)
and M : H, (3.9) K where and are the permeability and susceptibility of the material. K Notice that and are related to one another. Specifically, from Eqs. K (3.7), (3.8) and (3.9) we find that : ( ; 1), K
(3.10)
and : 9 1. (3.11) K When a field solution is sought in a region consisting of different materials, the appropriate boundary conditions must be applied at the material interfaces. In Section 2.2.3 we found that the normal component of B is continuous across an interface B :B . (3.12) L L We also found that the tangential component of H is discontinuous by an amount equal to the free surface current density J : LQ HR 9 H : J . (3.13) R LQ In the applications that we study, J : 0, and the tangential component LQ of H is continuous: H :H . (3.14) R R For many applications, we can determine the fields directly from the field equations. This is especially true when a high degree of symmetry is present. For example, we can use Eq. (3.4) to determine H when there is sufficient symmetry to evaluate H · dl around a closed path bounding the source J. The following examples illustrate this approach. Example 3.2.1 Determine the B-field outside an infinity long wire of radius R with current density J (Fig. 3.1).
100
CHAPTER 3 Field Analysis
FIGURE 3.1
Infinite wire.
Solution 3.2.1 Use cylindrical coordinates with the axis of the wire along the z-axis, J : J z . Apply Eq. (3.4) to a closed circular path of radius r R. From X symmetry we know that H is in the direction and that it has a constant magnitude H (r) at a fixed radius r. Thus, we obtain
L
H r d :
L 0
J r dr d, X
which reduces to or
H 2r : J R, X H (r) :
I , 2r
(3.15)
where I : J R is the current in the wire. In free space, B : H and therefore, X I B : . (3.16) 2r ) Example 3.2.2 Consider a long tightly wound cylindrical coil (solenoid) with N turns and with its axis along the z-axis (Fig. 3.2). Let I and L denote the
101
3.2 MAGNETOSTATIC ANALYSIS
FIGURE 3.2
Solenoid.
current through and length of the coil, respectively. Determine the B-field along the axis of the coil. Solution 3.2.2 Use cylindrical coordinates. Apply Eq. (3.4) to a rectangular path with one side of length S along the axis of the coil and the other parallel side just outside the coil. This gives
H · dl : N
S I, L
! where N(S/L)I is the total current passing through the rectangle. The field components along the radial edges of the rectangle and just outside the coil are negligible relative to H on the axis. Therefore, we ignore contributions along X these segments and obtain S H S:N I X L or NI H : . X L Because B : H, we have
N B : I, X L
where N/L is the number of turns per unit length.
(3.17) )
Example 3.2.3 Determine the H-field inside a toroid of material of height h and inner and outer radii R and R , respectively, which is wrapped by a tightly
102
CHAPTER 3 Field Analysis
Toroid with coil [9].
FIGURE 3.3
wound N turn coil carrying a current I. Assume that the material is linear with a permeability (Fig. 3.3). Solution 3.2.3 Use cylindrical coordinates with the toroid centered with respect to the z-axis. From symmetry we know that H : H . Apply Eq. (3.4) to a circular path inside the toroid and obtain
Thus,
H · dl : H 2r : NI.
H :
NI 2r
(R r R ),
(3.18)
NI . 2r
(3.19)
and B :
)
3.2.1 Vector potential In the previous examples we determined the fields by direct solution of the field equations. However, it is often more convenient to use the vector potential A as was discussed in Section 2.3. We repeat the derivation of A for convenience. Consider a stationary, homogeneous and isotropic material with a linear constitutive relation B : H. Apply Proposition A.5.1 to Eq. (3.2) and introduce the vector potential A · B : 0 $ B : ;A.
(3.20)
Substitute this into Eq. (3.1) and obtain A 9 ( · A) :9J.
(3.21)
103
3.2 MAGNETOSTATIC ANALYSIS
Next, uniquely specify A by imposing the Coulomb gauge condition · A : 0. Substitute this into Eq. (3.21), which gives A :9J.
(3.22)
Thus, we find that the magnetostatic field equations (3.1) and (3.2) and the constitutive relation (3.8) collectively reduce to a single second-order equation (3.22). The derivation of Eq. (3.22) is summarized in the following diagram: A9( · A):9J Coulomb Gauge $ A :9J. · A:0
;H:J B:;A $ $ · B:0 B:H
The solution to Eq. (3.22) can be expressed in integral form using the free-space Green’s function G(x, x) :9(1/4)(1/ x 9 x ) for the operator (Appendix B). Specifically, we obtain A(x) :
4
J(x) dv.
x 9 x
(3.23)
The integration in Eq. (3.23) is over the source region (region of nonvanishing J). Once A is determined, we can compute B using B : ;A, which gives B(x) :
4
J(x);(x 9 x) dv.
x 9 x
(3.24)
It is important to note that Eq. (3.24) applies to problems in which the source term J is specified in a homogeneous space characterized by a uniform permeability . It does not apply when material interfaces are present. The derivation of Eq. (3.24) is summarized in Eq. (3.25): A :9J $ A(x) :
4
4
J(x) dv
x 9 x
B : ;A
B(x) :
4
J(x);(x 9 x) dv.
x 9 x
(3.25)
104
CHAPTER 3 Field Analysis
In free space : , and Eqs. (3.23) and (3.24) reduce to J(x) A(x) : dv, 4
x 9 x 4 and
B(x) : 4
(3.26)
J(x);(x 9 x) dv.
x 9 x
(3.27) 4 For thin current filaments or wires, J dv ; Idl, and Eqs. (3.26) and (3.27) become I A(x) : 4
!
dl ,
x 9 x
(3.28)
and I B(x) : 4
dl;(x 9 x) ,
x 9 x
(3.29) ! where C is the circuit path. This is known as the Biot-Savart law. For surface currents, Jdv ; Kds, where K (A/m) is the surface current density. If K flows along a surface S, then Eqs. (3.26) and (3.27) reduce to A(x) : 4
K(x) ds ,
x 9 x
(3.30)
K(x);(x 9 x) ds .
x 9 x
(3.31)
1
and B(x) : 4
1 The following examples demonstrate the use of these equations. Example 3.2.4 Consider a current I flowing in a wire of length 2L along the z-axis as shown in Fig. 3.4. Determine the vector potential A(x, y, 0) at any point in the x-y plane. Use A to determine B : ;A. Solution 3.2.4 We apply Eq. (3.28) using cylindrical coordinates. In this case dl : dzz , and
x 9 x : (r ; z.
105
3.2 MAGNETOSTATIC ANALYSIS
Current along the z-axis.
FIGURE 3.4
We obtain I A(r) : 4
*
dz
\* (r ; z
z
I : [ln z ; (z ; r!] * z \* 4
I (L ; r ; L : ln z . 4 (L ; r 9 L
(3.32)
Next, we compute B, B : ;A : ;A z X A 1 A X r 9 X . : r r
(3.33)
From symmetry we know that A / : 0. Combining Eqs. (3.32) and (3.33) X we have B:9
I (L ; r ; L ln r 4 (L ; r 9 L
I L : . 2r (L ; r
(3.34)
106
CHAPTER 3 Field Analysis
Infinite current sheet with height 2h; (a) perspective of sheet; and (b) cross-sectional view. FIGURE 3.5
For a long wire r L, and Eqs. (3.32) and (3.34) reduce to
and
I A(r) : 9 ln(r) ; C z , 2
(3.35)
I B : , 2r
(3.36)
where C is a constant. These are the usual results for an infinite wire.
)
107
3.2 MAGNETOSTATIC ANALYSIS
Example 3.2.5 Consider an infinitely long current sheet of height 2h with a surface current density K : K z (Fig. 3.5); K has the units of A/m. Determine the vector potential A at any point in the x—y plane. Use A to determine B : ;A. Solution 3.2.5 We use the results of the previous example. Specifically, we treat the sheet as a collection of infinite line currents. The current for an infinitesimal section of height dy is given by I : K dy. From Eq. (3.35) we know that the contribution from this line current at y is I dA : 9 ln [(x ; (y 9 y)] X 2 K dy :9 ln [x ; (y 9 y)]. 4 therefore, K A (x, y) : 9 X 4
F
ln [x ; (y 9 y)] dy
\F K : 9 (y 9 y) ln [x ; (y 9 y)] 9 2(y 9 y) 4 ; 2x tan\
(y 9 y) F . x \F
This reduces to
K A (x, y) : 9 (h 9 y) ln [x ; (y 9 h)] ; (h ; y) ln [x ; (y ; h)] X 4 9 4h ; 2x tan\
(3.37)
(3.38)
2h x x ; y 9 h
.
Next, we evaluate B, B : ;A :
A A X x 9 X y y x
K x ; (y 9 h) 2h x : ln x ; 2 tan\ y . 4 x ; (y ; h) x ; y 9 h
)
108
CHAPTER 3 Field Analysis
Current loop.
FIGURE 3.6
Example 3.2.6 Determine the B-field along the axis of a circular current loop of radius R and current I (Fig. 3.6). Solution 3.2.6 We apply the Biot-Savart law (3.29) with dl : R d , x 9 x : zz 9 Rr , and x 9 x : (R ; z. We obtain IR B(z) : 4
L
d z , (R ; z)
which reduces to B(z) :
IR z . 2(R ; z)
B(z) :
m , 2(R ; z)
(3.39)
This can also be written as
where m : IRz is the magnetic dipole moment of the current loop.
)
The forementioned examples demonstrate the use of the vector potential in obtaining the B-field. However, it can also be used to determine the flux . Specifically, the flux through a surface S can be expressed in terms of the line integral of A around a path C bounding S.
109
3.2 MAGNETOSTATIC ANALYSIS
Specifically, :
B · ds
1
:
(;A) · ds
1
:
A · dl, (3.40) ! where in the final step we have used Stokes’ theorem. In two-dimensional (2D) problems the vector potential can be used to visualize flux lines. A flux line is a line that indicates the direction of the B-field at any point along its length. In addition, in 2D problems flux lines coincide with equipotential contours of A. To see this, consider a 2D problem in Cartesian coordinates. The vector potential is in the z-direction A : A (x, y)z . Consider the change in A along an infinitesiX mal segment dl : dxx ; dyy that lies along a flux line: dA : A · dl. (3.41) X Because dl is along a flux line, it is parallel and proportional to B. Therefore, dl : B : B x ; B y , V W where is a constant of proportionality. However,
(3.42)
B : ;A A (x, y) A (x, y) X x 9 X y . y x Therefore, from Eq. (3.42) we have :
dl :
A (x, y) A (x, y) X X x 9 y . y x
(3.43)
Substitute Eq. (3.43) into Eq. (3.41) and obtain dA :
A (x, y) A (x, y) A (x, y) A (x, y) X X X X 9 x y y x
: 0. This shows that for two-dimensional problems, A is constant along field lines, or equivalently that equipotential contours of A coincide with flux lines.
110
CHAPTER 3 Field Analysis
3.2.2 Force and torque So far, we have studied methods for determining the magnetic field for a given current source. In this section, we study methods for determining the force and torque on a current source given an external magnetic field. When a particle of charge q moves through an external field B it experiences a Lorentz force F : q(u;B ).
(3.44)
We generalize this to the case of currents. Consider charges per unit S volume moving with velocity u. This gives rise to a volume current density J : u. The force on each charge is given by Eq. (3.44), and S therefore the force per unit volume is f : J;B (force density),
(3.45)
where f is a force density (N/m). The total force on a conductor with a current density J is obtained by integrating f over the volume of the conductor: F:
f dv.
(3.46)
4
For thin current filaments or wires, J dv ; I dl, and Eq. (3.46) reduces to
F:I
dl;B .
(3.47)
This line integral is evaluated in the direction of the current flow over the length of the wire as described in Section A.4. The force on an infinitesimal length of the wire is dF : I dl;B .
(3.48)
In the special case when a wire of length l is perpendicular to B , the Lorentz force reduces to F : IlB .
(3.49)
111
3.2 MAGNETOSTATIC ANALYSIS
If we know the force density f, we can determine the torque using T: :
r;f dv
4
r;(J;B ) dv, (3.50) 4 where r is the vector from the point about which the torque is computed. We apply Eq. (3.50) to thin wire carrying a current I. The torque on a segment of the wire of length dl is dT : I[r;(dl;B )], and the total torque is obtained via integration, T:I
r;(dl;B ).
(3.51)
(3.52)
It is important to note than when evaluating Eqs. (3.51) and (3.52) the force term (dl;B ) must be evaluated before the cross product with r is taken. Also, Eq. (3.52) is evaluated in the direction of the current flow over the length of the wire as described in Section A.4. We demonstrate the use of Eq. (3.47) in the following example. Example 3.2.7 Determine the force per unit length between two parallel wires separated a distance d and carrying currents I and I , respectively (Fig. 3.7). Solution 3.2.7 We use Eq. (3.48) to compute the force. Choose a coordinate system with I along the z-axis. This current gives rise to a field B across I I B : 9 x . 2d The force is given by F : I z ;B I I : 9 y 2d
(3.53)
Notice that the force is attractive, it pulls the wires together. If the currents are in opposite directions the wires repel one another. ) Other common conductor configurations are shown in Fig. 3.8. Figure 3.8a shows two parallel current sheets carrying currents I and I ,
112
CHAPTER 3 Field Analysis
FIGURE 3.7
Force between two parallel wires.
respectively [1]. The force between the conductors is
h d h I I 9 ln 1 ; F : tan\ h d 2h d
(parallel current sheets).
Figure 3.8b shows two coplanar current sheets. The force between the conductors is
I I Q>U w ;x F: ln dx (coplanar current sheets). 2 w w x Q
3.2.3 Maxwell stress tensor In the previous section we used the Lorentz force to determine the force and torque on steady currents. An alternate and more general approach entails the use of the Maxwell stress tensor [1, 2]. In this approach, the force density equation (3.45) is given by f:
1 · T.
(3.54)
113
3.2 MAGNETOSTATIC ANALYSIS
Force between conductors: (a) parallel current sheets; and (b) coplanar current sheets [1].
FIGURE 3.8
Here, T is the Maxwell stress tensor T : T x x ; T x y ; T x z ; T y x ; T y y VV VW VX WV WW ; T y z ; T z x ; T z y ; T z z , WX XV XW XX
(3.55)
where x , y , and z are the Cartesian unit vectors. Such an expression is called a dyadic, and it is but one way of expressing the components of the second-order tensor T. From Eq. (3.54) we have f:
1
T T T T T T VV ; VW ; VX x ; WV ; WW ; WX y x y z x y z ;
T T T XV ; XW ; XX z . x y z
If we substitute J : ;H into Eq. (3.45), and consider linear media with a constitutive relation B : H, the components of T are as
114
CHAPTER 3 Field Analysis
follows [2, 3]: T
VV [T] : T WV T XV
T
VW T WW T XW
T (B 9 B ) B B B B V VX V W V X T : B B (B 9 B ) B B . W WX W V W X T B B B B (B 9 B ) X XX X V X W (3.56)
The force on a body is computed using Eq. (3.46), that is, F:
1
· T dv.
1
T · n ds,
4 This volume integral can be written as a surface integral using the Divergence theorem. Specifically, we find that F:
(3.57)
1 where is the permeability of the medium where the integration takes place, n is the outward unit normal to the bounding surface, and the integration is performed over a surface S immediately surrounding the body. In principle, S should be the surface of the body itself. However, in practice (especially in FEA), S usually encompasses the body but is slightly offset from its surface. To illustrate the use of Eq. (3.57), consider a surface with a unit normal n : n x ; n y ; n z . According to Eq. (3.57), the magnitude of V W X the surface force in the x direction is 1 F : V
(T n ; T n ; T n ) ds. VV V VW W VX X
(3.58)
1 Equation (3.57) is particularly useful for evaluating the force on soft magnetic materials that have a high permeability ( ) in free space. Because of the high permeability, the field components tangential to the surface are negligible. We choose a surface immediately surrounding the body so that the integration takes place in free space. Then Eq. (3.57) reduces to F:
1 2
1
B n ds, L
(3.59)
where B is the component of B normal to the surface. Notice that the L direction of F is along the outward normal to the surface [1]. The following example demonstrates the use of the Maxwell stress tensor approach.
3.2 MAGNETOSTATIC ANALYSIS
FIGURE 3.9
115
Magnetic circuit actuator: (a) perspective; and (b) side view.
Example 3.2.8 Consider the actuator shown in Fig. 3.9a. Determine the force on the bar. Assume that all materials are linear with permeability , and that there is no leakage flux. Solution 3.2.8 The magnetic circuit of Fig. 3.9a is analyzed in Example 5.8.1. The field in the gap is given by Ni B : , E 2g
(3.60)
where N is the number of turns, i is the current, and g is the gap length. This field is constant across the gap, and perpendicular to the surface of the bar. Thus, B : B in the gap and zero elsewhere as there is no leakage flux. Choose L E a rectangular surface of integration immediately outside the bar as indicated by
116
CHAPTER 3 Field Analysis
the dotted line in Fig. 3.9b. Substitute Eq. (3.60) into Eq. (3.59) and obtain
1 B n ds L 2 1 1 [2B A ]n : E E 2 NiA E n : 4g
F(g) :
(3.61)
where A is the cross-sectional area of the gap. Notice that the direction of the E force is the same as n (the outward normal to the surface of the bar in the gap region). This unit vector points toward the actuator, and hence the force tends to move the bar towards the actuator, thus reducing the gap. )
3.2.4 Energy A magnetostatic field contains energy. In the absence of dissipation mechanisms the energy stored in the field equals the energy required to create it. The energy supplied to a system of currents and magnetizable materials in creating a field B : ;A is given by [4] W : K
J · dA dv (J). 4 This can also be written in terms of the fields W : K
(3.62)
H · dB dv (J). (3.63) 4 If the system is linear in the sense that A is proportional to J, then Eqs. (3.62) and (3.63) reduce to 1 W : K 2
1 W : K 2
A · J dv (J),
(3.64)
B · H dv (J),
(3.65)
4
and
4
respectively. In linear, homogeneous and isotropic materials, B : H
117
3.2 MAGNETOSTATIC ANALYSIS
and Eq. (3.65) becomes 1 W : K 2
H dv (J),
1 W : K 2
B dv (J).
or
(3.66)
4
(3.67) 4 Therefore, the energy density for a linear system can be written as 1 w : B · H (J/m). K 2
(3.68)
Recall from Section 1.7 that a magnetized specimen possesses selfenergy. Specifically, if a specimen of volume V has a fixed magnetization M, then it possesses a self-energy W :9 Q 2
4
M · H dv (self-energy), +
(3.69)
where H is the field in the specimen due to M (i.e., due to the + continuum of dipole moments comprising the specimen). The selfenergy equation (3.69) can be considered to be the energy required to assemble a continuum of dipole moments in absence on an applied field [5, 6]. Notice that the energy density associated with W is Q (3.70) w :9 M · H . + Q 2 In addition to its self-energy, a magnetized specimen acquires a potential energy when it is subjected to an applied field H . This is ? given by W : 9 ?
4
M · H dv. ?
(3.71)
This can be viewed as the work required to move the specimen from an environment with zero field to a region permeated by H . From Eq. ? (3.71) we see that the energy density due to the coupling of M to H is ? w : 9 M · H ? ? : 9 MH cos(), (3.72) ? where is the angle between M and H . ?
118
CHAPTER 3 Field Analysis
3.2.5 Inductance In this section we introduce the concept of inductance. Consider a linear magnetic system in the form of an electrical circuit that consists of a voltage source V, a switch, resistor R and an ideal (no resistance) N turn coil (Fig. 3.10). When the switch is open there is no current and the system contains no energy. When the switch is closed, a current I flows, and an energy W is supplied by the source. A portion of this energy is Q dissipated in the form of heat W : IR, and the remainder is stored in the magnetic field W : K W :W ;W , Q K where 1 W : K 2
A · J dv (J)
(3.73)
4 (Section 3.2.4). For a linear system, both A and J are proportional to I and Eq. (3.73) can be written as 1 W : LI, K 2
(3.74)
where L is called the self-inductance of the coil. The unit of inductance is the Henry (H), which equals one volt second/ampere (1 H : 1 Vs/A). It is important to note that the expression (3.74) only applies to a magnetically linear system. There is no simple definition of inductance for nonlinear systems [2].
FIGURE 3.10
Series RL circuit.
119
3.2 MAGNETOSTATIC ANALYSIS
We can express the inductance in terms of other field variables by substituting either Eq. (3.65) or (3.73) for W in Eq. (3.74). Specifically, K we find that
L:
1 I 1 I
A · J dv 4
(linear system)
(3.75)
B · H dv. 4
The first integral in Eq. (3.75) is over the volume of the conductor (the region of nonvanishing J), whereas the second integral is over all space (the region of nonvanishing B and H). The second integral in Eq. (3.75) can be decomposed into two separate integrals, that is, one for the region V inside the conductor, and another for the region V outside the conductor. In this case, we have L:L ;L
where 1 L : I
B · H dv
(internal inductance),
(3.76)
4int
inside conductor
and 1 L : I
4ext
B · H dv
(external inductance);
(3.77)
outside conductor
L and L are called the internal and external self-inductance, respec tively. It can be shown that
1 L : I
B · ds
1
ext
outside conductor
:
" , I
where S is the surface area external to and bounded by the conductor, and " is the flux linkage "Y
1ext
B · ds
(flux linkage),
120
CHAPTER 3 Field Analysis
FIGURE 3.11
Magnetically coupled circuits.
" is the total flux linking the conductor. For many applications L L and we use the approximation LL
" : . I
(3.78)
So far, we have considered the inductance of a single circuit in isolation (self-inductance). However, two or more circuits can be magnetically coupled to one another giving rise to mutual inductance effects. For example, consider the two circuits shown in Fig. 3.11. The current I through circuit 1 gives rise to a flux " through circuit 2 and vice versa. If the circuits are magnetically linear, the coupling can be expressed in terms of mutual inductance coefficients M and M as follows: " M : , (3.79) I and " M : . I These coefficients can be determined using
1 M : A · J dv I I 4 1 : B · H dv, I I 4
(3.80)
121
3.2 MAGNETOSTATIC ANALYSIS
and
1 M : A · J dv I I 4 1 B · H dv. : I I 4 Moreover, it can be shown that M :M . The total magnetic energy of the circuits can be written as 1 1 W : L I ; L I ; MI I , K 2 2 where M : M : M . We demonstrate some of these concepts in the following examples. Example 3.2.9 Determine the internal self-inductance per unit length of an infinitely long straight wire of radius R. Assume that the wire has a permeability . Solution 3.2.9 Let I denote the total current through the wire. The current within a radius r R (inside the wire) is i(r) : I
r . R
Consider a length h of the wire. From Example 3.2.1 we know that H and B at a radius r inside the wire are H (r) :
i(r) , 2r
(3.81)
and i(r) B (r) : . 2r
(3.82)
Substitute Eqs. (3.81) and (3.82) into Eq. (3.76), and integrate over the length of wire h. This gives L : I
F
L
h : (H). 8
0 i(r) r dr d dz 2r (3.83)
122
CHAPTER 3 Field Analysis
Parallel wire transmission line: (a) reference frame; and (b) cross section ([2] of Chapter 2).
FIGURE 3.12
Thus, the internal self-inductance per unit length is L : (H/m). 8
(3.84)
Example 3.2.10 Determine the internal and external self-inductance per unit length of a transmission line consisting of two infinite parallel wires (Fig. 3.12). Solution 3.2.10 Consider a length h of the transmission line. We first compute the internal inductance. This is just the sum of the internal inductances of the individual wires. From Eq. (3.84) we have L : . 4
(3.85)
Next, we compute the external inductance. We need to determine the flux linkage between the two wires. This depends on the field component perpendicu-
123
3.2 MAGNETOSTATIC ANALYSIS
lar to the area between the wires. Therefore, we need B which is the sum of the W fields of the individual wires B : B ; B, W W W where I B : , W 2x and I . B : W 2(s 9 x) The total flux between the wires is given by
Q\?
(B ; B dx W W ? hI s9a : ln . a
":h
The external inductance per unit length is " L : hI
s9a : ln . a Thus, the total inductance per unit length is L:L ;L s9a : ; ln . 4 a
(3.86)
If the radii of the wires are small relative to the separation distance (a s), then Eq. (3.86) reduces to
s L ln a
(H/m).
(3.87) )
Example 3.2.11 Determine the self-inductance of a long, tightly wound solenoid with N turns, length L and radius a (Fig. 3.2).
124
CHAPTER 3 Field Analysis
Solution 3.2.11 From Example 3.2.2 we know that the field on the axis of the solenoid is N B: I . L The flux through the coil is Ba. Because each turn of the coil intercepts this flux, the flux linkage is " : NBa N : I a. L From Eq. (3.78) we find that the inductance of the coil is N L: a (H). L The inductance per unit length is L:
N a (H/m). L
)
Example 3.2.12 Determine the self-inductance of a tightly wound N turn coil wrapped around a toroid of material with a height h and inner and outer radii R and R , respectively. Assume that the material is linear with a permeability (Fig. 3.3). Solution 3.2.12 From Example 3.2.3 we know that the field inside the toroid is H :
NI 2r
(R r R ).
The flux linkage is
0 H dr 0 hNI 0 1 : dr 2 r 0 hNI R . : ln 2 R
" : h
125
3.2 MAGNETOSTATIC ANALYSIS
FIGURE 3.13
Coaxial solenoids ([3] of Chapter 2).
The inductance is estimated using Eq. (3.78), which gives L:
hN R ln 2 R
(H).
(3.88)
Example 3.2.13 Determine the mutual inductance beween the coaxial solenoids shown in Fig. 3.13. Solution 3.2.13 Consider the inner solenoid. It has a current I , length l , and N turns. The field along its axis is given by Eq. (3.17) (Example 3.2.2), N I . B : l Thus, the flux through the outer solenoid due to I is B A, where A : R (R is the radius of the inner solenoid). This flux passes through each of the N turns of the outer solenoid. Therefore, the flux linkage is " : B N A, N N I R. : l The mutual inductance is given by Eq. (3.79), " M : I N N (H). l Now consider the outer solenoid. It has a current I , length l and N turns. : R
126
CHAPTER 3 Field Analysis
The field along its interior due to I is N I . B : l This field passes through the interior of the inner solenoid, which has an area A : R. It creates a flux linkage with N l /l turns of the inner solenoid. Therefore, the flux linkage " is l " :B N A l N N I R. : l The mutual inductance M is " M : I N N : R (H). l Notice that M : M as expected. )
3.3 THE CURRENT MODEL The current model is used in the analysis of permanent magnets. In this model, the magnet is reduced to a distribution of equivalent current. This is then input into the magnetostatic field equations as a source term, and the field is obtained using standard methods for steady currents. The derivation of the current model starts with the magnetostatic field equations (3.1) and (3.2). Apply Proposition A.5.1 to the solenoidal field B and introduce the vector potential B : ;A.
(3.89)
Substitute Eq. (3.89) into Eq. (3.1), taking into account the constitutive relation (3.7). This yields A 9 ( · A) : 9 (J ; ;M). Next, impose the Coulomb gauge condition · A : 0 and obtain A : 9 (J ; ;M).
(3.90)
127
3.3 THE CURRENT MODEL
The form of Eq. (3.90) suggests the definition of an equivalent magnetic volume current density J Y ;M. Notice that because Eq. (3.90) is a K linear equation, the potential A (and hence B) can be obtained as a superposition of the solutions for J and J separately. K The derivation of the current model is summarized in the following diagram: A 9 ( · A):9 (J ; J ) K ;H : J B : ;A (J Y ;M) K $ $ ·B:0 B : (H ; M) Coulomb Gauge ·A:0
A : 9 (J ; J ). K If there is no free current (J : 0), and if we assume an infinite homogeneous material (no boundaries), then the solution to Eq. (3.90) can be written in integral form using the free-space Green’s function G(x, x) : 9(1/4)(1/ x 9 x ) for the operator (Appendix B). Specifically, we find that A(x) : 4
J (x) K dv.
x 9 x
(3.91)
Notice that in Eq. (3.91) the integration is over the region containing the source J . We compute B : ;A and obtain K (x 9 x) dv. (3.92) B(x) : J (x); K
x 9 x 4
To obtain Eq. (3.92) we have used J (x) 1 ; K : 9J (x); , K
x 9 x
x 9 x and
1 (x 9 x) :9 ,
x 9 x
x 9 x
where indicates differentiation with respect to the unprimed variables. If the magnetization M is confined to a volume V (of permeability ), and falls abruptly to zero outside of V, then Eqs. (3.91) and (3.92)
128
CHAPTER 3 Field Analysis
reduce to A(x) : 4
4
J (x) K dv ;
x 9 x 4
1
j (x) K ds,
x 9 x
(3.93)
and B(x) : 4
4
(x 9 x) dv ; J (x); K 4
x 9 x
1
(x 9 x) ds, j (x); K
x 9 x
(3.94)
respectively. In these expressions S is the surface of the magnet, and J K and j are equivalent volume and surface current densities. These are K defined in the following: Equivalent Currents J : ;M (A/m) (volume current density) K j : M;n (A/m) (surface current density). K
(3.95)
We demonstrate the current model in the following examples. Example 3.3.1 Determine the flux density along the axis of a cylindrical magnet that is polarized along its axis with a uniform magnetization M : M z . Q The magnet has a radius R and length L (Fig. 3.14a). Solution 3.3.1 We use cylindrical coordinates. The equivalent current densities are computed using Eq. (3.95). We readily determine that J : ;M : 0. (3.96) K To evaluate the surface term, we first evaluate the unit surface normals n : and then compute
z z : 0 r
r:R
9z z : 9L
j : M;n : M (r : R). (3.97) K Q The only nonvanishing source term is a surface current that circulates around the body of the cylinder (Fig. 3.14b). We substitute Eqs. (3.96) and (3.97) into
129
3.3 THE CURRENT MODEL
FIGURE 3.14
Cylindrical magnet: (a) physical magnet; and (b) equivalent surface
current.
Eq. (3.94) with x : zz , x : Rr ; ; zz and x 9 x given by Eq. (A.15) with p (0, 0, z) and p (R, , z). This yields L (zz 9 (Rr ; ; zz ))R d dz B(z) : M ; . Q 4 [R ; (z 9 z)] \* We are interested in the z-component of B, which is given by
B (z) : X 4
L
\* This is readily integrated yielding
M R d dz Q . [R ; (z 9 z)]
M z z;L B (z) : Q 9 . X 2 ((z ; L) ; R (z ; R
(3.98)
We demonstrate Eq. (3.98) with some sample calculations. Calculations: We apply Eq. (3.98) to a solid cylindrical magnet. First, we calculate B along the line 2 z 12 mm with R : 5 mm, L : 20 mm, and X M : 8;10 A/m (B : 10,000 G) (Fig. 3.15). Notice that there is a rapid Q P decrease in the field near the surface of the magnet. Next, we compute B at a X fixed height z : 5 mm and vary the magnet radius 2 R 30 mm (Fig. 3.16). Here, we set L : 20 mm and M : 8;10 A/m as before. Notice that B Q X
130
CHAPTER 3 Field Analysis
FIGURE 3.15
Component B vs z for a solid cylinder. X
initially increases with R, but then peaks (at R < 17 mm), and gradually decreases. The reason for this is as follows: As R increases, the projection of the field along the z-axis increases, but the field source (equivalent surface current) moves farther away from the observation point. For small values of R the first effect dominates and the field increases. After a point, the second effect dominates and the field decreases. )
FIGURE 3.16
(z : 5 mm).
Component B vs magnet radius R for a cylindrical magnet X
131
3.4 THE CHARGE MODEL
The current model is useful for computing the force and torque on permanent magnets. The force and torque can be determined using the basic relation for the force on a distribution of currents in an external field as described in Section 3.2.2. The procedure is as follows. 1. Use the current model to reduce the magnet to a distribution of equivalent volume and surface current densities J and j as K K specified in Eq. (3.95). 2. Determine the force using F:
4
J ;B dv ; K
1
j ;B ds. K
(3.99)
where V and S are the volume and surface of the magnet, respectively. 3. Determine the torque using T:
4
r;(J ;B ) dv ; K
1
r;(j ;B ) ds, K
(3.100)
where r is the vector from the point about which the torque is computed. It is important to note that when evaluating Eq. (3.100), the force terms (J ;B ) and ( j ;B ) must be evaluated before the cross product with K K r is taken. We will use Eqs. (3.99) and (3.100) in our study of electromechanical devices in Chapter 5.
3.4 THE CHARGE MODEL The charge model is another useful method for analyzing permanent magnets. In this model, a magnet is reduced to a distribution of equivalent ‘‘magnetic charge.’’ The charge distribution is used as a source term in the magnetostatic field equations, and the fields are obtained using standard methods. The derivation of the charge model is as follows: Start with the magnetostatic field equations for current-free regions ;H : 0 and · B : 0. Next, apply Proposition A.5.2 to the irrotational field H and introduce a scalar potential , K H : 9 . K
(3.101)
132
CHAPTER 3 Field Analysis
Finally, substitute Eq. (3.101) and the constitutive relation B : (H ; M) into · B : 0 and obtain : · M. (3.102) K In the absence of boundary surfaces we can represent the solution to Eq. (3.102) in integral form using the free space Green’s function G(x, x) for (Appendix B). In particular, we find that (x) : K
G(x, x) · M(x) dv
:9
1 4
· M(x) dv,
x 9 x
(3.103)
where x is the observation point, x is the source point, operates on the primed coordinates, and the integration is over the volume for which the magnetization exists. If M is confined to a volume V (of permeability ), and falls abruptly to zero outside of this volume, then Eq. (3.103) becomes 1 (x) : 9 K 4
1 · M(x) dv ; 4
x 9 x
M(x) · n ds,
x 9 x
(3.104)
4 1 where S is the surface that bounds V, and n is the outward unit normal to S. The form of Eq. (3.104) suggests the definitions of volume and surface charge densities given in Eq. (3.105): Charge Model : 9 · M (A/m) K : M · n (A/m) K
(volume charge density),
(3.105)
(surface charge density).
The derivation of the charge model is summarized in Eq. (3.06): ;H : 0 H : 9 : 9 K K K $ $ ·B:0 B : (H ; M) ( : 9 · M). K
(3.106)
If the magnet is in free space, B : H and from Eqs. (3.101) and (3.104) we have B(x) : 4
4
(x)(x 9 x) K dv ;
x 9 x 4
1
(x)(x 9 x) K ds.
x 9 x
(3.107)
133
3.4 THE CHARGE MODEL
The B-field due to an isolated permanent magnet with a known magnetization M in free space is obtained from Eq. (3.107) by first evaluating and , and then performing the indicated integrations. K K From the preceding analysis, we find that the field at a point x due to a ‘‘point charge’’ Q (x) at x is + Q (x)(x 9 x) B(x) : K . 4
x 9 x
(3.108)
where Q (x) : (x)V K K
(volume charge),
(3.109)
or Q (x) : (x)A (surface charge), (3.110) K K and V and A are the volume and area elements containing (x) K and (x), respectively. In one dimension, Eq. (3.108) reduces to K Q (x) K B(x) : . 4 (x 9 x)
(3.111)
We demonstrate the use of the charge model in the following example. Example 3.4.1 Compute the field above a rectangular bar magnet of width 2a, depth 2b, and height L. The magnet is polarized along its axis with uniform magnetization as shown in Fig. 3.17a. Assume that the magnetization is M : M z . (3.112) Q Solution 3.4.1 We first determine the charge densities. From Eq. (3.105) we find that : 9 · M : 0. K To evaluate we first compute the unit surface normals K z z : 0 n:
9z
z : 9L
<x
x :
From Eq. (3.112) and the fact that : M · n we have : M for the top K K Q surface (z : 0), and : 9M for the bottom surface (z : L). We evaluate B K Q X
134
FIGURE 3.17
CHAPTER 3 Field Analysis
Rectangular bar magnet, (a) physical magnet; and (b) equivalent
charge.
along the z-axis due to the top surface of the magnet (z : 0) using Eq. (3.107) with x : zz and x : xx ; yy , B (z) : X 4
?
@
\? \@
Mz Q dx dy. [x ; y ; z]
(3.113)
Notice that the integrand in Eq. (3.113) is an even function of x and y, and therefore B (z) : X
?
@
Mz Q dx dy [x ; y ; z]
Mz ? b : Q dy (y ; z)(b ; y ; z M by ? : Q tan\ z(b ; y ; z
M z(a ; b ; z : Q 9 tan\ 2 ab
.
In the last step we have used tan\(x) ; tan\(1/x) : /2. This is the field due to the top surface. A similar analysis applies to the bottom surface with z
135
3.4 THE CHARGE MODEL
replaced by z ; L. The total field is given by
M (z ; L)(a ; b ; (z ; L) B (z) : Q tan\ X ab 9 tan\
z(a ; b ; z ab
.
(3.114) )
3.4.1 Force The charge model is also useful for determining the force and torque on a magnet in an external field. An expression for the force in terms of magnetic charge can be derived from Eq. (3.99). Specifically, start with F: :
4
J ;B dv ; K
1
j ;B ds K
(;M);B dv ; (M;n );B ds, 4 1 and then apply the vector identity
(;U);V dv 9
4
(3.115)
(n ;U);V ds
1 :
[U;(;V) 9 U( · V)] dv ; (U · )V dv 4 4 to Eq. (3.115) with U : M and V : B . This gives F:
4
(M · )B dv,
(3.116)
where
(M · )B : M ;M ;M B x V V x W y X z
; M ;M ;M B y W V x W y X z ; M ;M ;M B z . X V x W y X z
To obtain Eq. (3.116) we have used the fact that · B : 0, which is
136
CHAPTER 3 Field Analysis
always true, and the fact that ;B : 0, which holds because the source of B does not overlap the region occupied by M. From Eq. (3.116) we find that f : (M · )B
(force density),
(3.117)
is a force density. Apply the identity
(U · )V dv : 9 4
( · U)V dv ;
4
(n · U)V ds
1
to Eq. (3.116) with U : M and V : B and obtain
F:9
4
( · M)B dv ;
1
(M · n)B ds.
From Eq. (3.105) this can be rewritten as
F:
4
B dv ; K
1
B ds. K
(3.118)
We can use Eq. (3.118) to compute the force on a magnet. The procedure is as follows: 1. given a magnet with a magnetization M, determine the equivalent charge densities : 9 · M and : M · n ; and K K 2. substitute and into Eq. (3.118) and determine the force F. K K We demonstrate this procedure in the following example. Example 3.4.2 Determine the activation force of the solenoidal actuator shown in Fig. 3.18a. Assume that the magnet has a fixed and uniform magnetization M : 9M z . Q Solution 3.4.2 We compute the activation force using Eq. (3.118). Consider a cylindrical magnet of radius R and length h (Fig. 3.18b). The equivalent charge densities are obtained from Eq. (3.105). We find that : 9 · M : 0. K
137
3.4 THE CHARGE MODEL
Solenoid with permanent magnet plunger: (a) device with circuitry; and (b) magnet in isolation.
FIGURE 3.18
To evaluate , we first determine the unit surface normals K
9z z : 0
n :
z z : h r
r : R.
Next, we evaluate : M · n and obtain K
M z:0 Q : K 9M z : h Q The field B
due to a solenoid was derived in Example 3.2.2. Specifically, B
N : i , L
where i is the current, N is the total number of turns, and L is the length of the solenoid. Only the north pole of the magnet is in the field of the solenoid.
138
CHAPTER 3 Field Analysis
The force is therefore F:
1
B ds K
N : i M L Q
0
L
r dr d
N : i M R. L Q
)
Often, it is not possible to evaluate Eq. (3.118) analytically. When this is the case, the force is obtained numerically by discretizing the volume V and surface S of the magnet F : (x )B (x )V ; (x )B (x )A , (3.119) K L L L K N N N L N where V and A are indexed volume and area elements, respectively. L N Combining this expression with Eq. (3.108) gives the force between two separate magnets
(x ) (x )(x 9 x ) H V V F: K L K H L (3.120) H L 4
x 9 x L H L H (x ) (x )(x 9 x ) I A V (3.121) ; K L K I L I L
x 9 x 4 L I L I (x ) (x )(x 9 x ) H V A ; K N K H N (3.122) H N 4
x 9 x N H N H (x ) (x )(x 9 x ) I A A , (3.123) ; K N K I N I N
x 9 x 4 N I N I where the j and k indices label the discretization of one magnet, and n and p label that of the other. Once two magnets have been reduced to discretized charge distributions, we use Eqs. (3.120) to (3.123) to compute the force between them. From this analysis, it follows that the force between two discrete ‘‘point charges’’ Q (x ) at x and Q (x ) at x is K K Q (x )Q (x )(x 9 x ) F : K K 4
x 9 x
(3.124)
139
3.4 THE CHARGE MODEL
where the charge Q is computed using either Eq. (3.109) or (3.110) K depending on whether it is volume charge or surface charge. In one dimension Eq. (3.124) reduces to Q (x )Q (x ) F : K K . 4 (x 9 x )
(3.125)
We demonstrate the use of Eq. (3.125) in the following example. Example 3.4.3 Consider two thin cylindrical magnets oriented along the x-axis as shown in Fig. 3.19a. Both magnets have a radius R and length L and are polarized to a level M . Assume R L and determine the force between the two Q magnets as a function of their separation d. Solution 3.4.3 We compute the force on the magnet to the right. Because the magnets are thin (R L), we can treat the magnets as dipoles to first order (Fig. 3.19b). The charge at each pole face is given by Eq. (3.110), which reduces to Q : <M R, K Q where the < sign refers to the north and south poles, respectively. The force is obtained by summing the forces between the charges as given by Eq. (3.125). We obtain
Q 1 2 1 F(d) : K 9 ; 9 . 4 d (L ; d) (2L ; d )
(3.126)
Force between two thin bipolar cylindrical magnets: (a) physical system; and (b) equivalent charge model.
FIGURE 3.19
140
CHAPTER 3 Field Analysis
FIGURE 3.20
Magnetic dipole.
The force on the magnet to the left is of the same magnitude but opposite direction as shown. ) Finally, consider a short thin cylindrical magnet of radius a, length d, and magnetization M as shown in (Fig. 3.20). From the analysis in Q Example 3.4.3 we know that there are surface charges of 9Q and ;Q K K at the left and right ends of the magnet, where Q : M a. These two K Q charges form a magnetic dipole with a dipole moment given by m:Q d (dipole moment). K The field of the dipole can be obtained from Eq. (3.108), which gives
r r Q B: K 9 ; , (3.127) 4 r r where r : (1/2)d ; r and r : 9(1/2)d ; r. When d r, Eq. (3.127) can be approximated by [7]
3(m · r)r m B: 9 . 4 r r
(3.128)
Consider two dipoles m and m separated by a distance r d (Fig. 3.21). The force imparted to m by m follows from Eq. (3.117). Specifi cally, we have F : (m · )B . We substitute Eq. (3.128) into Eq. (3.129) and obtain [1]
(3.129)
5(m · r)(m · r)r 3 F : (m · m )r ; (m · r)m ; (m · r)m 9 . 4 r r (3.130)
141
3.4 THE CHARGE MODEL
FIGURE 3.21
Two separated dipoles.
3.4.2 Torque The charge model is also useful for determining the torque on a magnet in an external field. We derive a formula for the torque as follows: Start with the torque formula (3.100) from the current model T: :
4 4
r;(J ;B ) dv ; K
1
r;(j ;B ) ds K
r;[(;M);B ] dv ;
1
r;[(M;n );B ] ds.
(3.131)
Next, apply the identity
r;[(;U);V] dv 9
4 :
r;[(n ;U);V] ds
1
r;[U · V ; U;(;V) 9 U( · V)] dv ;
4
U;V dv, 4
with U : M and V : B . This gives T:
r;(M · B ) dv ;
4
4
M;B dv.
(3.132)
Here, we have used · B : 0 and ;B : 0. From Eq. (3.132) we see that if B is uniform (B : 0), then the torque reduces to a couple T:
4
M;B dv
(uniform B ).
(3.133)
142
CHAPTER 3 Field Analysis
Finally, apply the identity
r;(U · V) dv; U;V dv:9 ( · U)(r;V) dv; (n · U)(r;V) ds 4 4 4 1 to Eq. (3.132) with U : M and V : B . This yields T:
(9 · M)(r;B ) dv ; (n · M)(r;B ) ds. 4 1 From Eq. (3.105) we have T:
4
(r;B ) dv ; K
1
(r;B ) ds. K
(3.134)
We can use Eq. (3.134) to compute the torque on a magnet. The procedure is as follows: 1. given a magnet with a magnetization M, determine the equivalent charge densities : 9 · M and : M · n ; and K K 2. substitute and into Eq. (3.134) and determine the torque T. K K We demonstrate this procedure in the following example. Example 3.4.4 Determine the torque on a bipolar cylinder in a uniform field (Fig. 3.22a). Assume that the cylinder has a magnetization M at an angle Q with respect to the x-axis (Fig. 3.22b). Solution 3.4.4 We use Eq. (3.134) to determine the torque. First, determine the equivalent charge densities Eq. (3.105). Let R and h denote the radius and length of the cylinder, respectively. The magnetization is given by M : M [cos()x ; sin()y ], Q where is the angular rotation of M relative to the x-axis (Fig. 3.22b). It follows that : 9 · M : 0. The surface charge is given by K : M · n K : M [cos() cos() ; sin() sin()], (3.135) Q
where we have used n : r : cos()x ; sin()y . The external field is B : B x : B [cos()r 9 sin() ].
(3.136)
3.4 THE CHARGE MODEL
143
Bipolar cylinder in an external field: (a) perspective of the cylinder; and (b) cross-sectional view.
FIGURE 3.22
Substitute Eqs. (3.135) and (3.136) into Eq. (3.134) and obtain
T:
F
L
M [cos() cos() ; sin() sin()] Q
;(Rr ;B [cos()r 9 sin() ])R d dz L : 9M hRB [cos() cos() sin() ; sin() sin()] dz Q : 9M hRB sin()z . (3.137) Q
This is the torque on the magnet as a function of its rotation angle .
)
144
CHAPTER 3 Field Analysis
3.5 MAGNETIC CIRCUIT ANALYSIS Magnetic circuits consist of field sources such as magnets and coils, and flux conduits made from soft magnetic materials that guide and direct the flux. In this section we develop equivalent lumped-parameter models for analyzing such circuits. These models follow from the integral form of the magnetostatic field equations
H · dl :
!
J · ds
1 :I ,
(3.138)
and
B · ds : 0. (3.139) 1 In Eq. (3.138), I is the total current passing through the area S, which is bounded by the contour C. It is important to note that in order to evaluate the left-hand side of Eq. (3.138), we need to know H · dl. That is, we need to know the orientation and magnitude of H relative to the integration path dl. For most practical applications, we choose dl based on the circuit geometry and assume that H is parallel to dl, which reduces H · dl to the scalar product H dl. This is one of the key assumptions of circuit analysis. Another is that the flux generated by the field sources is confined to the circuit with negligible leakage. Recall that the flux through a surface S is given by :
B · ds (magnetic flux). (3.140) 1 These assumptions are difficult to verify a priori, and need to be carefully weighed when analyzing a circuit. However, even when they are not strictly valid, the performance of a magnetic circuit can usually still be estimated with reasonable accuracy.
3.5.1 Current sources Consider a magnetic circuit consisting of a discrete set of segments, each with a different permeability, and a field source in the form of a coil. A circuit with four segments is shown in Fig. 3.23. We apply Eq. (3.138) to
145
3.5 MAGNETIC CIRCUIT ANALYSIS
FIGURE 3.23
Magnetic circuit with four segments.
a circuit with m such segments and obtain
K H · dl : H l : I . (3.141) GG ! G To simplify the analysis, assume that each element has a linear response B : H , and constant cross-sectional area A . Then, from Eq. (3.140) we G G G G have B : /A , and Eq. (3.141) becomes G G G K l G :I . (3.142) A G G G G or K R :I , G G G
where
l R: G , (3.143) G A G G is the reluctance of the i th element. If the cross-sectional area and permeability of an element vary over its length, then Eq. (3.143) generalizes to an integration RY
dl A
(reluctance),
where and A are the permeability and the cross-sectional area along the path. The reluctance of different rectangular sections are shown in Fig. 3.24. If there is no flux leakage, : and Eq. (3.142) reduces to G K R :I . (3.144) G G
146
FIGURE 3.24
CHAPTER 3 Field Analysis
Reluctance of rectangular sections: (a) solid element; and (b) air
gap.
In addition, if the current source is an n-turn coil carrying a current i, then I : ni, and Eq. (3.144) becomes :
ni . R ;R ;···;R K
(3.145)
We can use Eq. (3.145) to compute the flux density in the ith element. Specifically, B : /A , which gives G G G 1 ni B: , G A (R ; R ; · · · ; R ) G K
(3.146)
(recall that : ). G In most applications Eq. (3.146) simplifies because the reluctance of a few elements dominates the denominator. In particular, soft magnetic circuit elements have a high relative permeability when they operate P below saturation (Fig. 3.25). Therefore, these elements have a much lower reluctance than gap regions, and can usually be ignored when gaps are present. Magnetic circuits have many similarities to electrical circuits. Recall that electrical circuits are governed by Kirchhoff’s voltage and current laws R I : V , G G I G I
(3.147)
I : 0. G G
(3.148)
and
147
3.5 MAGNETIC CIRCUIT ANALYSIS
FIGURE 3.25
The B-H curve of soft magnetic material.
Kirchhoff’s voltage law (3.147) states that around any closed path the algebraic sum of the voltage drops equals the algebraic sum of voltage rises (applied voltage). The current law (3.148) states that the algebraic sum of the currents flowing into a circuit node is zero. The analogous laws for magnetic circuits are
and
R: n i , G G II G I
(3.149)
: 0. G G
(3.150)
The first law states that around any closed path the algebraic sum of the products of the reluctance and flux equals the algebraic sum of ampereturns (A · t). The second law states that the algebraic sum of the magnetic flux flowing into a circuit node is zero. An example of Eq. (3.150) is illustrated in Fig. 3.26. The magnetic circuit laws (3.149) and (3.150) are summarized in the following:
!
H · dl : I
B · ds : 0
$ $
R: n i , G G II G I : 0. G G
(3.151)
148
CHAPTER 3 Field Analysis
FIGURE 3.26
Algebraic sum of flux into a node is zero.
In keeping with the electrical analogy, we introduce the notation V K for the source term in a magnetic circuit; V is known as the magK netomotive force (mmf) and is given by V : ni K
(3.152)
for an n-turn coil. The unit of V is the ampere (A). However, it is K often specified in terms of ampere-turns (A · t) because of the form of Eq. (3.152). If a circuit contains one source, then the first equation in Eq. (3.151) reduces to R :V . G G K G
(3.153)
If there are N sources, then , R : V I , K G G G I
(3.154)
where V I is the magnetomotive force (mmf ) of the kth source. K Recall that resistors in parallel give rise to an equivalent resistance 1 1 1 : ; ;···. R R R If we work with conductance G : 1/R, then the equivalent conductance is G :G ;G ;···.
149
3.5 MAGNETIC CIRCUIT ANALYSIS
FIGURE 3.27
Permeance of gap regions: (a) angular gap; and (b) radial gap.
Similarly, when parallel reluctances are encountered, it is convenient to work with permeance P, P:
1 R
(permeance).
The permeance of a rectangular element of length l, cross-sectional area A, and permeability is P:
A . l
Other geometries of interest are shown in Fig. 3.27. Specifically, the permeance of an angular gap between radially oriented inclined surfaces (Fig. 3.27a) is given by w P : ln(r /r ). Similarly, the permeance of a radial gap between cylindrical surfaces at radii r and r with axial length w that subtend an angle is (Fig. 3.27b), P:
w . ln(r /r )
Other geometries are discussed in Reference 8.
150
CHAPTER 3 Field Analysis
FIGURE 3.28
Magnetic circuit with current source.
Example 3.5.1 Consider the magnetic circuit shown in Fig. 3.28. Determine the flux density in the gap, and the inductance of the coil. Assume that the core has a permeability and that there is no leakage flux. Solution 3.5.1 We determine the field in the gap first. Choose an integration path along the axis of the circuit as indicated by the dotted line. Apply Eq. (3.138) to the path assuming that H is parallel to it. This gives
H · dl : ni,
!
or H l ; H l : ni, (3.155) AA EE where l and l are the lengths of the path in the core and gap, respectively, and A E H and H are the fields in these sections. Notice that I : ni. Next, apply Eq. A E (3.139) at the core-gap interface. As there is no leakage flux, all the flux that passes through the core also passes through the gap : , A E B A :B A (no fringing flux at gap), (3.156) A A E E where A and A are the cross-sectional areas of the core and gap. Substitute Eq. A E (3.156) into Eq. (3.155), making use of B : H and B : H . This gives A A E E ni B : E (A /A )(l /) ; (l / ) E A A E ni : . (3.157) (A /A )( /)l ; l E A A E
3.5 MAGNETIC CIRCUIT ANALYSIS
151
If we assume that A /A 1 and / 1, then Eq. (3.157) reduces to E A B : ni/l . It is instructive to rederive Eq. (3.157) using Eq. (3.144). E E Specifically, ni , (3.158) R ;R A E where R : l /A and R : l / A . If we substitute R , R , and : B A A A A E E E A E E E into Eq. (3.158) we obtain Eq. (3.157). When R R , we obtain E niA E : ( ). (3.159) l E Inductance: Next, we determine the inductance. From Eqs. (3.156) and (3.157) we have :
ni . B : A ( /)l ; (A /A )l A A E E Thus, the flux through the core is
(3.160)
:B A A A A
niA A . (3.161) ( /)l ; (A /A )l A A E E Because this flux passes through each turn of the coil, the flux linkage is :
niA A . ( /)l ; (A /A )l A A E E Finally, the inductance of the coil follows from Eq. (3.78), ":
L:
(3.162)
" i
nA A . (3.163) ( /)l ; (A /A )l A A E E If there is no gap, l : 0 and we obtain E nA A L: (no gap). (3.164) l A On the other hand, if there is a gap, and the permeability of the core is high ( ), then L can be approximated by nA E. L: (3.165) l E :
152
CHAPTER 3 Field Analysis
Calculations: We demonstrate the theory with some sample calculations. Consider a magnetic circuit with the following parameters: n : 110, i : 1 A, l : 10 cm, l : 1 mm, : 1000 , and A : A : 1 cm. The fields and flux A E A E in the core and the gap are as follows: Core
Gap
100
100,000
H (A/m) B (T)
4;10\ 4;10\
(Wb)
4;10\ 4;10\
If the circuit has no gap (l : 0), then we find that E Core H (A/m)
1100
B (T)
1.38
(Wb)
1.38;10\
)
Example 3.5.2 Determine the inductance of the actuator shown in Fig. 3.29. Assume that the core has a permeability , and that there is no leakage flux. Solution 3.5.2 We follow the same analysis as in the previous example. First, apply Eq. (3.138) to a path around the actuator circuit (dotted line) assuming that H is parallel to the path. This gives
H · dl : ni
!
FIGURE 3.29
Magnetic circuit actuator with a current source.
3.5 MAGNETIC CIRCUIT ANALYSIS
153
or H l ; 2H x : ni. (3.166) AA E Next, apply Eq. (3.139) at the core-gap interface assuming that : , A E B A :B A (no fringing flux at gap). (3.167) A A E E Substitute Eq. (3.167) into Eq. (3.166) with B : H and B : H , and A A E E obtain ni B : E (A /A )(l /) ; (2x/ ) E A A ni : . (A /A )( /)l ; 2x E A A The flux through the circuit is : B A . Therefore, E E niA E : . (A /A )( /)l ; 2x E A A We are interested in the inductance of the coil. This can be determined from the flux linkage ". Because the flux passes through each turn of the coil, the flux linkage is " : n. The inductance follows from Eq. (3.78), and is given by L:
" i
(3.168)
nA E . (3.169) (A /A )( /)l ; 2x E A A Notice that the inductance in a function of the gap spacing x, L : L(x). Because / 1, we can estimate the inductance using nA E. L(x) ) 2x :
3.5.2 Magnet sources Permanent magnets are frequently used as source elements in magnetic circuits. The analysis of a circuit with a magnet source is similar to that of a current source. However, magnets are different than current sources in two respects. First, the mmf for a current source (coil) is known and is given by ni, but the mmf for a magnet is not known a priori. Instead, it depends on the magnet’s operating point (B , H ), which is a function K K
154
CHAPTER 3 Field Analysis
of the circuit itself. Second, the magnet itself is physically part of the circuit, and therefore represents a circuit reluctance. The procedure for the analysis of permanent magnet circuits is demonstrated in the following example. EXAMPLE 3.5.3 Consider the circuit of Fig. 3.30. Derive an expression for the operating point of the magnet (B , H ). Determine the flux density B in the K K E gap. Derive an expression for the magnetostatic energy in the gap. Assume that the core has an infinite permeability -, and that there is no leakage flux. Furthermore, assume that the magnet has a linear second quadrant demagnetization curve of the form B :B ; H , (3.170) K P K K where : B /H , and B and H and the residual magnetization and K P A P A coercivity, respectively. SOLUTION 3.5.3 Apply Eq. (3.138) to a path around the circuit and obtain
!
H · dl : H l ; H l ; H l : 0, A KK EE
(3.171)
or H l ; H l : 9H l , (3.172) A EE KK where H , H , and H are the fields in the magnet, core, and gap, respectively, K E and l , l , and l are the lengths of the paths in the respective sections. Here, K A E I : 0 and all of the flux is supplied by the magnet. In fact, the magnet can be replaced by an equivalent source equal to 9H l (because the magnet operates KK in the second quadrant, H is negative and 9H l is positive). We compare K KK Eq. (3.172) with Eq. (3.155) and find that 9H l plays the role of ni. However, KK
FIGURE 3.30 Magnetic circuit with magnet source.
155
3.5 MAGNETIC CIRCUIT ANALYSIS
H is not known a priori. Instead, it depends on the operating point of the K magnet. Operating point: The operating point (H , B ) is determined as follows. First, K K as the core has a high permeability, its field H is negligible compared to H K and H . Thus, the second term on the right-hand side of Eq. (3.171) can be E ignored, which gives H l : 9H l . EE KK
(3.173)
l B : 9 H K , E Kl E
(3.174)
It follows that
where we have used B : H . Because there is no flux leakage, : or E E K E B A :B A . K K E E
(3.175)
Combining Eqs. (3.174) and (3.175) we obtain A l B : 9 E K H K A l K K E
(load line).
(3.176)
slope of load line
Equation (3.176) defines the magnet’s load line. This line has a slope equal to 9 (A /A )(l /l ) as shown in Fig. 3.31. The operating point is the intersec E K K E
FIGURE 3.31
Demagnetization curve with load line.
156
CHAPTER 3 Field Analysis
tion of the load line with the second quadrant demagnetization curve (3.170) A l B ; H : 9 E K H . P K K A l K K E second quadrant demagnetization curve
(3.177)
load line
This gives 9B P H : . (3.178) K (B /H ; (A /A )(l /l )) P A E K K E Notice that as the gap goes to zero, the H-field in the magnet goes to zero, that is, lim H : 0. As the gap goes to infinity, H approaches the coercivity JE K K lim H : 9H . From Eq. (3.178) we determine the equivalent source term JE K A Bl PK 9H l : (equivalent source). K K (B /H ; (A /A )(l /l )) P A E K K E Gap field: To determine the field in the gap, substitute Eq. (3.178) into Eq. (3.174) which gives l B K, P B : E (B /H ; (A /A )(l /l )) l P A E K K E E or B P B : . (3.179) E ((l /l )(B /H ) ; (A /A )) E K P A E K Gap energy: The magnetostatic energy density in the gap is given by Eq. (3.68), : B · H . K E E Because B and H are colinear and constant throughout the gap, the total E E energy is (3.180) W : B H l A , K E EE E where l A is the volume of the gap. It is instructive to represent W in terms E E K of the magnet’s field. To this end, substitute Eqs. (3.173) and (3.175) into Eq. (3.180) and obtain W : 9B H l A K KK K K : 9B H ; (volume of magnet). K K
(3.181)
3.5 MAGNETIC CIRCUIT ANALYSIS
157
Notice that W 0 because H 0. Thus, for a fixed volume of magnet K K (volume : l A ), the gap energy is maximized when the magnet’s energy K K product B H is a maximum (B H ) ) K K K K 3.5.2.1 Energy product. From Example (3.5.3) we see that: (a) the operating point (B , H ) of the magnet depends on the circuit (gap dimensions, K K and so on); and (b) for a given volume of material the energy supplied to the gap is maximum when the energy product is a maximum. Thus, efficient circuit designs are realized when the load line of the circuit intersects the second quadrant demagnetization curve near (B H ) . K K 3.5.2.2 Equivalent circuit. When a magnet has a linear second quadrant demagnetization curve B :B ; H , (3.182) K P K K it can be replaced by an equivalent circuit element consisting of a source and reluctance term. To see this, consider the circuit of Example 3.5.3 (Fig. 3.30). From Eq. (3.182) and : B /H we have K P A B H : K9H . (3.183) K A K Substitute this into Eq. (3.173) and obtain B H l ; K l :H l . (3.184) EE K AK K If there is no flux leakage : B A : B A , and Eq. (3.184) reduces to E E K K (R ; R ) : H l . (3.185) E K AK Here, R : l / A is the reluctance of the gap, and E E E l (3.186) R : K K A K K is the equivalent reluctance of the magnet ( : B /H ). Comparing Eq. K P A (3.185) with Eq. (3.153) we see that the magnet can be thought of as an equivalent source with an internal reluctance:
V :H l (mmf), K AK Equivalent Circuit (3.187) $ l for Magnet R : K (reluctance). K A K K It must be emphasized that Eq. (3.187) only rigorously applies when the second quadrant demagnetization curve is of the form (3.182).
158
CHAPTER 3 Field Analysis
It is instructive to compute B using the equivalent circuit model E (3.187). From Eqs. (3.185) and (3.186) we obtain
or
Hl AK B A : E E (l / A ; l / A ) E E K K K A H E K A : , ((l /l ) ; (A /A )) E K K E K
B P B : , E ((l /l )(B /H ) ; (A /A )) E K P A E K which is the same result as Eq. (3.179) in Example 3.5.3. The equivalent circuits for both current and magnet sources are summarized in Figs. 3.32 and 3.33. A comparison of magnetic and
Magnetic circuits and field relations: (a) coil and core with no gap; (b) coil and core with a gap; and (c) magnet and core with a gap.
FIGURE 3.32
3.5 MAGNETIC CIRCUIT ANALYSIS
159
Magnetic circuits and equivalent lumped parameter models: (a) coil and core with no gap; (b) coil and core with a gap; and (c) magnet and core with a gap.
FIGURE 3.33
electrical circuits is summarized in Table 3.1. A magnetic circuit with both current and magnet sources is reduced to an equivalent lumpedparameter circuit in Fig. 3.34. Finally, for many applications, magnetic circuit analysis gives a reasonable estimate of circuit performance. Obviously, more accurate predictions require a more rigorous analysis that takes into account factors such as saturation, flux leakage, and so on. Frequently, an optimum design can be determined from intuition based on prior experience. For example, consider the magnetic circuits shown in Fig. 3.35. Which of these provides the strongest field in the gap region? The
160
CHAPTER 3 Field Analysis Comparison of magnetic and electrical circuits
TABLE 3.1
Magnetic circuit
Electric circuit
H B : H
E J : E
NI V : K Hl AK :
R:
B · ds dl A
V I:
R:
J · ds dl A
Magnetic circuit with a current and magnet source: (a) physical circuit; and (b) equivalent lumped-parameter circuit.
FIGURE 3.34
3.6 BOUNDARY-VALUE PROBLEMS
161
Magnetic circuits: (a) least efficient; (b) somewhat efficient; and (c) most efficient.
FIGURE 3.35
least efficient circuit is (a), which has the most leakage flux, the second most efficient circuit is (b), and the most efficient circuit is (c). In general, the closer the magnets are to the gap, the more efficient will be the device.
3.6 BOUNDARY-VALUE PROBLEMS Boundary-value problems arise naturally and frequently in the study of magnetic field theory. In problems of this type we seek the solution to a
162
CHAPTER 3 Field Analysis
partial differential equation in a given region, with constraints on the solution specified on the boundary of the region. In particular, the value or normal derivative of the solution might be specified on the boundary. Boundary-value problems can often be solved analytically when the boundary coincides with constant coordinate surfaces in an orthogonal curvilinear coordinate system. In this case, the solution can be separated into a product of single variable functions, and the partial differential equation can be reduced to a system of ordinary differential equations. This is called the method of separation of variables. For our purposes, boundary-value problems are classified as either Dirichlet, Neumann or mixed. In the Dirichlet problem, the value of the solution is specified on the boundary, whereas in the Neumann problem the normal derivative is specified. In mixed problems, the value of the solution is specified on some parts of the boundary, and the normal derivative is specified on other parts. In this section, we study boundary-value problems for Laplace’s equation : 0, in Cartesian, cylindrical and spherical coordinates.
3.6.1 Cartesian Coordinates We start with Laplace’s equation in Cartesian coordinates, ; ; : 0. x y z
(3.188)
Assume that (x, y, z) separates into a product of single variable functions (x, y, z) : X(x)Y(y)Z(z).
(3.189)
Substitute Eq. (3.189) into Eq. (3.188) and obtain Y(y)Z(z)
dX(x) dY(y) dZ(z) ; X(x)Z(z) ; X(x)Y(y) : 0. (3.190) dx dy dz
Next, divide Eq. (3.190) by X(x)Y(y)Z(z), which gives 1 dX(x) 1 dY(y) 1 dZ(z) ; ; : 0. X(x) dx Y(y) dy Z(z) dz
(3.191)
As each term in Eq. (3.191) is a function of a single variable, they are separately equal to a constant. If this were not the case, a variation in x
163
3.6 BOUNDARY-VALUE PROBLEMS
in the first term in Eq. (3.191) would require a change in the remaining terms to preserve the equality. However, this is not possible because the remaining terms are independent of x. Thus, Eq. (3.188) reduces to the following three ordinary differential equations: 1 dX(x) : 9k ; V X(x) dx
(3.192)
1 dY(y) : 9k ; W Y(y) dy
(3.193)
1 dZ(z) : 9k. X Z(z) dz
(3.194)
and
These equations need to be solved subject to the prescribed boundary conditions, and the constraint k ; k ; k : 0. The constants k, k, and V W X V W k can be positive, zero or negative, and therefore k , k , and k can be X V W X real or imaginary. The solutions to Eq. (3.192), (3.193), and (3.194) can be written as follows:
A x;A A sin(k x) ; A cos(k x) V V or X(x) : A eHIVV ; A e\HIVV A sinh(kx) ; A cosh(kx) or A eIV ; A e\IV B y;B B sin(k y) ; B cos(k y) W W or Y(y) : B eHIWW ; B e\HIWW B sinh(ky) ; B cosh(ky) or B eIW ; B e\IW
(k : 0) V (k 0) V , (3.195) (k 0 with k : ( k ) V V
(k : 0) W (k 0) W , (3.196) (k 0 with k : ( k ) W W
164
CHAPTER 3 Field Analysis
and
C z;C C sin(k z) ; C cos(k z) X X or Z(z) : C eHIXX ; C e\HIXX C sinh(kz) ; C cosh(kz) or C eIX ; C e\IX
(k : 0) X (k 0) X , (3.197) (k 0 with k : ( k ) X X
where j : (91, sinh(kz) : (eIX 9 e\IX)/2 and cosh(kz) : (eIX ; e\IX)/2. The constants A , B , and C (i : 1, . . . , 4) are determined from the G G G boundary conditions. A particular solution to Eq. (3.188) can be written as (x, y, z) : X (x)Y (y)Z (z). IVIWIX IV IW IX A general solution can be expressed as a superposition of all such , that is, IVIWIX
(x, y, z) : C (x, y, z). III III IVIWIX V W X V W X The constants C are determined from the boundary conditions. IVIWIX Often, the geometry of a system is such that one of its dimensions is much greater than the other two, and the field is essentially invariant along the greater dimension. When this is the case, the field problem reduces to a 2D problem: ; : 0. x y
(3.198)
The general solution to Eq. (3.198) can be written as (x, y) : (a x ; a )(b y ; b ) k : k :0 V W
; (c sin(k x) ; c cos(k x))(d eILW ; d e\ILW), (3.199) L L L L L L L k :k 0 L V
where k ; k : 0, and k 0. We demonstrate the theory in the followV W V ing examples.
165
3.6 BOUNDARY-VALUE PROBLEMS
FIGURE 3.36
Cross-sectional view of a rectangular magnet between flux plates.
EXAMPLE 3.6.1 Figure 3.36 shows a cross-sectional view of a magnetic circuit. The circuit consists of a magnet resting on (and attached to) a soft magnetic (flux) plate, and a second flux plate positioned a distance g above the magnet. Assume that the plates have infinite permeability and that they are magnetically connected with zero reluctance far away from the observation point (at infinity). Further assume that the magnet is ideal with a linear second quadrant demagnetization relation of the form B : (H ; M).
(3.200)
Determine B in the gap region above the magnet. W SOLUTION 3.6.1 From symmetry we know that there is no variation in the field in the x or z directions. Therefore, the problem reduces to a 1D Dirichlet boundary-value problem. Because there is no applied current, we use the magnetic scalar potential formulation (Section 3.4). Specifically, we use Eqs. (3.101) and (3.102), which reduce to d K, dy
(3.201)
dM d W. K: dy dy
(3.202)
H:9 and
There are two regions to consider, the magnet (region 1), and the gap (region 2). In these regions we have M:
M y Q 0
(region 1) (region 2).
(3.203)
166
CHAPTER 3 Field Analysis
We find that the volume charge density is zero in both regions dM W : 0. :9 K dy Therefore, Eq. (3.202) reduces to K : 0. y The general solutions to this equation is given by Eq. (3.196) with k : 0, that W is, (y) : a y ; b K
(region 1),
(3.204)
and (y) : a y ; b (region 2). K The coefficients are determined from the boundary conditions.
(3.205)
Magnet-plate interface (y : 0): We assume that the flux plates have infinite permeability ( -). Therefore, the H-field in them is negligible, and the potential is constant. As the plates are connected (at infinity), they are both at the same potential. Without loss of generality, we set this potential to zero, that is, : 0. As the potential is continuous at the magnet-plate interface, we have (0) : K : 0. From this we find that b : 0. Magnet-gap interface (y : lm): At the magnet-gap interface the potential and normal component of B are continuous (B : (H ; M )). These conditions L W Q give (l ) : (l ) K K K K
(continuity of ), K
and
d 9 K ;M Q dy From these we find that
WJK
d K : 9 dy
WJK
(continuity of B ). L
a l :a l ;b , K K or b : (a 9 a )l K
(3.206)
167
3.6 BOUNDARY-VALUE PROBLEMS
and or
9a ; M : 9a , Q
M :a 9a . Q Combining Eqs. (3.206) and (3.207) we obtain
(3.207)
b :M l . (3.208) QK Gap-plate interface (y :g ; lm): At the gap-plate interface the potential is continuous, that is, (l ; g) : K K : 0, which gives or
a (l ; g) ; b : 0, K
b . a :9 (l ; g) K Finally, substitute Eq. (3.208) into Eq. (3.209) and obtain Ml QK . a :9 (l ; g) K Therefore, the field in the gap is
(3.209)
(3.210)
d H :9 K E dy
or
l K , :M Q (l ; g) K
l K (gap). (3.211) B : M E Q (l ; g) K We can also determine the field in the magnet. Combining Eqs. (3.207) and (3.210) we have
l K a :M 19 Q (l ; g) K g :M . Q (l ; g) K
168
CHAPTER 3 Field Analysis
Therefore, d H :9 K
dy g : 9M (magnet). (3.212) Q (l ; g) K Notice that H is in the opposite direction to the magnetization. This is the
demagnetization field. Notice that this field goes to zero when the gap is reduced to zero, that is, lim H : 0. ) E EXAMPLE 3.6.2 Figure 3.37a shows a partial cross-sectional view of a magnetic circuit. The circuit consists of rectangular multipole magnet, the top of which is attached to a soft magnetic flux plate. A second flux plate is positioned a distance g below the magnet. Assume that the flux plates have infinite permeability, and that the magnet is ideal with a linear second quadrant
Magnetic circuit: (a) magnetization pattern with reference frame; and (b) reduced geometry.
FIGURE 3.37
169
3.6 BOUNDARY-VALUE PROBLEMS
demagnetization curve given by B : (H ; M). (3.213) Determine B in the gap region. W SOLUTION 3.6.2 This problem constitutes a 2D boundary-value problem. As there is no applied current, we use the magnetic scalar potential formulation H : 9 , K
(3.214)
and : · M, (3.215) K (Section 3.4). The 2D geometry can be reduced by exploiting the symmetry that results from the repeating magnetic structure. Specifically, the field is symmetric about the vertical center lines of each pole. We consider a reduced rectangular region extending from the center of one pole to the center of the neighboring pole as shown in Fig. 3.37b. The reduced geometry consists of three regions with
M y Q M : 9M y Q 0
(region 1) (region 2)
(3.216)
(region 3).
We find that : 9 · M : 0 in all three regions, and Eq. (3.215) reduces to K K; K : 0. x y The general solution to this equation is given by Eq. (3.199). We rule out the k : k : 0 solution because of the periodicity of the geometry. Thus, we obtain V W (x, y) : (S eIW ; S e\IW) sin(k x) K ; (C eIW ; C e\IW) cos(k x) (region 1), (3.217) (x, y) : (S eIW ; S e\IW) sin(k x) K ; (C eIW ; C e\IW) cos(k x)
(region 2),
(3.218)
(region 3),
(3.219)
and (x, y) : (S eIW ; S e\IW) sin(k x) K ; (C eIW ; C e\IW) cos(k x)
respectively. The coefficients in Eqs. (3.217) to (3.219) are determined by imposing the boundary conditions. Notice that the three regions in Fig. 3.37b are collectively bounded by four lines that form a rectangular area. Two sides
170
CHAPTER 3 Field Analysis
of the rectangle, that is, the y-axis and the line x : l, represent lines of symmetry along which the normal component H if zero. The boundaries that L form the top and bottom of the rectangle are bordered by a ferromagnetic material with infinite permeability ( : -). Consequently, the tangential component H vanishes along these boundaries. We evaluate the boundary R conditions region by region. Region 1: One of the boundaries that border region 1 is the y-axis along which the normal component of the field vanishes: H(0, y) : 9 L
(x, y) K x
:0
(g y h).
V
This condition yields S eIW : 9S e\IW. (3.220) Region 1 is also bounded by a flux plate ( : -) at y : h. The tangential component of the field is zero along this boundary, that is, H(x, h) : 9 R
(x, y) K x
:0
(0 x l/2).
WF
This condition is satisfied when C eIF : 9C e\IF. Applying Eqs. (3.220) and (3.221) to Eq. (3.217) gives
(3.221)
(x, y) : C sinh(k (y 9 h)) cos(k x), (3.222) K where C is a constant. Region 2: The conditions for region 2 are similar to those of region 1. Specifically, the normal component of the field is zero along the symmetry line defined by x : l, and the tangential field is zero along the upper boundary defined by y : h: H(l, y) : 9 L
(x, y) K x
:0
(g y h),
(3.223)
(l/2 x l).
(3.224)
VJ
and H(x, h) : 9 R
(x, y) K x
:0 WF Applying the first condition to Eq. (3.218) gives S eIW : 9S e\IW,
(3.225)
171
3.6 BOUNDARY-VALUE PROBLEMS
and n k : l
(n : 1, 2, 3, . . .).
(3.226)
Then, the second condition is satisfied when C eIF : 9C e\IF. Relations (3.225) to (3.227) reduce Eq. (3.218) to (x, y) : C sinh K
(3.227)
n n (y 9 h) cos x , l l
(3.228)
where C is a constant. Magnet-magnet interface: At the interface between the two magnets the normal component of B and the tangential component of H must be continuous, that is, B(l/2, y) : B(l/2, y) L L
(g y h),
(3.229)
H(l/2, y) : H(l/2, y) R R
(g y h),
(3.230)
and
where BG(l/2, y) : 9 L
G(x, y) K x
(i : 1, 2),
VJ
and HG(l/2, y) : 9 R
G(x, y) K y
(i : 1, 2). VJ
From these relations we obtain C :C ,
(3.231)
and k :k . (3.232) Therefore, the solutions to both regions are identical. We form a general solution for these regions using superposition, that is,
n n (x, y) : C sinh (y 9 h) cos x . (3.233)
L
l l L
172
CHAPTER 3 Field Analysis
Region 3: For region 3 there are two boundaries about which the field is symmetric and on which the normal component of H is zero. These boundaries are vertical lines defined by x : 0 and x : l. Therefore, H(0, y) : 0 L
(0 y g),
(3.234)
H(l, y) : 0 (0 y g). L When these conditions are imposed, we find that
(3.235)
and
S eIW : 9S e\IW,
(3.236)
and n k : l
(n : 1, 2, 3, . . .),
(3.237)
respectively. Also, as the bottom of the region borders a flux plate, the tangential component of the field vanishes there: H(x, 0) : 0 R
(0 x l).
(3.238)
This implies that C : 9C . (3.239) We call this region the gap, and form its general solution via superposition,
n n (y 9 h) cos x . (3.240) (x, y) : C sinh L l l L The coefficients C are determined at the magnet-gap interface. L Magnet-gap interface: At the interface between the magnet and the gap the tangential component of H must be continuous, that is, (x, g) : H (x, g) (0 x l). (3.241)
R R This condition, along with the fact that the functions sin(nx/l) are orthogonal on (0, l), implies that H
C
L
sinh(ng/l) :C . L sinh(n(g 9 h)/l)
(3.242)
In addition, the normal component of B must also be continuous, that is, B
L
(x, g) : B
L
(x, g)
(0 x l).
(3.243)
173
3.6 BOUNDARY-VALUE PROBLEMS
Because B
L
:
(H ;M ) L Q (H 9M ) L Q
(0x l/2) (l/2 x l)
(3.244)
Eq. (3.243) reduces to
nx M n Q C K(n, h, g, l) cos : l 9M l L Q L where K(n, h, g, l) : cosh
(0 x l/2) (l/2 x l)
ng ng n 9 sinh coth (g 9 h) . l l l
(3.245)
(3.246)
Again, exploiting the orthogonality of cos(nx/l) on (0, l) we find that
J
nx J nx dx 9 cos dx . (3.247) l l J After integrating Eq. (3.247) we obtain 2M Q C : L nK(n, h, g, l)
cos
4lM (91)L\ Q C :9 L (n)K(n, h, g, l)
(n : 1, 3, 5, . . .),
(3.248)
which fully defines the solution for the gap region. Specifically, substitute Eq. (3.248) into Eq. (3.240) and obtain (x, y) : 94lM Q
(91)L\ ny nx sinh cos . l l (n)K(n, h, g, l) L
(3.249)
From Eq. (3.249) we find that the vertical component of the field in this region is B
W
(x, y) : 94 M Q
(91)L\ ny nx cosh cos , (n)K(n, h, g, l) l l L
(3.250)
where g : the gap, h : g ; t , t : height of the magnet, and K(n, h, g, l) is K K as specified in Eq. (3.246). Calculations: We demonstrate the use of Eq. (3.250) with some sample calculations. Consider a structure with l:4.8 cm, t :1.28 cm, and M :3000 K Q (A/m). We compute the gap field B (x, y) at y : 0.375 mm and 0 x l, W for three different gaps g : 6, 8, and 10 mm (Fig. 3.38). A similar calculation is performed with y:2.5, 5 and 7.5 mm, for a fixed gap g:10 mm (Fig. 3.39). )
174
CHAPTER 3 Field Analysis
FIGURE 3.38
Component B in the gap vs gap height g. W
FIGURE 3.39
Component B in the gap vs y. W
175
3.6 BOUNDARY-VALUE PROBLEMS
3.6.2 Cylindrical coordinates We now consider Laplace’s equation in cylindrical coordinates, that is,
1 1 r ; ; : 0. r z r r r
(3.251)
Again, assume that separates into a product of single variable functions (r, , z) : R(r)()Z(z).
(3.252)
Substitute Eq. (3.252) into Eq. (3.251), divide by Eq. (3.252) and obtain
1 d dR 1 d 1 dZ r ; ; : 0. dr r d Z dz Rr dr
(3.253)
Notice that the last term in Eq. (3.253) is a function of z only. Therefore, from the same argument that follows Eq. (3.191) we have dZ 9 kZ : 0, dz
(3.254)
and
r d dR 1 d r ; kr ; : 0. R dr dr d
(3.255)
The last term in Eq. (3.255) is a function of only and we set it equal to a constant, that is, 9n. Thus, d ; n : 0, d
(3.256)
and Eq. (3.255) reduces to r
d dR r ; (kr 9 n)R : 0. dr dr
If there is no restriction on , we must have () : ( ; 2), which restricts n to integer values. The solutions for () and Z(z) are of the form () :
A ;A A sin(n) ; A cos(n)
(n : 0) (n " 0)
(3.257)
176 and
CHAPTER 3 Field Analysis
B z;B B sinh(kz) ; B cosh(kz) or Z(z) : B eIX ; B e\IX B sin(kz) ; B cos(kz) or B eHIX ; B e\HIX The radial solutions are of the form
(k : 0) (k 0) (3.258) (k 0 with k : ( k
C ln(r) ; C (k : 0, n : 0) C rL ; C r\L (k : 0, n " 0) R(r) : C J (kr) ; C J (kr) (k 0, n " 0, n " integer) (3.259) L \L C J (kr) ; C Y (kr) (k 0, n : 0 or integer) L L C I (#r) ; C K (#r) (k 0 ; k : j#) L L where J (kr) are Bessel functions of the first kind of order
!L
(91)K(kr/2)K!L , (kr) : m!$(m < n ; 1) K
(3.260)
cos(n)J (kr) 9 J (kr) L \L , sin(n)
(3.261)
and Y (kr) : L where $(p) :
xNe\V dx
(p 0),
is the gamma function $(p ; 1) : p! when p is an integer. The remaining functions I (#r) : j\LJ (j#r), L L and K (#r) : jL>[J (j#r) ; jY (j#r)] L L L 2
177
3.6 BOUNDARY-VALUE PROBLEMS
are the modified Bessel functions of the first and second kind, respectively. The functions Y (kr) and K (#r) diverge when their respective L L arguments are zero. Therefore, they are discarded when the solution region includes r : 0. When k is real (k 0) the two independent radial solutions are J (kr) L and Y (kr), or J (kr) and J (kr) depending on whether or not n is an L L \L integer. As noted in the preceding, n must be an integer when takes on its full range of values. However, if the solution region spans a restricted range of (e.g., wedge-shaped), then n need not be an integer. In either case, the asymptotic behavior of these functions is oscillatory. Thus, the solution (r, , z) is exponential in the z-direction and oscillatory in the r-direction. When k is imaginary (k 0) we define k : j# and the two independent radial solutions are I (#r) and K (#r). These functions exhibit L L exponential asymptotic behavior. Thus, the solution (r, , z) is oscillatory in the z-direction and exponential in the r-direction. For applications in which there is invariance along the z-axis, Eq. (3.251) reduces to a 2D equation:
1 1 r ; : 0. r r r r The foregoing analysis applies with k : 0, and a particular solution is of the form (r, ) : (a ln(r) ; a )(b ; b ) n:0
; (c rL ; c r\L)(d sin(n) ; d cos(n)).
(3.262)
n"0
If the region of interest includes the full angular range 0 2, then a general 2D solution is of the form (r, ) : a ;b ln(r); (a rL;b r\L)(c sin(n);d cos(n)). L L L L L
(3.263)
We demonstrate the two-dimensional theory in the following example. EXAMPLE 3.6.3 Determine the magnetic field of a bipolar cylindrical magnet that is polarized perpendicular to its axis (Fig. 3.40). Assume that the length of the cylinder is much greater than its diameter. This justifies a 2D analysis.
178
CHAPTER 3 Field Analysis
Bipolar cylinder: (a) perspective of cylinder; and (b) cross-sectional view with reference frame.
FIGURE 3.40
SOLUTION 3.6.3 We solve a 2D boundary-value problem. Choose a coordinate system with the magnetization along the x-axis, which gives M : M x Q : M [cos()r 9 sin() ]. Q
(3.264)
Because there are no current sources we can represent H in terms of a magnetic scalar potential H : 9 , K
(3.265)
where : 0, (3.266) K (Section 3.4). Let (r, ) and (r, ) denote the solutions to Eq. (3.266) K K inside and outside the magnet, respectively. From symmetry, these solutions must be even functions of , G(r, ) : G(r, 9) K K
(i : 1, 2).
(3.267)
They must also be well behaved at r : 0 and r : -, respectively, (0, ) -, K (-, ) -. K
(3.268) (3.269)
179
3.6 BOUNDARY-VALUE PROBLEMS
The general forms for (r, ) and (r, ) compatible with these conditions K K are (r, ) : a rL cos(n) K L L
(r a),
(3.270)
and (r, ) : b r\L cos(n) (a r), (3.271) K L L which follow from Eq. (3.263). The coefficients a and b are determined from L L the boundary conditions imposed at the surface of the magnet (r : a). Specifically, the potential (r, ) and the normal component of B must be continuous K at this surface, that is, (a, ) : (a, ), K K
(3.272)
and [9(a, ) ; M] · r : 9(a, ) · r , (3.273) K K respectively. These conditions yield the following expression for (r, ): K M a (r, ) : Q cos() (a r). K 2 r Outside the magnet B : H, and therefore M a Q cos() B (r, ) : P 2 r
(a r),
(3.274)
(a r).
(3.275)
(a r).
(3.276)
and B (r, ) :
M a Q sin() 2 r
Also, because B : ; A we have M a A(r, ) : Q sin()z 2 r
Calculations: We demonstrate the use of Eqs. (3.274) and (3.275) with some sample calculations. Consider a cylinder with a radius a : 2.5 mm, and a magnetization M : 4.3;10 (A/m) (B : 5400 G). The fields B (r, ) and Q P P B (r, ) are computed for r : 3.8 mm and 0 (Fig. 3.41). Notice that B (r, ) peaks at : 0. This shows that the radial field component peaks at the P center of the north pole as expected. )
180
CHAPTER 3 Field Analysis
FIGURE 3.41
magnet.
Components B and B vs for an infinite bipolar cylindrical P
3.6.3 Spherical coordinates Finally, we study Laplace’s equation in spherical coordinates:
1 1 1 r ; sin() ; : 0. (3.277) r r sin() r sin() r r We assume a solution of the form (r, , ) : R(r)()%().
(3.278)
Substitute Eq. (3.278) into Eq. (3.277) and obtain %
dR R d d% R% d ; sin() ; : 0. (3.279) dr r sin() d d r sin() d
Next, multiply Eq. (3.279) by r sin()/R(r)()%() and obtain
1 d dR 1 d d% 1 d r ; sin() ; : 0. R dr dr % sin() d d d
(3.280)
181
3.6 BOUNDARY-VALUE PROBLEMS
We reduce Eq. (3.280) to a system of equations by introducing separation constants m and n(n ; 1), d ; m : 0, d
(3.281)
d dR r 9 n(n ; 1)R : 0, dr dr
(3.282)
1 d d% m sin() ; n(n ; 1) 9 % : 0. d sin() sin() d
(3.283)
If the full range of is allowed (0 2), then m must be an integer to ensure that () is singled valued, that is, () : ( ; 2). Equation (3.283) is known as Legendre’s equation, and the solutions %() are finite at : 0 and only if n is an integer. When m and n are integers, the solutions are of the form () : A sin(m) ; A cos(m),
(3.284)
R(r) : B rL ; B r\L>,
(3.285)
and %() : C PK(cos()) ; C QK(cos()), L L where PK(cos()) and QK(cos()) are the associated Legendre functions of L L the first and second kind, respectively. The functions QK(cos()) are L singular at : 0 and , and therefore are excluded for problems that include these angles. The general solution to Eq. (3.277) for regions that include : 0 and encompass the full azimuthal range 0 2 is of the form L (r, , ) : [A sin(m);B cos(m)](C rL;D r\L>)PK(cos()). L K K L L L K (3.286) The sum over m ends at m : n because PK(cos()) : 0 for m n. L For problems with azimuthal symmetry, there is no dependence and m : 0. When this is the case, Eq. (3.277) reduces to
1 1 r ; sin() : 0. r r r r sin()
(3.287)
182
CHAPTER 3 Field Analysis
If the solution region includes : 0 or , the general solution to Eq. (3.287) is of the form (r, ) : (A rL ; C r\L>)P (cos()), (3.288) L L L L where P (cos()) : P (cos()) are called the Legendre polynomials. The L L first few Legendre polynomials are as follows: n
P (cos()) L
0
1
1
cos() (3 cos() 9 1) (5 cos() 9 3 cos())
2 3
EXAMPLE 3.6.4 Determine the H-field due to a uniformly magnetized bipolar sphere of radius a in free space (Fig. 3.42a) [7]. Assume that the sphere is polarized along its diameter and is made from a permanent magnet material with a second quadrant demagnetization curve of the form B : (H ; M). (3.289) SOLUTION 3.6.4 Use spherical coordinates with the sphere oriented with its magnetization along the z-axis as shown in Fig. 3.42a, M : M z . Q The H-field can be expressed in terms of a scaar potential H : 9 , (3.290) K Because : 9 · M : 0, we have K : 0 (3.291) K (Section 3.4). Given the orientation of the sphere, will have no dependence. K Let (r, ) and (r, ) denote the solutions to Eq. (3.291) inside and outside K K the magnet, respectively. These functions must be well behaved at r : 0, and r : -, respectively, (0, ) -, K
(3.292)
(-, ) -. K
(3.293)
and
183
3.6 BOUNDARY-VALUE PROBLEMS
Bipolar sphere: (a) magnetization and reference frame; (b) B-field; (c) H-field; and (d) internal fields.
FIGURE 3.42
The general forms for (r, ) and (r, ) compatible with these conditions K K are (r, ) : A rLP (cos()) K L L L
(r a),
and (r, ) : C r\L>P (cos()) K L L L We apply the boundary conditions
(r & a),
H(a, ) : H(a, ) F F
(H : H), R R
B(a, ) : B(a, ) P P
(B : B), L L
and
184
CHAPTER 3 Field Analysis
at the surface of the sphere. This gives (A aL 9 C a\L>) d L L P (cos()) : 0, a d L L
(3.294)
and (9 ; M z ) · r : 9 · r , K Q K respectively. The second condition (3.295) reduces to
(3.295)
0 : M cos() 9 C a\ Q 9 (A naL\ ; C (n ; 1)a\L>)P (cos()). (3.296) L L L L The Legendre polynomials P (cos ()) are linearly independent relative to one L another. Therefore, each of the terms containing P (cos()) or d/(d)P (cos()) L L in Eqs. (3.294) and (3.296) must separately be zero for each value of n. In particular, when n : 0 we have d P (cos()) : 0, d and C a\ : 0. Therefore, C : 0. The coefficient A is undetermined from these conditions, but it represents an additive constant that may be set to zero without affecting the fields. For n : 1 we have A 9 C a\ : 0 and M 9 A 9 2C a\ : 0, Q which we solve simultaneously and obtain C : M a Q
(3.297)
and A : M . (3.298) Q Continuing in this fashion we find that the remaining terms A : C : 0 for L L n & 2. Thus, : M r cos(), K Q
185
3.7 METHOD OF IMAGES
and
a : M cos(). K Q r The corresponding fields are H : 9M z Q
(r a),
(3.299)
and
a H : M [2 cos()r ; sin() ]. Q r
(3.300)
Notice that inside the sphere, H is in opposite direction to the magnetization. This is the demagnetization field (Section 1.8). Finally, we determine B from Eqs. (3.289) and (3.299): B : (H ; M z ) Q : M z . Q The B- and H-fields inside and outside the sphere are shown in Fig. 3.42b and c, respectively. The orientations of B, H, and M inside the sphere are shown in Fig. 3.42d. )
3.7 METHOD OF IMAGES The method of images is used to solve field problems when the source is near a material with a high permeability, and the solution region manifests a high degree of symmetry. In this method, the material is replaced by a number of carefully chosen image sources that render the appropriate boundary conditions at the interface. Both the position and magnitude of the image sources need to be determined. Once the image sources are known, a field solution is obtained by taking the superposition of the solutions for the individual source terms (real and image) as if they were in free space. The solution is only valid in the region outside the material (where the real source is located). The advantage of this method is that the original boundary-value problem (with material interfaces) is reduced to a free-space problem (without material interfaces). The validity of the method is based on the uniqueness of the field solution. Specifically, because the image sources are chosen to render the prescribed boundary conditions, and because the field solution satisfies these conditions and the field equation in the solution region, it is unique.
186
CHAPTER 3 Field Analysis
Wire above a ferromagnetic plate: (a) physical system; and (b) equivalent system.
FIGURE 3.43
We demonstrate the method of images using a simple example. Consider an infinite wire carrying a current i a distance h above of an infinite nonconducting plate that has infinite permeability ( -) (Fig. 3.43a). In the plate, H 0 but B is finite, and as the media is nonconducting there is no surface current. From Eqs. (3.12) and (3.14) we know that at the surface of the plate the normal component of B must be continuous and the tangential component H must be zero, that is, B : B, L L (3.301) H : 0, R where B and B are the normal components of B inside and outside L L the plate, respectively. These same boundary conditions are obtained if the plate is replaced by an image current directly beneath the source, a distance h into the plate as shown in Fig. 3.43b. The magnitude and direction of the image current needs to be the same as the source so that the boundary conditions (3.301) are satisfied at the surface of the plate. Therefore, the original problem is transformed into one in which there are two wires in free space separated by a distance 2h. The field solution to this equivalent problem is the superposition of the solutions of the individual wires, and is valid above the plate (in the region containing the real source). A similar analysis applies to the case of a current filament near wedge-shaped regions [1]. The source and image currents
3.7 METHOD OF IMAGES
187
Source and image currents for wedge-shaped regions: (a) rectangular corner section; and (b) 60° wedge [1].
FIGURE 3.44
for a rectangular corner section, and for a 60° wedge are shown in Fig. 3.44. The method of images can be summarized as follows: 1. Determine the magnitude and position of image sources that render the boundary conditions (3.301) at the material interface; 2. replace the material by the equivalent image sources; and 3. obtain the field solution as a superposition of the fields due to all the sources both real and image as if they were in free space. This method is demonstrated in the following examples.
188
CHAPTER 3 Field Analysis
EXAMPLE 3.7.1 Compute the force on a wire carrying a current i above a ferromagnetic plate as shown in Fig. 3.43a. Assume that the plate has an infinite permeability -. SOLUTION 3.7.1 We replace the plate by an image current as shown in Fig. 3.43b. The problem reduces to finding the force between two parallel wires carrying a current i and separated by a distance 2h. This was solved in Example 3.2.7 and the force is given by Eq. (3.53) with d : 2h i F : 9 z . 4h
(3.302) )
EXAMPLE 3.7.2 Compute the field due to a cylindrical magnet of radius R and length L that is polarized along its axis with uniform magnetization M : M z , Q The magnet is resting on an infinite plate. Assume that the plate has an infinite permeability - (Fig. 3.45a). SOLUTION 3.7.2 We replace the plate by an image magnet (Fig. 3.45b). The image magnet is identical to the source magnet, and positioned as shown. From symmetry, we know that the field due to the two magnets will satisfy the boundary conditions (3.301) at the interface. The resulting magnet structure is equivalent to a single magnet of length 2L (Fig. 3.45c). Thus, the field is given by Eq. (3.98) with L replaced by 2L:
z ; (2L) z M B (z) : Q 9 . X 2 ((z ; 2L) ; R (z ; R
(3.303) )
EXAMPLE 3.7.3 Determine the force on a cylindrical bar magnet a distance z above an infinite ferromagnetic plate that has an infinite permeability (Fig. 3.46a). The magnet has a radius R and length L and is magnetized to a level M along its axis. Assume R L. Q SOLUTION 3.7.3 We replace the plate by an image magnet as shown in Fig. 3.46b. The image magnet is identical to the source magnet, and it is easy to check that the field due to the two magnets will satisfy the boundary conditions (3.301). The force between the source and image magnet was determined in Example 3.4.3. It is given by Eq. (3.126) with d replaced by 2z,
Q 1 2 1 ; 9 , F(z) : K 9 4 (2z) (L ; 2z) 4(L ; z) where Q : M R. K Q
(3.304) )
3.7 METHOD OF IMAGES
189
Cylindrical magnet on a ferromagnetic plate: (a) physical system; (b) magnet and image; and (c) equivalent system.
FIGURE 3.45
EXAMPLE 3.7.4 Compute the field above a rectangular bar magnet of width 2a, depth 2b and length L, which is polarized along its axis with uniform magnetization M : M z . Q The magnet is resting on an infinite plate that has an infinite permeability - (Fig. 3.47a). SOLUTION 3.7.4 We replace the plate by an image magnet as shown in Fig. 3.47b. The image magnet is identical to the source magnet, and from symmetry we know that the field due to the two magnets will satisfy the boundary
190
CHAPTER 3 Field Analysis
Cylindrical magnet above a ferromagnetic plate: (a) physical system; and (b) magnet and image.
FIGURE 3.46
conditions (3.301) at the interface. The resulting magnet structure is equivalent to a single magnet of length 2L as shown in Fig. 3.47c. The field for this structure is given by Eq. (3.114) with L replaced by 2L,
(z ; 2L)(a ; b ; (z ; 2L) M B (z) : Q tan\ X ab 9tan\
z(a ; b ; z ab
.
(3.305) )
3.8 FINITE ELEMENT ANALYSIS Finite element analysis (FEA) is currently the most popular numerical method for the solution of electromagnetic field problems. There are two reasons for this. First, FEA is a powerful method that is readily adapted to a wide variety of applications. Second, there are numerous commercially available FEA software packages that enable a user to solve complex problems without having to develop his or her own application
3.8 FINITE ELEMENT ANALYSIS
191
Rectangular magnet on a ferromagnetic plate: (a) physical system; (b) magnet and image; and (c) equivalent system.
FIGURE 3.47
specific algorithms. In this section, we briefly review the FEA method, and present simple examples to demonstrate its use. More detailed presentations of electromagnetic FEA can be found in numerous excellent texts [10—15]. The finite element method entails the following steps: 1. Divide the solution region into a finite number of subregions called elements, and label the nodes that define each element; 2. select an approximating function for the field solution in each element (usually a polynomial);
192
CHAPTER 3 Field Analysis
3. express the solution in each element as a function of the nodal values and the spatial variables in that element; 4. define an energy functional for the field equation and evaluate this functional over each element. This gives an expression for the energy in each element in terms of its nodal values; 5. assemble a global energy expression as a sum of the energy of the individual elements. Reduce this expression to a function of the unknown nodal values eliminating redundant and prescribed nodal values; 6. minimize the global energy expression with respect to the unknown nodal values. This results in a system of algebraic equations for the nodal values; 7. solve the system of equations of Step 6 to determine the nodal values; and 8. reconstruct the desired solution from the nodal values using the expressions of Step 3. We demonstrate this procedure by solving a 2D magnetostatic field problem consisting of a rectangular conductor in free space (Fig. 3.48a). We solve the problem using a vector potential formulation (Section 3.2.1). The governing field equation is Eq. (3.22), which in two dimensions reduces to A (x, y) : 9 J (x, y). X X
(3.306)
Because A (x, y) and J (x, y) are scalars, we drop the subscript z for the X X remainder of this section, that is, A (x, y) ; A(x, y) and J (x, y) ; J(x, y). X X Step 1: Divide the solution region into triangular elements and label the vertex nodes of each element (Figs. 3.48b and 3.49). Step 2: Select an approximating function for the desired solution in each element. We assume that the potential is a linear function of x and y. Specifically, let AG(x, y) : a ; bx ; cy
(3.307)
represent the potential in the ith element. Step 3: Express AG(x, y) in terms of the spatial variables in the ith element and the corresponding nodal values AG (k : 1, 2, 3). Let (x , y ), I (x , y ), and (x , y ) denote the coordinates of the nodes for the ith
3.8 FINITE ELEMENT ANALYSIS
FIGURE 3.48
193
Conductor in free space: (a) rectangular conductor; and (b) FEA
mesh.
element. The potentials at these nodes can be written in the following matrix form: AG 1 x y a AG : 1 x y b . (3.308) AG 1 x y c An expression for the coefficients a, b, and c can be obtained by inverting
194
CHAPTER 3 Field Analysis
FIGURE 3.49
Triangular element with labeled nodes.
Eq. (3.308) as follows: 1 x y \ AG b : 1 x y AG . (3.309) c 1 x y AG We express AG(x, y) inside the ith element in terms of its nodal values by substituting Eq. (3.309) into Eq. (3.307). This gives a
1 x y \ AG AG(x, y) : [1 x y] 1 x y AG , 1 x y AG or AG(x, y) : AGG(x, y), I I I where 1 [(x y 9 x y ) ; (y 9 y )x ; (x 9 x )y], G(x, y) : 2G 1 G(x, y) : [(x y 9 x y ) ; (y 9 y )x ; (x 9 x )y], 2G and 1 G(x, y) : [(x y 9 x y ) ; (y 9 y )x ; (x 9 x )y], 2G
(3.310)
195
3.8 FINITE ELEMENT ANALYSIS
Here, G is the area of the ith element
1 x y 1 G : 1 x y . 2 1 x y The G are called element shape functions. It is easy to verify that I 0 k"j G(x , y ) : I H H 1 k:j
and that G(x, y) : 1. I I We represent the source J(x, y) in a similar fashion, that is, JG(x, y) : JGG(x, y), I I I where JG (k : 1, 2, 3) are known values of the source at the nodal points. I Step 4: Construct an energy functional compatible with Eq. (3.306) and evaluate it for a single element. In this case we use the energyrelated functional W(A) :
1 2
A ds 9
AJ ds.
(3.311)
We evaluate Eq. (3.311) over the ith element and obtain WG(AG) :
1 2
AG ds 9
AGJG ds.
(3.312)
Integration over ith
This can be written as
1 AG G · G ds AG I I J J 2 I J 9 AG G · G ds JG. I I J J I J We rewrite Eq. (3.313) in the following matrix form: WG(AG) :
WG(AG) : AG2KGAG 9 AG2PGJG,
(3.313)
196
CHAPTER 3 Field Analysis
where AG AG : AG AG AG2 : [AG AG AG] and KG : IJ
· ds I J
Integration over ith element
and PG : IJ
G · G ds. I J
Integration over ith element
Step 5: Next, assemble the energy of the elements, and form a global energy expression: , W(A) : WG(AG, AG, AG), (3.314) G where N is the total number of elements. The sum needs to be modified because some of the nodal values are redundant or prespecified. Specifically, nodal values that are common to more than one element are redundant, and nodal values on the boundary are prespecified. Let N denote the number of unknown nodal values and rewrite Eq. (3.314) in terms of these unknowns: W : W(A , A , . . . , A ). (3.315) , Step 6: Minimize Eq. (3.315) with respect to the unknown nodal values A : I W(A , A , . . . , A ) , :0 (n : 1, . . . , N). A L This leads to a system of equations for the unknown A : G , C A : 0, (3.316) GI I I
3.8 FINITE ELEMENT ANALYSIS
FIGURE 3.50
197
Flux plot for the rectangular conductor.
where the C are known. The N unknown nodal values are determined GI by solving Eq. (3.316). Step 7: Once the nodal values have been determined the solution is obtained using Eq. (3.310). A typical FEA output is shown in Fig. 3.50. This is a flux plot for the rectangular conductor problem. The lines represent equipotential contours of A (x, y). These are also known as flux X lines (Section 3.2.1). We now present a FEA of a permanent magnet structure. Consider the cylindrical (axisymmetric) structure shown in Fig. 3.51. This is an axially polarized permanent magnet shell with an attached flux plate. Such magnets are used for various applications including the focusing of charged particle beams [16], and have a field distribution in their interior similar to that of a solenoid.
FIGURE 3.51
Axially polarized cylindrical shell with attached flux plate.
198
CHAPTER 3 Field Analysis
Cross section of magnet structure: (a) right-hand side; and (b) finite element mesh.
FIGURE 3.52
The shell has inner and outer radii R and R , respectively, and is centered about the z-axis. We assume that the flux plate has a linear constitutive relation B : H, where . Because the structure is axisymmetric, it suffices to study the cross-sectional element shown in Fig. 3.52a. The full geometry is obtained by rotating this element 360° around the z-axis. The mesh for the element is shown in Fig. 3.52b. Notice that the mesh is refined in the interior of the magnet in order to render a more accurate field solution in this region. In general, a denser mesh gives a higher degree of accuracy. Figure 3.53 is a flux plot for the magnet. The full 3D flux distribution is obtained by rotating the crosssectional plot 360° around the z-axis. Finally, we evaluate the axial field
199
3.8 FINITE ELEMENT ANALYSIS
FIGURE 3.53
Flux plot.
component B along the interior of the magnet at two different radii r : 0 X and r : R /2 (Fig. 3.54). Notice that B is highest in the center of the X magnet as anticipated. At this point, we list several commercially available software packages for electromagnetic analysis. Some of these utilize the finite element method, while others utilize the boundary element method.
FIGURE 3.54
Component B vs z for two different radial values. X
200
CHAPTER 3 Field Analysis
1. Maxwell — Ansoft Corporation, Pittsburgh, Pennsylvania, USA. 2. OPERA — Vector Fields Inc., Aurora, Illinois, USA, or Vector Fields Ltd., Oxford, UK. 3. MagNet — Infolytica Corporation of Canada, or Infolytica Ltd. in the UK. 4. AMPERES — Integrated Engineering Software Corporation, Winnipeg, Manitoba, Canada. 5. Flux2D/3D — MAGSOFT Corporation, Troy, NY, USA. 6. ANSYS — Swanson Analysis Systems Inc., Houston, PA, USA. 7. COSMOS/M — Structural Research and Analysis Corporation, Los Angeles, CA, USA.
3.9 FINITE DIFFERENCE METHOD The finite difference method is one of the oldest and most developed numerical methods for the solution of partial differential equations. In this method, the solution region is discretized into a finite number of nodes, and the differential equation is approximated by a finite difference equation that relates the value of the solution at a given node to its values at the neighboring nodes. This discretization transforms the original boundary-value problem into a system of algebraic equations that can be solved using standard methods. In this section we briefly review the finite difference method. A comprehensive treatment of this method can be found in numerous texts [15, 17, 18]. The finite difference method entails the following steps: 1. Assign a grid of nodes (mesh) to the solution region and its boundary; 2. approximate the differential equation by a finite difference equation that relates the value of the solution at a given node to its values at neighboring nodes; 3. construct an algebraic system of equations for the nodal values taking into account the prescribed boundary values and/or initial conditions; and 4. solve the system of equations of Step 3, and obtain the nodal values of the solution. We demonstrate the finite difference method using Poisson’s equation in two dimensions: (x, y) : (x, y).
(3.317)
201
3.9 FINITE DIFFERENCE METHOD
FIGURE 3.55
Finite difference grid for a rectangular region.
This equation occurs frequently in magnetostatic field theory. In particular, it is the governing equation for the vector potential for 2D problems: A (x, y) : J (x, y), X X (Section 3.2.1). It is also the governing equation for the scalar potential in the magnetic charge model: (x, y) : 9 (x, y), K K (Section 3.4). Consider Eq. (3.317) in a rectangular solution region R in the x-y plane defined by (0 x L) and (0 y H). As a first step, discretize R. Choose discretization parameters x and y, and then assign a grid of nodal points for R and its boundary S. In this case, S consists of the four line segments (x:0, 0y H), (x:L, 0yH), (0xL, y:0), and (0 x L, y : H). Label the nodal points with indices (m, n) as shown in Fig. 3.55, where x : mx
(m : 0, 1, . . . , M),
y : ny
(n : 0, 1, . . . , N).
and
Notice that (x < x, y < y) : (m < 1)x, (n < 1)y), and so on. Thus,
202
CHAPTER 3 Field Analysis
coordinate points are identified with indices (x, y) ; (m, n) and (x < x, y < y) ; (m < 1, n < 1), and so on. It follows that functional values are also indexed (x, y); , KL and (x < x, y < y) ; , and so on. K!L! Next, we develop a finite difference equation for Eq. (3.317). To this end, consider the Taylor series expansion in x for the bivariate function (x, y): (x < x, y) : (x, y) < <
(x, y) 1 (x, y) x ; x x 2! x
1 (x, y) 1 (x, y) x ; x ; · · · . (3.318) 3! x 4! x
We can obtain three different approximations for (x, y)/x from Eq. (3.318). These follow from the ;x, 9x and the difference of the ;x and 9x expansions, respectively, that is, (x, y) (x ; x, y) 9 (x, y) : ; O(x) x x
(forward difference),
(x, y) (x 9 x, y) 9 (x, y) : ; O(x) x x
(backward difference),
and (x, y) (x ; x, y) 9 (x 9 x, y) : ; O(x) x 2x
(central difference),
where O(x) means that the error of the approximation is of the order of x, and so on. Clearly, the central difference approximation is the most accurate. We obtain similar expressions for (x, y)/y: (x, y) (x, y ; y) 9 (x, y 9 y) : ; O(y) x 2y
(central difference),
Once we have assigned a nodal grid to R, we can express the derivatives in terms of the nodal values, for example: (x, y) 9 K\L : K>L x 2x
(central difference).
Second-order derivatives are obtained in a similar way. For example, an expression for (x, y)/x is obtained from Eq. (3.318) by adding the
3.9 FINITE DIFFERENCE METHOD
203
;x and 9x expansions. This gives the central difference approximation: (x, y) (x ; x, y) 9 2(x, y) ; (x 9 x, y) . x x Using the second-order expressions we discretize Eq. (3.317) and obtain 9 2 ; 9 2 ; K>L KL K\L ; KL> KL KL\ : . (3.319) KL x y This can be further simplified by choosing x : y : h, which yields
K>L
; ; ; 9 4 : h . K\L KL> KL\ KL KL
(3.320)
Equation (3.320) represents a system of linear equations. To solve this system, we need to represent the unknown nodal values in the form of a column matrix (vector). To this end, we introduce a new index p Y m(N ; 1) ; n for m : 0, 1, . . . , M, and n : 0, 1, . . . , N. Using this index, Eq. (3.320) becomes ; ; ; 9 4 : h . N>,> N\,> N> N\ N N
(3.321)
This expression is only valid at the interior points m : 1, 2, . . . , M 9 1, and n : 1, 2, . . . , N 9 1, with ; and ; . It is K,>>L KL K,>>L KL instructive to examine the vector : ' , ,> ,> [] : ' ,>, ,> ' ' +,>,
(3.322)
204
CHAPTER 3 Field Analysis
In Eq. (3.322), the boundary values are identified by m : 0 $ ( , , . . . , ) (x : 0), , m : M $ ( , , . . . , ) (x : L), +,>> +,>> +,>>, n : 0 $ ( , , , . . . , ) (y : 0), +,> ,> ,> and n : N $ ( , , , . . . , ) (y : H). +,>>, , ,>>, ,>>, As a final step, assemble Eq. (3.321) into matrix form: (3.323) [A] · [ ] : [(], where [A] is a sparse matrix (many zero elements), [ ] is the vector of unknown nodal values at the interior points, and [(] is a vector containing the nodal source values and the known boundary condiN tions. There are many approaches to solving Eq. (3.323), and the most popular ones can be broadly classified as iterative methods and direct matrix methods. The iterative methods involve making an initial guess [ ], and then solving iteratively for successively more accurate solutions [ ]I. The direct matrix methods entail the direct solution of Eq. (3.323). For example, if [A] can be inverted, the solution is given by the following: [ ] : [A]\[(]. A detailed description of the various solution methods is beyond the scope of this book. However, they are described in detail in several texts [17, 18]. References 1. Moon, F. C. (1984). Magneto-solid Mechanics, New York: Wiley. 2. Lowther, D. A. and Silvester, P. P. (1986). Computer-aided Design in Magnetics, New York: Springer-Verlag. 3. Woodson, H. H. and Melcher, J. R. (1985). Electromechanical Dynamics Part II: Fields, Forces and Motion, Malabar, FL: John E. Krieger Publishing Co. 4. Johnk, C. T. A. (1988). Engineering Electromagnetic Fields and Waves, 2nd ed., New York: John Wiley and Sons. 5. Morrish, A. H. (1983). The Physical Principles of Magnetism, Malabar, FL: R. E. Krieger Publishing Co. 6. Craik, D. (1995). Magnetism: Principles and Applications, New York: John Wiley and Sons. 7. Reitz, J. R., Milford, F. J., and Christy, R. W. (1980). Foundations of Electromagnetic Theory, 3rd ed., Reading, MA: Addison-Wesley.
3.9 FINITE DIFFERENCE METHOD
205
8. McCaig, M. and Clegg, A. G. (1987). Permanent Magnets in Theory and Practice, 2nd ed., New York: John Wiley and Sons. 9. Zahn, M. (1987). Electromagnetic Field Theory: A Problem Solving Approach, Malabar, FL: R. E. Krieger Publishing Co. 10. Silvester, P. P. and Ferrari, R. L. (1990). Finite Elements for Electrical Engineers, 2nd ed., New York: Cambridge Univ. Press. 11. Sabonnadiere, J. C. and Coulomb, J. L. (1987). Finite Element Methods in CAD: Electrical and Magnetic Fields, New York: Springer-Verlag. 12. Hoole, S. R. H. (1989). Computer-aided Analysis and Design of Electromagnetic Devices, New York: Elsevier Science Publishing Co. 13. Salon, S. J. (1995). Finite Element Analysis of Electrical Machines, Boston: Kluwer Academic Publishers. 14. Sykulski, J. K. (ed.) (1995). Computational Magnetics, New York: Chapman & Hall. 15. Sadiku, M. N. O. (1992). Numerical Techniques in Electromagnetics, Boca Raton, FL: CRC Press. 16. Furlani, E. P. (1994). Analysis of a sintered NdFeB magnetic lens, J. Magn. Magn. Mat. 134: 117. 17. Smith, G. D. (1985). Numerical Solution of Partial Differential Equations: Finite Difference Methods, 3rd ed., Oxford Applied Mathematics and Computing Science Series, New York: Oxford University Press. 18. Ames, W. F. (1992). Numerical Methods for Partial Differential Equations, 3rd ed. New York: Academic Press.
a
This Page Intentionally Left Blank
CHAPTER
4
Permanent Magnet Applications
4.1 INTRODUCTION Permanent magnets in one form or another can be found in a wide variety of equipment ranging from consumer products, to industrial machinery, to research apparatus. In particular, they are used in audio/ video equipment, personal computers, printers, copiers, automobiles, household appliances, power tools, industrial motors and generators, and various biomedical apparatus. Permanent magnets have several advantages over conventional (current-driven) electromagnets. The fundamental advantage is that they can provide a relatively strong magnetic field over an extended spatial region for an indefinite period of time with no expenditure of energy. Of course, the field they provide is fixed whereas the field of an electromagnet can be changed by adjusting the current. Another advantage of permanent magnets is that they can be fabricated with a wide range of structural properties, geometric shapes, and magnetization patterns. They are also relatively inexpensive on a per unit basis depending on the material used. Permanent magnets have an additional advantage over electromagnets in that their performance scales well with size. Specifically, if we change all the linear dimensions L of an electromagnet, while keeping the field strength at all the rescaled observation points fixed, the current density must be adjusted by a factor of 1/L (Section 5.16). Therefore, as the size of an electromagnet is reduced, a point is reached where overheating precludes viability. Moreover, if the dimensions of the conductors are increased to compensate for this, the field strength at the
207
208
CHAPTER 4 Permanent Magnet Applications
rescaled observation points will decrease. By comparison, if we change all the linear dimensions of a permanent magnet, the field strength at all the rescaled observation points remains constant (assuming the magnetization is constant). Thus, there will be a dimension below which a permanent magnet will be the only viable field source. In this chapter we discuss various permanent magnet applications. We consider applications in which a magnet is used alone as a passive field source or in conjunction with other magnets as part of a functional device. Specifically, we discuss bias magnet structures, high field structures, latching magnets, magnetic suspensions, and magnetic gears and couplings. We also discuss various miscellaneous applications that utilize magnets including MRI, magnetooptical recording, electrophotography and free-electron lasers. Throughout the chapter, we present solved examples of various applications. Many of these contain formulas that are useful for design and optimization.
4.2 MAGNET STRUCTURES Bias magnets are used to provide a prescribed field distribution over a given region. Such magnets are usually either rectangular or cylindrical in shape and have bipolar or multipole magnetization patterns. In this section we study the most common bias magnet structures. We derive field solutions for these structures and demonstrate the solutions with practical calculations.
4.2.1 Rectangular structures The most common permanent magnet structure is the rectangular bar magnet. We developed a limited field solution for this structure in Section 3.4, where we used the charge model to obtain the B-field above the magnet along a line of symmetry (Example 3.4.1). In this section, we derive general field solutions for two different rectangular structures, an infinitely long magnet and a finite bar magnet. The field solutions are derived in the following examples. EXAMPLE 4.2.1 Determine the B-field outside a long rectangular magnet (Fig. 4.1). Assume that the magnet has a magnetization M : M y . Q
209
4.2 MAGNET STRUCTURES
Rectangular magnet.
FIGURE 4.1
SOLUTION 4.2.1 Assume that the magnet is infinitely long. In this case the analysis reduces to a 2D field problem. Choose a reference frame as shown in Fig. 4.2a. We use the current model to determine the field (Section 3.3). First, we reduce the magnet to an equivalent current distribution. We apply Eq. (3.95) and find that the volume current density is zero, J : ;M : 0. To determine K the surface current density j we need to identify the unit surface normals. K These are n : and n :
y (y : h) 9y (y : 9h)
x (x : w) 9x (x : 9w).
As j : M ; n and M : M y , we find that K Q 9M z (x : w) Q j : K M z (x : 9w). Q Thus, the magnet reduces to two parallel infinitely long current sheets of height 2h as shown in Fig. 4.2b. The field solution for a current sheet was derived in Example 3.2.5. We repeat it here for convenience,
K x ; (y 9 h) 2hx B (x, y) : ln x ; 2 tan\ y . 4 x ; (y ; h) x ; y 9 h (4.1) Here, K is the surface current density (A/m). The solution (4.1) applies to a current sheet at x : 0. However, for the magnet there are two current sheets,
210
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.2 Infinite rectangular magnet: (a) cross-sectional view with reference frame; and (b) equivalent surface currents.
one at x : 9w, and another at x : w. We obtain the field of the magnet by taking a superposition, B(x, y) : B (x ; w, y) 9 B (x 9 w, y).
(4.2)
Notice that the minus sign in the second term reflects the fact that the current at x : w is flowing in the negative z-direction (into the page as shown in
211
4.2 MAGNET STRUCTURES
Fig. 4.2b). We evaluate Eq. (4.2) and obtain the field components,
M (x ; w) ; (y 9 h) (x 9 w) ; (y 9 h) B (x, y) : Q ln 9 ln V 4 (x ; w) ; (y ; h) (x 9 w) ; (y ; h)
(4.3)
and
M 2h(x;w) 2h(x9w) B (x, y) : Q tan\ 9 tan\ W 2 (x;w);y9h (x9w);y9h
(4.4) where we have substituted K : M . Equations (4.3) and (4.4) give the B-field Q at any point outside the magnet. Calculations: We apply Eqs. (4.3) and (4.4) to a magnet with w : 10 mm, h : 5 mm, and M : 8.0; 10 A/m. This value of magnetization corresponds Q to sintered NdFeB with a remanence of B : 10,054 G. We evaluate B and B P V W along two different horizontal lines that span 92w x 2w. These are located 0.5 and 2.0 mm above the magnet, respectively (Fig. 4.3). Notice that closer to the magnet (0.5 mm above the surface), the vertical component B is higher near W the ends of the magnet than at its center (x : 0). However, farther away the opposite is true. The reason for this is as follows: Near the magnet, the field is predominantly vertical across the entire surface and has a greater magnitude closer to the current sheets, which are at the ends of the magnet. However, as you move away from the magnet, the vertical projection of the field decreases dramatically near the ends of the magnet and this results in a peak value in the center. Finally, notice that the horizontal component B peaks above the current V sheets (x : <w) as expected. ) EXAMPLE 4.2.2 Determine the B-field outside a rectangular bar magnet (Fig. 4.4a). Assume that the magnetization is M : M z . (4.5) Q SOLUTION 4.2.2 Choose a coordinate system centered with respect to the magnet (Figs. 4.4b and 4.5). Let (x , x ), (y , y ) and (z , z ) denote the positions of the edges of the magnet with respect to the x-, y-, and z-axes. We use the charge model to determine the field (Section 3.4). First, reduce the magnet to an equivalent charge distribution. From Eqs. (3.105) and (4.5) we find that the volume charge density is zero, : 9 · M : 0. However, there K is a surface charge density
M (z : z ) Q : K 9M (z : z ). Q
(4.6)
212
CHAPTER 4 Permanent Magnet Applications
The components B and B vs x: (a) y : h ; 0.5 mm; and (b) V W y : h ; 2 mm.
FIGURE 4.3
The B-field follows from Eq. (3.107), M B(x, y, z) : Q (91)I 4 I W V [(x 9 x)x ; (y 9 y)y ; (z 9 z )z ] dx dy I ; . [(x 9 x) ; (y 9 y) ; (z 9 z )] W V I
(4.7)
213
4.2 MAGNET STRUCTURES
FIGURE 4.4
Rectangular magnet: (a) geometry and polarization; and (b) refer-
ence frame.
The x-component: B follows from Eq. (4.7), V M W V (x 9 x) dx dy B (x, y, z) : Q (91)I . V 4 [(x 9 x) ; (y 9 y) ; (z 9 z )] I W V I Integration with respect to x gives
M W dy B (x, y, z) : Q (91)I>K . V 4 [(x9x );(y9y);(z9z )] I K W K I
214
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.5 Cross-sectional views of a bar magnet with reference frame: (a) x-y plane; and (b) x-z plane.
The remaining y integration can be evaluated by making a change of variable to : y 9 y. The resulting field expression is M B (x, y, z) : Q (91)I>K ln[F(x, y, z, x , y , y , z )], V K I 4 I K
(4.8)
where (y 9 y ) ; [(x 9 x ) ; (y 9 y ) ; (z 9 z )] K I F(x, y, z, x , y , y , z ) : . K I (y 9 y ) ; [(x 9 x ) ; (y 9 y ) ; (z 9 z )] K I
215
4.2 MAGNET STRUCTURES
The y-component: B also follows from Eq. (4.7), W M W V (y 9 y) dx dy B (x, y, z) : Q (91)I . W 4 [(x 9 x) ; (y 9 y) ; (z 9 z )] I W V I Integration with respect to y yields
M V dx B (x, y, z) : Q (91)I>K W 4 [(x 9 x) ; (y 9 y ) ; (z 9 z )] K I I K V . The remaining x integration is evaluated using a change of variable : x 9 x. The resulting field expression is M B (x, y, z) : Q (91)I>K ln[H(x, y, z, x , x , y , z )]. K I W 4 I K
(4.9)
where (x 9 x ) ; [(x 9 x ) ; (y 9 y ) ; (z 9 z )] K I . H(x, y, z, x , x , y , z ) : K I (x 9 x ) ; [(x 9 x ) ; (y 9 y ) ; (z 9 z )] K I The z-component: B is given by X M W V (z 9 z ) dx dy I B (x, y, z) : Q (91)I . X 4 [(x 9 x) ; (y 9 y) ; (z 9 z )] I W V I The x integration is performed using a change of variable : x 9 x,
M W V\V z9z I B (x, y, z) : 9 Q d dy X [ ; (y 9 y) ; (z 9 z )] 4 W V\V I M : Q (91)I>L>(z 9 z )(x 9 x ) I L 4 I L W dy ; , (4.10) W [(y 9 y) ; (z 9 zI)]((y 9 y) ; b
where b : (x 9 x ) ; (z 9 z ). The remaining y integration is performed L I using a change of variable : y 9 y. This gives M B (x, y, z) : Q (91)I>L>K X 4 I L K (x 9 x )(y 9 y ) L K g(x, y, z; x , y , z ) , (4.11) ;tan\ L K I (z 9 z ) I
216
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.6
Bar magnet.
where 1 g(x, y, z; x , y , z ) : . L K I [(x 9 x ) ; (y 9 y ) ; (z 9 z )] L K I The formulas (4.8), (4.9) and (4.11) give the field at any point outside the magnet. Calculations: We apply the theory to a magnet with a : 10 mm, b : 20 mm, and c : 10 mm. We set x : 0, x : a, y : 0, y : b, z : 9c, z : 0 mm, and M : 8.0;10 A/m (Fig. 4.6). We evaluate B across two different surfaces Q X above the magnet, one at z : 0.5 and the other at z : 2.0 mm. The computed data are shown in Fig. 4.7. Notice that closer to the magnet, B dips near the X center of the magnet, but farther away it peaks there. This can be understood from the current model. Specifically, the magnet can be reduced to equivalent current sheets along its edges. Near the magnet the field is predominantly vertical and has a greater magnitude closer to the current sheets, which are at the edges. However, as you move away from the magnet, the vertical projection of the field decreases dramatically near the ends of the magnet and this results in a peak value at the center. ) The field solution derived in Example 4.2.2 can be used to determine the field of multipole rectangular structures such as the quadrupole structure shown in Fig. 4.8. This structure consists of four block magnets and the field is given by B(x, y, z) : B (x, y, z), G G where B (x, y, z) is the field due to the ith block. G
4.2 MAGNET STRUCTURES
FIGURE 4.7
at z : 2 mm.
217
Magnetic field above a bar magnet: (a) B at z : 0.5 mm; and (b) B X X
218
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.8
Quadrupole magnet structure.
4.2.2 Cylindrical structures Cylindrical magnets are the second most common magnet structure. They are used in devices such as encoders, actuators, motors, electrophotographic copiers, and magnetic couplings and gears. Cylindrical magnets can have various magnetization patterns including bipolar with magnetization either through the cross section or along the axis, and multipole with radial or axial magnetization (Fig. 4.9). In Sections 3.3 and 3.6.2. we developed partial field solutions for the bipolar cylindrical magnets (Examples 3.3.1 and 3.6.3). In this section we consider the full range of magnetization patterns and derive general field solutions that
Cylindrical magnets: (a) bipolar magnetization; (b) multipole radial magnetization; and (c) multipole axial magnetization. FIGURE 4.9
219
4.2 MAGNET STRUCTURES
FIGURE 4.10
Bipolar cylinder: (a) geometry and magnetization; and (b) refer-
ence frame.
are valid at any point outside the magnet. The field solutions are derived in the following examples. EXAMPLE 4.2.3 Derive an expression for the B-field outside a bipolar cylindrical magnet of finite length (Fig. 4.10) [1]. Assume that the axis of the magnet is along the z-axis. Let R denote it radius, and let z and z denote the positions of its bottom and top edge, respectively. Furthermore, assume that the magnet has a magnetization M : M x . Q
220
CHAPTER 4 Permanent Magnet Applications
SOLUTION 4.2.3 We use the charge model of Section 3.4. First, we reduce the magnet to an equivalent charge distribution and then determine the field by treating the charge distribution as a field source. The equivalent charge distribution follows from Eq. (3.105). We find that the volume charge is density zero, : 9 · M : 9 · M x : 0, and that the surface charge density is K Q : M · n K : M x · r Q : M cos(). (4.12) Q The B-field is computed using
X L M cos() Q B(x) : 9 R d dz, 4
x 9 x X where acts on the unprimed variables. Radial component: The radial component is given by
M R X L B (r, , z) : Q cos()[r 9 R cos( 9 )] P 4 X ; g(r, , z; R, , z) d dz,
(4.13)
where g(r, , z; r, , z) :
1 (r ; r 9 2rr cos( 9 ) ; (z 9 z)
.
We evaluate the integration numerically using Simpson’s method and the z integration in closed form. We obtain M R , B (r, , z) : Q (91)I>S (n) cos((n)) P 2N L I ; [r 9 R cos( 9 (n))]I(r, , z; R, (n), z ), I where N is the Simpson’s method mesh number (even), (n) :
n 2 N
(n : 0, 1, 2, . . . , N),
(4.14)
221
4.2 MAGNET STRUCTURES
and
S (n) :
(n : 0) (n : 1, 3, 5, . . .) (n : 2, 4, 6 . . .) (n : N).
Function I in Eq. (4.14) is defined by
X
g(r, , z; r, , z) dz : (91)I>I(r, , z; r, , z ). I X I Evaluation of Eq. (4.15) gives
IY
where
(4.15)
F(r, , z; r, , z ) I if r ; r 9 2rr cos( 9 ) " 0 or 91 2(z 9 z ) I if
(4.16)
r : r, cos( 9 ) : 1, z " z I
(z 9 z )g(r, , z; r, , z ) I I . F(r, , z; r, , z ) : I r ; r 9 2rr cos( 9 ) Notice that the functional form of I depends on the relationship between the coordinates of the field point (r, , z) and the source point (r, , z). Specifically, I is given by F as long as r ; r 9 2rr cos( 9 ) " 0. If this condition is violated, then the lower ratio is used. Azimuthal component: The azimuthal field component is given by
M R X L B (r, , z) : Q cos() 4 X ; sin( 9 )g(r, , z; R, , z) d dz.
(4.17)
As in the radial case, the integration is evaluated using Simpson’s method, and the integration in z is expressed in terms of Eq. (4.16). The resulting
222
CHAPTER 4 Permanent Magnet Applications
expression for B is M R , B (r, , z) : Q (91)I>S (n) cos((n)) 2N L I ; sin( 9 (n))I(r, , z; R, (n), z ). (4.18) I Axial component: The derivation for the axial field component is similar to that for the previous two cases. Specifically,
M R X L cos()(z 9 z) B (r, , z) : Q X 4 X ; g(r, , z; R, , z) d dz.
(4.19)
Once again, the integration can be expressed as a discrete sum, M R , B (r, , z) : Q S (n) cos((n)) X 2N L X ; (z 9 z)g(r, , z; R, (n), z) dz. X The remaining integral can be evaluated analytically. This gives
(4.20)
M R , B (r, , z) : Q S (n) X 2N L ; (91)I cos((n))g(r, , z; R, (n), z ). (4.21) I I Calculations: We demonstrate the field solutions (4.14), (4.18) and (4.21) with some sample calculations. Consider a solid cylindrical magnet with R : 2.54 mm, L:50 mm (z :925 mm, z :25 mm), and M :4.3;10 A/m Q (B : 5400 G). We evaluate B and B at r : 3.8 mm for a series of angular P P values : 0, 10, 20, . . . , 180°. These data are shown in Fig. 4.11. Next, we analyze a cylindrical shell geometry with inner and outer radii R and R , respectively (Fig. 4.12). The hollow interior of the shell is taken into account using the principle of superposition. Specifically, we take the field due to a solid cylinder B (x), and subtract the field due to an oppositely magnetized inner 0 core B (x), 0 B(x) : B (x) 9 B (x). (4.22) 0 0 The cylindrical shell is symmetrically positioned with respect to the x-y plane
223
4.2 MAGNET STRUCTURES
FIGURE 4.11
The components B and B vs for a solid cylinder. P
(center at z : 0) with its top at z : 3 cm. The parameters used in the analysis are as follows: M : 4.3 ; 10 A/m Q R : 2.0 cm
(inner radius)
R : 4.0 cm
(outer radius)
L : 6.0 cm
(height).
The components B and B are computed at z:0 and 2.7 cm at a radius r:5 cm P for a series of angular values : 0, 10, 20, . . . , 180°. These data are shown in Fig. 4.13. Notice that the field decreases in magnitude as the observation point moves away from the center (z : 0 cm) towards the end of the cylinder. Finally, the field components B and B are evaluated at r : 5 cm and : 0° for a series P X of points z : 0, 0.25, 0.5, . . . , 5 cm. These data are shown in Fig. 4.14. Notice the behavior of the fields as the observation point passes over the end of the cylinder (z : 3 cm). )
224
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.12
Cylindrical shell geometry and reference frame.
EXAMPLE 4.2.4 Derive an expression for the field outside an infinitely long cylindrical magnet with a linear second quadrant demagnetization relation of the form B : H ; M (r, ). (4.23) Q The magnetization M (r, ) is well behaved but otherwise arbitrary [2]. Let R Q and R denote the inner and outer radii of the cylinder, respectively (Fig. 4.15). SOLUTION 4.2.4 Since the magnet is infinitely long the problem reduces to a 2D boundary value problem (BVP) in cylindrical coordinates. Boundary value problem: The magnetic fields H inside the magnet, and H and H in the regions V and V outside the magnet can be represented
4.2 MAGNET STRUCTURES
FIGURE 4.13
The components B and B vs for a cylindrical shell. P
FIGURE 4.14
The components B and B vs z for a cylindrical shell. P X
225
226
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.15
Cross section of an infinite cylinder with arbitrary magnetization.
in terms of scalar potentials , and where K K K H : 9 HI : 9I (k : 1, 2). (4.24) K K The BVP for and I follows from: (a) the constitutive relation B : K K H ; M (r, ); (b) the field equations Q · B : 0 · BI : 0 (k : 1, 2), (4.25) (c) the boundary conditions for B and H at the media interfaces; and (d) the constitutive relations BI : HI (k : 1, 2), (4.26) which apply outside the magnet. Specifically, inside the magnet we solve : 9 , K K where is an equivalent volume charge density, K : 9 · M (r, ). K Q
(4.27)
(4.28)
Outside the magnet we solve I : 0 K The boundary conditions are : I K K
at
(k : 1, 2). r:R I
(4.29)
(k : 1, 2),
(4.30)
and where
I K 9 K : I n n
at
r:R I
(k : 1, 2),
(4.31)
227
4.2 MAGNET STRUCTURES
I : nI · M (r, ). (4.32) K Q Here, nI : (91)Ir are the outward surface normals at r : R . We also I require that (0) -, K
(4.33)
and (-) -, (4.34) K which restrict the form of the solutions in regions V and V , respectively. Separation of variables: We employ the method of separation of variables to solve the BVP (4.27)—(4.34). Accordingly, we represent the solutions to Eqs. (4.27) and (4.29) as follows: (r, ) : [V (r) sin(i) ; U (r) cos(i)], K G G G (r, ) : rG[V sin(i) ; U cos(i)], K G G G
(4.35) (4.36)
and (r, ) : r\G[V sin(i) ; U cos(i)]. (4.37) K G G G The volume charge density and surface term can also be expressed as Fourier series, (r, ) : [C (r) sin(i) ; D (r) cos(i)], K G G G
(4.38)
and I() : [AI sin(i) ; BI cos(i)], (4.39) K G G G where C (r), D (r), AI and BI are determined once the magnetization M (r, ) G G G G Q is specified. Substituting Eqs. (4.35) and (4.38) into Eq. (4.27) yields the following ordinary differential equations for the unknown coefficients V (r) and G U (r), G dV 1 dV i G ; G 9 V : 9C (r), (4.40) G dr r dr r G
228
CHAPTER 4 Permanent Magnet Applications
and dU 1 dU i G ; G 9 U : 9D (r). G dr r dr r G
(4.41)
These equations are of the form df(r) 1 df(r) i ; 9 f(r) : g(r), r dr r dr with homogeneous solutions f (r) : rG and f (r) : r\G. Solution: We use the method of variation of parameters and obtain general solutions for Eqs. (4.40) and (4.41) of the form
f ())g()) P f ())g()) d) ; f (r) d), W(f , f )()) W(f , f )()) 0 0 where c and c are constants and W(f , f ) : f f 9 f f is the Wronskian. Application of this method to Eqs. (4.40) and (4.41) yields f(r) : c f (r) ; c f (r) 9 f (r)
P
1 P ) V (r) : E rG ; F r\G 9 G G G 2i 0
r G ) G 9 C ()) d) G ) r
(4.42)
and
1 P r G ) G U (r) : P rG ; Q r\G 9 ) 9 D ()) d). (4.43) G G G G 2i ) r 0 The unknown coefficients VI, UI (k : 1, 2), E , F , P , and Q are determined G G G G G G by imposing the boundary conditions (4.30) and (4.31). For example, the following group of simultaneous equations is obtained for the coefficients UI, G P , and Q : G G i UiR\G> ; P iRG\ 9 Q iR\G> 9 S : B, (4.44) G G G G R G UR\G : P RG ; Q R\G 9 S , (4.45) G G G G and URG : P RG ; Q R\G, G G G
(4.46)
UiRG\ 9 P iRG\ ; Q iR\G> : B, G G G G
(4.47)
and
229
4.2 MAGNET STRUCTURES
where
1 0 S : ) G 2i 0
R G ) G ; D ()) d), G ) R
and
1 0 R G ) G 9 S: ) D ()) d). G 2i G ) R 0 Equations (4.44)—(4.47) can be written in matrix form: 0 1
RG
9
0 1
9RG
91
1 0
91
9R\G R\G
1 0
9
U G U G : P G Q G
S B G; G RG> i R 9RG S G 0
.
(4.48)
B R\G> G i
A similar set of equations is obtained for the coefficients VI, E , and F . The G G G method for solving the BVP can now be summarized: Given M (r, ) inside the Q magnet, evaluate Eqs. (4.28) and (4.32) and determine the coefficients C (r), G D (r), and AI, BI (k : 1, 2) in Eqs. (4.38) and (4.39), respectively. Next, G G G substitute D (r) and BI into Eqs. (4.44)—(4.47), and C (r) and AI into the G G G G corresponding equations for VI, E , and F and obtain two sets of decoupled G G G (four by four) simultaneous equations for the unknown coefficients UI, P , Q , G G G and VI, E , and F . Finally, solve the equations for VI and UI and substitute G G G G G these into Eqs. (4.36) and (4.37) to obtain the desired field solution. Calculations: The analysis outlined above is applied to a cylinder with the magnetization as shown in Fig. 4.16. Such magnets are used in the electrophotographic process (Section 4.9) [2]. The magnetization is
9M r Q M r Q M : Q 9M r Q 0
(9 9 ) (9 ) ( ) ( 2 9 ). This can be expressed analytically as follows: M : M()r , Q
(4.49)
(4.50)
230
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.16
Magnetized cylinder and reference frame.
where
and
M() : M cos(i), G G
(4.51)
2M Q (2 sin(i ) 9 sin(i )). M: G i
(4.52)
A solution is sought for the region V outside the cylinder. Therefore, the coefficients of interest are V and U. From the symmetry of the problem, the G G coefficients V : 0 for all i. The coefficients U are determined from Eq. G G (4.48), which gives H(M , R , R , ) G U(M , R , R , ) : G G G G(R , R , ) G
(i : 1, 2, 3, . . .),
(4.53)
where
1 ; 1 K(M , R , R ) H(M , R , R , ) : RG> G G G R M R G ; G 19 9 1; i R M R G> G ;2 ; h* (M , R , R , ) , G G i R
(4.54)
231
4.2 MAGNET STRUCTURES
with
M R G G K(M , R , R ) : R 9 R , G i(i ; 1) R
(4.55)
and with
M 1 R G G 91 h* (M , R , R , ) : G G i R R 1 R G R 9 R 19i R ; R R ln R
and
R G G(R , R , ) : G R
(i 1) (4.56) (i : 1)
92 ;1 9
;2 ;1 . (4.57)
The field components follow from B : 9 . Specifically, K B(r, ) : ir\G>U(M , R , R , ) cos(i), P G G G and B (r, ) : ir\G>U(M , R , R , ) sin(i). G G G The parameters used in the analysis are as follows:
(4.58)
(4.59)
M : 2.0 ; 10 A/m Q R : 8.0 mm R : 12.0 mm : 30° : 60°. (4.60) The value of M corresponds to injection-molded barium ferrite material. We Q compute the field components B(r, ) and B (r, ) at r : 16 mm and : 0, P 5, 10, 15, . . . , 180° with : . Figure 4.17 shows the predicted field data along with the corresponding data from FEA. We also compute B (r, ) at the P same observation points as in the preceding and with : 1.0 , 1.5 , and
232
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.17
The components B and B vs (; : FEA). P
2.0 . These data are shown in Fig. 4.18. Notice that lower values of give rise to stronger fields outside the cylinder. ) EXAMPLE 4.2.5 Determine the B-field outside an infinitely long cylindrical magnet that has an alternating radial polarization and a linear second quadrant demagnetization curve of the form B : H ; M (). Q
(4.61)
Let R and R denote the inner and outer radii of the cylinder, respectively (Fig. 4.19). SOLUTION 4.2.5 As the magnet is infinitely long, the problem reduces to a 2D boundary value problem (BVP) similar to that of Example 4.2.4. We use the solution method presented there with the magnetization inside the cylinder given by M () : M()r , Q
(4.62)
233
4.2 MAGNET STRUCTURES
FIGURE 4.18
FIGURE 4.19
Component B vs . P
Infinite cylinder with radial magnetization.
234
CHAPTER 4 Permanent Magnet Applications
where M() : M cos(i), G G
(4.63)
and
M , (2k 9 1)i (2k ; 1)i M : Q (91)I sin 9 sin . (4.64) G i N N I In Eq. (4.64), N is the number of poles. The expressions (4.63) and (4.64) give the Fourier decomposition of M(). First, we determine the representations for () and I () by evaluating the coefficients C (r), D (r), AI and BI in K G G K G G Eqs. (4.38) and (4.39). From Eq. (4.28) we obtain () : 9 · M () K Q M() :9 . r From the orthogonality of sin(i) and cos(i) we find that C (r) : 0, and that G M D (r) : 9 G . G r Similarly, from Eq. (4.32) we have I() : (91)Ir · M () K Q : (91)IM().
Therefore AI : 0 and G (k : 1, 2). BI : (91)IM G G We seek a solution for the region V outside the cylinder. The coefficients of interest are V and U. From the orthogonality of sin(i) and cos(i) we find G G that V : 0 for all i. The coefficients U are given by Eq. (4.53): G G H(M , R , R , ) G U(M , R , R , ) : G (i : 1, 2, 3, . . .), (4.65) G G G(R , R , ) G where H(M , R , R , ) and G(R , R , ) are defined in Eqs. (4.54)—(4.57) G G G with M given by Eq. (4.64). Once the U are known, we compute the field G G components B(r, ) : ir\G>U(M , R , R , ) cos(i) P G G G
(4.66)
4.2 MAGNET STRUCTURES
FIGURE 4.20
235
The components B and B vs (r : 12 mm). P
and B (4.67) (r, ) : ir\G>U (M , R , R , ) sin(i). G G G Calculations: We demonstrate the field solution (4.66)—(4.67) via the analysis of an eight-pole cylindrical magnet with R : 1 mm, R : 10 mm, M :4.3 Q ; 10 A/m, and : . We evaluate B and B at r : 12 mm for 0 P 180°. These data are shown in Fig. 4.20. Here, : 0° corresponds to the center of a north pole. Notice that B peaks at the center of the pole, whereas B P peaks at the transition between neighboring poles at : 22.5°. ) EXAMPLE 4.2.6 Determine the radial and azimuthal B-field components outside a finite, radially polarized multipole cylinder (Fig. 4.21) [3]. Assume that the magnet has a magnetization M : <M r , (4.68) Q where the < term takes into account the alternating polarity of adjacent poles.
236
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.21
Radially polarized multipole cylinder.
SOLUTION 4.2.6 The geometry is three-dimensional and, therefore, we need to solve a 3D field problem. We use the current model of Section 3.3. From Eqs. (3.95) and (4.68) we find that the volume current density is zero J : ; M : 0. K
(4.69)
Therefore, Eq. (3.93) reduces to A(x) : 4
1
j (x) K ds,
x 9 x
(4.70)
where j : M ; n. K
(4.71)
The B-field is computed from Eq. (4.70) using B : ; A.
(4.72)
Instead of determining B for the entire magnet, it is simpler to evaluate Eq. (4.70) first for a single sector shown in Fig. 4.22, determine B for the sector, and then compute the total field as a superposition of the contributions from all the sectors. As a first step, we determine j for the various surfaces of the sector. K From Eqs. (4.68) and (4.71) it follows that j : 0 on the back (r : R ) and K front (r : R ) of the sector because the magnetization and surface normals are
237
4.2 MAGNET STRUCTURES
FIGURE 4.22
Sector geometry for a radially polarized cylinder.
either parallel or antiparallel for these surfaces. There are four remaining surfaces to consider:
r (1) r r (2) Q Q top surface : (1) (2) Q Q z : z (2), Q
(4.73)
r (1) r r (2) Q Q bottom surface : (1) (2) Q Q z : z (1), Q
(4.74)
r (1) r r (2) Q Q left side : : (1) Q z (1) z z (2), Q Q
(4.75)
238
CHAPTER 4 Permanent Magnet Applications
and
r (1) r r (2) Q Q right side : : (2) Q z (1) z z (2). Q Q
(4.76)
The unit normals for these surfaces are
n :
z
(top surface)
9z
(bottom surface)
9
(left side)
(4.77)
(right side).
Therefore, assuming the radial magnetization given in Eq. (4.68), the corresponding surface current densities are
9M Q M Q j : K 9M z Q M z Q
(top surface) (bottom surface)
(4.78)
(left side) (right side).
Taking into account Eqs. (4.73)—(4.78), we write Eq. (4.70) as M A (x) : Q (91)H Q 4 H ;
XQ PQ
XQ PQ
z dr dz
x 9 x : ( j) 9
Q
PQ
PQ
Q
Q
x 9 x
r d dr . XYXQH
(4.79)
In this expression, and for the remainder of this example, the subscript s denotes the contributions due to a single sector. Notice that the unit vector in the integrand of Eq. (4.79) is itself a function of , : 9sin()x ; cos()y .
(4.80)
239
4.2 MAGNET STRUCTURES
Therefore, Eq. (4.79) can be written in terms of Cartesian coordinates, M A (x) : Q (91)H Q 4 H XQ PQ z ; dr dz
x 9 x : ( j) Q XQ PQ P Q sin( ) x ; Q r d dr
x 9 x XYXQH PQ Q P Q cos() y 9 Q r d dr . (4.81)
x 9 x XYXQH PQ Q This can be rewritten in terms of cylindrical coordinates by computing the following projections:
A (x) : A (x) · r , PQ Q A (x) : A (x) · , Q Q
(4.82) (4.83)
and A (x) : A (x) · z . (4.84) XQ Q We apply these projections to Eq. (4.81) and obtain the following explicit forms: M A (x) : Q (91)H> PQ 4 H PQ Q sin( 9 ) r d dr, ;
x 9 x PQ Q XYXQH M A (x) : Q (91)H> Q 4 H PQ Q cos( 9 ) ; r d dr,
x 9 x PQ Q XYXQH
and
(4.85)
(4.86)
M A (x) : Q (91)H XQ 4 H XQ P 1 dr dz. (4.87) ;
x 9 x XQ PQ Y QH Equations (4.85), (4.86), and (4.87) give A (x) for a single sector. For the sector, Q B (x) is obtained using Eq. (4.72). As the total field is a superposition of the Q
240
CHAPTER 4 Permanent Magnet Applications
fields due to all the sectors, it can be expressed as , B(x) : (91)Q>B (x), (4.88) Q Q where the (91)Q> term takes into account the alternating polarity of adjacent poles. For the derivation of B and B we use the function P 1 . (4.89) g(r, , z; r, , z) : [r ; r 9 2rr cos( 9 ) ; (z 9 z)] Radial component: The radial field component due to a single sector follows from Eq. (4.72): 1 B (x) : (4.90) A (x) 9 A (x). PQ r XQ z Q Substituting Eqs. (4.86) and (4.87) into Eq. (4.90) yields M B (x) : Q (91)H> PQ 4 H XQ PQ r sin( 9 ) dr dz ;
x 9 x XQ PQ Y QH P Q (z 9 z) cos( 9 ) r d dr . (4.91) ; Q
x 9 x XYXQH PQ Q The integrations in r in the first term and in the second term can be written in terms of discrete sums using various numerical methods. We use Simpson’s method. Specifically, let N and N denote the number of mesh points in the r P and variables, respectively, and let r(n) and (m) denote the values at which Q the integrands are evaluated. As r (1)rr (2), and (1) (2), we Q Q Q Q have r (2) 9 r (1) Q r : Q (4.92) Q N P and (2) 9 (1) Q . : Q (4.93) Q N
The integration points are as follows:
and
n r (n) : r (1) ; (r (2) 9 r (1)) Q Q Q N Q P m (m) : (1) ; ( (2) 9 (1)) Q Q Q N Q
(n : 0, 1, 2, . . . , N ), P
(4.94)
(m : 0, 1, 2, . . . , N). (4.95)
241
4.2 MAGNET STRUCTURES
Finally, define the Simpson integration coefficients S (n) and S (m), P (n : 0) (n : 1, 3, 5 . . .) S (n) : P (n : 2, 4, 6 . . .) (n : N ) P and
(m : 0) (m : 1, 3, 5 . . .) S (m) : (m : 2, 4, 6 . . .) (m : N ). Using this scheme, Eq. (4.91) reduces to
(4.96)
(4.97)
M B (x) : Q (91)H> PQ 4 H , ; r P S (n)r (n) sin( 9 (j)) Q Q P Q L X ; Q g(r, , z; r (n), (j), z) dz Q Q XQ , ; S (m)(z 9 z (j)) cos( 9 (m)) Q Q Q K P ; Q g(r, , z; r, (m), z (j))r dr . Q Q PQ The remaining integrals in z and r can be expressed in closed form. In addition, the contributions from all the sectors need to be taken into account in accordance with Eq. (4.88). The resulting field expression is
M , B (r, , z) : Q (91)Q>>H>I P 4 Q H I ,P ; r S (n)r (n) sin( 9 (j))I (r, , z; r (n), (j), z (k)) Q Q Q P Q Q Q L , ; S (m)(z 9 z (j)) cos( 9 (m))I (r, , z; r (k), (m), z (j)) . Q Q Q Q Q Q K (4.98)
The function I in Eq. (4.98) is defined by XQ g(r, , z; r, , z) dz : (91)I>I (r, , z; r, , z (k)). (4.99) Q XQ I
242
CHAPTER 4 Permanent Magnet Applications
Evaluation of Eq. (4.99) gives
I Y
where
F(r, , z; r, , z (k)) Q if r ; r 9 2rr cos( 9 ) " 0 or 91 2(z 9 z (k)) Q if
(4.100)
r : r, cos( 9 ) : 1, z " z (k) Q
(z 9 z (k))g(r, , z; r, , z (k)) Q Q . F(r, , z; r, , z (k)) : Q r ; r 9 2rr cos( 9 ) The function I in Eq. (4.98) is defined by PQ g(r, , z; r, , z)r dr : (91)I>I (r, , z; r (k), , z). (4.101) Q PQ I Evaluation of Eq. (4.101) gives
K(r, , z; r (k), , z) Q if r sin( 9 ) ; (z 9 z) " 0 or 1 r r (1) (r 9 2r (k)) Q Q p: (91)N 2(r 9 r (k)) 2 r r (2) Q Q if r " r (k), cos( 9 ) : 1, z : z Q or I Y r ; 2r (k) Q 2(r ; r (k)) Q if r " r (k), cos( 9 ) : 91, z : z Q or
1 r (k) Q if r : 0, z : z
(4.102)
243
4.2 MAGNET STRUCTURES
where K(r, , z; r (k), , z) : Q
(r ; (z 9 z)) 9 rr (k) cos( 9 ) Q r sin( 9 ) ; (z 9 z)
;g(r, , z; r (k), , z). Q According to Eq. (4.100), the functional form of I depends on the relationship between the coordinates of the field point (unprimed) and the source point (primed). In particular, I is evaluated using F as long as r ; r 9 2rr cos( 9 ) " 0. However, if this condition is violated, then the lower ratio is used. The function I also depends on the relation between the source and field points. Specifically, if r sin( 9 ) ; (z 9 z) " 0 then the first term in Eq. (4.102) is used. When this condition is violated, the remaining three cases apply as indicated. Notice that the fourth term in Eq. (4.102) is well defined because r (k) " 0 for Q this case (the field and source points cannot coincide). Notice also that this analysis does not include the case in which the field and source point are one and the same. The parameters that appear in Eq. (4.98) are summarized here for convenience, r (2) 9 r (1) Q , r : Q Q N P (2) 9 (1) Q , : Q Q N and
(2s 9 3) (1) : Q N (2) : (2s 9 1) Q N s : 1, 2, . . . , N r (1) : R (inner radius) Q r (2) : R (outer radius) Q z (2) : top of sector Q z (1) : bottom of sector Q N : angular mesh number (even) N : radial mesh number (even). P
(4.103)
244
CHAPTER 4 Permanent Magnet Applications
Azimuthal component: The azimuthal field component also follows from Eq. (4.72), B (x) : A (x) 9 A (x). Q PQ z r XQ
(4.104)
Substitute Eqs. (4.85) and (4.87) into Eq. (4.104) and obtain M B (x) : Q (91)H Q 4 H PQ Q (z 9 z) sin( 9 ) ; r d dr
x 9 x PQ Q XYXQH XQ PQ r 9 r cos( 9 ) ; dr dz . (4.105)
x 9 x Y QH XQ PQ Similar to the radial case, the integrals can be evaluated numerically using Simpson’s method as outlined in Eqs. (4.92)—(4.97). This gives
M B (r, , z) : Q (91)H Q 4 H , ; S (m)(z 9 z (j)) sin( 9 (m)) Q Q Q K P ; Q g(r, , z; r, (m), z (j))r dr Q Q PQ , ; r P S (n)(r 9 r (n) cos( 9 (j))) Q Q P Q L X ; Q g(r, , z; r (n), (j), z) dz . Q Q XQ The remaining integrals in z and r are evaluated using Eqs. (4.100) and (4.102). The total azimuthal field component is obtained as a superposition of contributions from all the sectors,
M , B (r, , z) : Q (91)Q>H>I 4 Q H I , ; S (m)(z 9 z (j)) Q Q K ;sin( 9 (m))I (r, , z; r (k), (m), z (j)) Q Q Q Q
245
4.2 MAGNET STRUCTURES
, ; r P S (n)(r 9 r (n) cos( 9 (j))) Q Q P Q L
;I (r, , z; r (n), (j), z (k)) , Q Q Q
(4.106)
where I and I are defined in Eqs. (4.100) and (4.102), respectively. Calculations: We demonstrate the field solution with some sample calculations. First, we test Eqs. (4.98) and (4.106) using the 2D field solutions (4.66) and (4.67) from Example 4.2.5. Consider an infinitely long, radially polarized cylindrical shell with 10 alternating poles. Assume that the shell has inner and outer radii R : 2 cm and R : 4 cm, and a uniform magnetization M : 4.3 Q ; 10 A/m. The value of M corresponds to a bonded NdFeB material. For 3D Q analysis we need to set the length L of the shell. To obtain 2D results, the length must be much greater than the diameter of the shell (L 2R ). We set L : 80 cm with the bottom and top of the shell at z(1) : 940 cm and z(2) : 40 cm, respectively. The field components B and B are computed at a radius r : 5 cm P for a series of angular values : 0, 3, 6, . . . , 36° (i.e., from the center of one pole to the center of the neighboring pole). The data from the two models are compared in Fig. 4.23. Next, we apply the model to a cylindrical shell geometry with an axisymmetric magnetization (Fig. 4.24). We use the values R : 2 cm, R : 4 cm, L : 4 cm, and M : 4.3 ; 10 A/m. The cylinder is symmetrically Q positioned with respect to the x-y plane (z : 0 corresponds to the middle of cylinder). Field values are computed along several different lines. First, B is P computed at z : 0, 5, 10, . . . , 60 mm, with r : 50 mm (Fig. (4.25). Next, B P is computed along the line z : 40 mm at the points r : 0, 5, 10, . . . , 60 mm (Fig. 4.26). ) EXAMPLE 4.2.7 Determine the radial and axial field components at any point outside an axially polarized cylindrical shell (Fig. 4.27) [4]. Let R and R denote the inner and outer radii of the shell and z and z denote the axial position of the bottom and top of the shell. Assume that magnet has a second quadrant demagnetization curve B : (H ; M z ). (4.107) Q SOLUTION 4.2.7 We use the current model (Section 3.3). From Eqs. (3.95) and (4.107) we find that the volume current density is zero, J : ; M : K ; M z : 0. Therefore, Eq. (3.93) reduces to Q j (x) K A(x) : ds. (4.108) 4 x 9 x 1
246
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.23
FIGURE 4.24
B and B vs (r : 50 mm). P
Cylinder with axisymmetric magnetization.
247
4.2 MAGNET STRUCTURES
FIGURE 4.25
FIGURE 4.26
Component B vs z (r : 50 mm). P
The component B vs r (z : 40 mm). P
248
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.27
Axially polarized cylindrical shell with reference frame.
As for the surface current density, we find that j : 0 on the top and bottom K surfaces of the shell because the magnetization and surface normals are either parallel or antiparallel there. There are two remaining surfaces to consider,
(4.109)
(4.110)
r : R inner surface : 0 2 z zz
and r : R outer surface : 0 2 z zz .
The unit normals for these surfaces are n :
9r r
(inner surface) (outer surface).
(4.111)
249
4.2 MAGNET STRUCTURES
It follows that the surface current densities are
9M Q M Q Thus, Eq. (4.108) can be rewritten as j : K
(inner surface) (outer surface).
(4.112)
X L M R (j) d dz, (4.113) A(x) : Q (91)H A
x 9 x 4 PY0AH H X where R (1) : R , and R (2) : R . Recall that is a function of position, A A : 9sin()x ; cos()y . (4.114) Substitute Eq. (4.114) into Eq. (4.113) and obtain
M X L sin()x A(x) : Q (91)H 9 R (j) d dz A 4
x 9 x PY0AH H X X L cos()y ; R (j) d dz . A
x 9 x PY0AH X Note that A(x) has no z-component. It can be written in terms of cylindrical coordinates by computing the projections
A (x) : A(x) · r P
(4.115)
A (x) : A(x) · .
(4.116)
and
This gives
M X L sin( 9 ) A (x) : Q (91)H R (j) d dz P A 4
x 9 x PY0AH H X
(4.117)
and M A (x) : Q (91)H 4 H Radial field: The radial field
X L cos( 9 ) R (j) d dz. (4.118) A
x 9 x PY0AH X component is given by
B (x) : 9 A (x). P z
(4.119)
250
CHAPTER 4 Permanent Magnet Applications
Substitute Eq. (4.118) into Eq. (4.119) and take into account both the differentiation in z and the integration in z. This gives M B (r, , z) : Q (91)H>I P 4 H I ;
L
cos( 9 )g(r, , z; R (j), , z )R (j) d. A I A
(4.120)
where g(r, , z; r, , z) :
1 . (4.121) [r ; r 9 2rr cos( 9 ) ; (z 9 z)]
The remaining integral in can be written in terms of a discrete sum using a simple numerical scheme such as Simpson’s method. Specifically, let N denote the number of mesh points in the variable and let (m) denote the values at which the integrand is evaluated. As 0 2, we have (m) :
m 2 N
(m : 0, 1, 2, . . . , N).
(4.122)
The integration coefficients S(m) are specified in Eq. (4.97). Applying this scheme to Eq. (4.120) gives an equation for the total radial field component: M B (r, , z) : Q (91)H>IR (j) P A 2N H I , ; S (m) cos( 9 (m))g(r, , z; R (j), (m), z ). (4.123) A I K Axial field: The axial field component is obtained from
1 B (x) : (rA (x)) 9 A (x) . X P r r
(4.124)
Substituting Eqs. (4.117) and (4.118) into Eq. (4.124) yields
M X L r cos( 9 )9R (j) A B (x) : Q (91)H> R (j) d dz. X A 4
x 9 x PY0AH H X (4.125)
251
4.2 MAGNET STRUCTURES
Equation (4.125) can be simplified using the same numerical scheme as in the previous two sections. Specifically, we obtain M B (r, , z) : Q (91)H> X 4 H 2 , ; S(m)(r cos( 9 (m)) 9 R (j))R (j) A A N K X ; g(r, , z; R (j), (m), z) dz . A X The remaining integration in z can be evaluated analytically and will result in
M B (r, , z) : Q (91)H>I X 4 H I 2 , S (m)(r cos( 9 (m)) 9 R (j)) ; A N K
; R (j)I(r, , z; R (j), (m), z ) A A I
.
(4.126)
In Eq. (4.126), the function I(r, , z; r, , z ) is defined by I XI g(r, , z; r, , z) dz : (91)I>I(r, , z; r, , z ). (4.127) I XI I Evaluation of Eq. (4.127) gives
IY
where
F(r, , z; r, , z ) I if r ; r 9 2rr cos( 9 ) " 0 or 91 2(z9z ) I if r : r, cos( 9 ) : 1, z " z I
(z 9 z )g(r, , z; r, , z ) I I . F(r, , z; r, , z ) : I r ; r 9 2rr cos( 9 )
(4.128)
252
CHAPTER 4 Permanent Magnet Applications
Calculations: We demonstrate Eqs. (4.123) and (4.126) with some sample calculations. We use the following parameters: M : 4.3 ; 10 A/m Q R : 2.0 cm R : 4.0 cm h : 4.0 cm
(4.129)
where h is the height of the magnet (z : 92.0 cm, z : 2.0 cm). First, B and P B are computed at the points z : 0, 5, 10, . . . , 80 mm, for r : 0.25R , 0.5R , X and 0.75R . The computed data are compared with corresponding FEA data in Figs. 4.28 and 4.29. Next, B is computed along the axis of the magnet where X the radial component is zero (Fig. 4.30). Finally, B and B are computed along P X the line defined by z : 40 mm at the points r : 0, 5, 10, . . . , 80 mm. These data are shown in Fig. 4.31. )
FIGURE 4.28
The component B vs z (; : FEA). P
253
4.2 MAGNET STRUCTURES
FIGURE 4.29
The component B vs z (; : FEA). X
EXAMPLE 4.2.8 Derive an expression for the B-field outside an axially polarized multipole cylindrical disk magnet (Fig. 4.32) [5]. Assume that the magnet has a second quadrant demagnetization curve B : (H < M z ). (4.130) Q The < term takes into account the alternating polarity of adjacent poles. SOLUTION 4.2.8 We use the current model of Section 3.3. It follows from Eqs. (3.95) and (4.130) that the volume current density is zero, J : ; M : 0. K Therefore, Eq. (3.93) reduces to A(x) : 4
j (x) K ds.
x 9 x
(4.131)
(4.132)
1 The B-field is computed from Eq. (4.132) using B : ; A.
(4.133)
254
FIGURE 4.30
CHAPTER 4 Permanent Magnet Applications
The component B vs z along the axis of the cylinder (; : FEA). X
Instead of determining B for the entire magnet, we first evaluate Eq. (4.132) for a single sector shown in Fig. 4.33, determine B for the sector, and then obtain the total B-field as a superposition of the contributions from all the sectors. Equivalent current density: As a first step, we determine the surface current density j for the various surfaces of the sector shown in Fig. 4.33. The K magnetization for this sector is M : M z . Q
(4.134)
From Eqs. (3.95) and (4.134) it follows that j : 0 on the top and bottom of K the sector because the magnetization and surface normals are either parallel or antiparallel on these surfaces. There are four remaining surfaces to consider:
r : R inner surface : (1) (2) Q Q z (1) z z (2), Q Q
(4.135)
255
4.2 MAGNET STRUCTURES
FIGURE 4.31
The components B and B vs r (; : FEA). P X
r : R outer surface : (1) (2) Q Q z (1) z z (2), Q Q R r R left side : : (1) Q z (1) z z (2), Q Q
FIGURE 4.32
Axially polarized multipole disk magnet.
(4.136)
(4.137)
256
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.33
Sector of an axially polarized disk magnet.
and
R r R right side : : (2) Q z (1) z z (2). Q Q The unit normals for these surfaces are
n :
9r
(inner surface)
r
(outer surface)
9
(left side)
(4.138)
(4.139)
(right side).
The corresponding surface current densities are
9M Q M Q j : K M r Q 9M r Q
(inner surface) (outer surface) (left side) (right side).
(4.140)
257
4.2 MAGNET STRUCTURES
Vector potential: We use Eqs. (4.135)—(4.140) and rewrite Eq. (4.132), M A (x) : Q (91)H Q 4 H
XQ
Q
x 9 x
r (j) d dz Q
PYPQH Q X PQ r 9 Q dr dz .
x 9 x Y QH XQ PQ
XQ
(4.141)
In this expression, the subscript s denotes the contribution due to a single sector. The unit vectors r and in Eq. (4.141) are themselves functions of , and therefore need to be considered when performing the integration (Section A.2). From Table A.2 we have r : cos()x ; sin()y
(4.142)
: 9sin()x ; cos()y .
(4.143)
and
Therefore, Eq. (4.141) can be rewritten in terms of Cartesian components, M A (x) : Q (91)H Q 4 H
; 9
XQ
XQ
;
XQ
XQ
9
XQ
XQ
9
XQ
XQ
Q sin()x r (j) d dz Q
x 9 x PYP H Q Q Q cos()y
x 9 x r (j) d dz PYPQH Q Q PQ cos()x dr dz
x 9 x PQ Y QH PQ sin()y dr dz .
x 9 x PQ Y QH
Notice that A (x) has no z-component. It can be decomposed into its cylindrical Q components by computing the projections, A (x) : A (x) · r PQ Q
(4.144)
A (x) : A (x) · . Q Q
(4.145)
and
258
CHAPTER 4 Permanent Magnet Applications
The cylindrical components are M A (x) : Q (91)H PQ 4 H XQ Q sin( 9 ) r (j) d dz ; Q
x 9 x XQ Q PYPQH X PQ cos( 9 ) 9 Q dr dz ,
x 9 x XQ PQ Y QH
and
M A (x) : Q (91)H Q 4 H
XQ
(4.146)
Q cos( 9 ) r (j) d dz Q
x 9 x PYPQH XQ Q X PQ sin( 9 ) ; Q dr dz , (4.147)
x 9 x XQ PQ Y QH Equations (4.146) and (4.147) specify the components of the potential for a single sector. The field B (x) due to the sector can be obtained using Q B : ; A . Since the total field is a superposition of the fields due to all the Q Q sectors, it can be written as
, B(x) : (91)Q>B (x), (4.148) Q Q where (91)Q> takes into account the alternating polarity of adjacent poles. Radial component: The radial field component due to a single sector follows from Eq. (4.133), B (x) : 9 A (x). PQ z Q
(4.149)
Substituting Eq. (4.147) into Eq. (4.149) and taking into account both the differentiation in z and the integration in z yields M B (x) : Q (91)H>I PQ 4 H I Q cos( 9 )g(r, , z; r (j), , z (k))r (j) d ; Q Q Q Q P ; Q sin( 9 )g(r, , z; r, (j), z (k)) dr , Q Q PQ
(4.150)
259
4.2 MAGNET STRUCTURES
where g(r, , z; r, , z) :
1 . [r ; r 9 2rr cos( 9 ) ; (z 9 z)]
The remaining integrals in r and can be written in terms of discrete sums using Simpson’s method. Let N and N denote the (even) number of mesh P points in the r and variables, respectively. Let r(n) and (m) denote the Q values at which the integrands are evaluated. As r (1) r r (2) and Q Q (1) (2), we have Q Q r (2) 9 r (1) Q (4.151) r : Q Q N P and (2) 9 (1) Q . : Q Q N
(4.152)
The integration points are n r (n) : r (1) ; (r (2) 9 r (1)) Q Q Q N Q P
(n : 0, 1, 2, . . . , N ) P
(4.153)
and m ( (2) 9 (1)) (m) : (1) ; Q Q Q N Q The coefficients S (n) and S (m) are P S (n) : P and
(m : 0, 1, 2, . . . , N). (4.154)
(n : 0) (n : 1, 3, 5, . . .) (n : 2, 4, 6, . . .)
(4.155)
(n : N ), P
(m : 0) (m : 1, 3, 5, . . .) S (m) : (4.156) (m : 2, 4, 6, . . .) (m : N). Applying this scheme to Eq. (4.150) and summing the contribution from all the sectors in accordance with Eq. (4.148) gives an equation for the total radial field
260
CHAPTER 4 Permanent Magnet Applications
component, M , B (r, , z) : Q (91)Q>>H>I P 4 Q H I
, ; S (m) cos(9 (m))g(r, , z; r (j), (m), z (k))r (j) Q Q Q Q Q Q K
, ;r P S (n) sin( 9 (j))g(r, , z; r (n), (j), z (k)) , Q Q Q P Q Q L (4.157) where N : number of poles (1) : (2s 9 3) Q N (2) : (2s 9 1) Q N s : 1, 2, . . . , N r (1) : R (inner radius) Q r (2) : R (outer radius) Q z (2) : top of sector Q z (1) : bottom of sector Q N : angular mesh number (even) N : radial mesh number (even). P
(4.158)
Azimuthal component: The azimuthal component also follows from Eq. (4.133), B (x) : A (x). Q z PQ
(4.159)
Applying Eq. (4.159) to Eq. (4.146) and taking into account the differentiation
261
4.2 MAGNET STRUCTURES
in z and the integration in z yields M B (r, , z) : Q (91)H>I> Q 4 H I
Q sin( 9 )g(r, , z; r (j), , z (k))r (j) d Q Q Q Q P 9 Q cos( 9 (j))g(r, , z; r, (j), z (k)) dr . Q Q Q PQ Again, apply Simpson’s method and sum the contributions from all the sectors. This gives ;
M , B (r, , z) : Q (91)Q>H>I 4 Q H I
, ; S (m) sin(9 (m))g(r, , z; r (j), (m), z (k))r (j) Q Q Q Q Q Q K
, 9r P S (n) cos( 9 (j))g(r, , z; r (n), (j), z (k)) , Q Q Q P Q Q L (4.160) where the variables are defined in Eq. (4.158). Axial components: The axial component of a sector is given by
1 B (x) : (rA (x)) 9 A (x) . XQ Q r r PQ
(4.161)
Substituting Eqs. (4.146) and (4.147) into Eq. (4.161) gives M B (x) : Q (91)H> XQ 4 H
;
XQ
Q r cos( 9 ) 9 rQ (j) r (j) d dz Q
x 9 x XQ Q PYPQH X PQ r sin( 9 ) ; Q dr dz , (4.162)
x 9 x XQ PQ Y QH
262
CHAPTER 4 Permanent Magnet Applications
Equation (4.162) can be simplified using Simpson’s method as above, M B (r, , z) : Q (91)H> XQ 4 H
, ; S (m)[r cos( 9 (m)) 9 r (j)]r (j) Q Q Q Q K ;
XQ g(r, , z; r (j), (m), z) dz Q Q XQ
, ;r P S (n)r sin( 9 (j))) Q P Q L ;
XQ
XQ
g(r, , z; r (n), (j), z) dz Q Q
.
The remaining integration in z can be evaluated analytically. We sum the contributions from all the sectors and obtain B for the entire magnet, X M , B (r, , z) : Q (91)Q>>H>I X 4 Q H I
, ; S (m)[r cos( 9 (m)) 9 r (j)] Q Q Q K ;r (j)I(r, , z; r (j), (m), z (k)) Q Q Q Q
, ;r P S (n)r sin(9 (j)) I(r, , z; r (n), (j), z (k)) . Q Q P Q Q Q L (4.163) In Eq. (4.163), the function I is defined by
XQ g(r, , z; r, , z) dz : (91)I>I(r, , z; r, , z (k)). (4.164) Q XQ I
263
4.2 MAGNET STRUCTURES
Evaluation of Eq. (4.164) gives
IY
where
F(r, , z; r, , z (k)) Q if r ; r 9 2rr cos( 9 ) " 0 or 91 2(z9z (k)) Q if
(4.165)
r : r, cos( 9 ) : 1, z " z (k) Q
(z 9 z (k))g(r, , z; r, , z (k)) Q Q F(r, , z; r, , z (k)) : . Q r ; r 9 2rr cos( 9 ) The functional form of I depends on the relationship between the coordinates of the field point (unprimed) and the source point (primed). In particular, I is evaluated in terms of the function F as long as r ; r 9 2rr cos( 9 ) " 0. If this condition is violated, then the lower ratio is used. Calculations: We demonstrate the field solution via the analysis of a four-pole disk magnet with the following parameters: M : 4.3 ; 10 A/m R : 2.0 cm R : 4.0 cm h : 1.0 cm
(inner radius) (outer radius) (height).
(4.166)
The magnet is oriented with its upper surface in the x-y plane with one of its sectors symmetrically positioned in an angular sense with respect to the x-axis. The upper surface of this sector is taken to be a north pole. The field components B and B are evaluated along a radial line that is 1 cm above the surface of the P X magnet and centered with respect to this sector ( : 0). Field values are computed at r : 0, 5, 10, . . . , 45 mm. These data are compared with corresponding data from a 3D FEA as shown in Fig. 4.34. Similarly, the field components B and B are evaluated along an arc at the mean radius of the magnet, 1 cm X above its surface. The arc extends from the center of one pole to the center of the neighboring pole. Field values are computed at : 0, 5, 10, . . . , 90°. These data are shown in Fig. 4.35. )
264
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.34
The components B and B vs r (; : FEA). P X
FIGURE 4.35
The components B and B vs (; : FEA). X
4.3 HIGH FIELD STRUCTURES
265
4.3 HIGH FIELD STRUCTURES Conventional permanent magnet structures such as those studied in the preceding typically produce field strengths of 1.0 T in proximity to their surface. However, special structures can be fabricated that provide uniform multitesla fields throughout their interior. For example, fields of up to 4.0 T have been achieved in disk-shaped cavities 2.5 cm in diameter and 0.5 cm high with structures 15 cm in diameter [6—8]. Such structures can be used for numerous applications such as Hall effect studies. They can also be used to focus charged particle beams if through-ports are provided. The most popular high field structures are the ‘‘magic’’ cylinder and sphere [6—9]. These are made from high-strength rare-earth materials such as NdFeB. They are fabricated piecewise by cutting uniformly ` a magnetized block, and then orienting polarized slices of material from and assembling the constituent pieces so as to provide a predefined magnetization pattern throughout the assembled structure. The magnetization pattern is chosen to provide a desired field strength and uniformity in the interior of the structure. The ‘‘magic’’ cylinder is shown in Fig. 4.36. It has the form of a cylindrical shell with a magnetization that varies with the polar angle as shown. This structure can be built from a stack of cylindrical sections as shown in Fig. 4.37. These sections can be constructed as follows. First, cut a circular ring from a uniformly magnetized block of material. Next, cut the ring into the desired number of pie-shaped elements. Finally, reorder the pie-shaped elements around the circumference to obtain the desired angular magnetization distribution. The fabrication sequence is illustrated in Fig. 4.37. The flux density in the interior of a magic cylinder is given by B : B ln(r /r ), (4.167) P M G where B is the remanence of the material and r and r are the inner and P G M outer radii of the cylinder, respectively [6]. Notice that Eq. (4.167) implies that there is no theoretical upper limit to the field that can be obtained in this structure. However, the slow logarithmic dependency on r places M a practical upper limit of 2.5 T for cavities of more than a few centimeters in diameter. The mass of material required to obtain higher fields would be prohibitively large. The ‘‘magic’’ sphere is shown in Fig. 4.38. It has the form of a spherical shell in which the direction of magnetization varies with the polar angle as :2. This magnetization pattern is obtained by assembling a collection of melonlike segments. These segments can be constructed by
266
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.36
Magic cylinder: (a) idealized structure; and (b) physical structure.
beveling the ring structures used in the construction of the magic cylinder (Fig. 4.39). The flux density in the cavity of a magic sphere is given by B : B ln(r /r ), P M G
FIGURE 4.37
Fabrication sequence for a magic cylinder [9].
4.3 HIGH FIELD STRUCTURES
FIGURE 4.38
267
Magic sphere: (a) idealized structure; and (b) physical structure.
FIGURE 4.39
Fabrication sequence for a magic sphere [9].
268
CHAPTER 4 Permanent Magnet Applications
where r and r are the inner and outer radii of the shell, respectively [6]. G M Such structures provide uniform fields of 2.0 to 3.0 T in their interior when the outer radius of the shell is only three to four times the radius of the cavity. For example, a magic sphere with B : 1.2 T and an outer P radius of 3.0 cm produces a uniform field of 2.5 T in a cavity with a 1.0 cm radius [7].
4.4 MAGNETIC LATCHING Magnetic latching is one of the oldest and most common applications of permanent magnets. In latching applications, magnets are used to provide a force of attraction between two members. Latching magnets can be found in a variety of products including household appliances, tool equipment, electrical devices, and toys. Such magnets are usually made from ferrite materials because of their low cost. Latching applications fall into one of two categories: contact and noncontact. The distinction between these two is important because the design approach is different for each. In contact latching the magnet is attracted to and in contact with another member (either another magnet or a soft magnetic material such as iron). The latching magnet is often designed to have numerous closely spaced poles to enhance the latching force. In noncontact latching the magnet exerts a force of attraction over a distance. Here, the poles of the latching magnet need to be separated so that the field will span the gap between the magnet and the adhering surface in order to provide the latching force. Therefore, the magnetic structures for contact and noncontact latching are quite different. We consider contact latching first. Consider a magnet with a remanence B that is in contact with a plate of high-permeability material P such as iron that is unsaturated (Fig. 4.40). The force of attraction between the magnet and the plate can be estimated using the Maxwell stress tensor expression (3.59), which gives BA (4.168) F: P , 2 where A is the area of contact between the magnet and the plate. The force can be rewritten in terms of the flux : B A, P F: . (4.169) 2 A This expression implies that the latching force can be increased by
4.4 MAGNETIC LATCHING
FIGURE 4.40
269
Contact latching: (a) single-pole contact; and (b) multipole contact.
reducing the contact area A and/or increasing the flux . This can be accomplished by tapering the poles of the magnet, or by adding tapered iron pole pieces to the magnet. However, there is a practical limit to this effect because Eq. (4.169) only applies when the material being held is well below saturation. As A decreases the material saturates and the force no longer increases. This is especially true in applications involving thin plates that have a relatively low saturation threshold due to their limited cross-sectional area. One way to achieve a higher latching force in such applications is to use a magnet with numerous closely spaced poles as shown in Fig. 4.40b. If there are N such poles, and each exerts a force F, then the total force is F : NF as long as the material being held is below saturation. We now consider noncontact latching. In noncontact latching the two attracted members are separated by a gap, and the magnet must project its field across the gap to exert the force of attraction. Thus, the poles of the magnet must be effectively separated so that the field will span the gap. Noncontact latching is often implemented using a magnetic circuit as opposed to a standalone magnet. There are numerous latching circuits, and most of these can be designed and optimized using magnetic circuit theory (Section 3.5). We demonstrate the design method via the analysis of two practical latching circuits.
270
CHAPTER 4 Permanent Magnet Applications
Magnetic latching device: (a) magnet with two attached flux plates; and (b) latching circuit.
FIGURE 4.41
EXAMPLE 4.4.1 Consider the magnetic circuit of Fig. 4.41b. A magnet with attached flux plates is attracted to a wall made of soft magnetic material. Determine the force of attraction when a gap exists between the magnetic structure and the wall g 0. Assume that the flux plates and the wall have infinite permeability, and that the magnet has a linear second quadrant demagnetization curve of the form B :B ; H , K P K K
(4.170)
where : B /H . K P A SOLUTION 4.4.1 We use magnetic circuit theory (Section 3.5). As there are no currents we have w H · dl : 0, which gives H l ; 2H g : 0, (4.171) KK E where H is the field in the gap. The factor of 2 in Eq. (4.171) is due to the fact E that the path of` integration passes through the gap region twice as shown in Fig. 4.41b. In addition, notice that Eq. (4.171) ignores the path through the flux plates and the wall because H is negligible in these elements by assumption. As there is no fringing, w B · dS : 0, which implies that : . Thus, K E B A :B A (no fringing at gap), (4.172) K K E E
271
4.4 MAGNETIC LATCHING
where A : Lw is the area of the magnet, and A : Lw is the area of gap. K K E E We combine Eqs. (4.170), (4.171), and (4.172) and obtain A B P B (g) : K . E A (1 ; 2( / )(A /A )(g/l )) E K K E K Next, we determine the force using the Maxwell stress tensor (Section 3.2.3). We compute the force on the wall by applying Eq. (3.59) over an area bounding the wall (similar to the area indicated by the dotted line in Fig. 3.9 of Example 3.2.8). The only contributions to Eq. (3.59) come from the area in the gap regions. In these regions, B : B , and therefore Eq. (3.59) reduces to L E 1 F(g) : B(g)n ds L 2 1 1 (2B(g)A )n : E E 2 BA P K : n . (4.173) A (1 ; 2( / )(A /A )(g/l )) E K K E K The outward normal n of the wall points towards the magnet, and, therefore, the force is attractive. )
EXAMPLE 4.4.2 Determine the force of attraction for the latching circuit of Fig. 4.42b. Assume that the flux plate and the wall have infinite permeability and that the magnet has a linear second quadrant demagnetization curve of the form B :B ; H , K P K K
(4.174)
where : B /H . K P A SOLUTION 4.4.2 We use magnetic circuit theory (Section 3.5). First, apply w H · dl : 0 to the closed dotted line path on the right-hand side of the circuit. This gives H l ; 2H g : 0, (4.175) KK E where H is the field in the gap. The factor of 2 in Eq. (4.175) is due to the fact E that the path of integration passes through the gap region twice as shown. Notice that we have ignored contributions from the path through the flux plate and the wall as H is negligible in these elements. Next, because there is no fringing, : 2 and we have K E B A : 2B A (no fringing at gap), (4.176) K K E E
272
CHAPTER 4 Permanent Magnet Applications
Magnetic latching circuit: (a) magnet with attached flux plate; (b) latching circuit.
FIGURE 4.42
where A : Lw is the area of the magnet, and A : Lw is the area of the gap. K K E E The factor of 2 in Eq. (4.176) accounts for the fact that the flux through the magnet splits evenly between the two return paths as indicated by the dotted lines in Fig. 4.42b. We combine Eqs. (4.174), (4.175), and (4.176) and find that A B P B (g) : K . E A 2(1 ; ( / )(A /A )(g/l )) E K K E K
273
4.5 MAGNETIC SUSPENSION
We use the Maxwell stress tensor to obtain the force (Section 3.2.3). Specifically, we apply Eq. (3.59) over an area bounding the wall (similar to the area indicated by the dotted line in Fig. 3.9 of Example 3.2.8). The only contributions to Eq. (3.59) come from the area in the gap regions. There are three such regions — a central region immediately beneath the magnet and two end regions. To apply Eq. (3.59) we need to know B (the field component normal to the wall). In the L central gap region B : B : 2(A /A )B , whereas in the end regions B : B . L K E K E L E Substituting these into Eq. (3.59) gives
1 B(g)n ds, L 2 1 1 : [2B(g)A ; B A ]n E K K E 2 BA A ; 2A P K K E n . : (4.177) 2 (1 ; ( / )(A /A )(g/l )) 2A K K E K E Because the outward normal at the wall n points towards the magnetic circuit, the force is attractive. ) F(g) :
4.5 MAGNETIC SUSPENSION In magnetic suspension, magnets are used to provide passive levitation and/or positioning of a structure. The magnets can be arranged in myriad configurations and substantial levitation force can be realized when modern high-strength materials such as NdFeB are used. The two most common suspension configurations are the radial- and axial bearing configurations shown in Fig. 4.43 [10]. Magnetic bearings are often used to provide low friction support for rotating mechanical structures as illustrated in Fig. 4.44. They enable substantially higher operating speeds relative to conventional mechanical bearings. It is well known that a passive magnetic field alone cannot provide stable support of a structure. Such stability can only be achieved with some form of active control [11, 12]. This is a consequence of Earnshaw’s theorem, which states that ‘‘A body with steady charges, magnetizations, or currents placed in a steady electric or magnetic field cannot rest in stable equilibrium under the action of electric and magnetic fields alone’’ [13]. Earnshaw’s theorem can be understood from the following heuristic argument [14]. Let Q denote the magnetic scalar potential due to a K permanent magnet support structure. Assume that the structure is
274
CHAPTER 4 Permanent Magnet Applications
Cross-sectional view of axisymmetric magnetic bearings: (a) radial bearing configurations; and (b) axial bearing configurations.
FIGURE 4.43
designed to suspend a given magnet. Because the suspension point is outside the support structure, we have Q : 0. (4.178) K Now, the suspension force FQ is proportional to the field strength HQ : 9Q, K FQ . Q. K If the suspended magnet is to be in stable equilibrium, the net force must be zero, which implies that Q : 0. K
(4.179)
4.5 MAGNETIC SUSPENSION
FIGURE 4.44
275
Magnetic suspension with lateral mechanical constraint.
However, Eq. (4.179) indicates a potential maximum or minimum of Q. K For stable equilibrium, the potential must be a minimum that occurs when Q 0. (4.180) K However, Eq. (4.180) contradicts Eq. (4.178), which implies that it is impossible to provide stable suspension with a passive magnetic field. Therefore, to implement magnetic suspension, magnets are usually used along with some form of mechanical structure that constrains the motion of the supported member. An example of this is shown in Fig. 4.44. Here, opposing disk magnets levitate a rotating member, while its lateral motion is constrained by a fixed support with a conventional mechanical bearing that enables rotation. The force between the disk magnets is derived in the following example. EXAMPLE 4.5.1 Derive an expression for the levitation force between two identical, axially polarized cylindrical disks as shown in Fig. 4.45 [15]. The disks have inner and outer radii R and R , thickness t , and are separated by K a distance h (Fig. 4.46). Assume that both magnets have a linear second quadrant demagnetization curve of the form B : (H < M z ). (4.181) Q where the < takes into account the opposite polarizations of the opposing magnets.
276
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.45
FIGURE 4.46
Levitated magnetic disk.
Cross section of disk geometry with reference frame.
277
4.5 MAGNETIC SUSPENSION
SOLUTION 4.5.1 We use the charge model (Section 3.4). Specifically, we first reduce the magnets to equivalent charge distributions and then determine the force using the force formulas (3.120) to (3.123). The volume and surface charge densities and are given by Eq. (3.105). From Eq. (4.181) we have K K : 9 · M K : 9 · M z Q : 0, (4.182) and : M · n K : <M , (4.183) Q where the < sign takes into account the opposite polarity of the charge on opposing pole faces. Now, consider the force expressions (3.120) to (3.123) for a discretized magnet. Because : 0 in both magnets, the volume terms vanish K and the force between the magnets reduces to
(x ) (x )(x 9 x ) I A A . F: K N K I N (4.184) I N 4
x 9 x N I N I Instead of evaluating Eq. (4.184) directly, it is easier to derive the force in the following steps: First, obtain the force between two point sources; next, derive the force between opposing pole faces; and finally, obtain the total force by summing the contributions from all the surfaces. Point sources: We use cylindrical coordinates. Without loss of generality, one of the charges Q (x ) can be taken to be in the x-y plane at (r , , 0) while the K other Q (x ) is taken to be in a plane defined by z : h at (r , , h). The axial K force between these two point sources follows from Eq. (4.184): Q (r , , 0)Q (r , , h)h K K F : . (4.185) X 4 [r ; r 9 2r r cos( 9 ) ; h] Surfaces: The next step is to evaluate the axial force exerted by one pole face (surface) on an opposing pole face. Again, without loss of generality, one surface can be taken to lie on the x-y plane while the other can be taken to lie on the plane defined by z : h. The idea is to discretize each surface into a mesh of subsections (elements), define a point charge at the midpoint of each element in accordance with Eq. (3.110), and then compute the force exerted by one surface on the other by summing the contributions from the individual point charges. The area of each surface is (R 9 R). The first step is to divide this area into
278
CHAPTER 4 Permanent Magnet Applications
a set of indexed elements. These elements are taken to have equal areas so that they contain an equal amount of surface charge. Because the area of the surface is spanned by the two coordinate variables r and , two indices (i, j) are required. Let N and N be the number of mesh points in these two variables, P respectively. Each element will have the same area A, (R 9 R) . N N P Choose each element so that it spans an angular measure A :
:
2 . N
(4.186)
(4.187)
Let (r , ) denote the midpoint of the element A , and let R(i) and R(i ; 1) G H GH denote its radial boundaries. Then r : R(i ; 1) 9 R(i), G R(i ; 1) ; R(i) r: (i : 1, 2, . . . , N ), G P 2
(4.188) (4.189)
and , : (j 9 1) ; H 2
(4.190)
for j : 1, 2, . . . , N. The end points are computed using the following equation:
(R 9 R) (i : 1, 2, . . . , N ), (4.191) P N P where R(1) : R (inner radius). Equation (4.191) is used to generate all the end point values from which all the midpoint values are obtained. Therefore, Eqs. (4.189) to (4.191) define a scheme for discretizing the surface into elements and assigning values to the midpoints of each element. Next, consider the magnetic interaction between an element of one surface and an element of an opposing surface. Let the coordinates of the midpoints of the two elements be (r , , 0) G H and (r , , h), respectively. The magnetic charge associated with each element GY HY is R(i ; 1) :
R(i) ;
Q (r , , 0) : (r , , 0)A K G H G H
(4.192)
Q (r , , h) : (r , , h)A, K GY HY GY HY
(4.193)
and
279
4.5 MAGNETIC SUSPENSION
respectively. Substitute Eqs. (4.183) and (4.186) into these expressions and obtain Q : M K Q
(R 9 R) . N N P
(4.194)
The force is obtained by summing the contributions of all the elements in accordance with Eq. (4.185). By exploiting the symmetry of the problem the force reduces to
(R 9 R) F (h) : M X 4 Q N N P , , , h . (4.195) ;N P P [r ; r 9 2r r cos( ) ; h] GY G GY HY G GY HY G Force: There are two magnets to consider, which are labeled upper or lower in accordance with their respective positions. The upper magnet has its bottom surface in the plane defined by z : h and its top surface lies in the plane defined by z : h ; t (Fig. 4.46). The lower magnet has its top surface in the x-y plane K and its bottom surface in the plane defined by z : 9t . The axial force on the K top surface due to both the top and bottom surfaces of the upper magnet is F (h) : F (h) 9 F (h ; t ). X X X K Continuing in this fashion and taking into account the contributions from all the surfaces, the following formula is obtained for the total force between the magnets:
MN (R 9 R) F : Q X 4 N N P , , , (91)K h K K ; P P . [r ; r 9 2r r cos( ) ; h ] GY K K G GY HY G G GY HY where : (j 9 1) ; HY 2 :
2 N
(j : 1, 2, . . . , N)
(4.196)
280
CHAPTER 4 Permanent Magnet Applications
and r: G
R(i ; 1) ; R(i) 2
r : GY
R(i ; 1) ; R(i) 2
(i : 1, 2, . . . , N ) P (i : 1, 2, . . . , N ) P
(R 9 R) (i : 1, 2, . . . , N ) P N P (R 9 R) R(i ; 1) : R(i) ; (i : 1, 2, . . . , N ) P N P R(1) : R (inner radius). The terms h and are given by K K h : h ; mt (m : 0, 1, 2) K K h : separation of magnets R(i ; 1) :
R(i) ;
t : thickness of magnets K and
1
(m : 0)
: 2 K 1
(m : 1) (m : 2).
Notice that Eq. (4.196) contains nested summations of elementary functions that are easy to program. However, it is important to note that when programming nested summations, a higher degree of accuracy can be achieved with fewer mesh points (lower values N and N) if the total of a single P summation over one index is evaluated and saved separately (as an intermediate step), and then the totals of the successive summations are added together to obtain the final sum. For example, Eq. (4.196) could be evaluated as follows:
MN (R 9 R) F : Q S (m), X 4 N N P K where , S (m) : P S (i, m), G
4.5 MAGNETIC SUSPENSION
FIGURE 4.47
with
and
281
Levitation force vs separation distance (; : FEA).
, S (i, m) : P S (i, i, m), GY
, (91)K h K K S (i, i, m) : . [r ; r 9 2r r cos( ) ; h ] G GY K HY G GY HY Calculations: We demonstrate Eq. (4.196) with some sample calculations. We use the following parameters: M : 4.3 ; 10 A/m Q R : 2.0 cm R : 4.0 cm t : 1.0 cm. K Force values are computed for a series of separation distances with the mesh parameters set to N : 20 and N : 60. The computed data are compared with P corresponding FEA data in Fig. 4.47. )
282
CHAPTER 4 Permanent Magnet Applications
4.6 MAGNETIC GEARS Magnetic gears can be used in place of mechanical gears to reduce undesired vibrations and for applications that require torque coupling between separated members. There are many gear configurations. One such configuration consists of two separated, radially polarized cylindrical magnets that are constrained to rotate about their respective axes (Fig. 4.48). The magnets are magnetically coupled to one another and when one magnet rotates it imparts a torque to the other, causing it to rotate [16]. The coupling between the magnets is a function of several variables including the number of poles, material properties, dimensions, and separation. Furthermore, substantial torque can be realized if modern, rare-earth materials such as NdFeB are used. We derive a formula for the torque in the following example. EXAMPLE 4.6.1 Derive an expression for the torque between two separated radially-polarized multipole cylinders (Fig. 4.48) [16]. Assume that the magnets have identical geometric dimensions, and that their length is much greater than their diameter. Let N denote the number of poles, and let R and R denote the inner and outer radii of each magnet. Assume that the magnets have different second quadrant demagnetization relations. The demagnetization curve for the magnets in the primed and unprimed coordinate systems are B : H ; M,
FIGURE 4.48
(4.197)
Magnetically coupled cylinders with reference frames.
283
4.6 MAGNETIC GEARS
and B : (H ; M),
(4.198)
respectively, where M : <M r . (4.199) Q The < sign takes into account the alternating polarity of adjacent poles. SOLUTION 4.6.1 The gear consists of two magnets. We refer to the magnet centered with respect to the primed coordinate system (x, y) as the source magnet. We refer to the magnet centered with respect to the unprimed coordinate system (x, y) as the drive magnet. The analysis proceeds in two steps. First, we use boundary-value theory to obtain a free-space field solution for the source magnet. Second, the drive magnet is reduced to an equivalent current distribution and the torque is computed by considering this current to be in the external field of the source magnet. Source magnet: As the length of the source magnet is much greater than its diameter, we can determine its field using a 2D analysis. The field due to the source magnet was derived in Example 4.2.5. Specifically, the field components in the primed coordinates are given by B(r, ) : ir\G>U(M , R , R , ) cos(i) P G G G
(4.200)
and (4.201) B (r, ) : ir\G>U (M , R , R , ) sin(i). G G G The coefficients U(M , R , R , ) are defined by Eq. (4.53), with M given by G G G Eq. (4.64). The field components (4.200) and (4.201) constitute the external field in the following analysis. Therefore, in what follows we adopt the notation (r, ). B(r, ) : B(r, ) and B (r, ) : B PY P Y Drive magnet: To determine the torque on the drive magnet we first reduce it to a distribution of equivalent volume and surface current densities J and j , K K respectively (Section 3.3). Once this is done, we compute the torque using Eq. (3.100), T:
4
;
r(r, ) ; (J (r, ) ; B (r, ))r dr d dz K
1
r(r, ) ; (j (r, ) ; B (r, )) da, K
(4.202)
284
CHAPTER 4 Permanent Magnet Applications
where V and S denote the volume and surface of the magnet. Notice that B (r, ) is the field due to the source magnet represented in the unprimed coordinates. From Eq. (4.199) we have J : ; M : 0 and, therefore, the first K term in Eq. (4.202) is zero. As for the second term, there are N sectors, and each sector has two surfaces with current densities that contribute to the torque. These surfaces constitute the radial sides of the sector at angular positions and , respectively. If the magnet is rotated by an angle , then the surface current densities for the pth sector are R rR M z (p) : ; (1 ; 2p) Q N j (p, r, ) : (4.203) K R rR 9M z (p) : 9 (1 ; 2p), Q N where p : 0, 1, 2, . . . , N 9 1. Notice that : 0 occurs when the middle of the zeroth sector coincides with the x axis and its polarization is radially outward as shown in Fig. 4.48. It is assumed that the source magnet is held stationary with one of its sectors in a similar orientation with respect to the primed coordinate system and, therefore, is the relative angular offset of the two magnets. We substitute Eq. (4.203) into Eq. (4.202) and obtain
T() : 2M L Q
, \ 0 (91)N r[cos( (, p))B(r, (, p)) V 0 N ; sin( (, p))B(r, (, p))] dr, W (4.204)
where L is the length of the gear, and (, p) : ; (1 ; 2p). (4.205) N The coefficient 2M takes into account the fact that there are two surfaces at the Q interface between neighboring sectors (one for each sector). The integral in Eq. (4.204) can be evaluated numerically using Simpson’s method. The resulting torque formula is T() :
2M LR* , \ ,P Q (91)NS (q)r(q) P N P N O ;[cos( (, p))B(r(q), (, p)) V ;sin( (, p))B(r(q), (, p))], W
(4.206)
285
4.6 MAGNETIC GEARS
where N is the radial mesh coefficient (even), R* : R 9 R , and S (q) are the P P Simpson coefficient terms
S (q) : P
(q : 0) (q : 1, 3, 5, . . .) (q : 2, 4, 6, . . .)
(4.207)
(q : N ). P
The integration points are as follows: q r(q) : R ; (R 9 R ) N P
(q : 0, 1, 2, . . . , N ). P
(4.208)
To compute the torque we express the external field components B(r, ) PY and B(r, ) (Eqs. (4.200) and 4.201)) in terms of the unprimed coordinates, Y B(r, ) Y B(r(r, ), (r, )) cos((r, )) V PY 9 B (r(r, ), (r, )) sin((r, )) Y
(4.209)
and B(r, ) Y B(r(r, ), (r, )) sin((r, )) W PY ; B (r(r, ), (r, )) cos((r, )) Y
(4.210)
In Eqs. (4.209) and (4.210) we use the coordinate transformation relations r(r, ) : (r ; 2rd cos() ; d
(4.211)
and (r, ) : arctan
r sin() . r cos() ; d
(4.212)
The quantities B(r, (, p)) and B(r, (, p)) in Eq. (4.206) are V W evaluated using Eqs. (4.209) and (4.210). Calculations: We demonstrate the torque formula (4.206) via a practical example. To clarify the analysis, we distinguish the parameters for the source and drive magnets using superscripts s and d, respectively. For example, RQ and RQ refer to the inner and outer radii of the source magnet, and RB and RB refer to the inner and outer radii of the drive magnet. We use the following
286
CHAPTER 4 Permanent Magnet Applications
parameters: MQ : 7.2 ; 10 A/m Q MB : 7.2 ; 10 A/m Q d : 80 mm RQ : 10 mm RQ : 20 mm RB : 15 mm RB : 30 mm N : 4 NB : 4. (4.213) The values of MQ and MB are characteristic of sintered NdFeB material. The Q Q torque per unit length is computed with the source magnet fixed and the drive magnet rotated through a series of angular values : 0, 5, 10, . . . , 90°. The analysis is performed with N : 20 and : in Eq. (4.197). The torque data P are compared with corresponding FEA data in Fig. 4.49. Notice that the peak torque occurs at : 45° when the drive magnet is rotated half the angular span of one pole. An analysis is also performed to determine the decrease of peak torque ( : 45°) as a function of separation distance d for a series of values d : 60, 70, . . . , 160 mm. A first set of data is computed with : in Eq. (4.197). These data are compared with FEA data in Fig. 4.50. An additional calculation is performed with : 2 . This represents a steeper slope in the demagnetization curve (4.197), which results in a lower external field B and a reduced torque as shown. )
4.7 MAGNETIC COUPLINGS Synchronous magnetic couplings are used to transmit torque between two separated members. Couplings are used for various purposes including the activation of devices in sealed containers and the transmission of instrument readings from vacuum or pressure vessels. The two most common coupling configurations are radial and axial couplings [17]. Radial couplings consist of two coaxial radially polarized cylindrical magnets (Fig. 4.51). Axial couplings consist of two axially polarized multipole disks that face one another (Fig. 4.52). For both configurations, the transmitted torque is a function of several variables
4.7 MAGNETIC COUPLINGS
FIGURE 4.49
FIGURE 4.50
287
Torque per unit length vs rotation angle (FEA : ;).
Torque per unit length vs separation distance d (FEA : ●).
288
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.51
Synchronous radial coupling.
FIGURE 4.52
Synchronous axial coupling.
289
4.7 MAGNETIC COUPLINGS
including material properties, number of poles, dimensions, separation distance, and the relative angular offset of the magnets [18]. In this section, we derive formulas for the torque of radial and axial couplings. EXAMPLE 4.7.1 Derive an expression for the torque between two coaxial, radially polarized multipole cylinders (Fig. 4.51). Assume that the lengths of the magnets are much greater than their diameters. Further assume that the inner and outer magnets have linear second quadrant constitutive relations of the form B : H ; M,
(4.214)
B : (H ; M),
(4.215)
and
respectively, where M : <M r . (4.216) Q The < sign takes into account the alternating polarity of adjacent poles. SOLUTION 4.7.1 The coupling consists of an inner and outer magnet. We refer to the inner magnet as the source magnet and the outer magnet as the drive magnet. The analysis proceeds in two steps: First, we use boundary-value theory to obtain a free-space field solution for the inner magnet; and second, the outer magnet is reduced to a distribution of equivalent currents, and the torque is determined by considering the equivalent currents to be in the external field of the source magnet. Source magnet: Because the length of the source magnet is much greater than its diameter, we can determine its field using a 2D analysis. The field due to the inner magnet was derived in Example 4.2.5. The field components are given by B(r, ) : ir\G>U(M , R , R , ) cos(i) P G G G
(4.217)
and (4.218) B (r, ) : ir\G>U (M , R , R , ) sin(i). G G G The coefficients U(M , R , R , ) are defined by Eq. (4.53), with M given by G G G Eq. (4.64). In the following analysis the field components (4.217) and (4.218) constitute an external field. Therefore, we adopt the notation B(r, ) : P B(r, ) and B (r, ) : B (r, ). P Torque: We determine the torque on the outer magnet. Let N , M , and R Q and R denote the number of poles, magnetization, and inner and outer radii of
290
CHAPTER 4 Permanent Magnet Applications
this magnet, respectively. As a first step, we reduce the magnet to a distribution of equivalent volume and surface current densities J and j , respectively K K (Section 3.3). Once this is done, we compute the torque using Eq. (3.100). Specifically, T:
r(r, ) ; (J (r, ) ; B(r, ))r dr d dz K
;
4
1
r(r, ) ; (j (r, ) ; B(r, )) da, K
(4.219)
where V and S denote the volume and surface of the magnet. From Eq. (4.216) we have J : ; M : 0. Therefore, the first term in Eq. (4.219) is zero. As K for the second term, there are N sectors to consider, and each sector has two surfaces with current densities that contribute to the torque. These surfaces constitute the radial sides of the sector at angular positions and , respectively. If the magnet is rotated by an angle , then the surface current densities for the pth sector are determined using
j (p, r, ) : K
M z Q
9M z Q
R rR (p) : ; (1 ; 2p) N
R rR (p) : 9 (1 ; 2p), N
(4.220)
where p : 0, 1, 2, . . . , N 9 1. Notice that : 0 occurs when the middle of the zeroth sector coincides with the x-axis and its polarization is radially outward as shown in Fig. 4.53. It it is assumed that the source magnet is held stationary with one of its sectors in a similar orientation; therefore, represents the relative angular offset of the two magnets. We substitute Eq. (4.220) into Eq. (4.219) and obtain
, \ 0 (91)N rB(r, (, p)) dr, P N 0 where L is the length of the coupling and T() : 2M L Q
(, p) : ; (1 ; 2p). N
(4.221)
(4.222)
The coefficient 2M takes into account the fact that there are two surfaces at the Q
291
4.7 MAGNETIC COUPLINGS
FIGURE 4.53
Cross section of radial coupling with reference frame.
interface between neighboring sectors (one for each sector). The integral in Eq. (4.221) can be evaluated numerically using Simpson’s method. The resulting torque formula is 2M LR* , \ ,P Q (91)NS (q)r(q)B(r(q), (, p)), (4.223) P P N P N O where N is the radial mesh coefficient (even), R* : R 9 R , and S (q) are P P the Simpson coefficient terms T() :
S (q) : P The integration points are
(q : 0) (q : 1, 3, 5, . . . ) (q : 2, 4, 6, . . . )
(4.224)
(q : N ). P
q r(q) : R ; (R 9 R ) (q : 0, 1, 2, . . . , N ). N P P The field component B (r, ) is given by Eq. (4.217). P
(4.225)
Calculations: We demonstrate Eq. (4.223) with some sample calculations. To clarify the analysis, we distinguish the parameters for the inner (source) and outer (drive) magnets using superscripts in and out, respectively. For example, R and R refer to the inner and outer radii of the inner source magnet, and R and R refer to the inner and outer radii of the drive magnet. We consider
292
CHAPTER 4 Permanent Magnet Applications
a coupling with the following parameters: M : 5.0;10 A/m Q M : 5.0;10 A/m Q R : 2.0 cm R : 5.0 cm R : 6.0 cm R : 9.0 cm N : 6.0. (4.226) We compute the torque per unit length for a series of angular values : 0, 5, 10, . . . , 60° (rotation from the center of one pole to the next). The computed data are compared with corresponding FEA data in Fig. 4.54. We also determine the number of poles that render the maximum torque (all other parameters held constant). These data are shown in Fig. 4.55. It is clear that the maximum torque occurs when N : 8. Finally, we compute the peak torque as a function of the aspect ratio L/D of the coupling, where L is the length of the coupling and D : 2R is the diameter of the outer magnet. The same
FIGURE 4.54
Torque per unit length vs rotation angle.
293
4.7 MAGNETIC COUPLINGS
FIGURE 4.55
Torque per unit length vs number of poles.
parameters as in the preceding are used and L is varied from D to 10D. The percent difference between the scaled 2D data and a corresponding 3D analysis is shown in Fig. 4.56. According to this analysis, the scaled 2D value is within 7% of the actual value even when the aspect ratio is 1. ) EXAMPLE 4.7.2 Derive an expression for the torque between two coaxial radially polarized multipole cylinders (Fig. 4.51) [19]. Assume that both magnets have the same number of poles and linear second quadrant constitutive relations of the form B : (H ; M).
(4.227)
where M : <M r . (4.228) Q The < sign takes into account the alternating polarity of adjacent poles. SOLUTION 4.7.2 The coupling consists of an inner and outer magnet. We refer to the inner magnet as the source magnet and the outer magnet as the drive magnet. The analysis proceeds in two steps: First, we obtain a free-space field solution for the inner magnet; second, the outer magnet is reduced to a
294
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.56
Percent error of 2D model vs aspect ratio.
distribution of equivalent currents, and the torque is determined by considering these currents to be in the external field of the source magnet. Source Magnet: The field for the inner magnet was derived in Example 4.2.6. The only field component that contributes to the torque is the radial component, which is determined using M , B (r, , z) : Q (91)Q>>H>I P 4 Q H I , ; r P S (n)r (n) sin( 9 ( j)) Q Q P Q L ;I (r, , z; r (n), ( j), z (k)) Q Q Q , ; S (m)(z 9 z ( j )) cos( 9 (m)) Q Q Q K
;I (r, , z; r (k), (m), z ( j )) . Q Q Q
(4.229)
295
4.7 MAGNETIC COUPLINGS
The various functions and parameters in Eq. (4.229) are defined in Example 4.2.6. In the following analysis this component is an external field. Therefore, we adopt the notation B (r, , z) : B (r, , z). P P Torque: We determine the torque on the outer (drive) magnet. Let N , M , Q R , R and z and z denote the number of poles (even), magnetization, inner and outer radii, and end positions of this magnet, respectively. As a first step, we reduce the magnet to a distribution of equivalent volume and surface current densities J and j , respectively (Section 3.3). Once this is done, we compute K K the torque using Eq. (3.100), T:
4
r(r, );(J (r, );B(r, ))r dr d dz K
r(r, );(j (r, );B(r, ))da, (4.230) K 1 where V and S denote the volume and surface of the magnet. From Eq. (4.228) we have J : ;M : 0 and, therefore, the first term in Eq. (4.230) is zero. As K for the second term, there are N sectors to consider, and each sector has two surfaces with current densities that contribute to the torque. Consider the sector that is centered with respect to the x axis as shown in Fig. 4.53. This sector has surface currents along its radial sides at angular positions : 9/N and : ;/N If the sector is rotated by degrees the contributing surface current densities are given by: ;
R rR M z : ; /N Q z zz j (r, , z) : (4.231) K R rR 9M z : 9 /N Q z zz . It is assumed that the inner magnet is held stationary with one of its sectors centered about the x-axis. Therefore, represents an angular offset between the two magnets We substitute Eq. (4.231) into Eq. (4.230) and solve for the axial component. This gives
X
0
rB r, ; , z dr dz, (4.232) P N X 0 where B (r, , z) is given by Eq. (4.229). The coefficient 2N reflects the fact P that each sector contains two surface currents that contribute equally to the T () : 2N M Q X
296
CHAPTER 4 Permanent Magnet Applications
torque, and that the total axial torque is the same for each sector. The integrals in Eq. (4.232) can be evaluated using Simpson’s method. The resulting torque expression is M LR* ,X ,P Q S (p)S (q) X P NN P X N O ;r(q)B r(q), ; , z(p) , (4.233) P N where L : z 9 z , R* : R 9 R , and S (q) and S (p) are the Simpson P X coefficient terms T () : X
2N
and
S (q) : P
(q : 0) (q : 1, 3, 5, . . . ) (q : 2, 4, 6, . . . ) (q : N ) P
(p : 0) (p : 1, 3, 5, . . . ) S (p) : X (p : 2, 4, 6, . . . ) (p : N ) P The integration points are as follows:
and
q r(q) : R ; (R 9 R ) N P
(q : 0, 1, 2, . . . , N ), P
p z(p) : z ; (z 9 z ) (p : 0, 1, 2, . . . , N ). N X X Calculations: We demonstrate Eq. (4.233) via a practical design problem. Consider an application that requires a torque of T : 90 Nm to be transmitted through a 0.45-cm enclosure (i.e., the coaxial magnets must have a 0.45-cm radial gap between them). Assume that the magnets are made from bonded NdFeB with M : 4.3;10, and that they have the following dimensions: Q R : 1.0 cm (inner radius of inner magnet) R : 5.0 cm (inner radius of outer magnet) R : 8.0 cm (outer radius of outer magnet) L : 5.0 cm (length).
297
4.7 MAGNETIC COUPLINGS
FIGURE 4.57
Torque vs number of poles.
We compute the peak torque as a function of the number of poles N for a series of values for the outer radius of the inner magnet, R : 3, 3.5, 4, and 4.5 cm (when R : 4.5 cm there is a 0.5-cm radial gap between the inner and outer magnet). The computed data are shown in Fig. 4.57. The analysis shows that the peak torque (which occurs at an angular offset of : /N ) occurs at 4, 4, 6, and 10 poles, respectively. Notice that only the R : 4.5-cm data has a torque of 90 Nm. Thus, a ten-pole coupling with R : 4.5 cm is optimal for this application. ) EXAMPLE 4.7.3 Derive an expression for the torque between two identical axially polarized multipole cylindrical disks as shown in Fig. 4.58 [20, 21]. The disks have inner and outer radii R and R , respectively, and a thickness t . Let K N denote the number of poles on the surface of each disk, and let h denote the spacing between disks. Assume that both magnets have a linear second quadrant demagnetization curve of the form B : (H < M z ), (4.234) Q where the < sign takes into account the alternating polarity of adjacent poles.
298
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.58
Synchronous axial coupling.
SOLUTION 4.7.3 We use the charge model (Section 3.4). Our approach is as follows: First, reduce the magnets to equivalent charge distributions; next, determine the force imparted by one magnet to the other using Eqs. (3.120) to (3.123); Last, determine the torque. The volume and surface charge densities K and are given by Eq. (3.105). From Eq. (4.234) we find that K : 9 · (<M z ) : 0, (4.235) K Q and that : M · n : <M . K Q
(4.236)
299
4.7 MAGNETIC COUPLINGS
Now, consider the force formulas (3.120) to (3.123). Because : 0 in both K magnets, the volume terms vanish and the force between the magnets reduces to
(x ) (x )(x 9 x ) I A A . F: K N K I N (4.237) I N 4
x 9 x N I N I Instead of evaluating Eq. (4.237) directly it is easier to derive the force in the following steps. First, the force between two point sources is obtained. Next, the force between opposing pole faces (sectors) is derived. Third, an equation is derived for the force between a sector of one magnet and an entire surface of the opposing magnet. Last, the total force is obtained by summing the contributions from all of the sectors and surfaces. Because torque is obtained from the force, its derivation is presented as well. Point Sources: We use cylindrical coordinates with one of the charges Q (x ) K in the x-y plane at (r , , 0), and the other Q (x ) in a plane defined by z : h K at (r , , h). The axial force between Q (x ) and Q (x ) follows from Eq. K K (3.124), Q (r , , 0)Q (r , , h)h K K F : . (4.238) X 4 [r ; r 9 2r r cos( 9 ) ; h] The axial component of torque imparted to Q (x ) by Q (x ) is K K T : (r ;F ) · z , (4.239) X where F is the force imparted to Q (x ) by Q (x ), and r is the vector from K K the origin of the coordinate system to the first particle. The torque T is given by X Q (r , , 0)Q (r , , h)r r sin( 9 ) K . (4.240) T : K X 4 [r ; r 9 2r r cos( 9 ) ; h] Sectors: Next, we compute the axial force and torque imparted to one pole face (sector) by an opposing pole face. Again, without loss of generality, one sector is taken to lie on the x-y plane symmetrically positioned about the x-axis ( : 0), while the other is taken to lie on the plane defined by z : h, symmetrically positioned about an angular displacement : (Fig. 4.59). Q The idea is to discretize each sector into a mesh of subsections (elements), define a point charge at the midpoint of each element in accordance with Eq. (4.236), and then compute the force and torque imparted to one sector by the other by summing the contributions from the individual point charges. The area of each sector is area of sector :
(R 9 R) . N
(4.241)
300
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.59
Relative angular offset of two sectors.
In addition, each sector spans an angular measure 2 : . (4.242) N The first step is to discretize the area (4.241) into a set of indexed elements. These elements are chosen to have equal areas so that they contain an equal amount of surface charge. Because the area of the sector is spanned by the two coordinate variables r and , two indices (i, j) are required. Let N and N be P the number of mesh divisions in these two variables, respectively. Then each element will have the same area A given by (R 9 R) . (4.243) N N N P Furthermore, choose each element so that it spans an angular measure given by A :
:
. N
(4.244)
Let (r , ) denote the midpoint of the element A , and let R(i) and R(i ; 1) G H GH
301
4.7 MAGNETIC COUPLINGS
denote its radial boundaries. Then r : R(i ; 1) 9 R(i), G
(4.245)
and r: G
R(i ; 1) ; R(i) 2
(i : 1, 2, . . . , N ). P
(4.246)
As the sector is centered about : , Q
: 9 ; (j 9 1) ; H Q N 2
(4.247)
for j : 1, 2, . . . , N . The endpoints are computed using R(i ; 1) :
R(i) ;
(R 9 R) N P
(i : 1, 2, . . . , N ), P
(4.248)
where R(1) : R (the inner radius). Equation (4.248) is used to generate the endpoint values from which the midpoint values are obtained. Therefore, Eqs. (4.246)—(4.248) define a scheme for discretizing the sector into elements and assigning values to the midpoints of each element. Next, consider the magnetic interaction between an element of one sector and an element of an opposing sector. Let the coordinates of the midpoints of the two elements be (r , , 0) and G H (r , , h), respectively. The magnetic charge associated with each element is GY HY Q (r , , 0) : (r , , 0)A K G H G H
(4.249)
Q (r , , h) : (r , , h)A, K GY HY GY HY
(4.250)
and
respectively. From Eqs. (4.236) and (4.243) these charges reduce to
(R 9 R) . Q : <M K Q N N N P
(4.251)
where, once again, the < term takes into account the polarity of oppositely charged sectors. The force and torque are obtained by summing the contributions of all of the elements in accordance with Eqs. (4.238) and (4.240). The resulting
302
CHAPTER 4 Permanent Magnet Applications
expressions are
M (R 9 R) ,P , F (h, ) : Q X Q 4 N N N P G H , , h ; P , [r ; r 9 2r r cos( 9 ) ; h] GY GY HY G G GY H HY
(4.252)
and T
M (R 9 R) ,P , (h, ) : Q X Q 4 N N N P G H , , r r sin( 9 ) G GY H HY . ; P [r ; r 9 2r r cos( 9 ) ; h] GY G GY H HY GY HY G
(4.253)
Torque: Next, we determine the total axial force and torque taking into account the interaction of all the sectors of both magnets. There are two magnets that are labeled upper and lower in accordance with their relative positions. These magnets are identical and have a thickness t . Each magnet has two surfaces K (labeled top and bottom) and each surface contains N sectors. The upper magnet has its bottom surface in the plane defined by z : h and its top surface lies in the plane defined by z : h ; t . Assume that this magnet is rotated by K an angle as shown in Fig. 4.60. Under this condition, the angular positions
FIGURE 4.60
Rotated magnet surface.
303
4.7 MAGNETIC COUPLINGS
of the centers of its sectors (, n) are Q 2 (, n) : ; (n 9 1) (n : 1, 2, . . . , N ). (4.254) Q N The lower magnet is centered about : 0, and has its top surface in the x-y plane, and its bottom surface in the plane defined by z : 9t . Consider the K sector of this magnet on the top surface centered about : 0. Based on the results of the previous section, the axial force on this sector due to all of the sectors on both the top and bottom surfaces of the upper magnet is , F (h, ) : (91)L[F (h, (, n)) X X Q L 9F (h ; t , (, n))]. X K Q Continuing in this fashion and taking into account the contributions from all the sectors and surfaces, the following formulas are obtained for the total force and torque,
N M (R 9 R) , ,P , F () : Q X N N N 4 P K L G H , , (91)L>K h K K , ; P [r ; r 9 2r r cos( 9 (, n)) ; h ] GY K G GY H HY GY HY G and
(4.255)
N M (R 9 R) , ,P , T () : Q X 4 N N N P K L G H , , (91)L>K r r sin( 9 (, n)) K G GY H HY ; P , [r ; r 9 2r r cos( 9 (, n)) ; h ] GY K GY HY G G GY H HY where
(4.256)
:9 ; (j 9 1) ; , H N 2 (, n) : (, n) 9 ; (j 9 1) ; , HY Q N 2 and (, n) is given by Eq. (4.254). The terms h and are given by Q K K h : h ; mt (m : 0, 1, 2) K K h : separation of magnets
t : thickness of magnets K
304
CHAPTER 4 Permanent Magnet Applications
and
1 (m : 0)
: 2 (m : 1) K 1 (m : 2). The angle is the relative offset of the two magnets. The condition : 0 occurs when poles of one magnet align with the poles of opposite polarity on the other magnet (the torque is zero when : 0). Note that Eqs. (4.255) and (4.256) contain nested summations of elementary functions, which are easy to program. However, it is important to note that when programming nested summations, a higher degree of accuracy can be achieved with fewer mesh points (lower values N and N) if the total of a single summation over one index is evaluated and P saved separately (as an intermediate step), and then the totals of the successive summations are added together to obtain the final sum. For example, Eq. (4.256) can be evaluated as follows:
N M (R 9 R) , T () : Q S (, n, m), X N N N 4 P K L where , S (, n, m) : P S (, i, n, m), G with , S (, i, n, m) : S (, i, j, n, m), H and , S (, i, j, n, m) : P S (, i, i, j, n, m), GY and , (91)L>K r r sin( 9 (, n)) K G GY H HY S (, i, i, j, n, m) : . [r ; r 9 2r r cos( 9 (, n)) ; h ] GY K HY G G GY H HY Calculations: We demonstrate Eqs. (4.255) and (4.256) with a practical design. Consider an application that requires a torque of 1.0 Nm to be transmitted through a 10-mm enclosure. To begin the analysis, choose the
305
4.7 MAGNETIC COUPLINGS
FIGURE 4.61
Torque vs number of poles.
following nominal magnet dimensions, R : 1.0 cm R : 4.0 cm t : 1.5 cm K
(inner radius) (outer radius) (thickness).
The coupling is to be made from bonded NdFeB with a 10% concentration of binder by volume which renders a bulk magnetization M : 4.3;10 A/m. The Q peak torque is computed as a function of the number of poles N for a series of separations h : 10, 12, 14, and 16 mm. The analysis is performed with N P and N set to 12. We find that the peak torque (which occurs at an angular offset of : /N ) occurs at 6, 4, 4, and 4 poles, respectively (Fig. 4.61). Notice that the magnets can be separated by as much as 14 mm and still render the desired torque. To complete the analysis, the torque and force are computed as a function of for the h : 14 mm separation (Fig. 4.62). The maximum force is approximately 60 N and is attractive (negative) at : 0° and repulsive (positive) at 90°. )
306
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.62
Torque and force vs .
4.8 MAGNETIC RESONANCE IMAGING Magnetic resonance imaging (MRI) is widely used for noninvasive diagnosis of the human body. In this section, we briefly review the principles of MRI, and the role that permanent magnets play in this technology. The MRI process itself is relatively complex and we present a brief and simplified description of it here. The interested reader can find more detailed discussions of MRI in numerous texts [22—24]. In MRI, radio frequency (rf) signals are used to create sectional images of the body. To obtain an image, a patient is first placed in a cavity (imaging region) that is permeated by a strong magnetic field. The nuclear magnetic moments within the patient’s tissue tend to align either parallel or antiparallel to the field, with the former having a lower energy than the latter (Fig. 4.63b). In a given specimen there are a large number of magnetic moments and most of the parallel moments are cancelled by the antiparallel moments leaving a relatively small number of unpaired parallel moments. The interaction between the magnetic
4.8 MAGNETIC RESONANCE IMAGING
307
moments and the field causes them to precess about the field direction at an angular frequency (Larmor frequency), : B , (4.257) where is the nuclear gyromagnetic ratio and B is the applied flux density. The magnetic moment of the hydrogen nucleus (proton) is used for MRI. Hydrogen is abundant in biological tissues, and the proton has a relatively high gyromagnetic ratio ( : 42.6 MHz/T). The Larmor frequency for hydrogen is 64 MHz when B : 1.5 T. Once the nuclear moments are aligned, the specimen is subjected to a pulsed rf signal that is tuned to the Larmor frequency. The rf signal induces transitions between the lower- and higher-energy proton spin populations (Fig. 4.63c). As the moments transition from the higher- to the lower-energy state they generate an rf signal that is sensed by a coil
Orientation of nuclear magnetic moments: (a) random orientation, (b) alignment in a B-field; (c) reorientation due to absorption of rf energy; and (d) reorientation due to reradiation.
FIGURE 4.63
308
CHAPTER 4 Permanent Magnet Applications
and processed to create sectional 3D images (Fig. 4.63d). Spatial imaging can be accomplished by use of additional gradient field coils that have fields that superimpose with the primary field to produce a predefined variation in the bias field across the specimen. This gives rise to a spatial variation of the Larmor frequency, which in turn gives rise to a variation of the generated rf signal and which can be processed to form an image. There are essentially three different magnetic field systems used in MRI apparatus: 1. conventional electromagnets; 2. superconducting coils; and 3. permanent magnet structures. Conventional electromagnets consist of multiple turns of low gauge wire or conducting ribbons usually used in conjunction with soft magnetic cores and pole pieces. Electromagnet coils have a relatively high resistance and therefore they consume large amounts of energy and this causes heating. The heating sets a practical limit on the working current of such coils and, therefore, limits magnetic field strength. On the other hand, superconducting coils are capable of producing very stable and uniform fields of up to 4.0 T. However, they require liquid cryogens and cooling apparatus to suppress their resistance. Such coils are, therefore, expensive and require extensive preparation and maintenance. Permanent magnet systems are competitive with current-driven coils for MRI applications that require fields in the range of 0.1—0.5 T. They consume no energy and require little maintenance. When designing an MRI magnetic field system, the following parameters need to be considered: 1. 2. 3. 4. 5.
magnetic field strength; the homogeneity of the magnetic field; the temporal stability of the magnetic field; fringing fields; and patient access to the imaging volume.
A wide range of field strengths can be used for MRI. For example, whole body MRI has been achieved with fields ranging from 0.02 T (0.85 MHz) to 4.0 T (170 MHz). Higher field strengths are usually desirable because the number of transitions from high- to low-energy states scales approximately linearly with field strength. The three magnetic field systems can be ranked in terms of their field strength as follows: superconducting coil (3 T); conventional electromagnet (0.2—0.4 T); and permanent magnets (0.1—0.5 T).
4.8 MAGNETIC RESONANCE IMAGING
309
A high degree of field uniformity is required for MRI. This is defined as the normalized maximum deviation in field strength between two points in the imaging volume B B 9 B
. : B B The field homogeneity is usually specified in ‘‘parts per million’’ for a given diameter of spherical volume (dsv). It can also be expressed as a ratio of frequencies using the Larmor relation (4.257). In this case it is given by f 9f f
: . f f Typical requirements for MRI field homogeneity are 10 to 100 ppm over a spherical imaging volume with a diameter of 50 cm (50 dsv). Superconducting coil systems have the highest degree of field uniformity (10 ppm, 50 dsv), followed by permanent magnet systems (10 ppm, 20 dsv), and then conventional electromagnet systems (30 ppm, 40 dsv). In MRI, data acquisition occurs over time. Therefore, it is important that the magnetic field have a high degree of temporal stability. Instability can be caused by variations in the field source itself, or by time dependent factors such as the movement of external field sources or ferromagnetic materials. Superconducting coils are the most stable field source. On the other hand, conventional electromagnets are susceptible to variations in their power source. Permanent magnets also exhibit field instability due to fluctuations in the ambient temperature. In this regard, SmCo magnets have an advantage over NdFeB due to their lower temperature coefficient (Section 1.20). The magnetic fringe field is another factor that must be considered when evaluating MRI magnet systems. A related issue is magnetic shielding. The fringe field can have undesired effects on medical equipment in proximity to the MR imager. Conversely, external field sources or the movement of ferromagnetic materials can degrade MR image quality if the system is not properly shielded. These effects can be reduced using active or passive shielding. In active shielding, additional coil sets are arranged inside the imager and activated to counter the fringe field. In passive shielding, a soft magnetic enclosure is used to confine the field and to prevent stray fields from penetrating the imaging cavity. The final factor we consider is patient access to the imaging region. An MRI magnet system must provide an imaging region that can
310
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.64
Comparison of MRI magnetic field systems [25].
accommodate the patient (or body part), a supporting structure (couch), and the rf coils. This space must be sufficient to provide minimal physical and psychological comfort for the patient. For example, in whole body MRI the magnet system cavity is cylindrical with a diameter of 1.0 m. However, the free access diameter for imaging is approximately 0.75 m. In pediatric MRI, the corresponding diameters are 0.5 and 0.39 m, respectively. For research applications, the magnet cavity has a diameter ranging from 0.2 to 0.4 m, and the imaging region has a diameter of 0.135 to 0.31 m. The characteristics of the three MRI magnetic field systems are summarized in Fig. 4.64. As already noted, permanent magnet systems can be used for MRI with field requirements that range from 0.1 to 0.5 T. The main advantage of these systems is that they consume no power. Moreover, when rareearth magnets are used the size and weight of such structures are greatly reduced relative to conventional materials. The MRI permanent magnet systems are classified as either yokeless, yoked, or hybrid. Yokeless systems consist of permanent magnets alone. The imaging field is confined and focused by the magnetization pattern within the magnets. The ring dipole arrangement shown in Fig. 4.65a is an example of a yokeless system [25]. This is composed of eight permanent magnet segments with magnetizations indicated by the solid
4.8 MAGNETIC RESONANCE IMAGING
FIGURE 4.65
311
Two MRI magnet systems: (a) ring dipole; and (b) H-frame [25].
arrows. This assembly provides a transverse magnetic field across the imaging region as indicated by the dotted lines. The field generated by these systems is typically 0.3 T. In yoked systems a high-permeability material surrounds the permanent magnet structure. This enhances the field in the imaging region and minimizes the fringe field. The H-frame structure shown in Fig. 4.65b is an example of a yoked system [25]. In this structure, permanent magnet
312
CHAPTER 4 Permanent Magnet Applications
blocks are supported in a ferromagnetic frame called a yoke. The yoke not only supports the magnets but also serves as a flux return path for the imaging field. Soft magnetic pole pieces are also used to enhance the strength and uniformity of the imaging field. The H-frame structure is the design of choice for MRI despite its weight and size. Hybrid structures contain yoked and yokeless sections. The design of MRI magnet structures is complicated. In the design of conventional magnet structures, the analyst solves the ‘‘direct’’ problem in which the geometry and magnetization are given, and the field distribution is determined. In MRI, the analyst is faced with the more difficult ‘‘inverse’’ problem in which the field strength and uniformity are specified across the imaging region and the geometry and magnetization of the structure need to be determined. There is no unique solution to such problems in that a specified field distribution within a closed region can be obtained using an infinite number of different structures [26]. One design approach is to select one of the many ideal closed structures that give the desired interior field distribution and then remove a portion of the material to allow access to the cavity. These openings give rise to undesired field nonuniformities that must be compensated for. Remedies for this include the modification of the geometry or magnetization, and/or the insertion of additional components such as high-permeability materials.
4.9 ELECTROPHOTOGRAPHY Electrophotography is one of the most widely used methods for image reproduction and printing [27, 28]. The central element of this process is the photoconductor (PC), which consists of a thin sheet (10—50 m) of photosensitive material that is coated onto a ground plane. The PC generates electron-hole pairs as it absorbs incident light. Once generated, the electrons and holes separate and move under the influence of an internal electric field with the electrons moving to ground. This process forms the bases of electrophotographic image reproduction. The electrophotographic process comprises six distinct steps: Charge; Expose; Develop; Transfer; Fuse; and Clean (Figs. 4.66 and 4.67). We briefly review these steps in the order in which they occur: 1. Charge: The first step in the electrophotographic process entails charging the PC. The PC (which is initially uncharged) passes through a corona-charging device. This device creates ionized gas
4.9 ELECTROPHOTOGRAPHY
FIGURE 4.66
313
Electrophotographic system [27].
molecules, which are absorbed by the PC and thereby create a uniform charge distribution on its surface. This charge distribution gives rise to an electric field within the PC (Fig. 4.67(1)). 2. Expose: To form an image, the charged surface of the PC is exposed to light that is reflected from an existing image or produced by a laser writer. A portion of the incident light is absorbed in the PC and electron-hole pairs are created in the exposed regions. These charged carriers separate and drift under the influence of the PC’s internal field. The electrons move to ground and the holes move to the charged surface where they discharge the exposed regions. This produces a charge pattern on the exposed surface of the PC that mirrors the image. This is called the latent image (Fig. 4.67(2)). 3. Develop: During this step, charged pigmented toner particles are transported proximal to the latent image and are electrostatically attracted to regions of the PC in accordance with the local surface charge (Fig. 4.67(3)). This transforms the latent image into a real image.
314
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.67
Electrophotographic process.
4. Transfer: In the transfer step, the toner on the PC is electrostatically transferred to paper. This is accomplished by depositing a uniform charge of opposite polarity to that of the toner on the back side of the paper prior to it contacting the PC (Fig. 4.67(4)).
4.9 ELECTROPHOTOGRAPHY
315
5. Fuse: During this step, the image is permanently fixed to the paper by melting the toner onto the paper surface (Fig. 4.67(5)). 6. Clean: In this final step, the PC is discharged and cleaned of any excess toner using coronas, lamps, brushes, and/or scraper blades (Fig. 4.67(6)). Once this is done the PC is ready for the next imaging cycle. We are primarily interested in the development of the latent image since this is the most magnetically intensive step of the process. Most electrophotographic copiers and printers use ‘‘magnetic brush’’ technology to implement this step. While there are various implementations of this technology, they all perform essentially the same function, namely the transportation of developer material from a reservoir to the latent image. The developer material is usually either monocomponent or two component. In monocomponent systems, the toner particles are magnetic and on the order of 5—30 m in diameter. The magnetic transfer field couples directly to these particles. In two component materials, there are magnetic ‘‘carrier’’ particles and nonmagnetic toner particles. The carrier particles range from 70 to 250 m in diameter and are much larger than the toner particles. The magnetic transfer field couples to the carrier particles and several toner particles are transported by each carrier particle. A typical magnetic brush system is shown in Fig. 4.68. In this system the magnetic toner/carrier material is transported to the latent image by
FIGURE 4.68
Magnetic brush development station [28].
316
CHAPTER 4 Permanent Magnet Applications
riding on the surface of a conductive cylindrical shell. The conductive shell encloses a rotating radially polarized permanent magnet cylinder. The rotating magnet gives rise to a time-varying magnetic field at the surface of the shell, which is typically a few tenths of a tesla or lower. This time-varying field provides a magnetic transport force that causes the toner particles to move in the opposite direction as shown. In magnetic brush systems the magnetic transport force dominates the centripetal, gravitational and electrostatic forces. When two magnetized toner/carrier particles approach one another they exert a dipole-dipole force on each another. This force is far greater than all the other ‘‘ambient’’ forces, including the magnetic force exerted by the magnetic brush. Specifically, the dipole-dipole force is typically more than an order of magnitude greater than the force due to the magnetic brush [27]. Consequently, coaligned neighboring toner/carrier particles tend to adhere to one another, forming chains of particles (Fig. 4.69. During transport, these chains behave as a single mechanical entity. They ride on the surface of the conductive shell and cartwheel end over end as they are transported. The dynamics of this are complicated and beyond the scope of this text. We simply derive here the force on an isolated toner particle in the field of the magnetic brush. For this, we need the field components for an infinite, radially polarized multipole
FIGURE 4.69
Monocomponent toner transport [28].
317
4.9 ELECTROPHOTOGRAPHY
cylinder. These were derived in Example 4.2.5. The radial and azimuthal components are given by Eqs. (4.66) and (4.67), respectively, B (r, ) : ir\G>U (M , R , R , ) cos(i) P G G G
(4.258)
and B (r, ) : ir\G>U (M , R , R , ) sin(i), (4.259) G G G where U (M , R , R , ) is given by Eq. (4.65) with M given by (4.64) (the G G G superscript 2 has been dropped). In the following example we derive the force on a magnetic particle in the field of a magnetic brush. EXAMPLE 4.9.1 Determine the force on a spherical toner particle of radius R in the field of a cylindrical magnetic brush. Assume that the toner is made from a magnetically linear and isotropic material with permeability and susceptibility . K SOLUTION 4.9.1 The force on the toner particle is computed using Eq. (3.116). We assume that (M · )B is constant over the volume of the toner particle. This gives F:
(M · )B dv 4 (m · )B . where m is the magnetic dipole moment of the toner particle, and B is the field of the magnetic brush (multipole cylinder). Specifically, : B (r, )r ; B (r, ) , P where B (r, ) and B (r, ) are given by Eqs. (4.258) and (4.259), respectively. P From Example 1.8.1 we know that the dipole moment of the toner particle is B
4R( 9 ) H . ( ; 2 )
(4.260)
4 R( 9 ) (B · )B . ( ; 2 )
(4.261)
m: Thus, we find that F:
318
CHAPTER 4 Permanent Magnet Applications
We evaluate Eq. (4.261) and obtain F(r, ) :
4R( 9 ) ( ; 2 ) ;
B (r, ) P
; B (r, ) P
B (r, ) 1 B (r, ) P P ; B (r, ) r r r
B (r, ) 1 B (r, ) ; B (r, ) . r r
(4.262)
Finally, we substitute Eqs. (4.258) and (4.259) into Eq. (4.262) and find that 4 R( 9 ) ( ; 2 ) ; ii(i ; 1)r\G>GY>U U cos(i) cos(i) G GY G GY ; iiYr\G>GY>U U sin(i) sin(i) r G GY G GY ; ii( i ; 1)r\G>GY>U U cos(i) sin(i) G GY G GY 9 iir\G>GY>U U sin(i) cos(i) , (4.263) G GY G GY where U (M , R , R , ) is given by Eq. (4.65) with M given by Eq. (4.64). G G G )
F(r, ) :9
4.10 Magnetooptical Recording Magnetooptical (M-O) recording is routinely used for digital data storage [29—31]. Just as in magnetic storage, the information is recorded along tracks on data disks. A track is a narrow annulus region a distance r from the center of the disk. Each track is divided into a number of smaller segments or sectors and each segment stores a single block of data of typically either 512 or 1024 bytes. In conventional M-O systems, data retrieval/storage is accomplished using a read/write head that is held stationary while the disk spins (Fig. 4.70). Typical disk rotation speeds are 1200 and 1800 rpm for slower drives, and 3600 rpm for high data-rate systems.
4.10 MAGNETOOPTICAL RECORDING
FIGURE 4.70
319
Magnetooptic recording system.
An M-O storage disk consists of a multilayer stack of thin films. The conventional quadrilayer structure is shown in Fig. 4.71. This structure is optically designed to provide near-optimum readback. The functional storage layer in this stack is a thin magnetic film (250 Å) that is currently made from a rare-earth transition-metal (RE-TM) material, preferably an alloy based on Tb, Fe, and Co. This material is characterized by a vertical magnetic anisotropy and a high magnetic coercivity on the order of several tenths of a tesla (several kilo-gauss at room temperature). Because of the high coercivity, the film is magnetically very stable and
FIGURE 4.71
An M-O recording thin film stack [31].
320
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.72
Magnetooptic recording process.
can retain a recorded magnetization pattern indefinitely under normal operation. However, this layer tends to be chemically unstable due to the high affinity of rare-earth elements for oxygen. In practice, it is sandwiched between two chemically stable dielectric layers as shown. Data is recorded in the RE-TM layer using a thermomagnetic process. In conventional systems, recording occurs with the media moving (linear velocity < 10 m/s) between a stationary laser source and a fixed bias field source (Fig. 4.72). The film is initially polarized in one direction (e.g., the down direction). To record information, the laser is pulsed (duration < 50 ns) and focused so as to heat a local spot of the film (diameter < 1m). As the temperature of the spot increases the coercivity within it decreases. Eventually, the spot reaches the Curie temperature (T < 473 K) at which point the local magnetization reverses A and aligns with the bias field, which is typically 0.02—0.04 T across the film. The temperature at the center of the spot can reach a value as high as 500 K, which, while sufficient to reverse the magnetization, is not high enough to damage the film. When the laser is turned off, the temperature decreases back to room temperature, and the magnetization within the spot remains reversed. Thus, the spot represents a recorded bit of information. By repeating this process, regions of the disk are selectively reversed-magnetized and the resulting magnetization pattern represents the recorded information. Data readout in M-O systems is based on the magnetooptic Kerr effect. Specifically, when linearly polarized light is normally incident on a perpendicular magnetic medium its plane of polarization undergoes a slight rotation upon reflection (Fig. 4.73). This rotation is known as the I polar Kerr effect and the direction of rotation depends on the orientation
4.10 MAGNETOOPTICAL RECORDING
321
Kerr effect: (a) reflection from a north pole; and (b) reflection from a south pole [30].
FIGURE 4.73
of the magnetization within the film. To read data, a polarized laser beam is applied to the disk as it rotates and the reflected light is collected and processed. The same laser source is usually used for both the read and write process. However, the output power is substantially reduced for readout so as to avoid erasure. To erase information, the bias field is reversed and the film is heated at a constant power level until an entire sector is erased. The disk makes at least one revolution between the erasure of an old block and the recording of a new one. Therefore, the bias field reversal required for the erase and rerecord sequence can be implemented over a relatively long
322
CHAPTER 4 Permanent Magnet Applications
FIGURE 4.74
Cross section of a rotary actuator.
time frame (16 ms at 3600 rpm). It is convenient and common to use permanent magnet actuators for the bias field mechanism. In these actuators the magnet that provides the bias field is also a functional part of the actuator. A cross-sectional view of a rotary bias field actuator is shown in Fig. 4.74. This consists of a rotating bipolar cylindrical bias magnet surrounded by a stationary concentric conductive shell, and a drive coil. When the coil is energized it rotates the magnet to one of two equilibrium positions, thereby producing a bias field of desired polarity across the recording media. Rotation of the magnet gives rise to eddy currents in the conductive shell that impart a drag torque and reduce oscillations of the magnet about its equilibrium positions. We present an analysis of this actuator in Example 5.11.1. We work through a design of the bias magnet in the following example [32]. EXAMPLE 4.10.1 Determine the dimensions of a cylindrical bias field magnet for use in an M-O system (Fig. 4.10). The magnet must provide a bias field of between 0.02 and 0.04 T along a radial length of the disk which is taken to be 100 mm. Assume that the axis of the magnet is positioned 4 mm below the recording layer and aligned with the radial length of the disk. Assume that the magnet has a second quadrant constitutive equation B : (H ; M ). (4.264) Q Let R, z and z denote the radius and respective end points of the cylinder. SOLUTION 4.10.1 For the erase/rerecord sequence the magnet needs to be rotated between two equilibrium positions. We choose these positions to be when the magnetization is at <85° with respect to the x-axis (Fig. 4.74). The magnetization needs to be offset from the vertical in this fashion so that it can couple to the coil’s field (which is predominately vertical).
323
4.10 MAGNETOOPTICAL RECORDING
Field Solution: A field solution for this magnet was derived in Example 4.2.3. Only the radial component is of interest, and it is repeated here for convenience, M R , B (r, , z) : Q (91)I>S (n) cos((n)) P 2N L I ;[r 9 R cos( 9 (n))]I(r, , z; R, (n), z ), I where N is the Simpson’s method mesh number (even), (n) :
n 2 N
(4.265)
(n : 0, 1, 2, . . . , N),
and
(n : 0) (n : 1, 3, 5, . . . ) S (n) : (n : 2, 4, 6, . . . ) (n : N). The function I(r, , z; R, , z ) is given by I F(r, , z; r, , z ) I if
IY
where
r ; r 9 2rr cos( 9 ) " 0 or 91 2(z 9 z ) I if r : r, cos( 9 ) : 1, z " z , I
(z 9 z )g(r, , z; r, , z ) I I . F(r, , z; r, , z ) : I r ; r 9 2rr cos( 9 ) We want to evaluate the field along the y-axis. For a given observation point (0, y, z) on this axis the field is determined by substituting r : y, and : /2 9 in Eq. (4.265), that is, K B (y, , z) : B y, 9 , z , (4.266) W K P K 2
where is the angular position of the north pole of the magnet relative to the K
324
CHAPTER 4 Permanent Magnet Applications
x-axis ( is positive in the counterclockwise direction). We use Eqs. (4.265) K and (4.266) to design the magnet. Calculations: For the design, we assume that the magnet is made from sintered NdFeB material with a uniform magnetization of M : 7.2;10 A/m Q (B : 9000 G). First, we determine a viable magnet radius R. The length of the P magnet is set to 100 mm (z : 950 mm, z : 50 mm) and Eq. (4.265) is evaluated for a range of R values, 0.5 mm R 1.5 mm with : 85°, r : 4 K mm, and z : 0. A plot of B vs R is shown in Fig. 4.75. From this plot it is P obvious that a magnet with R : 1 mm will suffice because its field is 0.028 T, which is within the desired range. Having determined a viable radius, we examine the angular dependency of the field. Specifically, we compute the field with R : 1 mm, r : 4 mm, and z : 0 for a series of angular offsets relative to the y-axis, 0 90° 9 20°. This corresponds to a magnet rotation K 70° 90°. These data are shown in Fig. 4.76. Note that the field is within K the desired operating range over this entire angular range. Thus, there is considerable tolerance in the angular position of the magnet. The next step is the determination of a viable length for the magnet. For this, we set R : 1 mm
FIGURE 4.75
The M-O bias field vs magnet radius R (h : 100 mm).
4.11 FREE-ELECTRON LASERS
325
The M-O bias field 4 mm above the magnet vs angular offset 90° 9 (z : 0). K
FIGURE 4.76
and vary the length h as follows: h : 100, 104, 108 mm (954 mm z 950 mm, and 50 mm z 54 mm). The bias field is computed with : 85° at K a distance r : 4 mm above the magnet along a series of points that span its length. A plot of B for h : 100, 104, 108 mm is shown in Fig. 4.77. Note that P the field data is plotted for positive values of z only (the data for negative z are the mirror image). We find that a magnet length of h : 108 mm will suffice because the field is virtually constant along the specified length of 100 mm (0 z 50 mm). In summary, the analysis shows that a sintered NdFeB magnet with a radius R : 1 mm and length h : 108 mm rotated an angle of <85° with respect to the x-axis will provide a uniform bias field of <0.028 T over a radial length of 100 mm. )
4.11 FREE-ELECTRON LASERS A free-electron laser (FEL) is a device that generates tunable, coherent and high-power electromagnetic radiation from a beam of high velocity
326
FIGURE 4.77
CHAPTER 4 Permanent Magnet Applications
The M-O bias field 4 mm above the axis of the magnet vs z.
electrons that moves through a spatially periodic magnetic field (Fig. 4.78). The periodic magnetic field is provided by a structure that is typically a few meters in length and is referred to as a wiggler or undulator. The beam is created using a variety of apparatus including rf linear accelerators, microtrons, storage rings, and electrostatic accelerators. The basic operation of an FEL is relatively straightforward [33, 34]. The electron beam is injected at one end of the wiggler and travels down its length. Once inside the wiggler, the electrons experience a spatially varying magnetic field that causes them to oscillate (wiggle or undulate) back and forth in a plane perpendicular to the magnetic field. This oscillatory motion causes the electrons to radiate electromagnetic energy and as the electrons are randomly positioned in the beam, the radiation is incoherent. This corresponds to spontaneous radiation in a conventional laser. Moreover, just as in a conventional laser, the radiation is amplified and made coherent by placing the wiggler between two parallel mirrors that form a resonant cavity for the emitted radiation` The electromagnetic wave set up by the mirrors interacts with the electron
327
4.11 FREE-ELECTRON LASERS
FIGURE 4.78
Free-electron laser [33].
beam to promote amplification and coherence of the radiation. One of the mirrors is partially transparent, and the radiation exits this mirror in the form of a coherent laser beam. The wavelength of a FEL depends on the energy of the electrons, and the period and magnitude of the magnetic field. We derive an expression for the wavelength by considering the behavior of a single electron as it moves through the wiggler (without the mirrors). Let L denote the U length of the wiggler, and let denote the spatial period of its magnetic U field ( is the distance over which the magnetic field undergoes one U complete cycle and returns to its original value). The number of cycles of the field along the wiggler is N : L / . U U U The electron moves with a speed v : #c where # is a constant and c is the speed of light. Once it enters the wiggler, the electron radiates energy down the wiggler at the speed of light. At any time t, the wavefront of the radiation is a distance d : (c 9 v)t ahead of the electron. The electron travels the length of the wiggler in a time : L /v. U As it passes through the wiggler it undergoes N oscillations, each of U which generates one cycle of radiation. Therefore, when the electron exits the wiggler the wave packet of radiation that it generates contains N cycles and spans a distance d : (c 9 v) in front of it. The wavelength U
328
CHAPTER 4 Permanent Magnet Applications
of the radiation is therefore *
(c 9 v) N U (c 9 v) : U L U (1 9 #) : . U #
: *
(4.267)
As the electron is moving close to the speed of light, # 1 and Eq. (4.267) can be rewritten as : U, * 2
(4.268)
where :
1 . (1 9 #)
The derivation of Eq. (4.268) ignores the additional path length that the electron travels due to its oscillatory motion. In wigglers with moderate to strong fields this can be significant. The longer path length results in a reduced velocity v, which produces a longer wavelength than predicted by Eq. (4.268). A detailed analysis of this effect yields a wavelength
m c eB M U U : U 1; , (4.269) * 2 E 2m c where E, e, and m are the energy, charge, and rest mass of the electron, respectively, and B : +B, is the rms average magnetic field in the U wiggler (the average is taken along the axis of the wiggler). To obtain Eq. (4.269) we have used the relation E : m c. The laser frequency is f : c/ . * * From Eq. (4.269) we see that the wavelength (frequency) of an FEL can be tuned by changing the period and magnitude of the wiggler field, or the electron energy. For example, if : 10 cm, and eB /2m c : 1, U U U the FEL emission can range from the infrared (IR) to the ultraviolet (UV) by changing E from 100 to 1000 MeV. In practice, the wiggler field is made as strong as possible to maximize the electron’s oscillatory motion and radiation. A typical wiggler field is on the order of 0.5 T for : 2 U cm. In practice, the term eB /2m c is on the order of unity. U U
4.11 FREE-ELECTRON LASERS
FIGURE 4.79
329
Cross section of a permanent magnet wiggler.
Wigglers can be made using permanent magnets or electromagnets. Permanent magnet wigglers are made using rare-earth materials such as SmCo or NdFeB. One such wiggler is shown in Fig. 4.79. This configuration of magnets is often referred to as the Halbach configuration [35]. In this wiggler there are two rows of magnets arranged parallel to one another and separated by a distance 2h. The electrons travel along the axis of the wiggler (dotted line), and oscillate in a plane perpendicular to the paper. It is desirable to have the magnetic field uniform over the length of oscillation and, therefore, we assume that the magnets are of infinite extent in this direction (perpendicular to the page). The periodic wiggler field is obtained by rotating the magnetization of adjacent magnets by 2/N, where N is the number of magnets per wiggler period in one array. In total, 2N magnets are needed to form one period of the wiggler (N magnets in each row). The magnetic field for the Halbach wiggler configuration can be obtained using complex variable theory [36, 37]. The field solution is B* (x, y) : B (x, y) 9 iB (x, y) U V W sin(-n/N) : 2iB cos(nk (x ; iy)) P U n/N N ;[1 9 exp(9nk L)] exp(9nk h), U U
(4.270)
330
CHAPTER 4 Permanent Magnet Applications
where B is the remanent magnetization of the magnets, k : 2/ , P U S n : 1 ; pN, and i : (91. It should be noted that our notation is slightly different than that found in the literature. Specifically, the variable N that we use is denoted by M in the literature [33, 35]. The field solution (4.270) is dominated by the fundamental term in the series (p : 0), B*(x, y) : 2iB U P
sin(-/N) /N
;[1 9 exp(9k L)] exp(9k h) cos(k (x ; iy)). (4.271) U U U In the limit as L ; -, N ; - and - : 1, B* (x, y) obtains a maximum U on-axis value, B : 2B exp(92h/ ). U P U Another example of a permanent magnet wiggler is shown in Fig. 4.80. This wiggler consists of a sequence of cylindrical ring structures that are formed using pie-shaped magnet segments, each with a different orientation of magnetization. Each magnet segment spans the same angular measure (in this case 2/8 radians). The ring structures are sequentially rotated by this angular measure as shown. This produces a helical field variation along the axis of the wiggler where the electron beam flows. There are also hybrid wiggler configurations that are formed using permanent magnets and soft magnetic materials. One such wiggler is shown in Fig. 4.81. The soft magnetic pole pieces are made from a high-
FIGURE 4.80
Permanent magnet wiggler with a helical field [33].
331
4.11 FREE-ELECTRON LASERS
FIGURE 4.81
Hybrid wiggler [33].
grade material such as Vanadium permendur, which has high permeability and flux capacity. The pole pieces are designed to operate at saturation, which tends to smooth out the nonuniformities in the permanent magnets. Hybrid wigglers provide more field uniformity and a higher peak field than permanent magnet wigglers. Finally, wiggler fields can also be produced by (current-driven) electromagnets [33]. In fact, the first FEL used a superconducting wiggler. Electromagnets have advantages over permanent magnets in that they are adjustable. However, their disadvantages include cost, complexity, maintenance, and the fact that their fields do not scale well for short wiggler periods. References 1. Furlani, E. P., Reznik, S., and Janson, W. (1994). A three-dimensional field solution for bipolar cylinders, IEEE Trans. Mag. 30 (5): 2916. 2. Mayergoyz, I. D., Furlani, E. P., and Reznik, S. (1993). The computation of magnetic fields of permanent magnet cylinders used in the electrophotographic process, J. Appl. Phys. 73 (10): 5440. 3. Furlani, E. P., Reznik, S., and Kroll, A. (1995). A three-dimensional field solution for radially polarized cylinders, IEEE Trans. Mag. 31 (1): 844. 4. Furlani, E. P. (1995). Formulas for computing the field distributions of cylindrical shells with axial polarization, Int. J. Appl. Elec. Mech. 6 (2): 103. 5. Furlani, E. P. (1994). A three-dimensional field solution for axially-polarized multipole disks, J. Magn. Magn. Mat. 135: 205.
332
CHAPTER 4 Permanent Magnet Applications
6. Leupold, H. A. and Potenziani, II, E. (1991). Augmentation of field uniformity and strength in spherical and cylindrical magnetic field sources, J. Appl. Phys. 70 (10): 6621. 7. Leupold, H. A., Tilak, A. S., and Potenziani, II, E. (1993). Multi-tesla permanent magnet field sources, J. Appl. Phys. 73 (10): 6861. 8. Leupold, H. A., Tilak A. S., and Potenziani, II, E. (1993). Adjustable multitesla permanent magnet field sources, IEEE Trans. Mag. 29: 2902. 9. Leupold, H. A., Tilak, A. S., and Potenziani, II, E., (1994). Laminar construction of spheroidal field sources with distortion-free access, J. Appl. Phys. 76 (10): 6859. 10. Yonnet, J. P. (1981). Permanent magnet bearings and couplings, IEEE Trans. Mag. 17 (1): 1169. 11. Frazier, R. H., Gilinson, P. J., and Oberbeck, G. A. (1974). Magnetic and Electric Suspensions, Cambridge, MA: MIT Press. 12. Sinha, P. K. (1987). Electromagnetic Suspension: Dynamics & Control, IEE Control Engineering Series 30, London, UK: Peter Peregrinus Ltd. 13. Moon, F. C. (1984). Magneto-solid Mechanics, New York: Wiley. 14. McCaig, M. and Clegg, A. G. (1987). Permanent Magnets in Theory and Practice, 2nd ed., New York: John Wiley and Sons. 15. Furlani, E. P. (1993). A formula for the levitation force between magnetic disks, IEEE Trans. Mag. 29 (6): 4165. 16. Furlani, E. P. (2000). Analytical analysis of magnetically coupled multipole cylinders, J. Phys. D: Appl. Phys. 33: 28. 17. Furlani, E. P. (1996). Analysis and optimization of synchronous magnetic couplings, J. Appl. Phys. 79 (8): 4692. 18. Charpentier, J. F. and Lemarquand, G. (1999). Optimal design of cylindrical air-gap synchronous permanent magnet couplings, IEEE Trans. Mag. 35 (2): 1037. 19. Furlani, E. P. (1995). A three-dimensional model for computing the torque of radial couplings, IEEE Trans. Mag. 31 (5): 2522. 20. Furlani, E. P. (1993). Formulas for the force and torque of axial couplings, IEEE Trans. Mag. 29 (5): 2295. 21. Wang, R. and Furlani, E. P. (1994). Design and analysis of a permanent magnet axial coupling using 3d finite element field computations, IEEE Trans. Mag. 30 (4). 22. Carlton, R. R. and Adler, A. M. (1992). Principles of Radiographic Imaging: An Art and a Science, Albany, New York: Delmar Publishers. 23. Foster, M. A. (1984). Magnetic Resonance in Medicine and Biology, NY: Pergamon Press. 24. Kuperman, V. (2000). Magnetic Resonance Imaging: Physical Principles and Applications, New York: Academic Press. 25. Thomas, S. R. (1993). Magnets and gradient coils: types and characteristics, The Physics of MRI, M. J. Bronskill and P. Sprawls, eds., New York: American Institute of Physics, 56.
4.11 FREE-ELECTRON LASERS
333
26. Abele, M. G. (1993). Structures of Permanent Magnets: Generation of Uniform Fields, New York: John Wiley & Sons. 27. Williams, E. M. (1984). The Physics and Technology of Xerographic Processes, New York: John Wiley & Sons. 28. Schein, L. B. (1988). Electrophotography and Development Physics, 2nd ed., New York: Springer Verlag. 29. Marchant, A. B. (1990). Optical Recording: A Technical Overview, Reading, MA: Addision-Wesley Pub. Co. 30. Mansuripur, M. (1995). The Physical Principles of Magneto-optical Recording, Cambridge, England: Cambridge Univ. Press. 31. McDaniel, T. W. and Victora, R. H. (1997). Handbook of Magneto-Optical Recording; Materials, Subsystems, Techniques, Westwood, New Jersey: Noyes Pub. 32. Furlani, E. P. and Williams, C. (1995). Formulas for optimizing the bias field for magneto-optic recording, J. Magn. Magn. Mat. 148: 475. 33. Brau, C. A. (1990). Free-Electron Lasers, New York: Academic Press. 34. Svelto, O. (1989). Principles of Lasers, 3rd ed., New York: Plenum Press. 35. Halbach, K. (1985). Application of permanent magnets in accelerators and electron storage rings, J. Appl. Phys. 57 (1): 3605. 36. Halbach, K. (1980). Design of permanent magnets with oriented rare earth cobalt material, Nucl. Instum. Method 169: 1. 37. Halbach, K. (1981). Physical and optical properties of rare earth cobalt magnets, Nucl. Instum. Method 187: 109.
a
This Page Intentionally Left Blank
CHAPTER
5
Electromechanical Devices
5.1 INTRODUCTION Electromechanical devices convert electrical energy to mechanical energy and vice versa. Common examples of these include generators, actuators, transducers, and motors. Such devices can be found in a wide variety of products including audio and video devices, appliances, computers, automobiles, automated machinery, and power tools. In this chapter, we study the theory of electromechanical devices, as well as various practical applications. Electromechanical devices are governed by a coupled system of electrical and mechanical equations. The electrical equations follow from quasi-static field theory, and the mechanical equations follow from Newton’s laws. In the following sections, we derive the equations of motion for electromechanical devices that execute either linear or rotational motion. We apply these equations to the analysis of various practical devices including magnetic circuit actuators, linear, rotational and resonant actuators, axial-field motors, and stepper motors. The chapter concludes with a brief discussion of magnetic microactuators. The emphasis of this chapter is on magnetically linear, singly excited (single pair of electrical terminals), electromechanical devices with a single degree of mechanical freedom (single coordinate variable). The interested reader can find discussions of multiply excited devices with multiple degrees of freedom in various other texts [1—5].
5.2 DEVICE BASICS We begin with a brief overview of electromechanical devices. These typically consist of three subsystems: the electrical drive circuitry, an electromechanical coupling subsystem, and a mechanical subsystem. The 335
336
CHAPTER 5 Electromechanical Devices
interrelation of these subsystems is depicted in the following diagram: Electrical Electromechanical Mechanical drive circuit coupling subsystem subsystem In many devices, a given component can be considered to be an element of two or more subsystems. For example, in a moving coil actuator the coil can be considered to be an element of all three subsystems. For our purposes, electromechanical devices can be classified into one of the following two categories: 1. Devices in which all the electrical circuit elements are stationary, and the moving element is either a soft magnetic member or a permanent magnet. These are sometimes referred to as magnetic circuit actuators, or moving magnet actuators, respectively. 2. Devices in which the moving member is a coil that is also part of the electrical circuit. These are sometimes referred to as moving coil actuators. An example of the first type of device is shown in Fig. 5.1a. This is a magnetic circuit actuator. In this device the electrical drive circuit consists of a voltage source V , an external resistor R that limits the Q current, and a lossless coil. All of these circuit elements are stationary. The electromechanical coupling subsystem is a magnetic circuit that consists of the lossless coil, a soft magnetic core, and a movable soft magnetic bar that is separated from the core by a gap. The mechanical subsystem consists of the movable bar that is represented by a mass m, and an attached spring. The magnetic circuit is housed in a frictionless frame. The core and spring are fixed to the frame, and the bar is free to move as shown. To activate the device, a voltage is applied to the coil, creating a magnetic field through the circuit. This, in turn, gives rise to a force that acts on the bar to pull it toward the core. The motion is resisted by the force exerted by the spring. An example of the second type of device is shown in Fig. 5.1b. In this device all the circuitry except for the coil is stationary (only the coil is shown). The leads to the coil are flexible and allow the coil to rotate relative to the circuit. The coil is positioned in an external field that is provided by a wedge-shaped magnet. When the coil is energized, its side elements experience a Lorentz force. This force gives rise to a torque that causes the coil to rotate as shown. Usually, a mechanical component such as a spring is used to provide a restoring torque.
5.3 QUASI-STATIC FIELD THEORY
337
Electromechanical devices: (a) magnetic circuit actuator; and (b) moving coil actuator. FIGURE 5.1
5.3 QUASI-STATIC FIELD THEORY In this section we study quasi-static field theory and its application to electromechanical devices. Quasi-static field theory governs the behavior of slowly varying fields, and applies to conventional electromechanical devices that operate at low frequencies and relatively low velocities. This section is divided into two parts. In the first part, we consider quasistatic field theory in a stationary reference frame and derive Kirchhoff’s circuit laws for RLC circuits. In the second part, we consider quasi-static field theory as observed in two different inertial reference frames (in motion relative to one another). We obtain the equations for transforming the fields from one reference frame to the other. We then use these
338
CHAPTER 5 Electromechanical Devices
to determine the voltage induced in a conductor (coil) as it moves through a stationary magnetic field.
5.3.1 Stationary Reference Frames We have already studied quasi-static field theory in a stationary reference frame in Section 2.4. We restate the key results here for convenience. In quasi-static field theory the displacement current is negligible (D/t : 0), and Maxwell’s equations reduce to: Quasi-Static Field Theory Differential Form
Integral Form
;H : J
H · dl :
!
·B:0 ;E : 9
B t
·D:
J · ds 1
B · ds : 0
1
E·dl:9
!
1
D · ds :
1
B · ds t
(5.1)
dv.
4
For electromagnetic devices with time varying or steady currents the free charge density is negligible and we neglect D altogether. The first equation in (5.1) is a generalization of Ampere’s law for slowly varying currents. It implies that · J : 0,
(5.2)
which is Eq. (2.6) for the conservation of charge with the /t term set to zero. Using Gauss’ law, we rewrite Eq. (5.2) in integral form,
J · ds : 0, (5.3) 1 where S is a closed surface. If we apply Eq. (5.3) to an electrical circuit we obtain Kirchhoff’s current law, which states that the algebraic sum of the currents entering a circuit node must be zero. Specifically, if there are n wires at a node then i :0 L L
(Kirchhoff’s current law),
(5.4)
339
5.3 QUASI-STATIC FIELD THEORY
FIGURE 5.2
The RLC circuit.
where i is the current flowing into the node in the nth wire (i is positive L L into the node and negative out of the node). Next, we derive Kirchhoff’s voltage law for the RLC circuit shown in Fig. 5.2. First, apply Faraday’s law to a closed contour C that passes through the circuit,
E · dl : 9
B · ds. t
(5.5) ! 1 We evaluate w E · dl clockwise around the circuit. In doing so, we ! decompose the integral into integrals over the individual components,
> \
E · dl ; Q
source
E · dl ;
resistor
E · dl ;
inductor
E · dl : 9
1
B · ds. t
(5.6)
capacitor
The line integrals in Eq. (5.6) are evaluated as described in Section A.4. We assume that the source is ideal with no internal resistance. The field E in the source can be represented as the gradient of a potential Q E : 9. Therefore, Q > > E · dl : 9 · dl Q \ \
source
where and > \
: 9( 9 ), > \ are the potentials at the positive and negative
340
CHAPTER 5 Electromechanical Devices
terminals of the source, respectively. The source voltage V is defined as Q the potential difference V : 9 and, therefore, Q > \ > V Y9 E · dl. (5.7) Q Q \ Thus, Eq. (5.6) becomes
V : Q
E · dl ;
resistor
E · dl ;
capacitor
E · dl ;
1
B · ds. t
(5.8)
inductor
In the resistor, J : E (Ohm’s law) and, therefore,
E · dl :
resistor
J · dl
resistor
: iR.
(5.9)
The same analysis applies to the inductor,
E · dl :
inductor
J
· dl
inductor
: iR , (5.10) where and R are the conductivity and resistance of the inductor. For the capacitor, we have
E · dl :
1 C
i dt.
(5.11)
capacitor
Substituting Eqs. (5.9), (5.10) and (5.11) into Eq. (5.8) yields 1 V : i(R ; R ) ; Q C
i dt ;
B · ds. t
(5.12)
1 Now, consider the last term in Eq. (5.12). This can be written as d dt
B · ds : 1
d" , dt
(5.13)
341
5.3 QUASI-STATIC FIELD THEORY
where " is called the flux linkage. That is, " is the total flux through (linking) the circuit, "Y
B · ds (flux linkage). 1 The term d"/dt represents the voltage induced in the circuit by a time rate of change of magnetic flux through it. For a simple RLC circuit (with no moving members), " is a function of the current only and, therefore, d" d"(i) di : . dt di dt
(5.14)
We define the circuit inductance L to be LY
d"(i) di
(inductance),
and then obtain d" di :L . dt dt
(5.15)
Finally, substitute Eq. (5.15) into Eq. (5.12) to obtain 1 V : i(R ; R ) ; Q C
i dt ; L
di . dt
(5.16)
This is Kirchhoff’s voltage law for an RLC circuit. In electromechanical devices, the flux linkage " is no longer a function of the current only. Instead, it is a function of both the current and the position of a moving member. Thus, for the analysis of electromechanical devices the preceding derivation needs to be generalized to include the effects of electromechanical coupling. We discuss this in detail in Section 5.4.1.
5.3.2 Moving Reference Frames In our study of electromechanical devices we will need to determine the voltage induced in a coil as it moves through a stationary magnetic field. This requires an understanding of quasi-static field theory as observed in two different inertial reference frames (in motion relative to one another). For our purposes, an inertial reference frame is one that moves with constant velocity. In this section, we derive the transformation laws
342
CHAPTER 5 Electromechanical Devices
FIGURE 5.3
Inertial reference frames in relative motion along one axis.
for quasi-static fields as observed in two different inertial reference frames. Consider two inertial reference frames, a stationary ‘‘laboratory’’ reference frame O with coordinates (x, y, z, t), and a moving reference frame O with coordinates (x, y, z, t). Frame O moves with velocity u with respect to O as shown in Fig. 5.3. Observers in O and O measure different values for the electromagnetic fields. Rigorous comparisons of these measurements can be made if one knows how the fields transform from O to O or vice versa. The transformation laws follow from the theory of special relativity. A detailed presentation of this theory is beyond the scope of this text and is also unnecessary for our presentation. For our purposes, it suffices to state the underlying principles of the theory, and then simply write down the coordinate transform relations. The interested reader can find detailed presentations of special relativity in numerous excellent texts [6]. The basic postulates of special relativity are as follows: 1. The laws of physics are the same in all inertial reference frames. 2. Observers in inertial reference frames measure the same value for the speed of light. The first postulate implies that the quasi-static field equations must have the same form in both O and O as shown in Eqs. (5.17):
343
5.3 QUASI-STATIC FIELD THEORY
Quasi-Static Equations Reference Frame O
Reference Frame O
;H : J
;H : J
·B:0 B ;E : 9 t ·D:
· B : 0 B ;E : 9 t · D : .
(5.17)
Here, is the del operator with respect to the primed coordinate, :
x ; y ; z . x y z
Our goal is to determine the transformation laws for the fields going from O to O and vice versa. We can determine these if we know the transformation relations for the coordinate variables. These follow from the second postulate of special relativity and are given by t : (t 9 #x/c) x : (x 9 #ct) y : y
(5.18)
z : z, where :
1 (1 ; #
,
and u #: . c Here, c is the speed of light in a vacuum. We are interested in low-velocity electromechanical devices for which u c. Therefore, #0
(5.19)
1.
(5.20)
and
We substitute Eqs. (5.19) and Eq. (5.20) into Eq. (5.18), and find that for low velocities the Lorentz transformation relations Eq. (5.18) reduce to
344
CHAPTER 5 Electromechanical Devices
FIGURE 5.4
Inertial reference frames in relative motion.
the Galilean transform relations, t : t x : x 9 ut y : y
(5.21)
z : z. This analysis generalizes to the case of an arbitrary velocity as shown in Fig. 5.4. For this case we find that t : t and x : x 9 ut, or x : x 9 u t V y : y 9 u t (5.22) W z : z 9 u t. X Now that we know the transform relations for the coordinates, we can determine the corresponding relations for the field equations. These will ultimately enable us to compare the fields in the two reference
345
5.3 QUASI-STATIC FIELD THEORY
frames. We start by considering a scalar-valued function f of the primed coordinates, f (x, y, z, t). By virtue of Eq. (5.22), f can be considered to be a function of the unprimed coordinates, f (x, y, z, t) : f (x(x, t), y(y, t), z(z, t), t(t)). We use the chain rule to determine the derivatives of f with respect to the unprimed coordinates. We consider spatial derivatives first and find that f f x : x x x :
f . x
It follows that f : f .
(5.23)
Next, we consider the time derivative. From the chain rule we have f f t f x f y f z : ; ; ; y t z t t t t x t :
f f f f 9 u 9 u 9 u . t x V y W z X
This can be rewritten as f f : 9 u · (f), t t or, alternatively, as f f : ; u · (f). t t
(5.24)
The relations (5.23) and (5.24) relate the derivatives in O and O for scalar-valued functions. A similar analysis applies to vector-valued functions. Let A(x, y, z, t) be a vector-valued function of the primed coordinate variables. This can also be considered to be a vector-valued function of the unprimed coordinates A(x, y, z, t) : A(x(x, t), y(y, t), z(z, t), t(t)). It is easy to check that · A : · A,
(5.25)
346
CHAPTER 5 Electromechanical Devices
and that ;A : ;A.
(5.26)
Similarly, by taking the time derivative we find that A A : ; (u · )A. t t This can be rewritten as A A : ; u( · A) 9 ;(u;A), t t
(5.27)
where we have used the identity ;(A;B) : (B · )A 9 (A · )B ;A( · B) 9 B( · A). Finally, substitute Eqs. (5.25), (5.26), and (5.27) into the quasi-static equations for O in Eq. (5.17) and obtain ;H : J · B : 0 ;(E 9 u;B) : 9
B t
(5.28)
· J : 0. We have used · B : 0 in the third equation of Eq. (5.28). By comparing Eqs. (5.17) and (5.28) we find that the fields transform as follows: Transformation Relations H : H
(5.29)
B : B
(5.30)
E : E ; u;B
(5.31)
J : J.
(5.32)
In Eqs. (5.29)—(5.32) the variables (H, B, E, u, J) are measured in the stationary frame O, and the variables (H, B, E, J) are measured in the moving frame O [4]. For these relations to hold, the variables must be compared at the same physical point in space. For example, from Eqs. (5.22) and (5.31) we have E(x, y, z, t) : E(x 9 u t, y 9 u t, z 9 u t, t) V W X ;u;B(x 9 u t, y 9 u t, z 9 u t, t). V W X
347
5.3 QUASI-STATIC FIELD THEORY
Notice that we have assumed that u is a constant (independent of position). The relations (5.29)—(5.32) give the transformations for the fields and the source. However, we still need to know the constitutive relations in O and O. We apply a general principle that says that the constitutive relations for moving media are the same as for stationary media when they are written in terms of the fields defined in an inertial reference frame moving with the media [4]. For example, for a linear, homogeneous and isotropic media at rest in O we have J : E and B : H. However, in O (which is moving with respect to the media) the relations are modified by virtue of Eqs. (5.29)—(5.32). Specifically, we find that Constitutive Relations Reference Frame O
Reference Frame O
J : (E ; u;B)
J : E
B : H
B : H.
(5.33)
In our study of electromechanical devices we will need to determine the voltage induced in a coil as it moves through a stationary magnetic field. Therefore, we discuss this in some detail. Consider a reference frame O in which there is a static (but not necessarily uniform) B-field. Consider an open conductor moving with velocity u relative to O. Assume that the conductor has a conductivity (as measured by observers in both O and O). We want to determine the voltage induced across the terminals of the conductor as measured in O. We apply Faraday’s law to a stationary contour C that passes through the conductor at a given (but arbitrary) instant in time,
E · dl : 9
d dt
B · ds. (5.34) ! 1 We label the terminals a and b, and then decompose the left-hand side of Eq. (5.34) into integrals across the terminals and the conductor itself,
E · dl : !
?
E · dl ;
@
E · dl,
?
terminals
: 9V ;
@
conductor
@
E · dl,
?
conductor
(5.35)
348
CHAPTER 5 Electromechanical Devices
where V is the terminal voltage (as measured in O). As for the remaining integral, from Eqs. (5.31) and (5.32) we have E : E 9 u;B and E : J/ : J/. Therefore, J E : 9 u;B
(moving conductor).
(5.36)
Notice that all the terms in Eq. (5.36) are measured in O (the conductor is moving with respect to O). As the conductor is open, there is no current through it (J : 0) and, therefore, E : 9u;B
(5.37)
Combining Eqs. (5.34), (5.35) and (5.37) gives us
V:9
@
(u;B) · dl ;
d dt
B · ds. (5.38) ? 1 Because B, C, S, and ds are constant in O, the last term in Eq. (5.38) is zero and the terminal voltage reduces to
V:9
@
(u;B) · dl. (5.39) ? It is important to note that if there were a time-varying current in the conductor, the flux through the circuit due to the current would add to the voltage via the last term in Eq. (5.38). We demonstrate this theory in the following examples. EXAMPLE 5.3.1 A conductive bar is in sliding contact with a pair of stationary conducting rails as shown in Fig. 5.5. The bar is moving through a constant uniform B-field with a constant velocity u relative to the rails. Determine the voltage across the rails. Assume that the bar and rails are good conductors. SOLUTION 5.3.1 We choose a reference frame O at rest with respect to the rails, and a stationary contour C (dotted line) that runs clockwise through the rails, and passes through the bar at a single instant in time (Fig. 5.5). Apply Faraday’s law (5.5) to C,
E · dl : 9
d dt
B · ds.
(5.40)
! 1 As B, C, S, and ds are constant in O we have d dt
1
B · ds : 0.
(5.41)
349
5.3 QUASI-STATIC FIELD THEORY
FIGURE 5.5
Conductive bar moving along conductive rails.
Substitute this into Eq. (5.40) and decompose the right-hand side into integrals over the various segments and obtain,
>
E · dl ;
\
terminals
E · dl ;
upper rail
E · dl ;
moving bar
E · dl : 0
(5.42)
lower rail
As the upper and lower rails are good conductors ( -) we have E 0 and, therefore,
E · dl :
upper rail
E · dl : 0.
(5.43)
lower rail
In the moving bar we have
E · dl :
(E 9 u;B) · dl,
moving bar
but E : J/ : J/ 0 and, therefore,
E · dl : 9 (u;B) · dl.
(5.44)
moving bar
To evaluate Eq. (5.44), we use u : ux , B : Bz , and dl : dyy (Fig. 5.5). Therefore, 9 (u;B) · dl : 9 9uB dy F : 9uBh,
350
CHAPTER 5 Electromechanical Devices
or
E · dl : 9uBh.
(5.45)
moving bar
Finally, substitute Eqs. (5.43) and (5.45) into Eq. (5.42) and obtain V : 9uBh, where V is the terminal voltage,
V:9
>
E · dl.
)
\
EXAMPLE 5.3.2 Determine the lift and drag force on a wire carrying a current i and moving with velocity u relative to a conductive sheet of thickness as shown in Fig. 5.6a [7]. SOLUTION 5.3.2 We choose a reference frame at rest with respect to the wire (Fig. 5.6b). In this reference frame, the magnetic field B of the wire is constant (time invariant), and the sheet moves with velocity u : 9ux . The sheet experiences a time varying magnetic field. The current density induced at a point x in the sheet is J (x) : 9uBR(x), W X where BR(x) is the total magnetic field at x, that is, W BR(x) : B(x) ; BG(x). W W W Here, B (x) and BG (x) are the fields due to the current in the wire, and the W W induced current in the sheet, respectively. Specifically, we have i x B(x) : . W 2 x ; h and BG(x) : W 2
K ()) X d), x9) \
(5.46)
where K (x) : 9uBR(x) (5.47) W X is the current density per unit length ( is the thickness of the sheet). Notice
351
5.3 QUASI-STATIC FIELD THEORY
Wire moving above a conductive plate: (a) wire moving and plate stationary; and (b) wire stationary and plate moving [7].
FIGURE 5.6
that Eq. (5.46) contains a singularity. As K ()) is continuous, we evaluate the X integral using the Cauchy Principal value. Substituting Eq. (5.46) into Eq. (5.47) yields the following integral equation for K (x), X u K ()) 2u X K (x) : 9 d) 9 B(x), (5.48) X w x9) w W \ where w : 2/( ). We solve Eq. (5.48) using the Fourier transforms, eIVK (x) dx, F[K ] : X X \ and
F
K ()) 1 X d) : F[K ]F . X x9) x \
352
CHAPTER 5 Electromechanical Devices
The second transform follows from convolution theory. In addition, notice that F
1 i k 0 : x 9i k 0.
We apply these to Eq. (5.48) and obtain
F[B] u 1 2u W 19 F . F[K ] : 9 X w x w 1 ; (u/w) Next, we apply an inverse transform to Eq. (5.49) and find that
(5.49)
2u 1 u B()) W K (x) : 9 B (x) 9 d) . (5.50) X w (1 ; u/w) W w )9x \ The last term in Eq. (5.50) can be rewritten in a more useful form using B : ;A. After some analysis we find that
2u 1 u B(x) ; B(x) , K (x) : 9 X w V w (1 ; u/w) W
(5.51)
where i h B(x) : , V 2 x ; h and i x . B(x) : W 2 x ; h Notice that Eq. (5.51) is not limited to a single moving current filament. Rather, it is valid for a system of moving field sources provided they are of limited extent, produce a one-dimensional (1D) current pattern, and travel at the same velocity. The lift and drag forces on the wire are equal and opposite to the forces on the sheet. The force density on the moving sheet is the Lorentz force density (2.23), f : K (x)z ;(B(x)x ; B(x)y ). V W X Therefore, f : 9K (x)B(x) W V X and f : K (x)B(x). V W X
353
5.3 QUASI-STATIC FIELD THEORY
We integrate over the sheet to obtain the total lift and drag force
>
K (x)B(x) dx V X \ ih > K (x) X dx : x ; h 2 \ i u : 4h w ; u
F : *
:F
u w ; u
(Lift force)
and w F : F " u *
(Drag force),
where i F : . 4h A normalized plot of these forces is shown in Fig. 5.7. Notice that the drag force reaches a maximum when u : w. )
FIGURE 5.7
Normalized plot of lift and drag forces vs normalized velocity [7].
354
CHAPTER 5 Electromechanical Devices
5.4 ELECTRICAL EQUATIONS The electrical equations for an electromechanical device follow from circuit theory, which in turn is based on quasi-static field theory. In this section, we derive a generalized version of Kirchhoff’s voltage law that takes into account the effects of electromechanical coupling. We study two different types of circuits, stationary and moving coil. In a stationary circuit, all the circuit components are at rest with respect to one another. For these circuits, we apply quasi-static field theory in a reference frame fixed with respect to the circuit. The electromechanical coupling is accounted for by a flux linkage term that takes into account the voltage induced in the circuit by a time rate of change of magnetic flux through it. This change in flux is due to a change in the magnetic field that can be caused by a change in the current and/or by the movement of a ferromagnetic member that alters the field. Mechanical motion is taken into account by considering the flux to be a function of both the circuit current and mechanical displacement. In a moving coil circuit, a coil is moving with respect to an external magnetic field. To analyze this kind of circuit, we apply quasi-static field theory in a reference frame that is at rest with respect to the magnetic field. Mechanical motion is taken into account by the voltage induced in the coil as it moves through the field. Finally, it is important to note that the following analysis applies to magnetically linear devices.
5.4.1 Stationary Circuits In this section we study electromechanical devices with stationary electrical circuits. Such a device is shown in Fig. 5.8. It consists of a voltage source V , a resistor R, and a magnetic circuit with a fixed coil Q and a moving soft magnetic bar that alters the flux through the coil when it moves. We analyze the electrical circuit using quasi-static field theory. Specifically, we choose a reference frame at rest with respect to the circuit and a fixed integration contour C along the circuit. At any point along the circuit the electric field is given by E:9
A 9 , t
(5.52)
where 9 is due to the accumulation of charge along the circuit, and A/t is due to the time-varying current. We evaluate the line integral
355
5.4 ELECTRICAL EQUATIONS
Electromechanical actuator circuit.
FIGURE 5.8
of Eq. (5.52) clockwise along C and obtain
A · dl 9 t
E · dl : 9
· dl. ! ! ! The first term can be decomposed into integrals over the individual components
>
E · dl ;
\ source
E · dl : 9
E · dl ;
resistor
A · dl 9 t
· dl.
(5.53)
coil
Notice that w · dl : 0 because it entails the integration of a conservative vector field around a closed path. Also notice that although w A/t · dl is evaluated around the entire path, the major contribution to it is from the inductor. We assume that the source is ideal with no internal resistance and define the voltage impressed at the terminals in the usual way:
V Y9 Q
> \
E · dl. Q
(5.54)
Thus, Eq. (5.53) becomes V : Q
E · dl ;
resistor
E · dl ; coil
A · dl. t
(5.55)
356
CHAPTER 5 Electromechanical Devices
In the resistor J : E (Ohm’s law). Therefore,
E · dl :
resistor
J · dl,
(5.56)
resistor
: iR. Similarly, in the coil J :
E, and, therefore,
E · dl : iR
.
(5.57)
coil
Substituting Eqs. (5.56) and (5.57) into Eq. (5.55) yields V : i(R ; R ) ; Q
A · dl. t
(5.58)
Now, consider the last term in Eq. (5.58). We can rewrite this as follows:
A · dl : t :
1
1 d : dt :
A ; · ds t
B · ds t B · ds
1
d" , dt
(5.59)
where "Y
B · ds (flux linkage). 1 The symbol " denotes the flux linkage, that is, the total flux through (linking) the circuit. To determine " we need to evaluate B · ds over an 1 area bounded by the circuit path in accordance with Stokes’ theorem. However, this integral can be difficult to evaluate depending on the topology of the circuit. For a simple single-turn current loop, the flux is denoted by and, therefore, " : . For a tightly wound N turn coil the flux linkage is " : N
(N turn coil),
where is the flux through each turn.
357
5.4 ELECTRICAL EQUATIONS
We digress briefly to discuss the second and third step in Eq. (5.59). Notice that we have replaced the partial time derivative inside the integral by a total time derivative outside the integral. We do this because inside the integral, B is a function of both time-independent coordinate variables and other time-dependent variables (including t itself). However, the integration effectively eliminates the time-independent variables, and leaves only the time-dependent variables. Therefore, the total time derivative is appropriate outside the integral. We return to the derivation. Substitute Eq. (5.59) into Eq. (5.58) and obtain V : i(R ; R ) ; Q
d" . dt
(5.60)
mechanical coupling
This is a generalization of Kirchhoff’s voltage law. The term d"/dt in Eq. (5.60) requires some discussion because it is key to the electromechanical coupling. This term represents the voltage induced in the circuit by a time rate of change of magnetic flux through the circuit. The change in flux can be due to a change in circuit current and/or the movement of a ferromagnetic member that alters the field. To account for this we write " as a function of both the circuit current and the position of the member. Let x and denote the position of the member when it executes linear and rotational motion, respectively. Then ":
"(i, x) (linear motion) "(i, ) (rotational motion).
Given this functional dependence, the total time derivative of " is
"(i, x) di "(i, x) dx ; (linear motion) i dt x dt
d" : dt "(i, ) di "(i, ) d ; (rotational motion) i dt dt or
d" : dt
L
di "(i, x) ; v dt x
di "(i, ) L ; dt
(linear motion) (5.61) (rotational motion)
358
CHAPTER 5 Electromechanical Devices
where L is the inductance and which is defined by LY
" i
(inductance).
(5.62)
The variables v : dx/dt and : d/dt are the linear and angular velocities, respectively. It is important to note that the definition of inductance Eq. (5.62) applies only to magnetically linear devices. There is no simple definition of L for nonlinear devices [8]. Finally, substitute Eq. (5.61) into Eq. (5.60) and obtain
Electrical $ Equations
di(t) "(i, x) V (t) : i(t)(R ; R ) ; L ; v(t) Q dt x (linear motion)
di(t) "(i, ) V (t) : i(t)(R ; R ) ; L ; (t) Q dt
.
(5.63)
(rotational motion)
These are the electrical equations of motion for stationary circuits with time varying currents and a moving ferromagnetic member.
5.4.2 Moving Coils In this section we derive Kirchhoff’s voltage law for circuits that have a coil moving through an external magnetic field B . All the other circuit components are stationary with respect to the field. Let O denote the stationary reference frame. The coil moves with velocity u through B , which is constant (but not necessarily uniform) with respect to O. Such a circuit is shown in Fig. 5.9. This figure shows a linear actuator consisting of a conductive bar of mass m in sliding contact with a pair of stationary conducting rails. The rails are connected to a voltage source V and a resistor R that limits the current. The bar is moving through a Q constant uniform B-field with a time-dependent velocity u(t) relative to the rails. This circuit is analyzed in Example 5.6.2. We derive Kirchhoff’s voltage law for a circuit with a moving coil by applying Faraday’s law to a stationary contour C that passes through the circuit and coincides with the coil for a given position of its travel (the dotted line in Fig. 5.9). Thus, C coincides with the entire circuit at a single instant in time. We evaluate
E · dl : 9 !
d dt
1
B · ds
(5.64)
359
5.4 ELECTRICAL EQUATIONS
Moving conductor actuator.
FIGURE 5.9
clockwise around the contour and obtain
>
E · dl ;
\
source
E · dl ;
resistor
E · dl : 9 coil
d dt
B · ds.
(5.65)
The integrals over the fixed components (source and resistor) follow from Eqs. (5.54)—(5.57). We obtain V : iR ; Q
E · dl ;
d dt
B · ds. (5.66) The coil is moving with velocity u through B . The E-field in it as measured in O is given by Eq. (5.36), J E : 9 u;B . Therefore,
where R and obtain
E · dl :
J 9 u;B · dl,
: iR 9
(u;B ) · dl, (5.67) is the resistance of the coil. Substitute Eq. (5.67) into Eq. (5.66)
V : i(R ; R ) 9 Q
d (u;B ) · dl ; dt
B · ds.
(5.68)
Now consider the last term in Eq. (5.68). The dominant contribution to this term comes from the coil. Moreover, as the system is magnetically linear we can write B as a superposition of two fields: the field B due G
360
CHAPTER 5 Electromechanical Devices
to the current i and the external field B , B:B ;B . G Therefore, the last term in Eq. (5.68) can be written as d dt
B · ds : :
d dt
d B · ds ; G dt
d d " (i) ; dt dt
:L
di d ; dt dt
B · ds
B · ds
B · ds, (5.69) where " (i) : B · ds, and we have used the definition of inductance G " (i) L : . i Finally, we substitute Eq. (5.69) into Eq. (5.68) and find that di V : i(R ; R ) ; L 9 Q dt
d (u;B ) · dl ; dt
B · ds. (5.70) This is a generalization of Kirchhoff’s voltage law that takes into account the voltage induced in a coil as it moves through B . If B is static, d B · ds : 0 and Eq. (5.70) reduces to dt di V : i(R ; R ) ; L 9 (u;B ) · dl. (B static). (5.71) Q dt It is important to note that the line integral in Eq. (5.71) is evaluated as described in Section A.4 with the direction of integration in the direction of conventional current flow. The voltage equation (5.71) applies to a coil in linear motion. However, it can be generalized to the case of a coil that executes rotational motion. Specifically, if a segment dl of the coil is rotating with an angular velocity at a distance r with respect to a fixed reference frame, then its linear velocity is
u : ;r.
(5.72)
We substitute Eq. (5.72) into Eq. (5.71) and obtain the circuit equation for rotational motion, di V (t) : i(R ; R ) ; L 9 Q dt
[(;r);B ] · dl
(B static). (5.73)
361
5.5 MECHANICAL EQUATIONS
It is important to note that when evaluating Eq. (5.73), the velocity (;r) must be evaluated before the cross product with B is taken. If the rotation is with respect to a specific axis (which we label the z-axis), then : z , and Eq. (5.73) can be rewritten as di V (t) : i(t)(R ; R ) ; L 9 Q dt
[(;r );B ] · dl ,
(B static), (5.74)
where r is a vector from (and perpendicular to) the z-axis to the , segment dl.
5.5 MECHANICAL EQUATIONS In this section we review the mechanical equations of motion for electromechanical devices that execute linear or rotational motion with a single degree of freedom. Such linear or rotational motion is described by a single coordinate variable that we denote by x or , respectively. The equations for these variables follow from Newton’s second law of motion, m
dx : F(i, x) dt
(linear motion),
(5.75)
(rotational motion),
(5.76)
or d J : T(i, ) K dt
where m is the mass, J is the moment of inertia about the axis of K rotation, and F(i, x) and T(i, ) are the force and torque. The moment of inertia of a point mass m at a fixed distance r about the axis of rotation is J : mr. If the mass is distributed, then J is computed via integraK K tion, J : K
4
r dv,
where and V are the density and volume of the mass, respectively,
and r is the distance of the element dv from the axis of rotation. For
362
CHAPTER 5 Electromechanical Devices
example, if a rigid body of density is rotating about the z-axis, its
moment of inertia is J : K
(x ; y) dv. (5.77)
4 Notice that in Eqs. (5.75) and (5.76) we have chosen the current i and position x or as the dependent variables. This is often a convenient choice, but other variables may be more convenient depending on the application. Regardless of the choice of variables, for lumped-parameter analysis we need to determine analytical expressions for F(i, x) and T(i, ). This is usually accomplished using one of the following methods: 1. Derive an analytical solution for the magnetic field and determine the force density using the Lorentz force formula or the Maxwell stress tensor. Then, perform the appropriate volume integration to obtain the total force or torque. This approach is most useful in the analysis of moving coil actuators. It is discussed in what follows. 2. Derive an analytical expression for the field energy (or coenergy) and then determine the force or torque via differentiation. This approach is especially useful in the analysis of magnetic circuit or moving magnet devices. It is discussed in Section 5.7. 3. Discretize the independent variables into a discrete set of values, use a numerical method such as FEA to obtain force or torque data for these values, and then fit the data to an analytical formula. This is called a hybrid approach and is discussed in Section 5.15. The first method is the most direct approach for determining the force or torque. It is especially useful in the analysis of moving coil actuators. The force and torque imparted to the coil are obtained using the Lorentz force formulas Eqs. (3.47) and (3.52) (Section 3.2.2). Specifically, if a coil carries a current i and moves with velocity u through an external field B , the force and torque imparted to it are F:i
(dl;B ),
(5.78)
and T:i
r;(dl;B ), (5.79) where r is the vector from the origin about which the torque is computed
5.6 ELECTROMECHANICAL EQUATIONS
363
and integration is over the length of the coil. When evaluating Eq. (5.79), the force (dl;B ) must be evaluated before the cross product with r is taken. Equations (5.78) and (5.79) are vector equations. We are interested in both the force in the x direction, and the torque about a given axis, which we choose to be in the z-axis. The magnitudes of force and torque in these directions are F :i V
(dl;B ) · x ,
(5.80)
[r ;(dl;B )] · z , ,
(5.81)
and T :i X
where r is a vector from (and perpendicular to) the z-axis to the , segment dl.
5.6 ELECTROMECHANICAL EQUATIONS In this section we summarize the equations of motion for electromechanical devices. We have seen that the electrical and mechanical equations are separately second-order equations in the independent variables i and x or , respectively. These equations are coupled and therefore represent a second-order system of equations. These coupled equations can sometimes be solved directly. However, for practical analysis it is often easier to reduce them to a first-order system of equations. Such systems lend themselves to various numerical timestepping solution procedures such as the Euler or Runge-Kutta methods. These are discussed in Appendix C.
5.6.1 Stationary Circuits In this section we summarize the equations of motion for electromechanical devices with stationary circuits. These include magnetic circuit and moving magnet actuators. The electrical and mechanical equations for these devices were derived in the foregoing text. We repeat
364
CHAPTER 5 Electromechanical Devices
them here for convenience,
Electrical $ Equations
and
di(t) "(i, x) V (t) : i(t)(R ; R ) ; L ; v(t) Q dt x (linear motion)
di(t) "(i, ) V (t) : i(t)(R ; R ) ; L ; (t) Q dt
.
(5.82)
(rotational motion)
dx(t) : F(i, x) (linear motion) dt Mechanical $ . d(t) Equations J : T(i, ) (rotational motion) K dt m
(5.83)
Since the same variables i and x or appear in both sets of equations, they need to be solved simultaneously. Thus, the analysis of an electromechanical device entails the solution of the following coupled second-order equations: Linear Motion di(t) "(i, x) ; v(t) V (t) : i(t)(R ; R ) ; L Q x dt dx(t) m : F(i, x) dt
(5.84)
or Rotational Motion di(t) "(i, ) V (t) : i(t)(R ; R ) ; L ; (t) Q dt d(t) J : T(i, ), K dt
(5.85)
For practical analysis is convenient to rewrite the second-order equations in terms of the following first-order systems,
365
5.6 ELECTROMECHANICAL EQUATIONS
Linear Motion di(t) 1 : dt L
V (t) 9 i(t)(R ; R ) 9 Q dv(t) 1 : F(i, x) dt m dx(t) : v(t) dt
(5.86)
(5.87)
"(i, x) v(t) x
and Rotational Motion di(t) 1 : dt L
V (t) 9 i(t)(R ; R ) 9 Q d (t) 1 : T(i, ) J dt K d(t) : (t). dt
"(i, ) (t)
Equations (5.86) and (5.87) constitute initial-value problems that need to be solved subject to the following initial conditions, i(t ) : i x(t ) : x v(t ) : v
(linear motion)
(5.88)
and i(t ) : i (t ) : (rotational motion) (5.89) (t ) : , respectively. Occasionally, the initial-value problems can be solved analytically. More often, their solution requires the application of a numerical time-stepping procedure such as the Euler or Runge Kutta method (Appendix C). EXAMPLE 5.6.1 Apply Euler’s method to Eq. (5.87). SOLUTION 5.6.1 We reduce Eq. (5.87) to a set of algebraic equations using Euler’s method as discussed in Appendix C. In Euler’s method, derivatives are
366
CHAPTER 5 Electromechanical Devices
approximated by a simple difference quotient, for example, df f(t ; t) 9 f(t) . t dt
(5.90)
We apply Eq. (5.90) to Eq. (5.87) and obtain i(t ; t) : i(t) ;
t "(i, ) V (t) 9 i(t)(R ; R ) 9 (t) Q L
t T(i, ) J K (t ; t) : (t) ; t (t).
(t ; t) : (t) ;
(5.91)
To solve Eq. (5.91) we need expressions for "(i, ) , L:
"(i, ) , i
(5.92) (5.93)
and T(i, ).
(5.94)
Once these are known, Eq. (5.91) can be solved using the following iterative scheme: i(n ; 1) : i(n) ;
t "(i(n), (n)) V (n) 9 i(n)(R ; R ) 9 (n) Q L t T(i(n), (n)) J K (n ; 1) : (n) ; t (n).
(n ; 1) : (n) ;
(5.95)
The solution method is as follows: Choose a value for t. Then, given the initial conditions [i(0), (0), (0)] use Eq. (5.95) to obtain [i(1), (1), (1)], which gives the solution at t : t. Repeat this process to obtain [i(2), (2), (2)], [i(3), (3), (3)], . . . , [i(n), (n), (n)], [i(n ; 1), (n ; 1), (n ; 1)]. These are the solution values at t : 2t, 3t, . . . , nt, (n ; 1)t, . . . , respectively. )
5.6.2 Moving coils In this section we summarize the equations of motion for electromechanical devices with coils that move through an external field B
367
5.6 ELECTROMECHANICAL EQUATIONS
(moving coil actuators). The electrical and mechanical equations were derived in the preceding; we repeat them here for convenience:
Electrical $ Equations
di(t) V (t) : i(t)(R ; R ) ; L 9 Q dt
(u ; B ) · dl
(linear motion) di(t) 9 V (t) : i(t)(R ; R ) ; L Q dt
(rotational motion)
[(;r);B ] · dl
(5.96)
and
Mechanical $ Equations
m
dx(t) :i dt
d(t) J :i K dt
(dl;B ) · x ; F (x)
(linear motion)
[r;(dl;B )] · z ; T ().
(rotational motion) (5.97)
In Eq. (5.97) the term T () represents the mechanical torque imposed
on the system. The electrical and mechanical equations are coupled and can be grouped into the following second-order systems of equations, Linear Motion di(t) V (t) : i(t)(R ; R ) ; L 9 Q dt m
dx(t) : i(t) dt
(u ; B ) · dl
(dl ; B ) · x ; F (x)
(5.98)
368
CHAPTER 5 Electromechanical Devices
and Rotational Motion di(t) V (t) : i(t)(R ; R ) ; L 9 Q dt d(t) J : i(t) K dt
[( ; r) ; B ] · dl
(5.99)
[r ; (dl ; B )] · z ; T ().
The line integrals in Eqs. (5.98) and (5.99) are evaluated as described in Section A.4. The direction of integration follows the current i. In addition, in Eq. (5.99) the velocity ( ; r) and force i(dl ; B ) are evaluated before the remaining cross products are taken. For conventional rotary devices, the analysis is usually performed using cylindrical coordinates with the rotation taken to be about the z-axis, : (t)z . The current flow is often taken to be along the z-axis with dl : dzz . When this is the case, Eq. (5.99) reduces to the following form: Rotational Motion di(t) V (t) : i(t)(R ; R ) ; L ; K (t) Q C dt J
(5.100)
d(t) : i(t)K ; T (),
K dt R
where K is called the electrical constant (or back emf constant) and K is C R called the torque constant. In the MKS system K : K . We derive Eq. C R (5.100) for two different rotational devices in Examples 5.6.3 and 5.11.1. These equations are also derived for a device with a different configuration in Example 5.9.1. For practical analysis, it is convenient to rewrite Eqs. (5.98) and (5.99) in terms of the following first-order systems: Linear Motion di(t) 1 : [V (t) 9 i(t)(R ; R ) ; dt L Q du(t) 1 : [i(t) dt m
(u ; B ) · dl]
(dl ; B ) · x ; F (x)]
dx(t) : u(t) dt
(5.101)
369
5.6 ELECTROMECHANICAL EQUATIONS
and Rotational Motion di(t) 1 : [V (t) 9 i(t)(R ; R ) ; L Q dt
[( ; r) ; B ] · dl]
(5.102)
d (t) 1 : [i(t) [r ; (dl ; B )] · z ; T ()]
J dt K d(t) : (t). dt
Equations (5.101) and (5.102) constitute initial-value problems that need to be solved subject to the initial conditions (5.88) and (5.89), respectively. We demonstrate the use of these equations in the following examples. EXAMPLE 5.6.2 Consider a linear actuator consisting of a conductive bar of mass m in sliding contact with a pair of stationary conducting rails as shown in Fig. 5.9. The rails are connected to a voltage source V , and a resistor R that Q limits the current. The bar is moving through a constant uniform B-field with a time-dependent velocity u(t) relative to the rails. Determine the equations of motion for the system. Assume that the rails are good conductors with negligible resistance. SOLUTION 5.6.2 This device has a moving conductor and is governed by the linear motion equations (5.101). In this case the coil is the moving bar, and there is no mechanical force (F (x) : 0). Therefore, Eq. (5.101) becomes
di(t) 1 : V (t) 9 i(t)(R ; R ) ; L Q dt du(t) i(t) : dt m
dx(t) : u(t), dt
(u ; B ) · dl
(dl ; B ) · x (5.103)
where L is the inductance of the complete circuit. We need to evaluate the integrals in Eq. (5.103). We use a Cartesian reference frame O at rest with respect to the rails as shown in Fig. 5.9.
370
CHAPTER 5 Electromechanical Devices
Induced voltage: To evaluate the induced voltage, we need to determine (u ; B ) · dl along the bar. From Fig. 5.9 we have B : 9B z , (5.104) u(t) : u(t)x , (5.105) and dl : dyy . (5.106) Therefore (u ; B ) · dl : u(t)B dy. (5.107) We integrate Eq. (5.107) along the bar in the direction of current. This gives
(u ; B ) · dl :
F
(u ; B ) · dl
: B u(t)
dy
F : 9B hu(t). (5.108) Force: Next we evaluate the force. For this we need (dl ; B ) · x along the bar. From Eqs. (5.104) and (5.106) we have (dl ; B ) · x : 9B dy. We integrate this along the bar in the direction of current flow and obtain
(dl ; B ) · x :
F
(dl ; B ) · x
: 9B
dy
F : B h. (5.109) Therefore, the Lorentz force on the bar is F : iB h. Equations of motion: Substitute Eqs. (5.108) and (5.109) into Eq. (5.103), which gives di(t) 1 : [V (t) 9 i(t)(R ; R ) 9 B hu(t)] L Q dt du(t) 1 : i(t)B h dt m dx(t) : u(t). dt
5.6 ELECTROMECHANICAL EQUATIONS
FIGURE 5.10
371
Rotating coil device: (a) perspective of device; (b) end view; and
(c) side view.
These equations need to be solved subject to initial conditions (5.88).
)
EXAMPLE 5.6.3 Consider the rotary device shown in Fig. 5.10a. This represents a basic direct-current (dc) motor. In this device, a magnetic field B (assumed constant) is produced between two pole pieces by a field current i through a D coil that is wrapped around the pole pieces. A rectangular coil is mounted between the pole pieces and is free to rotate about the z-axis. When a current i passes through the coil it experiences a torque that causes it to rotate as shown.
372
CHAPTER 5 Electromechanical Devices
A split ring with brushes is connected to the terminals of the rotating coil so that the current through it reverses direction every half turn. This ensures that the torque will always be in the same direction, thereby producing continuous rotation. Write the equations of motion for this device. SOLUTION 5.6.3 This device has a moving coil and is governed by the rotational motion equations (5.102). There is no mechanical torque and Eqs. (5.102) reduce to di(t) 1 : [V (t) 9 i(t)(R ; R ) ; dt L Q
[( ; r) ; B ] · dl]
d (t) i(t) : [r ; (dl ; B )] · z J dt K d(t) : (t). dt
(5.110)
where L and R are the inductance and resistance of the coil, and J is the K moment of inertia of the coil about the z-axis. We need to evaluate the induced voltage and torque integrals in Eq. (5.110). We use a Cartesian reference frame at rest with respect to the pole pieces as shown in Fig. 5.10a. We choose a rotation angle , which is measured with respect to the y-axis in a counterclockwise sense about the z-axis as shown in Fig. 5.10b. Induced voltage: We determine the voltage induced at the terminals of the coil by summing the contributions from its various segments. The coil has four segments; relatively long left and right side segments, and two shorter end segments (Fig. 5.10b,c). We evaluate the side segments first. To evaluate the induced voltage we need [( ; r) ; B ] · dl along these segments. We know that, (t) : (t)z , and from Fig. 5.10b we find that
9b sin()x ; b cos()y b sin()x 9 b cos()y
r:
(left side) (right side).
Therefore, ;r:
9 (t)b cos()x 9 (t)b sin()y (t)b cos()x ; (t)b sin()y
Note that B is in the x-direction, B : B x .
(left side) (right side).
373
5.6 ELECTROMECHANICAL EQUATIONS
Therefore,
(t)bB sin()z ( ; r) ; B : 9 (t)bB sin()z
(left side) (right side).
Because dl : dzz for both segments we have
(t)bB sin() dz [( ; r) ; B ] · dl : 9 (t)bB sin() dz
(left side) (right side).
(5.111)
We evaluate Eq. (5.111) along the side segments with the integration in the direction of the current. This gives
[( ; r) ; B ] · dl :
(t)bB sin() dz
?
left side
: 9 (t)abB sin(),
(5.112)
and
[( ; r) ; B ] · dl : 9
?
(t)bB sin() dz
right side
: 9 (t)abB sin(). Next, consider the end segments. For these we have ( ; r) ; B . z , and dl . r. It follows that [( ; r) ; B ] · dl . z · r : 0
(end segments).
Therefore,
[( ; r) ; B ] · dl : 0. end segments
(5.113)
374
CHAPTER 5 Electromechanical Devices
This shows that the end segments make no contribution to the induced voltage. The induced voltage is the sum of Eqs. (5.112) and (5.113). Specifically,
[( ; r) ; B ] · dl : 92 (t)abB sin()
: 9 (t)A B sin(), (5.114) where A : 2ab is the area of the loop. If there are n loops in the coil, then Eq. (5.114) is multiplied by n. Torque: We now determine the torque. Only the left- and right side segments of the coil contribute to the torque. We make a brief argument to show that the end segments make no contribution. Consider the torque on an element of an end segment, dT : r ; (dl ; B ) · z . X On these segments, dl dr and, therefore, dl ; B . dr ; B . z .
This gives
dT . (r ; z ) · z : 0. X Thus, the end segments generate no torque. Now consider the left- and right side segments. For these we have dl : dzz . Therefore, dl ; B : B dzy
It follows that
(side segments).
9bB sin() dz z (left side) (5.115) bB sin() dz z (right side). We evaluate the integral of Eq. (5.115) along the side segments, integrate in the direction of current flow, and obtain r ; (dl ; B ) :
[r ; (dl ; B )] · z : 9
?
bB sin() dz
left side
: abB sin(),
and
[r ; (dl ; B )] · z :
?
(5.116)
bB sin() dz
right side
: abB
sin().
(5.117)
5.6 ELECTROMECHANICAL EQUATIONS
375
The torque on the coil is the sum of Eqs. (5.116) and (5.117),
[r ; (dl ; B )] · z : 2abB sin() : A B sin().
(5.118)
If there are n loops in the coil then Eq. (5.118) is multiplied by n. Moment of inertia: Next, we determine the moment of inertial. We use Eq. (5.77), which we repeat here for convenience, J : K
(x ; y) dv.
4
The inertia of each side segment is J : K
?
bA dz
: abA : m b, where A and are the area and density of the wire, respectively, and
m : aA is the mass of each side segment. The inertia of each end segment is J : 2 K
@
A d
: bA : m b, where m : 2bA is the mass of each end segment. Summing the contributions from all segments gives the total moment of inertia, J : 2J ; 2J K K K : (2m ; m )b. If there are n loops in the coil then Eq. (5.119) is multiplied by n.
(5.119)
376
CHAPTER 5 Electromechanical Devices
Equations of motion: We are finally ready to write the equations of motion. Substitute Eqs. (5.114) and (5.118) into Eq. (5.110) and obtain di(t) 1 : [V (t) 9 i(t)(R ; R ) 9 A B sin((t)) (t)] dt L Q d (t) 1 : i(t)A B sin((t)) J dt K d(t) : (t). dt These equations are solved subject to the initial conditions (5.89). Notice that the electrical constant K equals the torque constant K (see Eq. (5.100)). C R Specifically, K : K : A B sin((t)). C R If the coil had n turns all at the same angle , then K and K would be C R multiplied by n. If the turns were distributed in an angular sense, then K and C K would be obtained by integrating the contributions of the individual turns R over the angular span of the coil. )
5.7 ENERGY ANALYSIS In this section we discuss an energy method for the analysis of electromechanical devices. This method, based on the conservation of energy, is used to determine the force or torque on a moving member. It is especially useful for the analysis of magnetic circuit actuators as we shall see. We start with the basic energy balance relation: electrical energy energy energy mechanical input at : dissipated by ; stored in ; energy device terminals device magnetic field output Apply this to a system that is subjected to an infinitesimal (virtual) displacement and obtain dW C
electrical energy input
: dW
energy dissipated
; dW
DJB
stored field energy
; dW
mechanical energy
(5.120)
377
5.7 ENERGY ANALYSIS
The term dW accounts for energy dissipation due to loss mechan isms such as ohmic heating and mechanical friction. For our purposes, we consider the system to be lossless and treat the loss mechanisms as external to the system. For example, a real inductor can be treated as a lossless coil in series with an external resistance. Thus, without loss of generality, we restrict our attention to lossless systems in which dW : 0. For these systems, energy balance equation (5.120) becomes dW : dW ; dW . C DJB
(5.121)
We consider the various terms in Eq. (5.121). Recall from Eq. (5.60) that mechanical motion couples to the electrical equations via flux linkage ". Moreover, the voltage e induced across the terminals of an electromechanical component such as a coil is e : d"/dt. Therefore, the electrical energy input into the terminals of the component during a time dt is d" dW : i dt C dt : d" i.
(5.122)
Substitute Eq. (5.122) into Eq. (5.121) and obtain dW
DJB
: d" i 9 dW
.
(5.123)
(linear motion) (rotational motion),
(5.124)
The mechanical energy output is given by dW
:
f dx DJB T d DJB
where f and T are the force and torque imparted to the mechanical DJB DJB member by the magnetic field. It is important to note that a positive force (or torque) is assumed to be in the same direction as a positive displacement dx (or d). From Eqs. (5.123) and (5.124) we have dW
DJB
:
id" 9 f dx DJB id" 9 T d DJB
(linear motion) (rotational motion).
(5.125)
We desire relations for f and T in terms of the field energy W . The DJB DJB DJB form of Eq. (5.125) suggests that we can obtain these if we choose " and
378
CHAPTER 5 Electromechanical Devices
x or as the independent variables. With this choice, W is of the form DJB W (", x) (linear motion) DJB W : DJB W (", ) (rotational motion). DJB The total derivative of W is DJB W (", x) W (", x) DJB DJB d" ; dx (linear motion) " x (5.126) dW : DJB W (", ) W (", ) DJB DJB d" ; d (rotational motion). "
Now, we compare Eq. (5.125) with Eq. (5.126) and find that W
(", x) DJB x W (", ) DJB T :9 DJB f :9 DJB
(linear motion) (5.127) (rotational motion),
which are the desired results. Recall from Eq. (3.74) in Section 3.2.5 that for linear systems the energy stored in the magnetic field is W : Li (linear systems). DJB Moreover, from Eq. (3.78) we know that L:
" . i
Therefore, W
DJB
:
"i . 2
(5.128)
We use Eq. (5.128) to express W in terms of the independent variables DJB " and x or . This gives
"i(", x) " : 2 2L(x)
W : DJB "i(", ) " : 2 2L()
(linear motion) (5.129) (rotational motion).
379
5.7 ENERGY ANALYSIS
We can use Eq. (5.129) to determine f and T . First determine L(x) or DJB DJB L(), substitute these into Eq. (5.129), and then obtain f and T using DJB DJB Eq. (5.127). Instead of choosing flux linkage " and position (x or ) as the independent variables, it is often more natural and convenient to work with current i and position. When this is the case, the energy has a functional dependency of the form W (i, x) or W (i, ), and Eq. (5.125) DJB DJB becomes
id"(i, x) 9 f dx DJB dW : DJB id"(i, ) 9 T d DJB
(linear motion) (rotational motion),
(5.130)
where d"(i, x) :
"(i, x) "(i, x) di ; dx i x
(linear motion)
d"(i, ) :
"(i, ) "(i, ) di ; d i
(rotational motion). (5.131)
Similarly,
W (i, x) DJB W : DJB W (i, ) DJB
(linear motion) (rotational motion),
which gives
W (i, x) W (i, x) DJB DJB di ; dx i x
(linear motion)
dW : DJB W (i, ) W (i, ) DJB DJB di ; d i
(5.132) (rotational motion).
The functional form of these relations motivate the definition of the coenergy function W A , DJB W A (i, x) : i"(i, x) 9 W (i, x) DJB DJB (linear motion) (Coenergy)
(5.133)
W A (i, ) : i"(i, ) 9 W (i, ) DJB DJB (rotational motion). We can determine the force and torque directly from the coenergy. To
380
CHAPTER 5 Electromechanical Devices
this end, first take the total derivative of W A (i, x). This gives DJB d(i"(i, x)) 9 dW (i, x) (linear motion) DJB dW A : DJB d(i"(i, )) 9 dW (i, ) (rotational motion) DJB or
"(i, x)di ; id"(i, x) 9 dW (i, x) DJB dW A : DJB "(i, )di ; id"(i, ) 9 dW (i, ) DJB
(linear motion) (5.134) (rotational motion).
Substitute Eqs. (5.123) and (5.124) into Eq. (5.134) and obtain
"(i, x)di ; f dx (linear motion) DJB (5.135) dW A : DJB "(i, )di ; T d (rotational motion). DJB Now dW A is formally given by DJB W A (i, x) W A (i, x) DJB DJB di ; dx (linear motion) x i dW A : (5.136) DJB W A (i, ) W A (i, ) DJB DJB di ; d (rotational motion). i
We compare Eqs. (5.135) and (5.136) and find that ":
W A (i, x) DJB , i
(5.137)
and that f : DJB
W A (i, x) DJB x
W A (i, ) DJB T : DJB
(linear motion) (5.138) (rotational motion).
The relations (5.138) are the desired expressions for force and torque. Last, we obtain an explicit expression for W A from Eqs. (5.129) and DJB (5.133),
"(i, x)i L(x)i : 2 2
WA : DJB "(i, )i L()i : 2 2
(linear motion) (5.139) (rotational motion).
We can use Eq. (5.139) to determine both force and torque. First
381
5.7 ENERGY ANALYSIS
FIGURE 5.11
Electromechanical system: (a) actuator; and (b) energy conversion
diagram.
determine L(x) or L(), substitute these into Eq. (5.139), and then obtain f and T using Eq. (5.138). DJB DJB EXAMPLE 5.7.1 Set up the equations of motion for the actuator circuit of Fig. 5.11 [1]. SOLUTION 5.7.1 This is a magnetic circuit actuator. Choose a convention in which f (force imparted to the moving member by the field) is positive in the DJB direction of increasing displacement x as shown. For the equations of motion use the first-order system (5.86), which is written in terms of the independent variables (i, x),
1 "(i, x) di(t) : V (t) 9 i(t)(R ; R ) 9 v(t) L(x) Q x dt dv(t) 1 : F(i, x) dt m dx(t) : v(t). dt
(5.140)
382
CHAPTER 5 Electromechanical Devices
To solve Eq. (5.140) we need expressions for the terms L(x), "(i, x)/x, and F(i, x). As the independent variables are i and x, we work with the coenergy W A (i, x). From Eq. (5.139) we have DJB L(x)i W A (i, x) : . (5.141) DJB 2 The magnetic circuit of Fig. 5.11 was analyzed in Example 3.5.2 from Section 3.5.1. The inductance L(x) was found to be NA E , (5.142) (A /A )( /)l ; 2x E A A where l is the path length in the core, and A and A are the cross-sectional A A E areas of the core and gap, respectively. Substitute Eq. (5.142) into Eq. (5.141) and obtain L(x) :
NA i E W A (i, x) : . (5.143) DJB 2((A /A )( /)l ; 2x) E A A Next, use Eqs. (5.137) and (5.138) to obtain " and f . We find that DJB W A (i, x) DJB "(i, x) : i :
NA i E , ((A /A )( /)l ; 2x) E A A
(5.144)
and that f (i, x) : DJB
W A (i, x) DJB x
NA i E . (5.145) ((A /A )( /)l ; 2x) E A A The minus sign in Eq. (5.145) implies that the direction of f is opposite to the DJB direction of increasing x (i.e., toward the actuator and opposite to the direction of increasing air gap). The total force F(i, x) on the moving member is a sum of the forces due to the field and spring :9
F(i, x) : f
(i, x) ; f (x), (5.146) DJB Q where f (x) is the force due to the spring. In its initial state, the actuator is Q unenergized with i(0) : 0, and the mass is at rest in an equilibrium position x . In this position the spring force is zero, f (x ) : 0. Thus, when the mass is Q
383
5.8 MAGNETIC CIRCUIT ACTUATORS
at a position x the spring force is f (x) : 9k(x 9 x ). (5.147) Q Notice that when x x (x x ) this force tends to move the mass towards (away from) the magnetic circuit. Substitute Eqs. (5.144), (5.145), (5.146), and (5.147) into Eq. (5.140) and obtain di(t) ((A /A )( /)l ; 2x) E A A : dt NA E
2 NA i E ; V (t)9i(t)(R;R ); v(t) ((A /A )( /)l ;2x) Q E A A dv(t) NA i 1 E :9 ; k(x 9 x ) dt m ((A /A )( /)l ; 2x) E A A dx(t) : v(t). (5.148) dt
This nonlinear first-order system has to be solved subject to the initial conditions x(0) : x , v(0) : v and i(0) : i . ) Finally, for linear systems the energy and coenergy are numerically equal, W : W A . This follows from the fact that " : Li for such DJB DJB systems, which gives " Li : . 2L 2 However, for nonlinear systems " is not proportional to i and W and DJB W A are not numerically equal. The difference between W and W A for DJB DJB DJB linear and nonlinear systems is illustrated in Fig. 5.12.
5.8 MAGNETIC CIRCUIT ACTUATORS In this section we study magnetic circuit actuators. In these actuators the moving member is part of the magnetic circuit, and its motion alters the flux through the electrical circuit. The energy method is especially useful for analyzing these actuators. We demonstrate this in the following examples. EXAMPLE 5.8.1 Consider the magnetic actuator circuit shown in Fig. 5.13a. Assume that the magnetic circuit operates in a linear region of its B-H curve
384
CHAPTER 5 Electromechanical Devices
Graphical representation of energy and coenergy: (a) linear system; and (b) nonlinear system.
FIGURE 5.12
with , and that there is no flux leakage. Determine the force on the plunger [2]. SOLUTION 5.8.1 We determine an analytical expression for the force on the plunger using both energy and coenergy approaches. Choose a convention in which a positive f is in the direction of increasing displacement x as shown. DJB Energy: In the energy approach, " and x are the independent variables and the energy is determined using Eq. (5.129), " W : . DJB 2L(x)
(5.149)
5.8 MAGNETIC CIRCUIT ACTUATORS
385
Actuator with moving plunger: (a) actuator circuit; and (b) reference frame for plunger [2].
FIGURE 5.13
We need to determine the inductance L(x). We use the inductance relation (3.75). L:
1 i
B · H dv.
2 i
B · H dv,
(5.150)
4 As , the core and plunger have a high permeability and the H-field in these elements is negligible. Therefore, Eq. (5.150) reduces to an integration over the gap regions, L:
4g
(5.151)
386
CHAPTER 5 Electromechanical Devices
where V : wg(d 9 x) is the volume of the gap. Factor 2 in Eq. (5.151) takes E into account the identical integrations over the gap regions above and below the plunger. To evaluate Eq. (5.151) we need to know the fields B and H in the E E gap. To this end, apply Eq. (3.141) to a path around the circuit through the plunger and obtain 2H g : ni, E which gives ni H : . E 2g Further, B : H . Substitute B and H into Eq. (5.151) and obtain E E E E nw L(x) : (d 9 x). (5.152) 2g Then, substitute this into Eq. (5.149), which gives g" W (", x) : . DJB nw(d 9 x) Finally, obtain the force using Eq. (5.127), f
DJB
:9
W (", x) DJB x
:9
g" . nw(d 9 x)
(5.153)
(5.154)
As 0 x d and " 0, the force is negative, which means that it acts to pull the plunger towards the core (in the direction of decreasing x). Formula (5.154) would be more useful if we had an explicit expression for ". For an n-turn coil we have " : n where is the flux through each turn of the coil. As there is no flux leakage, : : B A where A : w(d 9 x). Therefore E E E E ni ni : w(d 9 x) and " : w(d 9 x). 2g 2 g Coenergy: In the coenergy approach i and x are the independent variables and the coenergy is given by Eq. (5.139), WA : DJB
L(x)i . 2
(5.155)
5.8 MAGNETIC CIRCUIT ACTUATORS
387
Substitute Eq. (5.152) into Eq. (5.155) and obtain in WA : w(d 9 x). DJB 4g
(5.156)
The force follows from Eq. (5.138), inw f :9 . DJB 4g
(5.157)
Notice that the force is independent of x. The two force expressions (5.154) and (5.157) are equivalent. To see this, substitute Eq. (3.78) (i.e., " : L(x)i) into Eq. (5.154), which gives Eq. (5.157). ) EXAMPLE 5.8.2 Consider the cylindrical actuator shown in Fig. 5.14. Assume that the magnetic circuit is linear with , and that there is no flux leakage. Set up the equations of motion for the actuator [2]. SOLUTION 5.8.2 First, choose a convention in which a positive f is in the DJB direction of increasing x as shown. As the motion is linear, we use the first-order
FIGURE 5.14
Cross section of a cylindrical actuator circuit [2].
388
CHAPTER 5 Electromechanical Devices
system of equations (5.86),
di(t) 1 "(i, x) : V (t) 9 i(t)(R ; R ) 9 v(t) dt L(x) Q x dv(t) 1 : F(i, x) m dt
(5.158)
dx(t) : v(t). dt To solve Eq. (5.158) we need expressions for L(x), "(i, x)/x, and F(i, x). As the independent variables are i and x, we work with the coenergy W A (i, x). DJB From Eq. (5.139) we have W A (i, x) : DJB
"(i, x)i 2
:
L(x)i . 2
(5.159)
Inductance: We determine the inductance using Eq. (3.75), L:
1 i
B · H dv. (5.160) 4 As the core and plunger have a high permeability, the H-field in these elements is negligible, and Eq. (5.160) reduces to an integration over the nonmagnetic sleeve regions, L:
1 i
B · H dv ;
1 i
B · H dv,
(5.161)
4 4 where the subscripts top and bot refer to the top and bottom sleeve regions, respectively. Notice that V dgh and V dg(h 9 x). We need to determine H and H . To this end, apply Eq. (3.141) to either of the dotted paths in Fig. 5.14. This gives g ; H g : ni. (5.162) In addition, as there is no flux leakage : , or B A :B A , (5.163) where A : dh and A : d(h 9 x). From Eq. (5.163) and the fact that H
5.8 MAGNETIC CIRCUIT ACTUATORS
B
: H and B : H , we find that h H :H . (h 9 x)
389
(5.164)
Substitute Eq. (5.164) into Eq. (5.162) and obtain H
:
ni (h 9 x) . g (2h 9 x)
(5.165)
Finally, substitute Eqs. (5.164) and (5.165) into Eq. (5.161), which gives the inductance ndh (h 9 x) . L(x) : g (2h 9 x)
(5.166)
Coenergy: Once we know the inductance, the coenergy is easily obtained from Eq. (5.159), indh (h 9 x) W A (i, x) : . DJB 2g (2h 9 x)
(5.167)
We use Eqs. (5.137) and (5.138) to obtain " and f , and find that DJB W A (i, x) DJB ": i indh (h 9 x) : , g (2h 9 x)
(5.168)
and f
DJB
:
W A (i, x) DJB x
ind h :9 . (2h 9 x) 2g
(5.169)
As x h, Eq. (5.169) is negative, which implies that the direction of f is DJB opposite to the direction of increasing x (i.e., opposite to the direction of increasing air gap). The total force F(i, x) is a sum of the forces due to the field and spring, F(i, x) : f where
DJB
(i, x) ; f (x). Q
(5.170)
f (x) : 9k(x 9 x ). (5.171) Q Here, x is the initial position of the plunger, which is assumed to be the equilibrium position for the spring. When x x (x x ), f (x) tends to move Q
390
CHAPTER 5 Electromechanical Devices
the mass into (out of) the solenoid. Finally, substitute Eqs. (5.168), (5.169), (5.170), and (5.171) into Eq. (5.158) and obtain
di(t) (2h9x) ind h g : V (t)9i(t)(R;R ); v(t) Q dt ndh (h9x) g (2h9x) dv(t) 1 ind h :9 ; k(x 9 x ) m dt 2g (2h 9 x)
dx(t) : v(t). dt
(5.172)
This nonlinear first-order system is valid for 0 x h. It is solved subject to the initial conditions x(0) : x , v(0) : v and i(0) : i . ) EXAMPLE 5.8.3 Set up the equations of motion for the actuator shown in Fig. 5.15. Assume that all magnetic materials are linear with . The nonmagnetic sleeves of thickness s provide low friction support for the moving plunger [4]. SOLUTION 5.8.3 Set up a coordinate system with f positive in the direction DJB of increasing x as shown. As the motion is linear, we use the first-order system of equation (5.86)
di(t) 1 "(i, x) : V (t) 9 i(t)(R ; R ) 9 v(t) dt L(x) Q x dv(t) 1 : F(i, x) dt m dx(t) : v(t). dt
FIGURE 5.15
(5.173)
Cross section of an actuator circuit [4].
391
5.8 MAGNETIC CIRCUIT ACTUATORS
To solve Eq. (5.173) we need expressions for L(x), "(i, x)/x and F(i, x). As the independent variables are i and x, we work with the coenergy W A (i, x). From DJB Eq. (5.139) we have "(i, x)i W A (i, x) : DJB 2 :
L(x)i . 2
(5.174)
Inductance: We determine the inductance using Eq. (3.75), L:
1 i
B · H dv. (5.175) 4 As the core and plunger have a high permeability, H is negligible in these elements and Eq. (5.175) reduces to an integration over the gap region and the nonmagnetic sleeve regions, L:
1 i
B · H dv ;
2 i
B · H dv, (5.176) 4E 4Q where V and V are the volumes of the gap and sleeve regions, respectively. The E Q factor of 2 in the second integral takes into account the identical integrations over each sleeve region. Notice that V : A x and V : A s, where A and A E E Q Q E Q are the cross-sectional areas of the gap and sleeve, respectively. We need to determine H and H . To this end, apply Eq. (3.141) to the dotted path in Fig. E Q 5.15. This gives H x ; H s : ni. (5.177) E Q Further, as there is no flux leakage : 2 , or E Q B A : 2B A . (5.178) E E Q Q Combining Eqs. (5.177) and (5.178) with B : H and B : H gives E E Q Q ni H : , (5.179) E (x ; (A /2A )s) E Q and A ni H : E . (5.180) Q 2A (x ; (A /2A )s) Q E Q Next, substitute Eqs. (5.179) and (5.180) into Eq. (5.176), which gives the inductance L(x) :
nA E . (x ; (A /2A )s) E Q
(5.181)
392
CHAPTER 5 Electromechanical Devices
Coenergy: Once we know the inductance, the coenergy is obtained easily from Eq. (5.159) inA E W A (i, x) : . DJB 2[x ; (A /2A )s] E Q Use Eqs. (5.137) and (5.138) to obtain " and f . We find that DJB W A (i, x) DJB ": i :
inA E , [x ; (A /2A )s] E Q
(5.182)
(5.183)
and f
DJB
:
W A (i, x) DJB x
inA E . (5.184) 2[x ; (A /2A )s] E Q Notice that f is negative, which implies that the direction of f is opposite to DJB DJB the direction of increasing x (i.e., opposite to the direction of increasing air gap). The total force F(i, x) is a sum of the forces due to the field and the spring, :9
F(i, x) : f
DJB
(i, x) ; f (x), Q
(5.185)
where f (x) : 9k(x 9 x ). (5.186) Q Here, x is the initial position of the plunger, which is assumed to be the equilibrium position for the spring. When x x (x x ), f (x) tends to move Q the mass into (out of) the circuit. Finally, substitute Eqs. (5.181), (5.184), (5.185), and (5.186) into Eq. (5.173) to obtain
di(t) (x ; (A /2A )s) inA E Q E : V (t) 9 i(t)(R ; R ) ; v(t) Q dt nA [x ; (A /2A )s] E E Q dv(t) 1 inA E :9 ; k(x 9 x ) dt m 2[x ; (A /2A )s] E Q dx(t) : v(t). (5.187) dt
This nonlinear first-order system is solved subject to the initial conditions x(0) : x , v(0) : v , and i(0) : i . )
393
5.9 AXIAL-FIELD ACTUATORS
5.9 AXIAL-FIELD ACTUATORS Axial-field actuators are moving coil actuators. These actuators are well suited and widely used for applications that require controllable bidirectional rotation. For example, they are routinely used for positioning read/write heads in computer disk drives. These actuators consist of pie-shaped magnets that are positioned above (and/or below) a coil that is free to rotate as shown in Fig. 5.16. When the coil is energized, it experiences a torque and rotates either clockwise or counterclockwise depending on the direction of current. A mechanical component such as a spring is often used to provide a restoring torque, and to fix the unenergized position of the coil (Fig. 5.17). Axial-field actuators can be designed and optimized prior to fabrication using lumped-parameter analysis [9, 10]. We develop a model for performing such analysis in the following example. EXAMPLE 5.9.1 Determine the equations of motion for the axial-field actuator shown in Fig. 5.16. Assume that the coil has n turns. SOLUTION 5.9.1 This device is a moving coil actuator and is governed by Eqs. (5.102),
di(t) 1 : V (t) 9 i(t)(R ; R ) ; dt L Q
[( ; r) ; B
· dl
1 d (t) : i(t) [r ; (dl ; B )] · z ; T ()
dt J K d(t) : (t). dt
FIGURE 5.16
Drive magnet and coil for an axial-field actuator.
(5.188)
394
CHAPTER 5 Electromechanical Devices
FIGURE 5.17
Axial-field actuator with spring mechanism.
Here, L and R are the inductance and resistance of the coil, and J is its K moment of inertia above the z-axis. We need to evaluate the induced voltage and torque integrals in Eq. (5.188). We use a cylindrical coordinate system at rest with respect to magnets as shown in Fig. 5.16. Before we begin, we state some simplifying assumptions. Assumptions: The first assumption is that the magnetic field due to the magnet is essentially uniform and constant across the coil, that is,
9B z 0
K (5.189) B z 9 0,
K where 9 and ; define the angular positions of the radial edges of the K K B :
395
5.9 AXIAL-FIELD ACTUATORS
FIGURE 5.18
Electrical connection for a single turn of the coil.
magnet. In reality, the field distribution above the magnet varies with height, and from point to point at a fixed height. However, an average value of B
that is adequate for this analysis can be obtained using three-dimensional FEA, closed-form analysis [11], or measured data if the fabricated magnet is available. The second assumption is that each of the coil’s two radial edges rotates over a separate magnetic pole throughout the entire range of motion. Induced voltage: We determine the voltage induced at the terminals of the coil by summing the voltage induced across its various segments. For reference purposes, Fig. 5.18 shows the electrical connection of a single turn. The coil has four segments: an upper and a lower segment, and two side segments (Fig. 5.19). We consider a single turn of the coil and evaluate the side (radial) segments first. We need to determine ( ; r) ; B along these segments. First, we know that (t) : (t)z . Then from Fig. 5.19 we have r : rr . Thus, ; r : (t)r Note that B is in the z-direction, 9B z B : B z
(side segments).
(left side) (right side).
396
CHAPTER 5 Electromechanical Devices
FIGURE 5.19
Top view of the coil with the magnetic field.
Therefore, ( ; r) ; B
:
9 (t)rB r (t)rB r
(left side) (right side).
(5.190)
We evaluate
[(;r);B ] · dl
along the left and right sides where dl : dr r
(side segments).
We obtain
[(;r);B ] · dl : 9 left side
0 (t)B r dr 0
R 9 R , : 9 (t)B 2
(5.191)
397
5.9 AXIAL-FIELD ACTUATORS
and
[( ; r) ; B ] · dl :
0 (t)B r dr 0
right side
R 9 R . : 9 (t)B 2
(5.192)
Now we consider the top and bottom segments. For these we have ( ; r) ; B r . Because dl , it follows that [( ; r) ; B ] · dl r · : 0
(top/bottom segments).
Therefore,
[( ; r) ; B ] · dl : 0, top/bottom segments
which shows that these segments do not contribute to the induced voltage. Finally, add Eqs. (5.191) and (5.192), and multiply by n (the number of turns) to obtain an expression for the induced voltage
[( ; r) ; B ] · dl : 9 (t)nB (R 9 R).
(5.193)
Torque: Only the side segments of the coil contribute to the torque. We make a brief argument to show that the upper and lower segments make no contribution. Consider the torque on an element of the upper or lower segment, dT : [r ; (dl ; B )] · z , X where r : rr and dl . . Notice that dl ; B
. ; B . r .
This gives dT . [r ; r ] · z : 0, X which shows that the upper and lower segments generate no torque. Now
398
CHAPTER 5 Electromechanical Devices
consider the left- and right side segments. For these, we have dl : drr , and
B dr dl ; B : 9B dr
(left side) (right side).
Therefore,
B r dr z r ; (dl ; B ) : 9B r dr z
(left side) (right side).
(5.194)
We evaluate
[r ; (dl ; B )] · z
along the left- and right side segments with the integration in the direction of the current flow. This gives
[r ; (dl ; B )] · z :
0 B r dr 0
left side
:B and
(R 9 R) , 2
(5.195)
0 B r dr [r ; (dl ; B )] · z : 9 0 right side
:B
(R 9 R) . 2
(5.196)
The torque on the coil is ni times the sum of Eqs. (5.195) and (5.196),
[r ; (dl ; B )] · z : niB (R 9 R). (5.197) Equations of motion: We are finally ready to write the equations of motion. Substitute the integrals (5.193) and (5.197) into Eq. (5.188) and obtain di(t) 1 : [V (t) 9 i(t)(R ; R ) 9 nB (R 9 R) (t)] dt L Q d (t) 1 : [i(t)nB (R 9 R) ; T ()]
dt J K d(t) : (t). dt
(5.198)
399
5.9 AXIAL-FIELD ACTUATORS
These equations are solved subject to the initial conditions of Eq. (5.89). Finally, from Eq. (5.198) we see that the electrical constant K equals the torque constant C K (see Eq. (5.100)). Specifically, R K : K : nB (R 9 R). C R Calculations: We apply Eq. (5.198) to an actuator with the mechanical mechanism shown in Fig. 5.17. Here, the mechanical torque is supplied by the spring. To perform the analysis we need expressions for the coil inductance L and the mechanical restoring torque T ().
Inductance: The inductance of the coil can be estimated by considering a short circular coil with a height h equal to that of the actuator coil, and with A a radius r , A R 9R . r : A 2
(5.199)
To first order, the flux through the center of the coil (through each turn) is (i)
inr A . (h ; (2r )) A A
(5.200)
Therefore, the flux linkage is "(i) : n(i). Recall that the inductance is d"(i) , di
(5.201)
nr A . (h ; (2r )) A A
(5.202)
L: which gives L
This approximation is appropriate if the mean circumferential arc subtended by the coil is approximately equal to R 9 R , which is the case for many practical designs. However, if this is not the case the inductance can be determined empirically, or calculated using three-dimensional FEA. Mechanical Torque: The mechanical torque is provided by the spring mechanism shown in Fig. 5.17. The restoring force is F(Y) : k (Y 9 Y ) ; F , Q where k is the spring constant, Y is the initial spring length, F is the force Q
400
CHAPTER 5 Electromechanical Devices
FIGURE 5.20
Applied voltage V (t). Q
when Y : Y , and Y is the stretched length. The length Y is defined by the angle #. From geometry we have Y : (X ; D 9 2XD cos(# ; ). The restoring torque is given by () : DF(Y()) cos(# ; ). (5.203)
Analysis: We substitute Eqs. (5.202) and (5.203) into Eq. (5.198) and perform the analysis. The applied voltage V (t) is shown in Fig. 5.20 and the remaining Q parameters are as follows: T
B : 0.5 T
R;R :3 n : 100 turns R : 10 mm R : 15 mm
401
5.9 AXIAL-FIELD ACTUATORS
FIGURE 5.21
Current i(t).
D : 10 mm # : 45° h : 1.5 mm A k : 1.5 N/mm Q J : 6 ; 10\ kg · m. K The moment of inertia is computed using Eq. (5.77), which gives J : (R 9 R) A ; (R 9 R) A , K where and A are the density and area of the wire and is the
angular span of the coil (left- to right side) in radians. The response of the actuator is computed for the initial conditions i(0) : 0A, (0) : 0°, and (0) : 0 rad/sec. The current profile i(t) is shown in Fig. 5.21. The rotation angle (t) and angular velocity (t) are shown in Fig. 5.22. Notice that the angular velocity (t) passes through zero and turns negative as the angular position (t) peaks at approximately 6.5 ms. )
402
CHAPTER 5 Electromechanical Devices
FIGURE 5.22
Angular position (t) and velocity (t).
5.10 RESONANT ACTUATORS Resonant actuators are used as drive mechanisms for scanning elements in optical scanning systems [12]. An example of such an actuator is shown in Fig. 5.23. In this actuator a wedge-shaped magnet provides an axial bias field, and a coil rotates above the magnet once it is energized. A torsional pivot mechanism provides the restoring torque. When power consumption is an issue, the torsional mechanism can be designed to render the actuator resonant at the scanning frequency. When this is the case, the actuator can be brought to a resonant oscillation with a sequence of low-energy pulse excitations, thus minimizing power consumption. We demonstrate this in the following example. EXAMPLE 5.10.1 Develop a model for the design of the axial-field resonant actuator shown in Fig. 5.23. SOLUTION 5.10.1 The behavior of this device is governed by the equations for an axial-field actuator as derived in Example 5.9.1. We repeat them here for
403
5.10 RESONANT ACTUATORS
FIGURE 5.23
Resonant actuator.
convenience, di(t) 1 : [V (t) 9 i(t)(R ; R ) 9 nB (R 9 R) (t)] dt L Q d (t) 1 : [i(t)nB (R 9 R) ; T ()]
dt J K d(t) : (t). dt
(5.204)
Mechanical torque: The restoring torque is provided by a torsional pivot mechanism T () : 9K . (5.205)
N Here, K is the spring constant of the pivot. This simple linear relation is N typically only valid for limited rotations, for example, 912° 12°. The
404
CHAPTER 5 Electromechanical Devices
resonant frequency f for this mechanism is f:
1 2
K N. J K
(5.206)
Calculations: We demonstrate the use of Eq. (5.204) with some sample calculations. We numerically integrate these equations using the fourth-order Runge-Kutta method (Appendix C). The following parameters are used for the analysis: B
: 0.55 T
R;R :6 n : 100 turns R : 3.2 mm R : 11.5 mm L : 0.4 mH J : 0.3 ; 10\ kg · m K f : 100 Hz The field B : 0.55 T can be achieved using sintered NdFeB magnets above and below the coil with soft-magnetic flux plates attached to the outer surface of each magnet. The flux plates enhance the field across the coil. We apply a periodic step function voltage V (t) with a magnitude of <3 V and a period of Q 10 ms as shown in Fig. 5.24. It only requires two cycles to bring the system to resonance after which the circuit is assumed to be open (i(t) : 0 for t 20 ms). The actuator response is computed with the following initial conditions: i(0) : 0 A, (0) : 0°, and (0) : 0 rad/s. The circuit current is plotted in Fig. 5.25. The rotation angle (t) and back voltage V (t) are plotted in Fig. 5.26. The back voltage is the voltage induced in the coil as it moves. It is given by V (t) : nB (R 9 R) (t). Note that V (t) passes through zero as the angular displacement peaks ( (t) : 0), as it should. Because there are no damping terms, the actuator continues in a resonant mode with an angular oscillation of : <12°. This type of performance is consistent with many such systems where the only appreciable dissipation results from windage and eddy current losses. )
405
5.10 RESONANT ACTUATORS
FIGURE 5.24
Applied voltage V (t). Q
FIGURE 5.25
Current i(t).
406
CHAPTER 5 Electromechanical Devices
FIGURE 5.26
V (t) and (t) vs t.
5.11 MAGNETOOPTICAL BIAS FIELD ACTUATOR Magnetooptical (M-O) recording was discussed in Section 4.10. The basic elements of an M-O system are shown in Fig. 5.27. Recall that M-O systems require oppositely oriented bias fields for the recording and erasure processes [13—15]. In this section we discuss an actuator that serves the dual function of providing the bias field and implementing the bias field reversal. Specifically, we analyze the rotary permanent magnet actuator shown in Fig. 5.28. This actuator consists of a cylindrical bipolar bias magnet (Fig. 5.29), surrounded by a stationary concentric conductive shell and a drive coil. When the coil is energized it rotates the magnet to one of two equilibrium positions, thereby producing a bias field of desired polarity across the recording media. Rotation of the magnet gives rise to eddy currents in the conductive shell that impart a drag torque and reduce oscillations of the magnet about its equilibrium positions. We develop a model for this actuator in the following example [16, 17].
5.11 MAGNETOOPTICAL BIAS FIELD ACTUATOR
FIGURE 5.27
Basic elements of a magnetooptic recording system.
FIGURE 5.28
Cross section of a rotary bias field actuator.
FIGURE 5.29
Cylindrical bipolar bias magnet.
407
408
CHAPTER 5 Electromechanical Devices
EXAMPLE 5.11.1 Set up the equations of motion for the rotary bias actuator shown in Fig. 5.28. The magnet has a radius a and length h, with a h. The coil has n turns and has a length l . A SOLUTION 5.11.1 This is a moving magnet actuator and is governed by the equations for rotational motion (5.87), which we repeat here for convenience:
di(t) 1 "(i, ) : V (t) 9 i(t)(R ; R ) 9 (t) dt L Q
d (t) 1 : T(i, , ) J dt K d(t) : (t). dt
(5.207)
These equations contain nonlinear terms and can be solved numerically once expressions for L, "(i, )/, and T(i, , ) have been determined. We have added an dependence in T(i, , ) to account for the effects of eddy currents in the conductive shell. Inductance: The inductance of the coil can be estimated using the result of Eq. (3.87) for two parallel wires (Example 3.2.10). Taking into account the fact that there are n turns, the inductance is approximated by
s nl , (5.208) L A ln r U where s is the mean separation of the coils, l is the length of each side of the A coil (into the page), and r is the radius of the wire. U Induced voltage: A voltage V is induced in the coil as the magnet rotates. It is given by V
(t) :
"(i, ) (t).
(5.209)
The flux linkage "(i, ) can be written as the sum of two terms, "(i, ) : " (i) ; " ().
The first term " (i) is the flux linkage due to the current through the coil itself. The second term " () is the flux linkage due to the magnet. Notice that
only the second term has an angular dependence and, therefore, "(i, ) " () : .
409
5.11 MAGNETOOPTICAL BIAS FIELD ACTUATOR
To determine this term we need an expression for the field of the magnet. For this actuator, the length of the magnet is much greater than its diameter, and a two-dimensional analysis suffices (Fig. 5.27). A two-dimensional field solution for the magnet was derived in Example 3.6.3. The vector potential was found to be M a A (r, ) : Q sin()z
2 r
(a r).
(5.210)
Recall that Eq. (5.210) applies when the magnet is positioned with its north pole aligned with : 0 as defined by Eq. (3.264). If the magnet is rotated by an angle then Eq. (5.210) is modified by substituting ; 9 , that is, A
M a (r, , ) : Q sin( 9 )z .
2 r
(5.211)
The flux linkage due to the magnet is "
() : n
:n
1
B (r, , ) · ds
A
(r, , ) · dl.
Therefore, " ()
:n
A (r, , ) · dl
\JA :n A (R , 0, ) dz ; A (R , , ) dz
A
A \JA a : n M l cos(), (5.212) Q AR A where is the angular position of the magnet relative to the x-axis, and R is A the radial position of the sides of the coil. The coil has n side elements (into the page) that are positioned at r : R , and : 0 and , respectively. We A substitute Eq. (5.212) into Eq. (5.209) and obtain
V (t) : K ((t)) (t), C where K is the electrical constant C a K ((t)) : n M l cos((t)). C Q AR A
(5.213)
(5.214)
410
CHAPTER 5 Electromechanical Devices
Torque: The torque on the magnet can be written as a superposition of two terms, T(i, , ) : T
(i, ) ; T ( ). (5.215) The terms T (i, ) and T ( ) are the torques due to the coil and eddy currents induced in the surrounding conductive shell, respectively. Drive torque: We consider the torque imparted to the coil by the magnet. This is given by Eq. (5.79), T:i
r ; (dl ; B ), (5.216) where r is the distance from the current element dl to the axis of rotation and integration is in the direction of current flow. Here, r : R r , and dl : dzz . A Therefore, r ; (dl ; B ) : R B dz z , A P where B is given by Eq. (3.274), P M a B (r, , ) : Q cos( 9 ). P 2 r
(5.217)
In Eq. (5.217), (r, ) is the observation point and is the rotational position of the magnet with respect to the x-axis. The coil has n pairs of wires positioned as described in the foregoing. Therefore, the axial component of the torque is
M a \JA T () : ni(t) Q cos(0 9 ) dz ; cos( 9 ) dz X 2 R \JA A a : 9ni(t) M l cos((t)). (5.218) Q AR A This is the torque imparted to the coil by the magnet. The torque on the magnet is of equal magnitude but in the opposite direction. Therefore, the torque imparted to the magnet by the coil is a (i(t), (t)) : ni(t) M l cos((t)). Q AR A By comparing Eq. (5.214) with Eq. (5.219) we find that T
(i(t), (t)) : K ((t))i(t), R where K ((t)) is the torque constant and K ((t)) : K ((t)). R R C T
(5.219)
(5.220)
5.11 MAGNETOOPTICAL BIAS FIELD ACTUATOR
411
Eddy currents: As the magnet rotates it induces eddy currents in the conductive shell that give rise to a drag torque. To analyze this, we choose a reference frame that is at rest with respect to the shell, and with its origin on the axis of the magnet. We assume that the shell is thinner than the skin depth . associated with the desired field reversal time. For example, . 25 mm for a field reversal of 10 ms. Based on these assumptions, the magnetic field induced in the shell is negligible compared to the magnet’s field, which is obtained from the vector potential equation (5.211). Therefore, the electric field induced at a position (r, ) in the shell is E(r, , (t)) : 9 :9
A
(r, , (t))
t
A
(r, , (t))
(t).
(5.221)
This gives rise to an induced current density in the shell, J(r, , (t)) : E(r, , (t)),
(5.222)
where is the conductivity of the shell. Substituting Eqs. (5.211) and (5.221) into Eq. (5.222) gives M a J (r, , (t)) : Q cos( 9 (t)) (t). X 2 r
(5.223)
Thus, the current through an infinitesimal cross section r dr d of the shell is di(t) : J (r, , (t))r dr d. (5.224) X The drag torque is obtained by combining Eqs. (5.216), (5.217) and (5.224) and then integrating over the cross-sectional area of the shell (we assume that the shell is as long as the magnet). This gives T
M L 0 1 Q (t) cos( 9 (t)) dr d 2 r 0 M R : ha Q ln (t). (5.225) 2 R
( (t)) : ha
We write this as T
( (t)) : 9K (t),
(5.226)
where K : Sgn( (t))ha
M R Q ln , 2 R
(5.227)
412
CHAPTER 5 Electromechanical Devices
and Sgn( (t)) :
1 91
(t) & 0 (t) 0.
The term Sgn( (t)) takes into account the fact that the torque due to the eddy current acts to oppose the rotation. Notice that we have ignored the current induced in the conductive shell by the drive coil, and the voltage induced in the coil by the eddy currents within the shell. These effects are second order [17]. Equations of motion: Finally, substitute Eqs. (5.213), (5.220), and (5.226) into Eq. (5.207) and obtain the equation of motion for the actuator, di(t) 1 : [V (t) 9 i(t)(R ; R ) 9 K ((t)) (t)] C L Q dt d (t) 1 : [K ((t))i(t) 9 K (t)] R dt J K d(t) : (t), dt
(5.228)
where L is given by Eq. (5.208). Calculations: To demonstrate the theory we consider an M-O system that requires a bias field in the range 0.02—0.04 T along a radial length of 100 mm, and a field-reversal time of 12 ms. A magnet for this system was designed in Example 4.10.1. Specifically, we use a sintered NdFeB magnet (M : 7.2;10 Q A/m) with a radius R : 1 mm, and a length h : 108 mm. This magnet needs to be rotated : <85° in 12 ms. Preliminary analysis: Before we design the actuator we need to estimate nominal values for several key parameters. First, we determine the moment of inertia of the magnet. This is given by J : 1/2hR, where : 7.52 ; 10 K kg/m is the density of sintered NdFeB, and R and h are as given in the preceding text. Using these values we obtain J : 1.275 ; 10\ kg/m. Second, K we estimate the number of amp turns required to reverse the field in 12 ms. This is a three-step process: First, we estimate the angular acceleration : 2/t required to rotate the magnet through an angle of : in less than half of the reversal time, t : 5 ; 10\ s. We use this reduced time because the acceleration must be great enough to compensate for the drag due to the conductive shell. We calculate : 2.5 ; 10 rad/s. Next, we estimate the drive torque T : J , which yields T : 3.18 ; 10\ Nm. Finally, we estimate the amp K turns using Eq. (5.219). We use the following nominal values: (t) : /3, l : h, and R : 3 ; 10\ m. This value of R (radial position of the coil) places A A A
413
5.12 LINEAR ACTUATORS
the coil far enough away from the magnet to accommodate the conductive shell. Using these values we obtain T R A M ha cos(/3) Q : 20.
ni :
(5.229)
If we choose a nominal value of current i : 200 ma, then n : 100 turns. We are finally ready to perform the actuator design. Actuator design: Our goal is to determine the radial thickness of the conductive shell that renders the desired field reversal time. We integrate Eq. (5.228) using the fourth-order Runge-Kutta method with the inner radius of the shell set to R : 2.0 mm (1.0 mm from the surface of the magnet) and the outer radius set to R : 2.5, 3.0 and 3.5 mm, respectively. Further, we assume that the shell is made from aluminum, which has a conductivity : 3.82 ; 10 m\. For these calculations, the radial position of the coil is 0.5 mm beyond the outer surface of the shell, that is, R : 3.0, 3.5 and 4.0 mm, A respectively. The excitation voltage is held constant at V : 12 V and 36-gauge Q wire is used with n : 100 turns. The response of the actuator is computed subject to the following initial conditions: i(0) : 0 A, (0) : 985°, (0) : 0 rad/s. The resistance and inductance of the coil compute to 32.6 and 2.0 mH, respectively. A parametric plot of (t) vs t with R : 2.5, 3.0 and 3.5 mm is shown in Fig. 5.30. Notice that as R increases, the eddy current damping increases and the magnitude of the oscillations about the equilibrium position of 90° decrease. When R : 3.5 mm, the magnet stabilizes at 90° within the desired reversal time of 12 ms. Thus a shell with R : 2.0 mm and R : 3.5 mm will suffice for this application. Additional plots of i(t) and V (t) : K((t)) (t) for R : 3.0 mm are shown in Fig. 5.31. Last, we note that in practice, the equilibrium position of the magnet is offset from 90° (e.g., 85°) so that it can be rotated by the coil. This offset is usually accomplished using magnetic detent in which a soft-magnetic element is fixed in proximity to the magnet. The presence of this element gives rise to second-order effects, and will not alter the analysis appreciably. )
5.12 LINEAR ACTUATORS Linear actuators are used for applications that require smooth and controllable bidirectional motion. For example, they are used as radial access actuators in optical and magnetic disk drives. In this section we
414
CHAPTER 5 Electromechanical Devices
FIGURE 5.30
Parametric plot of (t) vs t with R : 2.5, 3.0 and 3.5 mm.
FIGURE 5.31
Current and induced voltage vs t with R : 3.0 mm.
415
5.12 LINEAR ACTUATORS
FIGURE 5.32
Linear actuator: (a) perspective of actuator; and (b) cross-sectional
view.
study the moving coil linear actuator shown in Fig. 5.32a. This actuator consists of a stator and a moving coil that executes bidirectional motion. The stator consists of a closed soft magnetic core with a permanent magnet mounted on the inside surface of the base (Fig. 5.33a). The coil encloses the top segment of the stator and is free to move along it. The operation of the actuator is straightforward. The magnet produces a field B in the gap region of the stator, and when the coil is E energized, the wires that pass through the gap experience a Lorentz force that causes the coil to move (Fig. 5.32b). This is a moving coil actuator and its behavior is governed by the equations for linear motion (5.98), di(t) V (t) : i(t)(R ; R ) ; L 9 Q dt dx(t) m : i(t) dt
(u ; B ) · dl
(dl ; B ) · x ; F (x).
(5.230)
416
CHAPTER 5 Electromechanical Devices
Linear actuator stator: (a) perspective of stator; and (b) crosssectional view.
FIGURE 5.33
Here m, R , and L are the mass, resistance, and inductance of the coil. The terms m and R are easily determined given the number of turns n of the coil and the wire gauge. To solve this system we need expressions for L, the induced voltage (u ; B ) · dl, and the Lorentz force i (dl ; B ) · x . We derive these under the assumption that the stator is operating in a linear region of its B-H curve with B : H . First, we determine the inductance L. Consider the total flux (i, x) through the top segment of the stator. As the magnetic circuit is linear, (i, x) is the sum of the flux due to the coil (i), and the flux due to the magnet (x),
(i, x) : (i) ; (x).
The flux linkage can also be separated into two components, "(i, x) : " (i) ; " (x).
(5.231)
417
5.12 LINEAR ACTUATORS
The self inductance of the coil is L:
d" (i) . di
(5.232)
The magnetic circuit is essentially a coil wrapped around a closed core and was analyzed in Example 3.5.1. Thus, if we assume negligible leakage flux, L is given by Eq. (3.164), L:
nA A, l A
(5.233)
where and A are the permeability and cross-sectional area of the core A and l is the average path length around the core. A Next, we determine the induced voltage (u ; B ) · dl. We use the reference frame shown in Fig. 5.33a. Specifically, we have u : u(t)x , B : B y , and dl : dzz . Only the wires in the gap contribute to the E induced voltage. Therefore,
(u ; B ) · dl : n
\UA u(t)B dz E
: 9nw B u(t), A E
(5.234)
where w is the width of the core (Fig. 5.33a). It remains to determine B . A E This follows from the results of Example 3.5.3. Specifically, we apply the analysis presented there to one half of the stator circuit (dotted line in Fig. 5.33b), assuming that there is no leakage flux and that the magnet has a linear demagnetization curve of the form, B :B ; H . K P K K From Eq. (3.179) we have B l P K, B : E ( ; l /g) g K K
(5.235)
where : B /H and A : A . Substitute Eq. (5.235) into Eq. (5.234) K P A E K and obtain
l nw B K u(t). A P (u ; B ) · dl : 9 ( ; l /g) g K K
(5.236)
This is the desired expression for the induced voltage. If : , Eq. K
418
CHAPTER 5 Electromechanical Devices
(5.236) reduces to
nw B A P u(t). (u ; B ) · dl : 9 (g/l ; 1) K Last, we determine the Lorentz force. Again, we have dl : dz z and B : B y . Therefore, E \UA B dz, i (dl ; B ) · x : 9i n E :inw B . (5.237) A E Substitute Eq. (5.235) for B in Eq. (5.237) and obtain E nw B l A P K. (5.238) i (dl ; B ) · x : i ( ; l /g) g K K This is the Lorentz force on the coil. We are finally ready to write the equations of motion. Substitute Eqs. (5.233), (5.236), and (5.238) into Eq. (5.230). This gives
V (t) : i(t)(R ; R ) ; Q
nA di(t) dx(t) A ;K dt l dt A
(5.239)
dx(t) m : Ki(t), dt where we have used u(t) : dx(t)/dt and n w B l A P K. (5.240) ( ; l /g) g K K Equations (5.239) are solved subject to initial conditions i(0):i , x(0):x , and u(0) : dx(t)/dt : u . R The coupled equations (5.239) can be rewritten as a first-order system and solved numerically using techniques such as the Euler or the Runge-Kutta method (Appendix C). Rapid parametric simulations are possible once drive voltage V (t) and resistance R are specified. Q The response time of a linear actuator is limited by the time constant of the coil : L/R . However, the response time can be reduced by putting a conductive sleeve around a segment of the stator as shown in Fig. 5.34. The sleeve is referred to as a shorted turn [18]. Consider an equivalent magnetic circuit for the coil-stator-sleeve system (Fig. 5.35). This is basically a transformer with the drive coil as the primary, and the shorted turn as the secondary. The electrical equations for this system K:
419
5.12 LINEAR ACTUATORS
FIGURE 5.34
Linear actuator with shorted turn.
are di di ; M QR V(t) : i R ; L dt dt di di 0 : i R ; L QR ; M , QR QR QR dt dt
(5.241)
where R and L are the resistance and self-inductance of the coil, R QR and L are the resistance and self-inductance of the shorted turn, and M QR is the mutual inductance. The mutual inductance can be expressed as M : (L L , (5.242) QR where 1 is a coupling constant. We desire an expression for the
FIGURE 5.35
shorted turn.
Equivalent magnetic circuit for the stator with the drive coil and
420
CHAPTER 5 Electromechanical Devices
response time of this system, that is, the time dependence of i . Take the Laplace transform of Eq. (5.241), which gives V(s) : I (s)R ; sL I (s) ; sMI (s) QR (5.243) 0 : I (s)R ; sL I (s) ; sMI (s), QR QR QR QR where all initial conditions have been set to zero. We solve Eq. (5.243) for I (s) and obtain V(s) s ; 1 QR I (s) : , (5.244) R s (1 9 ) ; s( ; ) ; 1 QR QR where : L /R and : L /R are the time constants for the coil QR QR QR and shorted turn, respectively. Since 1 we have
(1 9 ) ( ; ). QR QR The denominator in Eq. (5.244) can be simplified using the approximation sa ; sb ; 1 (sa/b ; 1)(sb ; 1), which applies when a b. Using this approximation reduces the denominator to s (1 9 ) ; s( ; ) ; 1 (s ; 1)(s( ; ) ; 1), (5.245) QR QR P QR where (1 9 ) : QR . P ( ; ) QR Substitute Eq. (5.245) into Eq. (5.244) and obtain s ; 1 V(s) QR I (t) : . R (s ; 1)(s( ; ) ; 1) P QR We examine the response to a step function voltage
0 t0 V 0 t. /s. Transform Eq. (5.246) back to the time domain and V(t) :
Thus, V(s) : V obtain I
(5.246)
1 V (t) : R ( ; 9 ) QR P ; [( 9 )(1 9 e\ROP) ; (1 9 e\RO >OQR)]. QR P
421
5.13 AXIAL-FIELD MOTORS
Since ( ; ), the time response is dominated by (1 9 e\ROP) and P QR the initial current surge can be approximated by V ( 9 ) P (1 9 e\ROP). (t) QR R ( ; ) QR /n we have I
Since L L QR
(1 9 ) . P QR 1 ; R /nR QR In practice, 1 and, therefore, 9 . Thus, the initial current QR P QR surge through the coil when a shorted turn is present is V R (t) (1 9 e\ROP) (shorted turn). R R ; nR QR The response of the coil without the shorted turn is I
I
V (t) : (1 9 e\RO ) R
(no shorted turn).
(5.247)
(5.248)
However, R (1 9 ) , R ; nR P QR and R (1 9 ) 1. R ; nR QR Therefore, . By comparing Eqs. (5.247) and (5.248) we find that P the response time of an actuator with a shorted turn is much shorter than without one. Finally, an alternate linear motor configuration with a shorted turn is shown in Fig. 5.36. The response time of this actuator has been analyzed by Wagner [18].
5.13 AXIAL-FIELD MOTORS Permanent magnet axial-field motors are used for low-torque, servo and speed control applications. They are also referred to as brushless dc motors, and are commonly used as spindle motors for both disk and cassette tape drives [19]. A typical motor configuration is shown in Fig. 5.37. The motor consists of a set of equally spaced stator coils, fixed
422
CHAPTER 5 Electromechanical Devices
Linear actuator: (a) perspective of core; and (b) cross-sectional view with drive coil.
FIGURE 5.36
beneath a cylindrical multipole rotor magnet with flux return plates positioned above the magnet and below the coils to enhance the field in the motor cavity. During normal operation, the coils are energized in a synchronous fashion as they pass under the transition region between adjacent rotor poles. They are synchronized so as to impart a continual steady torque to the rotor. Hall effect sensors are mounted on the stator near the coils to sense the rotor’s magnetic field and synchronize the activation of the coils. The numbers of rotor poles and coils are
423
5.13 AXIAL-FIELD MOTORS
FIGURE 5.37
Permanent magnet brushless dc motor.
chosen to ensure proper start up and steady drive torque with no dead zones. The behavior of the motor is relatively straightforward [20]. When a voltage V is applied, the rotor accelerates and eventually reaches a final steady-state angular velocity . The equations governing the steady state are V : R i ; K , C C
(5.249)
T : K i, R
(5.250)
and
where R and i are the operating resistance and current and K and K C C R are electrical and torque constants. If there is substantial drag torque, it can be taken into account by modifying Eq. (5.250) as follows: T:K i9T . R B
(5.251)
Drag torque T is due to various loss mechanisms such as friction, air B resistance, etc.
424
CHAPTER 5 Electromechanical Devices
FIGURE 5.38
Performance curves of an axial-field motor [19].
We combine Eqs. (5.249) and (5.250) and obtain RT V 9 C . (5.252) K K K C R C Notice that for a fixed voltage V the speed is a linear function of T as shown in Fig. 5.38. There are two parameters that define this line, the no-load speed , and the stall-torque T , where ,* Q V : , (5.253) ,* K C and :
KV T : R . (5.254) Q R C We will show how to compute these parameters in text that follows. The motor efficiency is P T ) : : . P Vi As shown in Fig. 5.38, the efficiency depends dramatically on the operating point of the motor. A good design must take this into account. The design of an axial field motor is usually an iterative process. The designer usually starts with some basic system requirements such as operating speed and voltage, and then determines motor parameters compatible with these specifications. While a complete description of
425
5.13 AXIAL-FIELD MOTORS
this process is beyond the scope of this book, nevertheless a brief review of a typical design cycle is instructive. Assume that a motor is needed and that the operating voltage V, speed , and torque T have been specified along with some basic size constraints. As a first step, we decide on a rotor magnet and drive circuitry. We choose an eight-pole magnet (material and size), and a Y-connected three-phase motor driver circuit with an operating voltage V and resistance R . We choose a total C of six equally spaced drive coils with two coils per phase. Two phases will be active at a time to ensure proper start up and smooth drive. Therefore, the operating voltage V is applied across four coils in series. Next, we design the coils. We start with the operating resistnce R . C This is the total (series) resistance of all active coils. In our case, two phases are active at a time so there are four active coils (two per phase). Once R is known, the coils can be specified (number of turns and wire C gauge). In particular, the number of turns and length of each coil are related to the operating resistance as follows: N N l R : ?A U , C A U
(5.255)
where N is the number of active coils, N is the number of turns in ?A each coil, l is the length of the wire in each coil, and and A are the U U conductivity and cross-sectional area of the wire. In our case there are 4 active coils, N : 4. Notice that Eq. (5.255) constitutes one equation with ?A three unknowns: N , l , and A . Therefore, additional information is U U needed. One approach to obtaining the unknowns is to specify the orientation, shape, and dimensions of the coil. To this end, we specify the motor cavity height h, which puts an upper limit of h 9 t on the K height of the coils (Fig. 5.37). We choose the cross-sectional area A and A mean length per turn l/ (Fig. 5.39). Given the coil dimensions, we write R the following additional relations, N A A : U A f N
(5.256)
and l : f N l/ , (5.257) U E R where f is a wire packing factor and f is a geometric factor. For N E precision wound coils f 0.69. The three unknowns N , l , and A U N U are obtained via the simultaneous solution of Eqs. (5.255), (5.256), and (5.257).
426
FIGURE 5.39
CHAPTER 5 Electromechanical Devices
Coil geometry: (a) top view of the stator; and (b) perspective of
coil.
Once the orientation and geometry of the coils are known we compute the torque constant K . To this end, consider the force on an R infinitesimal segment of coil wire dl. We assume that the rotor field is in the z direction and obtain dF : idl ; B : i(drr ; r d ) ; B z E : iB (r dr 9 dr ), (5.258) E where B is the rotor field in the gap region. The torque is given by E dT : r ; dF, and if we substitute Eq. (5.258) and take the z-component we find that dT : 9iB r dr. X E This is the torque imparted to the segment of the coil by the rotor magnet. Notice that only current in the radial direction contributes to the axial torque. If we consider a typical coil geometry, there are two sides that have an effective length l in the radial direction (Fig. 5.39b). The C torque on the rotor due to an active coil can be estimated as T
: 2iN B R/ l , E C
(5.259)
427
5.13 AXIAL-FIELD MOTORS
where factor 2 takes into account the two sides of the coil and R/ is the mean radius of the magnet, R ;R . R/ : 2 The effective length l can be determined from the coil geometry. The C total torque is simply Eq. (5.259) times the number of active coils, T : 2iN N B R/ l . ?A E C Therefore, the torque constant is K : 2N N B R/ l . (5.260) R ?A E C Furthermore, in the MKS system K : K . At this point, all the terms in C R Eq. (5.260) are known except for B , the field in the gap region. This can E be determined using the results of Examples 5.13.1 or 5.13.2 [21—23]. Once we know R , V, K , and K we use Eqs. (5.253) and (5.254) to C R C determine the no-load speed and stall torque T . We then plot the ,* Q speed-torque curve vs T (Fig. 5.38). As the operating speed is given, we can determine both the operating point of the motor and its efficiency. EXAMPLE 5.13.1 Develop a two-dimensional model for the field in the gap region of an axial-field motor (Fig. 5.37). Assume the magnet has a linear second quadrant demagnetization curve of the form B : H ; M , Q
(5.261)
where B : P . (5.262) H A SOLUTION 5.13.1 We reduce the three-dimensional motor geometry to an approximate two-dimensional geometry [21]. This is done by introducing a cylindrical cutting plane at the mean radius of the magnets. This cylinder is then imagined to be unrolled into a two-dimensional surface of infinite extent into and out of the page, as illustrated in Fig. 5.40a. This two-dimensional geometry can be reduced further by exploiting the symmetry that results from the repeating magnetic structure. Specifically, the field is symmetric about the vertical center lines of each pole. The field solution for the reduced geometry is similar to that of the periodic geometry of Example 3.6.2. Therefore, we follow closely the solution procedure presented there. As depicted in Fig. 5.40b, there are three regions to consider. In this figure, l is the arc length from the center
428
FIGURE 5.40
CHAPTER 5 Electromechanical Devices
Motor geometry: (a) approximate 2D geometry; and (b) reduced
geometry.
of one pole to the next at the mean radius, and x is the distance along the ‘‘straightened’’ arc. Therefore, if R and R are the inner and outer radii of the magnet, respectively, and N : the number of poles, then (R ; R ) . l: N The variable x is related to the angular measure along the arc as follows: x 2 . lN The magnetization for the three regions is :
M y Q M : 9M y Q Q 0
(5.263)
(Region 1) (Region 2) (Region 3).
(5.264)
429
5.13 AXIAL-FIELD MOTORS
Because M is constant in each region, the equation for the magnetic scalar Q potential reduces to : 0 K
(5.265)
for all three regions. The three regions under consideration are collectively bounded by four lines that form a rectangular area. The vertical sides of the rectangle, which are defined by the y-axis and the line x : l, represent lines of symmetry for the field, along which the normal component H is zero. Moreover, L the boundaries that form the top and bottom of the rectangle are bordered by soft-magnetic materials (flux return plates). For the sake of simplicity, we assume that the flux plates have infinite permeability ( : -). Consequently, the tangential component of the field vanishes along these boundaries. When we impose these boundary conditions we find that the general solutions for the magnet and gap regions are
n(y 9 h) nx (x, y) : C sinh cos
L l l L
(5.266)
and
ny nx (x, y) : C sinh cos , L l l L
(5.267)
respectively, where C and C are constants that need to be determined.
L L At the interface between the magnet and the gap, the tangential component of H must be continuous: H
R
(x, g) : H (x, g) R
(0 x l).
(5.268)
This condition, along with the fact that the functions sin(nx/l) are orthogonal on (0, l), implies that sinh(ng/l) C :C .
L L sinh(n(g 9 h)/l)
(5.269)
In addition, the normal component of B must also be continuous B
L
(x, g) : B (x, g) L
(0 x l).
(5.270)
430
CHAPTER 5 Electromechanical Devices
As B
L
:
H ; M
L Q H 9 M
L Q
(0 x l/2) (l/2 x l),
(5.271)
then Eq. (5.270) reduces to
M Q
(0 x l/2)
nx n C K(n, h, g, l, ) cos : L l l L 9 M Q where
(l/2 x l)
ng ng n K(n, h, g, l, ) : 9 cosh ; sinh coth (g 9 h) . (5.272) l l l Again, by exploiting the orthogonality of the functions cos(nx/l) on (0, l), it follows that C : L
(91)L\ 4l M Q (n)K(n, h, g, l, )
(n : 1, 3, 5, . . .).
(5.273)
Substituting Eq. (5.273) into Eq. (5.267) gives the complete solution for the gap region (x, y) :
4l M Q
(91)L\ ny nx sinh cos . l l (n)K(n, h, g, l, ) L
Therefore, the vertical or axial component of the field in this region is B (x, y) : W
4M Q
(91)L\ (n)K(n, h, g, l, ) L
;cosh
ny nx cos , l l
(5.274)
where x equals the horizontal distance along a circumferential arc at the mean radius of the magnet, y equals the vertical distance above the lower flux plate, g is the gap height, t is the thickness of the magnet, h : g ; t , l : K K (R ; R )/N , and R and R are the inner and outer radii of the magnet.
431
5.13 AXIAL-FIELD MOTORS
Calculations: We apply Eq. (5.274) to a motor with the following parameters: M : 4.0 ; 10 A/m Q g : 6.0 mm t : 4.0 mm K 40 R : mm 120 R : mm l : 20 mm N : 8. (5.275) Field values are computed with : 1.0 , 1.5 , and 2.0 , and with the vertical position in the gap set to y : 2 mm. Ten values are computed for each value of . Specifically, field values are computed at : 0, 5, 10, . . . , 45° (i.e., from the center of one pole to that of its neighbor). These data are shown in Fig. 5.41. Notice that the field decreases with increasing . )
FIGURE 5.41
Axial field B vs . X
432
CHAPTER 5 Electromechanical Devices
FIGURE 5.42
Motor geometry for 3D analysis.
EXAMPLE 5.13.2 Develop a three-dimensional field solution for the gap region of an axial-field motor (Fig. 5.42). Assume that the magnet has a second quadrant demagnetization curve B : (H < M z ), Q
(5.276)
where the < term takes into account the alternating polarity of adjacent poles [22, 23]. SOLUTION 5.13.2 For the three-dimensional model we approximate the magnetic structure of the motor in terms of the magnetic circuit shown in Fig. 5.43. Specifically, we assume that the upper and lower flux plates have infinite permeability ( : -), and are of infinite extent in both the vertical and horizontal directions. The field in the gap region of this idealized geometry can be determined using the method of images (Section 3.7). That is, the field in the gap can be expressed as a superposition of the fields from a double infinite sum of image magnets. Let B (r, , z) denote the field due to the motor magnet
X in free space. We studied this geometry in Example 4.2.8. The axial field B (r, , z) is given by Eq. (4.163). Notice that B (r, , z) can be
X
X considered to be a function of the spatial coordinates (r, , z), and the axial position of the magnet, B (r, , z) : B (r, , z; z , z ),
X
X
(5.277)
5.13 AXIAL-FIELD MOTORS
FIGURE 5.43
433
Cross section of a motor model with infinite flux plates.
where z : z (1) and z : z (2) are the positions of the bottom and top of the Q Q magnet on the z-axis, respectively. Image magnets: To obtain a field solution we need to take into account the fields from the image magnets. These are located in the regions occupied by the flux plates. For example, the images for an infinitesimal magnetic charge Q on the bottom surface of the motor magnet are shown in Fig. 5.44. Each image magnet has a field component B (r, , z; z , z ), and the positions z and z
X are chosen so that the tangential component of H is zero and the normal component of B is continuous at the surface of the flux plates. Each image magnet in one flux plate gives rise to a mirror image magnet in the other flux plate. Therefore, there is a double infinite set of image magnets. We write the
434
CHAPTER 5 Electromechanical Devices
FIGURE 5.44
Image charges for Q.
total field as a superposition of all the magnets, real and image, B (r, , z) : X
[B (r, , z; Z (n), Z (n))
X L
;B
X
(r, , z; Z (n), Z (n))].
(5.278)
5.13 AXIAL-FIELD MOTORS
435
The axial positions are given by the following recursive relations: Z (n ; 2) : Z (n) ; 2(h ; t ) K Z (n ; 2) : Z (n ; 2) ; 2t K and Z (n) : 9Z (n) Z (n) : 9Z (n), where Z (1) : h, Z (1) : h ; 2t (n : 1, 3, 5, . . .). According to these rela K tions, the first few terms are as follows: Z (3) : h ; 2(h ; t ) K Z (3) : (3h ; 2t ) ; 2t K K Z (5) : (3h ; 2t ) ; 2(h ; t ) K K Z (5) : (5h ; 4t ) ; 2t K K and Z (1) : 9(h ; 2t ) K Z (1) : 9h Z (3) : 9(3h ; 4t ) K Z (3) : 9(3h ; 2t ) K Z (5) : 9(5h ; 6t ) K Z (5) : 9(5h ; 4t ). K The first and second terms in Eq. (5.278) correspond to the images in the upper and lower flux plates, respectively. If the lower flux plate is absent, the field reduces to B (r, , z) : B (r, , z; h, h ; 2t ). (5.279) X X K Calculations: We apply Eq. (5.278) to a motor geometry with the following parameters: M : 7.2 ; 10 (A/m) Q N :4 R : 0.5 cm R : 2.5 cm.
436
CHAPTER 5 Electromechanical Devices
FIGURE 5.45
B vs h. X
First, we compute B at the midpoint of the gap (r : (R ; R )/2, : 0°, and X z : h/2) for a range of gap heights h : 4, 5, . . . , 10 mm, and magnet heights t : 4, 6, 8 and 10 mm. These data are shown in Fig. 5.45. Notice that the field K drops off sharply as the gap h increases. This analysis is performed with N : 40 and N : 20, and there is no improvement in accuracy beyond the first P five terms in Eq. (5.278). From these initial calculations, one can select a gap height h and magnet thickness t that render the desired field strength. For K example, h : 6 mm and t : 10 mm, which yields a field of 0.428 T at the K midpoint of the gap. Next, we fix h : 6 mm and t : 10 mm, and then perform K an analysis to determine the variation of B in the radial direction. Specifically, X B is computed along two different radial lines located midheight in the gap X (z : h/2) and spanning the surface of the magnets from R r R (Fig. 5.46). These lines are positioned at : 0° (center line of a pole) and : 30°, respectively. Notice that at : 0, B & 0.35 T for 10 mm r 23 mm, and X that it drops off rapidly outside of this interval. There is also a considerable drop in field strength as one moves in an angular sense from the center of the pole : 0. To see this, the field is computed along three different circumferential arcs spanning an angular measure of 90°, that is, from the center of one pole to the center of the neighboring pole at radii r : 10, 15 and 20 mm, respectively,
437
5.14 STEPPER MOTORS
FIGURE 5.46
B vs r. X
with (z : h/2) (Fig. 5.47). This analysis shows that the field drops off sharply for 30°. In summary, we find that the region of maximum field strength for this motor geometry is 10 mm r 23 mm with 930° 30°. )
5.14 STEPPER MOTORS Stepper motors execute rotational motion in a sequence of discrete angular steps. Well suited for open-loop digital control, they are widely used for precision motion and positioning. These motors can be found in numerous industrial and consumer products including computers and peripherals, disk drives, printers, robotic devices, and in electronic cameras where they are used for lens focusing. Stepper motors come in a variety of configurations and designs. In this section we study a permanent magnet stepper motor known as a ‘‘can-stack’’ motor. This motor consists of two separate bobbin-wound drive coils that enclose axial sections of a cylindrical radially polarized multipole permanent magnet rotor (Fig. 5.48). The rotor is usually made from an isotropic grade of ceramic ferrite, Sm Co or Nd Fe B.
438
CHAPTER 5 Electromechanical Devices
FIGURE 5.47
FIGURE 5.48
B vs . X
Exploded view of a permanent magnet stepper motor.
5.14 STEPPER MOTORS
439
This is a two-phase motor with each coil comprising a separate drive phase. Each coil is mounted on a soft-magnetic stator element that consists of two opposing toothed components and a flux return ring. The number of teeth on a stator equals the number of poles on the rotor. The drive coils are assembled by first sliding the coil and flux return ring on one of the toothed components and then inserting the remaining toothed element into the assembly in such a way that the stator teeth are equally spaced. A cross section of the interior of an assembled four-pole stator is shown in Fig. 5.49a. An ‘‘unrolled’’ perspective of an assembled stator element (without the coil) is shown in Fig. 5.50.
Cross section of stepper motor stator: (a) unenergized stator (rotor not shown); (b) energized stator (rotor not shown); (c) rotor at position of maximum drive torque; and (d) rotor rotated to position of zero drive torque.
FIGURE 5.49
440
CHAPTER 5 Electromechanical Devices
FIGURE 5.50
Magnetic circuit of stator element.
In the can-stack motor each rotation step is initiated by activating a single phase. Specifically, when one phase is activated (with the other phase off) the rotor experiences a torque that rotates it into magnetic alignment with that phase. The stator teeth of the two phases are offset from one another by half-a-tooth (or pole) pitch in an angular sense. Therefore, by sequentially activating the two phases the rotor can be repetitively stepped an angular measure equal to half a pole pitch. The torque developed by the motor can be understood by considering the activation of a single phase as depicted in Fig. 5.49. The activation sequence is as follows: (a) the coil is unactivated (rotor not shown); (b) the coil is activated and a magnetic field is created in the stator cavity (rotor not shown); (c) the rotor is initially in a position where it experiences maximum drive torque that is due to the interaction of the radially directed stator field and equivalent current sheets of the rotor (surface currents exist at the transitions between neighboring poles); and (d) the rotor has rotated into magnetic alignment with the phase where it experiences zero torque. This occurs when the equivalent current sheets of the rotor are centered with respect to the gap as shown. When
441
5.14 STEPPER MOTORS
the rotor is in this position there is no drive torque because the stator field across the equivalent current sheets is parallel to the direction of rotation as shown in Fig. 5.49b. The alignment of the rotor to a given phase is enhanced by the lower reluctance of the rotor/stator circuit in the aligned position (Fig. 5.49d). When the rotor aligns with one phase it is in a position maximum torque for the other phase because the stator teeth for that phase are offset by half-a-tooth (or pole) pitch. Thus, the rotor can be continuously rotated (stepped) by activating the two phases in sequential fashion. In the following example, we derive an expression for the drive torque of a ‘‘can-stack’’ stepper motor. EXAMPLE 5.14.1 Derive an expression for the drive torque of the stepper motor shown in Fig. 5.48. Assume that the rotor is a radially polarized multipole cylinder with a magnetization M : <M r , (5.280) Q where < takes into account the alternating polarity of adjacent poles. Further assume that each stator is operating below saturation in a linear region of its B-H curve with . SOLUTION 5.14.1 The torque is derived in two steps. First, we obtain a field solution in the interior of an energized stator with no rotor present. Second, we reduce the rotor to an equivalent current distribution and then compute the torque on it using a Lorentz force approach with the stator field as an external field. Stator field: Let L and R denote the length and inner radius of the stator, Q respectively. Each stator has N teeth, which equals the number of poles on the rotor. The stator is enclosed by an n-turn coil and a flux return ring. A cross section of the interior of a four-pole stator is shown in Fig. 5.49a. Let R denote the angular span of each tooth and denote the angular expanse of the E gap between each tooth. These angles are related by the following equation, 2 : 9 . (5.281) E N R We seek a field solution for the interior of the stator. To obtain an analytical solution, we assume that the stator is infinitely long. This assumption is not as radical as it might first seem because the stator tends to confine the flux and minimize leakage etc. In the interior of the stator, ; H : 0.
442
CHAPTER 5 Electromechanical Devices
Thus, we can represent H in terms of a scalar potential, H : 9 (Section K 3.4). The field problem reduces to a boundary value problem (BVP) in cylindrical coordinates (Section 3.6.2). Specifically, we solve : 0. (5.282) K The general solution to Eq. (5.282) that is well behaved at r : 0 is of the form (r, ) : a rN sin(p) ; b rN cos(p). (5.283) K N N N We choose a coordinate system in which : 0 coincides with the center of a stator tooth (Fig. 5.51a). In this coordinate system (r, ) must be an even K function of . Thus, we discard the sin(n) terms in Eq. (5.283) and obtain (r, ) : b rN cos(p). (5.284) K N N We determine the coefficients b by matching the boundary conditions at the N
Stepper motor stator: (a) reference frame; and (b) profile of scalar potential along the stator surface.
FIGURE 5.51
443
5.14 STEPPER MOTORS
interior surface of the stator. We assume that the potential (R , ) along this K Q surface is constant across each tooth and alternates from to 9 on
successive teeth. As the gap between neighboring teeth is small ( ), we E R assume that the potential is linear between the teeth and this implies a profile for (R , ) as shown in Fig. 5.51b. To solve the BVP we need a Fourier series K Q representation for (R , ). First, we derive an expression for in terms K Q
of the motor’s physical parameters. Let g denote the gap between adjacent stator teeth, and let H denote the field in the gap region (Figs. 5.49b and 5.50). We E assume that the potential is linear between adjacent teeth. Therefore, 1 K H :9 E R Q 2 :9 , g
(5.285) (5.286)
where : and g : R . We apply Eq. (3.138) from Section 3.5 to the E Q E dotted path in Fig. 5.50 and obtain ni H : . E g
(5.287)
Therefore,
:9
ni . 2
(5.288)
Now that we know , we can represent the potential along the stator
(R , ) in terms of the Fourier series K Q (R , ) : A cos( ), (5.289) K Q I I I
where
8
[cos(k) 9 1] cos R A :9 I I kN 2 E 4 ni : [cos(k) 9 1] cos R , (5.290) I 2 kN E and : kN /2. We evaluate Eq. (5.284) at r : R , compare this with Eq. I Q (5.289), and obtain b : A R\N N I Q
p:
kN . 2
444
CHAPTER 5 Electromechanical Devices
Thus, the potential in the interior of the stator is
r ?I A cos( ). (5.291) I R I I
Q The field components are obtained from B : 9 . For example, K I A r? \ cos( ). B (r, ) : 9 (5.292) P I R?I I I I
Q Drive torque: To compute the torque we represent the rotor as a distribution of equivalent currents. Recall that the rotor is radially polarized with (r, ) : K
M : <M r . Q Let R and R denote its inner and outer radius, respectively. Assume that it is initially positioned with one of its north poles centered about 0° (i.e., each pole is initially centered with respect to a stator tooth). The equivalent current densities are given by Eq. (3.95). We find that the volume current density is zero, J : ; M : 0, and that the surface current density is K j : M ; n K : <M z . Q These surface currents are located at the edges of each sector at angular positions
1 2 : n; (n : 0, 1, 2, . . . , N 9 1). L 2 N It is important to note that there are two such currents at these positions, one along the surface of each of the neighboring sectors. If the rotor is rotated by an angle , the positions of the surface currents change to
1 2 (n : 0, 1, 2, . . . , N 9 1). () : ; n ; L 2 N The surface current densities at these locations are
M z (n : 0, 2, 4, . . .) Q j (n) : K 9M z (n : 1, 3, 5, . . .). Q The torque is given by Eq. (3.100) from Section 3.3, that is, T() : 2L
, \ 0 r ; [j (n) ; B (r, ())] dr. K L L 0
(5.293)
(5.294)
445
5.14 STEPPER MOTORS
Factor 2 takes into account the contribution of each surface of the neighboring sectors. From Eq. (5.294) we find that the component of torque about the z-axis is
, \ 0 (91)L B (r, ())r dr. P L L 0 Finally, substituting Eq. (5.292) into Eq. (5.295) gives T () : 2LM X Q
, \ (91)L L I
(R?I> 9 R?I>) I ;A cos[ ()], I ;1 I L R?I Q I where L and R are the length and interior radius of the stator and Q 4 ni A : [cos(k) 9 1] cos R . I kN I 2 E Here,
(5.295)
T () : 2 LM X Q
(5.296)
: I
kN , 2
and
1 2 () : ; n ; . L 2 N Calculations: We demonstrate Eq. (5.296) with some sample calculations. The following parameters are used in the analysis: M : 4.3 ; 10 A/m Q N : 10 n : 100 turns i : 0.25 A R : 6.0 mm R : 8.0 mm R : 8.1 mm Q L : 2.0 mm. The angular span of each stator tooth is taken to be : 27° (75% of the span R
446
CHAPTER 5 Electromechanical Devices
of a rotor pole). The angular span of the gap between neighboring teeth is : 9° ( : 2/N 9 ). There are two phases to consider, which are E E R labeled 1 and 2, respectively. Recall that the stator teeth of the two phases are offset from one another by an angular span equal to one half of the pole pitch, which in this case is 18°. The drive torque T () for each phase is shown in Fig. X 5.52. This plot shows the torque on the rotor as it rotates the angular span of one pole (036°). When :0, the rotor is positioned as in Fig. 5.49c and d for phase 1 and phase 2, respectively. In this initial position, phase 1 exerts almost maximum torque, and phase two exerts zero torque. The peak torque for phase 1 is 1500 N · m, which occurs at : 12°. When the rotor is rotated to : 18° (half a pole pitch) as shown in Fig. 5.49d, the torque of phase 1 is zero and the torque of phase 2 is near a maximum. For 18 36°, the torque of phase 1 reverses direction as shown. )
5.15 HYBRID ANALYTICAL-FEM ANALYSIS In the previous sections we derived analytical lumped-parameter models for the analysis of various electromechanical devices. Such models can be developed for some devices, but certainly not all. Geometric asymmetries and/or material nonlinearities often preclude a full analytical treatment. However, in such cases it is still possible to perform a lumped-parameter analysis using a hybrid analytical-FEM approach. In this section we describe the approach and demonstrate it via application to a practical device. In the hybrid analytical-FEM method we set up the equations of motion and identify which terms, if any, can be obtained in analytical form. We derive expressions for these terms using the methods presented in the previous sections. Next, we consider the remaining terms. We obtain expressions for these using parametric FEA. The method is as follows. First, specify a discrete set of values for the independent variables i and x or that span the working range of the device. Next, perform a series of FEA calculations to determine values for the unknown terms for each value of the independent variables. The computed data is then fit to a smooth curve, which renders the unknown terms in analytical form. For example, suppose we are analyzing a rotary actuator and need an expression for the torque T() as a function of rotation angle . We first specify a discrete set of angular values ( , , . . . , , . . . , ) G , that span the range of motion of the rotor. Next, we perform FEA calculations and determine T() at these values. This results in a series
5.15 HYBRID ANALYTICAL-FEM ANALYSIS
FIGURE 5.52
Drive torque vs : (a) phase 1; and (b) phase 2.
447
448
CHAPTER 5 Electromechanical Devices
Rotary permanent magnet actuator: (a) actuator perspective; and (b) cross-sectional view with reference frame.
FIGURE 5.53
of data ( , T( )) (i : 1, 2, . . . , N). This data is then fit to a finite polyG G nomial of the form, ) T() : a I. I I This renders T() in a form suitable for a lumped-parameter analysis. We demonstrate the hybrid analytical-FEM approach via application to the rotary actuator shown in Fig. 5.53. This type of actuator is used to implement rotational motion that is typically 45°. The actuator consists of a drive coil that is wrapped around a soft-magnetic core and a
449
5.15 HYBRID ANALYTICAL-FEM ANALYSIS
cylindrical bipolar magnet mounted for rotation in the gap of the core. When the coil is energized, the field generated by the coil imparts a torque to the magnet that causes it to rotate clockwise. As it rotates, the magnet experiences an additional counterclockwise torque due to the core itself. This is called reluctance torque. When the coil is turned off, the reluctance torque rotates the magnet back to its initial position where it is held in place against a mechanical stop mechanism. We develop a model for analyzing this actuator in the following example [24]. EXAMPLE 5.15.1 Derive the equations of motion for the actuator shown in Fig. 5.53. SOLUTION 5.15.1 This is a moving magnet actuator and its behavior is governed by rotational motion equations (5.87):
di(t) 1 "(i, ) : V (t) 9 i(t)(R ; R ) 9 (t) , dt L Q d (t) 1 : T(i, ), dt J K d(t) : (t). dt
(5.297)
To solve Eq. (5.297) we need expressions for "(i, ) , L:
"(i, ) , i
(5.298) (5.299)
and T(i, ).
(5.300)
Assumptions: In order to simplify the analysis we make the following assumptions. First, we assume that the core is operating in a linear region of its magnetization curve with a permeability . B : H (core). (5.301) Thus, the magnetic circuit is linear, and the effects of coil and magnet can be separated, B:B ;B ,
"(i, ) : " (i) ; " (),
(5.302) (5.303)
450
CHAPTER 5 Electromechanical Devices
and T(i, ) : T
(i, ) ; T (),
(5.304)
where B : field due to the coil B : field due to the magnet
T (i, ) : torque on the magnet due to the coil T () : torque on the magnet due to the core " () : flux linkage due to the magnet
" (i) : flux linkage due to the current. (5.305) Notice that " () and T () are functions of only. If the magnetic circuit
is nonlinear, then a nonlinear analysis needs to be performed for each time stepped integration of Eq. (5.297). The second assumption is that the magnet has a linear second-quadrant demagnetization curve of the form B : (H ; M ), (5.306) Q where M is uniform and fixed. Q Flux linkage: The flux linkage is given by Eq. (5.303). Therefore, Eqs. (5.298) and (5.299) reduce to "(i, ) d" () : , d
(5.307)
and L:
d" (i) , di
(5.308)
respectively. Flux linkage due to the magnet: The flux linkage due to the magnet is determined using FEA, FEA $ "
(). (5.309)
Specifically, FEA calculations are performed for a discrete set of angles I (k : 1, 2, . . .) that span the anticipated rotation of the magnet. The flux linkage " ( ) is computed (with the coil turned off) for each value of ,
I I " ( ) :
I
B 1
( ) · ds.
I
integration over coil
5.15 HYBRID ANALYTICAL-FEM ANALYSIS
451
Typical FEA flux analysis: (a) rotor at : 0°; (b) rotor at : 45°; (c) rotor at : 90°; and (d) normalized flux vs .
FIGURE 5.54
Typical FEA calculations are shown in Fig. 5.54. Specifically, flux plots are shown for (a) : 0°, (b) : 45°, and (c) : 90°. Notice that the maximum flux linkage occurs when : 0°. A normalized plot of () (the average flux through the coil) is shown in (d). From these calculations we obtain the data ( , " ( )) and then fit it to a polynomial, I I , " () : a I.
I I
(5.310)
452
CHAPTER 5 Electromechanical Devices
From this we obtain () ,
: a kI\ (FEA), I d I which is the desired analytical expression for Eq. (5.307). d"
(5.311)
Flux linkage due to the coil: The flux linkage due to the coil can be determined using FEA or magnetic circuit analysis. As the circuit is linear, the flux linkage is proportional to the current and the constant of proportionality C can be computed using FEA, "
(i) : niC (FEA). (5.312) On the other hand, if the gap g in the circuit is small relative to the dimensions of the cross-sectional area, then the flux (i) through each turn of the coil can be determined using magnetic circuit analysis as in Example 3.5.1. This gives niA , (i) : g
(5.313)
where g and A are the length and area of the gap. As the coil has n turns, the flux linkage is "
(i) : n (i) niA . : g
In either case, from Eq. (5.308) we have L:
nC nA g
(FEA), (5.314) (circuit analysis).
Torque: As noted already, the torque on the magnet can be separated into two components T(i, ) : T
(i, ) ; T (). (5.315) The term T (i, ) is the torque on the magnet due to the current through the coil. This is referred to as the drive torque. The other term, T () is the torque on the magnet due to the presence of the magnetic circuit (core). It is independent of the current, and is referred to as the reluctance torque.
5.15 HYBRID ANALYTICAL-FEM ANALYSIS
453
Drive torque: The drive torque can be computed using the formula for the torque on a bipolar cylinder in an external field (Example 3.4.4), : M hR B sin()z , Q are the length and radius of the magnet. In addition, T
where h and R
ni B : g
(5.316)
is the field in the gap due to the coil (Example 3.5.1). Therefore, ni sin(). T (i, ) : M hR Q g
(5.317)
This result is valid when magnetic circuit analysis applies. When this is not the case, an analytical expression for T (i, ) can be determined using FEA. Specifically, choose a reference current i and perform a single FEA to compute at a series of points (R , ) that span the field components B and B
H P the circumference of the magnet, (R , ) : B (R , )r ; B (R , ) . (5.318) H P H H This data is then substituted into Eq. (3.134), which in turn can be evaluated numerically for a sequence of angles that span the angular rotation of the I magnet. The resulting data ( , T ( )) is then fit to a finite polynomial I I i , T (i, ) : d I, (5.319) I i I where the multiplier i/i reflects the linearity of the circuit. Reluctance torque: The reluctance torque is computed using FEA B
FEA $ T
(). (5.320) This torque is computed for a set of angular values that span the angular I rotation of the magnet. The computed data ( , T ( )) is then fit to a finite I I polynomial, , () : c I. (5.321) I I A sample plot of T () (obtained from FEA) is shown in Fig. 5.55; it is negative because it acts to rotate the magnet counterclockwise (clockwise rotation is taken to be positive). Notice that T () is maximum at : 45°. T
454
CHAPTER 5 Electromechanical Devices
FIGURE 5.55
Plot of reluctance torque T
() vs .
Calculations: We demonstrate the hybrid approach with some sample calculations. The following parameters are used in the analysis: R : 3.0 n : 500 turns M : 4.3 ; 10 A/m Q R : 4.0 mm
g : 10.0 mm h : 25.0 mm J : 6 ; 10\ kg · m. K The width of the core is 10 mm. Therefore, A : 250 mm. First, determine Eqs. (5.311), (5.314), (5.317), and (5.321) and then substitute these into the
455
5.16 MAGNETIC MEMS
FIGURE 5.56
Normalized coil current vs t.
equations of motion (5.297). We apply a step function voltage pulse of the form V (t) : Q
V 0
(0 t T ) (T t -),
where V :12 V and T :2.0 ms. The initial conditions are i :0 A, :0.175 radians (10°), and : 0 rad/s. We solve Eq. (5.297) numerically using Euler’s method. A normalized plot of i(t) is shown in Fig. 5.56. Notice that the current begins decaying after T when the applied voltage becomes zero. Normalized torque profiles are shown in Fig. 5.57. Finally, a plot of (t) is shown in Fig. 5.58. As expected, (t) reaches a maximum and then returns to its initial value due to the reluctance torque. )
5.16 MAGNETIC MEMS In the previous sections we have studied conventional electromechanical devices. These are macroscopic in size, typically greater than one cubic
456
CHAPTER 5 Electromechanical Devices
FIGURE 5.57
Normalized torque profiles vs t.
FIGURE 5.58
Rotation angle (t) vs t.
457
5.16 MAGNETIC MEMS
centimeter in volume. In this section we study micro-electromechanical devices. These are typically less than one cubic millimeter in volume, and can have features that are on the order of microns. Such devices fall under the general category of micro-electromechanical systems (MEMS) [25]. These MEMS devices are fabricated using techniques derived from the microelectronics industry. A typical fabrication cycle includes steps such as mask design and generation, photolithography, and bulk and surface micromachining [26]. These methods enable a high degree of integration, and a low manufacturing cost with tens to hundreds of devices realized on a single silicon wafer. Advances in microsystems technology have spawned the rapid development of a variety of devices for both research and commercial use. These include accelerometers, light modulators, microfluidic devices, micromotors, molecular filters, and various actuators and sensors [25, 27]. To date, the majority of MEMS transducers have been electrostatically driven. There are at least two reasons for this. First, electrostatic actuators lend themselves to microelectronic fabrication methods. Second, the electrostatic force scales relatively well at the micro domain. We demonstrate this using a model of a parallel plate capacitor. The force of attraction between the plates is
+ A F : E, C 2 where + is the permittivity of free space, A is the area of each plate, and E is the electric field between the plates. If we keep the E-field constant, F scales as L where L is the characteristic dimension of the capacitor. C This follows from the L dependency of A. Thus, if the dimensions of a capacitor are reduced by a factor of ten, the electrostatic force decreases by a factor of one hundred. The magnetostatic force for current-driven actuators does not scale as well as the electrostatic force. To see this, consider the force between a pair of parallel wires, s F : I I . A 2 d Here, I and I are the currents in each wire, s is the length of the wires, and d is the separation distance. The current is I : JA where J is the current density and A is the area of the wire. The force can be rewritten as s F : J J A A . A 2 d
(5.322)
458
CHAPTER 5 Electromechanical Devices
From Eq. (5.322) we see that if the current density is kept constant, the force scales as L, where L is the characteristic dimension of the circuit. Thus, if the dimensions of a circuit are reduced by a factor of 10, the force decreases by 10,000, which is two orders of magnitude weaker than the electrostatic case. If we allow the current density to increase, but keep the heat flow through the surface of the wires constant, the force scales as L. Or, if we allow the current density to increase, but keep the maximum temperature rise constant, the force varies as L [28]. This last case scales as well as the electrostatic case, but it will likely give rise to deleterious thermal effects. We can perform the same analysis for the force between a magnet and a conductor. Consider a conductor carrying a current I with a length L perpendicular to a B-field. The force on the conductor is : ILB.
The B-field of the magnet is independent of size. That is, if we change all the linear dimensions of a permanent magnet, the field strength at all the rescaled observation points remains constant. However, the current I scales as L for the constant current density case, L for the constant heat flow case, and L for the constant temperature rise case [28]. Therefore, F scales as L for the constant current density
case, L for the constant heat flow case, and L for the constant temperature rise case. Again, the last case scales as well as the electrostatic force, but there could be substantial issues regarding the management of thermal energy. It is instructive to consider how the field of a current source scales at the micro domain. Consider an infinite straight wire. The B-field at a distance r from the wire is F
I B(r) : . 2r As I : JA we have JA B(r) : , 2 r which scales as L (B L). Thus, if we reduce the dimensions of a circuit, the current density J must increase by 1/L if field strength at the rescaled observation points is to remain constant. By comparison, if we change all the linear dimensions of a permanent magnet, the field strength at all the rescaled observation points remains constant (assuming that the magnetization is constant). Moreover, there is no power consumption.
459
5.16 MAGNETIC MEMS
Therefore, permanent magnets have advantages over conventional current sources in terms of the fields they provide at the micro domain. In spite of the relatively poor scaling of the magnetic force, there is a substantial and growing interest in magnetic microactuators. Over the last few years several research groups have demonstrated viable devices [29—31]. There has also been recent progress towards the commercialization of such devices [32]. However, most of these devices are not fabricated in a purely batch mode. Instead, they typically require a modest amount of manual assembly. It is likely that magnetic microactuators will gain in importance as methods for depositing rare-earth materials such as NdFeB onto microstructures improve. In the remainder of this section we discuss the operation of a batch-fabricated magnetic microactuator. A batch-fabricated magnetic microactuator is shown in Fig. 5.59. It consists of a soft magnetic (NiFe) plate supported on one side by a narrow torsion beam that is anchored at either end to a silicon substrate. The fabrication process for this structure is shown in Fig. 5.60a—e [30]. The operation of the actuator is as follows: When a uniform field H is applied, it induces a magnetization in the plate and a magnetic torque develops that acts to rotate the plate up, away from the substrate. This rotation is resisted by a restoring torque that results from the torsion of the support beam. The plate rotates to an equilibrium deflection angle where the magnetic and restoring torques are in balance (Fig. 5.61b). To determine we need to compare the two torques. We consider the restoring torque first. This is due to the torsion of the support beam. Since the midportion of the beam is attached to the plate, it does not twist. Therefore, the restoring torque is due to the torsion of the two end portions of the beam (between the plate and the anchor points). Each of these can be considered as a separate beam with length l and cross-sectional dimensions w and t. We assume that both
FIGURE 5.59
Magnetic MEMS actuator.
460
CHAPTER 5 Electromechanical Devices
FIGURE 5.60
Fabrication process for actuator: (a)—(e) [30].
beams obey Hooke’s law T () : k, K where T is the mechanical restoring torque and k is the torsional K stiffness. This is given by k : 2
KG ; J , l
(5.323)
461
5.16 MAGNETIC MEMS
FIGURE 5.61
A MEMS actuator: (a) top view; and (b) side view [30].
where K is a cross-section dependent factor, G is the shear modulus, is the residual stress in the beam, and J is the polar moment of inertia. The term J in Eq. (5.323) is usually negligible compared to KG. The formulas for K, G, and J are as follows: K:
1 nb tanh n 2a L
E G: , 2(1 ; v)
ab 192a 19 3 b
,
(5.324) (5.325)
and J:
1 (wt ; wt), 12
(5.326)
where E is Young’s modulus, v is Poisson’s ratio, a : min(w, t), and b : max(w, t). To get a sense of the magnitude of these terms we analyze a polysilicon beam with E:170 GPa, l : 400 m, w : 2.2 m, and t : 2.2 m. We evaluate Eqs. (5.324)—(5.326) and find that K : 3.3 ; 10\ m, J : 3.9 ; 10\ m, and k : 2 nN-m/rad [30].
462
CHAPTER 5 Electromechanical Devices
We now consider the magnetic torque. If there is no external field (H : 0) the plate is unmagnetized, and it remains at rest in a down position. When an external field is applied (H 0), it induces a magnetization M in the plate, which couples to H and this gives rise to a torque T . The torque is given by Eq. (3.133), which in this case & reduces to T : V M ; H
& : V MH sin(), (5.327)
where V is the volume of the magnet and is the angle between M
and H . Before we can determine T we need an expression for the & magnetization M. The magnetization of the plate depends on H as specified by its M vs H plot (Section 1.11). However, H (the field inside the plate) depends on M. Recall from Section 1.8 that the field inside a magnetized specimen is H:H ;H ? B : H 9 N M, ? + where H is the applied field and H : 9N M is the demagnetization ? B + field (N is the demagnetization factor). Thus, M is functionally depend+ ent on itself, which makes it difficult to determine. Often, we obtain M by making simplifying assumptions. For example, if the material is below saturation, we sometimes assume that it is linear (Example 1.8.1), or that it has an infinite permeability - (Example 3.2.8). On the other hand, if the material is saturated we can assume that it has a fixed magnetization M . For this application we model the magnetization of Q the plate below saturation with an anisotropy model, and above saturation we use a uniform magnetization M . Q Below saturation, the magnetization model for the plate is based on uniaxial anisotropy (Section 1.9). Specifically, we assume that the plate has a single easy axis. When an external field is applied it induces a magnetization M and also rotates M an angle away from the easy axis. Consequently, the plate acquires an energy density w : K sin(). ? ? The constant K can have contributions from stress, crystalline and shape ? anisotropy (Section 1.9). In this case, the plate is a low-stress polycrystalline thin film, and therefore the stress and crystalline anisotropy are negligible compared to the shape anisotropy. The shape-induced easy
463
5.16 MAGNETIC MEMS
axis runs the length of the plate, parallel to its surface. The constant K is ? M K : Q (N 9 N ), ? R J 2 where N and N are the anisotropy constants along thickness and length R of the plate, respectively [33]. There is a shape-induced torque on M which is given by w ? T () : 9V
? : 9V K sin(2), (5.328)
? where V is the volume of the plate. The negative sign in Eq. (5.328)
indicates that T () acts to oppose an increase in . That is, T () acts to ? ? align M with the easy axis. There is an equal and opposite torque 9T () ? on the plate itself. This acts to rotate the plate towards M, away from the substrate (Fig. 5.62). We use Eq. (5.328) to determine the equilibrium rotation of the magnetization. However, to determine this we still need an expression for M. We obtain this by considering the components of H along the direction of M. Specifically, we have H0 : H0 ; H0 , ? B where H0 and H0 are the applied and demagnetization field components ? B parallel to M (Section 1.8). Let denote the fixed angle between H and the initial (unactivated) direction of the easy axis, and let denote the equilibrium rotation of the easy axis (i.e., of the plate) (Fig. 5.62). The
FIGURE 5.62
Torque diagram with reference frame [30].
464
CHAPTER 5 Electromechanical Devices
angle between M and H is : 9 9 . Therefore, H0 can be written as ? H0 : H cos( 9 9 ). ? The component H0 is given by B H0 : 9N ()M, B + where N () is the demagnetization factor that accounts for the rotation + of M. Specifically, if M is at an angle with respect to the easy axis then Therefore,
N () : N cos() ; N sin(). + J R
H0 : 9(N cos() ; N sin() M. B J R At this point, we have expressions for H0 and H0 , but still no ? B expression for M. We obtain this by making a further assumption: When the plate is subjected to H , the domain structure within it changes in such a way as to minimize the magnetostatic energy. This gives H 0 inside the plate, which implies that H0 : 9H0 , ? B
or
H cos( 9 9 ) M : . (5.329) N () + We use Eq. (5.329) for M when the plate is below saturation. Above saturation we assume that M : M . Therefore, the magnetization can be Q written as H cos( 9 9 ) M : min ,M . (5.330) Q (N cos() ; N sin() J R We are finally ready to determine the equilibrium rotation angle of the plate. Equilibrium occurs when the magnetic and restoring torques are in balance. We consider the torque on both M and the plate (Fig. 5.62). The torque on M is
where and
T (, , M) : T (, , M) 9 T (), + & ? T : V MH sin( 9 9 )
&
(5.331)
T () : V K sin(2).
? ?
(5.332)
465
5.16 MAGNETIC MEMS
FIGURE 5.63
Rotation angle vs applied field H [30].
We use Eq. (5.330) for M in Eq. (5.331). The torque on the plate is (, ) : T () 9 T (). ? K In equilibrium, the net torque on both M and the plate is zero. This occurs when T
T (, ) : T () & ?
(5.333)
and T () : T (). ? K Combining Eqs. (5.333) and (5.334) gives
(5.334)
T (, , M) : T () : T (). (5.335) & ? K The relation (5.335) represents a set of transcendental equations for , , and M. These equations can be solved numerically for the three variables. A typical response curve for the actuator is shown in Fig. 5.63 [30]. If a time-varying field is applied the plate can be made to oscillate up and down [34].
466
CHAPTER 5 Electromechanical Devices
References 1. Brown, D. and Hamilton III, E. P. (1984). Electromechanical Energy Conversion, New York: Macmillan Publishing Co. 2. Fitzgerald, A. E., Kingsley, C., and Umans, S. D. (1983). Electric Machinery, 4th ed., New York: McGraw-Hill. 3. Krause, P. C. and Wasynczuk, O. (1989). Electromechanical Motion Devices, New York: McGraw-Hill. 4. Woodson, H. H. and Melcher, J. R. (1968). Electromechanical Dynamics, Part I: Discrete Systems, New York: John Wiley & Sons. 5. Woodson, H. H. and Melcher, J. R. (1985). Electromechanical Dynamics, Part II: Fields, Forces and Motion, Malabar, FL: John E. Krieger Publishing Co. 6. Taylor, E. F. and Wheeler, J. A. (1992). Spacetime Physics: Introduction to Special Relativity, 2nd ed., New York: W. H. Freeman and Co. 7. Moon, F. C. (1984). Magneto-solid Mechanics, New York: Wiley. 8. Lowther, D. A. and Silvester, P. P. (1986). Computer-aided Design in Magnetics, New York: Springer-Verlag. 9. Furlani, E. P. and O’Brien, M. (1994). Analysis of axial-field actuators. IEEE Trans. Mag. 30(6): 4323. 10. Furlani, E. P. and O’Brien, M. (1996). Predicting the dynamic behavior of axial-field actuators. IEEE Trans. Mag. 32(2): 302. 11. Furlani, E. P. (1994). A three-dimensional field solution for axially-polarized multipole disks. J. Magn. Magn. Mat. 135(2): 205. 12. Furlani, E. P. (1996). An analytical method for predicting the transient behavior of resonant electromagnetic actuators. Int. J. Appl. Elec. Mech. 9: 101. 13. Marchant, A. B. (1990). Optical Recording: A Technical Overview, Reading, MA: Addison-Wesley Pub. Co. 14. Mansuripur, M. (1995). The Physical Principles of Magneto-optical Recording, Cambridge, England: Cambridge Univ. Press. 15. McDaniel, T. W. and Victora, R. H. (1997). Handbook of Magneto-Optical Recording; Materials, Subsystems, Techniques, Westwood, NJ: Noyes Pub. 16. Furlani, E. P. and Williams, C. (1995). Formulas for optimizing the bias field for magneto-optic recording. J. Magn. Magn. Mat. 148: 475. 17. Furlani, E. P. (1997). Field analysis and simulation of a permanent magnet bias field device for magneto-optic recording. J. Phys. D: Appl. Phys. 30: 1846. 18. Wagner, J. A. (1982). The shorted turn in the linear actuator of a high performance disk drive. IEEE Trans. Magn. 18(6): 1770. 19. Kenjo, T. and Nagamori, S. (1984). Permanent Magnet and Brushless DC Motors, Oxford: Oxford Univ. Press. 20. Campbell, P. (1974). Principles of a permanent magnet axial-field dc machine. IEE Proc. 121: 1489. 21. Furlani, E. P. (1994). Computing the field in permanent-magnet axial-field motors. IEEE Trans. Mag. 30(5): 3660.
5.16 MAGNETIC MEMS
467
22. Furlani, E. P. and Knewtson, M. A. (1997). A three-dimensional field solution for permanent-magnet axial-field motors. IEEE Trans. Mag. 33(3): 2322. 23. Furlani, E. P. (1997). Field analysis and optimization of NdFeB axial-field permanent-magnet motors. IEEE Trans. Mag. 33(5): 3883. 24. Furlani, E. P., Lee, J. K., and Dowe, D. (1993). Predicting the dynamic behavior of moving magnet actuators. J. Appl. Phys. 73(7): 3555. 25. Kovacs, G. T. A. (1998). Micromachined Transducers Sourcebook, Boston: McGraw-Hill. 26. Madou, M. (1997). Fundamentals of Microfabrication, Boca Raton, FL: CRC Press. 27. Trimmer, W. S. (1997). Micromechanics and MEMS; Classical and Seminal Papers to 1990, New York: IEEE Press. 28. Trimmer, W. S. N. (1989). Microrobots and micromechanical systems. Sensors and Actuators 19: 267. 29. Wagner, B., Benecke, W., Engelmann, G., and Simon, J. (1992). Microactuators with moving magnets for linear, torsional or multiaxial motion. Sensors and Actuators A (Physical) 32: 598. 30. Judy, J. W. and Muller, R. S. (1996). Magnetic microactuation of torsional polysilicon structures. Sensors and Actuators A (Physical) 53: 392. 31. Ruan, M., Shen, J., and Wheeler, C. B. (2000). Latching micro electromagnetic relays. Tech. Dig. Solid-State Sensor and Actuator Workshop (Hilton Head 2000) 146. 32. Dewa, A. S., Orcutt, J. W., Hudson, M., Krozier, D., Richards, A., and Laor, H. (2000). Development of silicon two-axis micromirror for an optical crossconnect. Tech. Dig. Solid-State Sensor and Actuator Workshop (Hilton Head 2000) 93. 33. Osborn, J. A. (1945). Demagnetizing factors of the general ellipsoid. Phys. Rev. 67: 351. 34. Judy, J. W. and Muller, R. S. (1996). Batch-fabricated, addressible, magnetically actuated microactuators. Tech. Dig. Solid-State Sensor and Actuator Workshop (Hilton Head 1996) 187.
a
This Page Intentionally Left Blank
Appendix A: Vector Analysis
Vector analysis provides a succinct and elegant description of electromagnetic field theory and is essential for the analysis of permanent magnet and electromechanical devices. In this appendix we review the fundamentals of vector analysis. We start with a discussion of vector calculus in the Cartesian, cylindrical and spherical coordinate systems. Next, we review the basics of vector integration. Following this, we state some key theorems and identities of vector analysis. Last, we discuss coordinate transformations. The interested reader can find comprehensive discussions of vector analysis in numerous texts [1—3].
A.1 CARTESIAN COORDINATES For problems with rectangular geometries we use the Cartesian coordinates x, y, and z with unit vectors x , y , and z (Fig. A.1). Throughout the text, italicized symbols such as x and f denote scalars (numbers, functions etc.), while boldface symbols such as x, A, and B denote vectors. Thus, a vector A with components A , A and A has the representation V W X A : A x ; A y ; A z , V W X and magnitude A : (A ; A ; A, (A.1) V W X which is a scalar. We work with scalar-valued functions such as f(x, y, z), and vector-valued functions such as A(x, y, z) : A (x, y, z)x ; A (x, y, z)y ; A (x, y, z)z . V W X 469
470
APPENDIX A Vector Analysis
Cartesian coordinates: (a) reference frame; (b) constant coordinate surfaces; (c) unit vectors; and (d) infinitesimal volume element [3].
FIGURE A.1
Note that each component of A(x, y, z) is itself a scalar-valued function. Such functions are also referred to as scalar and vector fields. Given two vectors A and B, we form the sum/difference A < B : (A < B )x ; (A < B )y ; (A < B )z V V W W X X and dot product, A·BYA B ;A B ;A B (dot product). V V W W X X Notice that the dot product gives a scalar. Two vectors are said to be orthogonal if their dot product is zero, A · B : 0 $ A orthogonal to B.
471
A.1 CARTESIAN COORDINATES
The unit vectors are orthonormal, x · x : y · y : z · z : 1, x · y : x · z : y · z : 0. In addition to the dot product, there is also the vector (cross) product, which is given by
x
y
z
A;B: A A A V W X B B B V W X : (A B 9 A B )x ; (A B 9 A B )y W X X W X V V X ; (A B 9 A B )z (cross product). (A.2) V W W V Notice that the vector product gives another vector with an orientation as shown in Fig. A.2. We use vectors to evaluate the distance between two points. Specifically, let p : (x , y , z ) and p : (x , y , z ) be arbitrary points, and let x : x x ; y y ; z z and x : x x ; y y ; z z be vectors from the origin to p and p , respectively. The distance between the two points is given by x 9 x , where
x 9 x Y ((x 9 x ) ; (y 9 y ) ; (z 9 z ) (distance from p : (x , y , z ) to p : (x , y , z )).
(A.3)
Infinitesimal displacements give rise to the following expressions for
FIGURE A.2
Vector product.
472
APPENDIX A Vector Analysis
differential length dl, area ds, and volume dv, dl : dxx ; dyy ; dzz ,
(A.4)
dy dz
(constant x)
ds : dx dz
(constant y)
dx dy
(constant z),
and
(A.5)
dv : dx dy dz,
(A.6)
respectively. Next we review vector calculus. Let f(x, y, z) and A(x, y, z) be scalar and vector-valued functions, respectively. We introduce the differential (del) operator Y
x ; y ; z , x y z
(A.7)
and define the gradient, divergence, and curl f Y
f f f x ; y ; z x y z
(gradient),
(A.8)
A A A V; W; X x y z
(divergence),
(A.9)
·AY and
x y z ;AY x y x A A A V W X A A A A X 9 W x ; V 9 X y : y z z x
;
A A W 9 V z x y
(curl).
(A.10)
We will also make use of the Laplacian operator , which is defined on scalar-valued functions as follows: f : · f :
f f f ; ; x y z
(Laplacian).
(A.11)
A.2 CYLINDRICAL COORDINATES
473
This operator can also be applied to vector-valued functions. Since x , y and z have no functional dependence, the spatial derivatives in apply to the vector components only (leaving the unit vectors unchanged). For example, A : (A )x ; (A )y ; (A )z . (A.12) V W X This operation is more complicated in other coordinated systems in which the unit vectors are themselves spatially varying functions. These include the cylindrical and spherical coordinate systems that we study next.
A.2 CYLINDRICAL COORDINATES For cylindrical geometries we use cylindrical coordinates r, , and z with unit vectors r , , and z as shown in Fig. A.3. The unit vectors for this system are orthonormal r · r : · : z · z : 1, r · : r · z : · z : 0. It is important to note that r and have different orientations at different points in space. Specifically, from Table A.2 we have r : cos()x ; sin()y and : 9sin()x ; cos()y . Thus, we find that r : and : 9r .
(A.13)
Because of this, care is needed when performing vector calculus in this coordinate system. A vector is written as A : A r ; A ; A z , P X with A < B : (A < B )r ; (A < B) ; (A < B )z . P P X X
474
APPENDIX A Vector Analysis
Cylindrical coordinates: (a) reference frame; (b) constant coordinate surfaces; (c) unit vectors; and (d) infinitesimal volume element [3].
FIGURE A.3
The dot and cross products are given by and
A · B : A B ; A B ; A B X X P P
r
z
A ; B : A A A P X B B B P X
: (A B 9 A B)r ; (A B 9 A B ) ; (A B 9 A B )z , (A.14) X X X P P X P P respectively.
475
A.2 CYLINDRICAL COORDINATES
We determine the distance between two points p : (r , , ) and p : (r , , ) using the Cartesian formula (A.3). First, transform these points into Cartesian vectors x : x x ;y y ;z z and x : x x ;y y ;z z using Table A.1. This gives x : r cos( ), y : r sin( ) for i : 1, 2. G G G G G G Next, substitute these relations into Eq. (A.3) and obtain
x 9 x : (r ; r 9 2r r cos( 9 ) ; (z 9 z ) (distance from p : (r , , z ) to p : (r , , z )).
(A.15)
The differential length, area, and volume are given by dl : drr ; r d ; dzz ,
r d dz
(constant r) (constant )
ds : dr dz
and
(A.16)
r dr d
(A.17)
(constant z),
dv : r dr d dz,
(A.18)
respectively. The operations (A.8), (A.9), and (A.10) take the form f : ·A:
f 1 f f r ; ; z , r r z
(A.19)
1 (rA ) 1 A A P ; ; X, r r r z
(A.20)
and
r r z ;A: r z A rA A P X 1 A A A A X 9 r ; P 9 X : r z z r
;
1 (rA) A 9 P z . r r
(A.21)
476
APPENDIX A Vector Analysis
The Laplacian equation (A.11) is f :
1 f 1 f f r ; ; . r r r r z
(A.22)
We apply to a vector field A taking into account the functional dependency of the unit vectors and obtain
1 2 A r A : A 9 A 9 P r P r
; A 9
1 2 A P ; A z , A ; X r r
(A.23)
where A and A are given by Eq. (A.22) with f replaced by the P respective vector component.
A.3 SPHERICAL COORDINATES For spherical geometries we use spherical coordinates r, , and with unit vectors r , , and . These unit vectors are orthonormal r · r : · : · : 1 r · : r · : · : 0. They are also vector-valued spatially varying functions (Fig. A.4). From Table A.3 we have r : sin() cos()x ; sin() sin()y ; cos()z , : cos() cos()x ; cos() sin()y 9 sin()z , : 9sin()x ; cos()y ,
(A.24)
and, therefore, : 9r sin() 9 cos() r :
: 9r
: cos()
r : sin().
(A.25)
A.3 SPHERICAL COORDINATES
477
Spherical coordinates: (a) reference frame; (b) constant coordinate surfaces; (c) unit vectors; and (d) infinitesimal volume element [3].
FIGURE A.4
We must be mindful of Eq. (A.25) when performing vector calculus in this system. Vectors have representations A : A r ; A ; A , P F with A ; B : (A ; B )r ; (A ; B ) ; (A ; B) . P P F F The dot and cross products are given by A · B : A B ; A B ; A B P P F F
478
APPENDIX A Vector Analysis
and
r
A;B: A P B P
A A F B B F
: (A B 9 A B )r ; (A B 9 A B) ; (A B 9 A B ) , F F P P P F F P
(A.26)
respectively. We evaluate the distance between two points p : (r , , ) and p : (r , , ) using Eq. (A.3). We transform p and p into Cartesian vectors x : x x ; y y ; z z and x : x x ; y y ; z z using Table A.1. This gives x : r sin( ) cos( ), y : r sin( ) sin( ), and z : r cos( ) for G G G G G G G G G G G i : 1, 2. These are substituted into Eq. (A.3), which yields
x 9 x : (r ; r 9 2r r sin( ) sin( ) cos( 9 ) (distance from p : (r , , ) to p : (r , , )).
(A.27)
The differential length, area, and volume are given by dl : drr ; r d ; r sin() d ,
ds :
r sin() d d
(constant r)
r sin() dr d
(constant )
r dr d
(constant ),
(A.28)
(A.29)
and dv : r sin() dr d d,
(A.30)
respectively. The vector operations (A.8), (A.9), and (A.10) are as follows: f :
·A:
f 1 f 1 f r ; ; , r r r sin()
1 (rA ) 1 (sin()A ) 1 A P ; F ; , r r r sin() r sin()
(A.31)
(A.32)
479
A.4 INTEGRALS OF VECTOR FUNCTIONS
and
r r r sin() ;A: r A rA r sin()A P F (A sin()) A 1 A 1 1 (rA) P9 : 9 F r ; r sin() r sin() r
;
1 (rA ) A F 9 P . r r
(A.33)
The Laplacian equation (A.11) is given by f :
f 1 f 1 f 1 r ; sin() ; . (A.34) r r sin() r sin() r r
We apply the operator to a vector field A taking into account the functional dependency of the unit vectors (A.25) and obtain
2 2 A 2 cos() 2 A F9 A : A 9 A 9 A 9 r P r P r r sin() F r sin()
1 2 A 2 cos() A P9 A ; ; A 9 F r sin() F r r sin() ; A 9
1 2 A 2 cos() A P; F , (A.35) A ; r sin() r sin() r sin()
where A , A , and A are given by Eq. (A.34) with f replaced by P F the respective vector component.
A.4 INTEGRALS OF VECTOR FUNCTIONS In our study of magnetic structures and devices we frequently encounter integrals with vector-valued integrands. In this section we review such integrals, and the procedures for evaluating them. The first integral we study is a line integral of the form
f dl. (A.36) ! In this integral f(x, y, z) is a scalar-valued function, C is the path of
480
APPENDIX A Vector Analysis
FIGURE A.5
Line integration path and line element.
integration, and dl is a differential line element along C (Fig. A.5). Notice that this integral yields a vector. For example, in Cartesian coordinates we use the line element (A.4), which gives
f dl :
!
f(x, y, z)[dxx ; dyy ; dzz ]
!
:
f(x, y, z) dx x ;
!
f(x, y, z) dy y ;
!
f(x, y, z) dz z , (A.37)
!
which is obviously a vector. If the integral is to be evaluated from point P to point P we write . f dl. If integration path C is closed we write . w f dl. Notice that in the second step in Eq. (A.37) we have taken the unit ! vectors outside the integral because they are constant in both magnitude and direction. This is true only in the Cartesian system. Care must be taken when evaluating such integrals using cylindrical and spherical coordinates because the direction of the unit vectors in these systems varies from point to point. EXAMPLE A.4.1 Evaluate the line integral of f(x, y) : x ; y along the three paths shown in Fig. A.6. SOLUTION A.4.1 Along path 1 we have
!
W V y dy y ; (x ; y) dx x x y ; y x ; y . :x 3 3
(x ; y) dl :
481
A.4 INTEGRALS OF VECTOR FUNCTIONS
FIGURE A.6
Line integration paths.
To integrate along path 2 we need to relate y and x. Along this path y y: x x and y dy : dx. x Therefore dl : dxx ; (y /x ) dxy . The integration reduces to
(x ; y) dl :
!
V
x ;
y x x
y dxx ; dxy x
y x y y x x ; 1 ; y . x 3 x 3
: 1; Last, along path 3 we have
(x ; y) dl : !
V x dx x ;
W
(x ; y) dy y
x y : x ; y x ; y . 3 3
)
482
APPENDIX A Vector Analysis
The second integral we consider is a line integral of the form
A · dl. (A.38) ! Here, A : A (x, y, x)x ; A (x, y, x)y ; A (x, y, x)z is a vector-valued funcV W X tion and A · dl is the projection of A along the path of integration C. This integral results in a scalar. For example, in Cartesian coordinates we have
A · dl :
[A (x, y, x) dx ; A (x, y, x) dy ; A (x, y, x) dz]. V W X
! ! EXAMPLE A.4.2 Evaluate the line integral of A(x, y) : 2xyx ; xy along the three paths shown in Fig. A.6. SOLUTION A.4.2 Along path 1 we have
!
A(x, y) · dl :
W
0 dy ;
: xy .
V 2xy dx
Along path 2 we have y y: x x and y dy : dx, x with dl : dxx ; (y /x ) dxy . Thus, the integration reduces to V V y A(x, y) · dl : 2xy dx ; x dx x ! y V y V :2 x dx ; x dx x x : xy . Last, along path 3 we obtain
!
V W 2x0 dx ; x dy : xy .
A(x, y) · dl :
)
483
A.4 INTEGRALS OF VECTOR FUNCTIONS
The third integral we study is a volume integral of the form
A dv.
(A.39)
4
We integrate the vector field A over a volume V here, which results in a vector. To see this, consider the integral in Cartesian coordinates,
A dv : 4
4
:
[A (x, y, x)x ; A (x, y, x)y ; A (x, y, x)z ] dx dy dz V W X
4
;
A (x, y, z) dx dy dz x ; V
4
4
A (x, y, z) dx dy dz y W
A (x, y, z) dx dy dz z . X
The last integral we study is a surface integral of the form
A · n ds.
(A.40)
1
In this integral A is a vector field, n is the unit normal to a surface S, and the projection A · n is integrated over S. The integral (A.40) is also written as
A · ds,
(A.41)
1
where ds : n ds. To evaluate Eq. (A.40) we need to know the sense of n . The following convention is adopted: If S is a closed surface (enclosing a volume V) then the positive direction of n is in the outward direction from the volume (i.e., pointing away from V). If S is an open surface bounded by a curve C, then the positive direction of n depends on the direction in which C is traversed. This is specified in terms of the right-hand rule (Fig. A.7). Specifically, if the fingers of the right hand follow the direction of travel around the path C, then the thumb points in the direction of positive n . EXAMPLE A.4.3 Evaluate the integral w A · ds where A(x, y) : rr ; #zz , and 1 S is the closed surface of the cylinder shown in Fig. A.8. Here, and # are constants.
484
APPENDIX A Vector Analysis
FIGURE A.7
Orientation of a normal vector for a bounded surface.
SOLUTION A.4.3 We use cylindrical coordinates. First, we determine the unit normals on the surface of the cylinder. We find that
n :
FIGURE A.8
z
z:L
r
r:a
9z
z : 0.
Cylindrical surface of integration.
485
A.5 THEOREMS AND IDENTITIES
Next, we evaluate A · ds over S using Eq. (A.17) and obtain
#Lr dr d
(ds : r dr d)
z:L
A · ds : a d dz
(ds : a d dz)
r:a
(ds : r dr d)
z : 0.
Therefore,
0
A · ds : #L
1
L ?
r dr d ; a
: #La ; a2L
* L
d dz
: (2 ; #)aL.
)
A.5 THEOREMS AND IDENTITIES In this section we review some important theorems, propositions, and identities of vector calculus. There are two key integral theorems that are fundamental to the study of electromagnetic field theory. The first is Stokes’ theorem. This theorem equates the integral of the curl of a vector field over an open surface S to the line integral of the vector field around the curve C, which forms the boundary of S (Fig. A.9),
A · dl : !
Integral around closed curve C bounding surface S
( ; A) · n ds
(Stokes’ theorem).
(A.42)
1 Integral over surface S
We digress briefly to discuss the terms in this equation. Notice that to evaluate Eq. (A.42), we need to know the orientation of the unit normal n relative to the direction taken for the line integral around curve C. This is specified in terms of the right-hand rule (Fig. A.7). Specifically, if the fingers of the right hand follow the direction of travel around the path C, then the thumb points in the direction of positive n . EXAMPLE A.5.1 Verify Stokes’ theorem for the vector field A(x, y) : (2x 9 y)x 9 yzy ; yzz where S is the half-surface of the sphere x ; y ; z : a and C is its bounding curve in the x-y plane (Fig. A.10).
486
APPENDIX A Vector Analysis
FIGURE A.9
Geometry for Stokes’ theorem.
SOLUTION A.5.1 First, we evaluate ( ; A) · n ds. It is easy to check that 1 ; A : z . In this case, the open surface S is defined by x ; y ; z : a with z & 0. The unit vector n on S is given by n :
xx ; yy ; zz . a
Therefore, z ( ; A) · n : . a
FIGURE A.10
Hemisphere geometry with reference frame.
487
A.5 THEOREMS AND IDENTITIES
We want to evaluate (z/a) ds over the hemisphere. We use spherical coordi1 nates in which z : r cos(), and ds : r sin() d d. On S we have r : a and, therefore,
z L L ds : a cos() sin() d d a 1 : a.
(A.43)
Next, we evaluate w A · dl using cylindrical coordinates. We transform A(x, y) ! into cylindrical coordinates using Table A.1. This gives A(r, ) : (2r cos() 9 r sin())(cos()r 9 sin() ) 9 r sin()z(sin()r ; cos() ) ; r sin()zz . On the curve C, r : a, z : 0, and dl : a d . Therefore, A(r, ) · dl : (92a cos() sin() ; a sin()) d and
!
A · dl :
L
(92a cos() sin() ; a sin()) d
: a.
(A.44)
Notice that Eq. (A.44) equals Eq. (A.43), which verifies Stokes’ theorem.
)
The second integral theorem that we study is the Divergence theorem. Before we discuss this, we define the flux of a vector field A through a surface S, Flux of A : through surface S
A · n ds. 1 The Divergence theorem equates the integral of the divergence of a vector field over a volume V to the total flux passing through the closed
488
APPENDIX A Vector Analysis
Geometry for Divergence theorem.
FIGURE A.11
surface S that forms the boundary for V (Fig. A.11),
· A dv :
4
A · n ds
(Divergence theorem).
(A.45)
1
Integral over volume V
Integral over surface S enclosing V
In Eq. (A.45) n is the outward unit normal to the surface S. EXAMPLE A.5.2 Verify the Divergence theorem for the vector field A(x, y) : rr ; #zz where V and S are the volume and surface of the cylinder shown in Fig. A.8. Here, and # are constants. SOLUTION A.5.2 We have already evaluated w A · n ds in Example A.4.3. 1 Specifically, we have
A · ds : #La ; a2L
1 : (2 ; #)aL.
(A.46)
It remains to evaluate · A dv. We use cylindrical coordinates and obtain 4 ·A:
1 (r) (#z) ; r r z
: 2 ; #.
489
A.5 THEOREMS AND IDENTITIES
Therefore,
· A dv : (2 ; #)
4
dv 4 : (2 ; #)aL, )
which is the same as Eq. (A.46). Throughout this text we make use of the following definitions: DEFINITION A.5.1 A vector field V is said to be solenoidal if ·V:0
(solenoidal field).
DEFINITION A.5.2 A vector field V is said to be irrotational if ;V:0
(irrotational field).
PROPOSITION A.5.1 A solenoidal vector field V can be written as the curl of a vector field, that is, ·V:0$V:;A for some A. PROPOSITION A.5.2 An irrotational vector field V can be written as the gradient of a scalar-valued function, that is, ; V : 0 $ V : 9 for some . The following theorems are important in our study of Maxwell’s equations. THEOREM A.5.1 A vector is uniquely specified by giving its divergence and curl within a region, and its normal component over the boundary of the region. Let be a region in space and let f and K be given scalar and vector-valued functions, respectively. Then a vector field V is uniquely specified if ·V:f ;V:K
(A.47)
and the normal component V is specified on the boundary of . L THEOREM A.5.2 Helmholtz’s Theorem: A vector field V satisfying Eq. (A.47) with both f and K vanishing at infinity may be written as the sum of two parts,
490
APPENDIX A Vector Analysis
one that is irrotational, and another that is solenoidal, that is, V : 9 ; ; A.
(A.48)
Given Eqs. (A.47) and (A.48), it can be shown that
(x) :
1 4
f(x) dv
x 9 x
(A.49)
A(x) :
1 4
K(x) dv.
x 9 x
(A.50)
and
Last, we list some useful identities:
Vector Algebra A ; B : 9B ; A (A ; B) · C : A · (B ; C) : (C ; A) · B
(A.51)
A ; (B ; C) : B(A · C) 9 C(A · B),
Differential Identities · ( ; A) : 0 · (A ; B) : B · ( ; A) 9 A · ( ; B) ; (A ; B) : (B · )A 9 (A · )B ; A( · B) 9 B( · A) ( ; A) ; A : (A · )A 9 (A · A) ; ; A : ( · A) 9 A · (fA) : f · A ; (A · )f ; (f) : 0 (fg) : fg ; gf,
(A.52)
491
A.6 COORDINATE TRANSFORMATIONS
and Integral Identities
f(g) · ds :
1
[fg ; (f) · (g)] dv 4
[fg 9 gf)] · ds :
1
(A.53) [fg 9 gf] dv.
4
A.6 COORDINATE TRANSFORMATIONS This section contains transformation relationships for the Cartesian, cylindrical and spherical coordinate systems. A vector A has the following representations: A : A x ; A y ; A z V W X
(Cartesian)
: A r ; A ; A z P X
(cylindrical)
: A r ; A ; A P F
(spherical).
We compute the components of A in a given coordinate system by taking the inner product with the appropriate unit vectors. For example, given the Cartesian representation for A we obtain the radial component in cylindrical coordinates as follows: A : A · r P : A x · r ; A y · r ; A z · r . V W X We summarize the transform relations in the tables that will follow. To avoid confusion, we use r and r to denote the cylindrical and spherical A Q radial coordinate variable, respectively. Table A.1 is used to convert from cylindrical or spherical coordinates to Cartesian coordinates. EXAMPLE A.6.1 Consider a vector A : A r ; A ; A z at a point (r , , z) P A X A in cylindrical coordinates. Determine its Alocation and representation in Cartesian coordinates.
492
APPENDIX A Vector Analysis Coordinate transformations for Cartesian coordinates
TABLE A.1
Cartesian
Cylindrical
x
: cos()r 9 sin() A
y z x y z A V
: : : : : :
A W
:
A X
:
Spherical
: sin() cos()r ; cos() cos() Q 9sin() sin()r ; cos() : sin() sin()r ;cos() sin() ;cos() A Q z : cos()r 9 sin() r cos() : r sin() cos() A Q r sin() : r sin() sin() A Q z : r cos() Q A cos()9A sin() : A sin() cos() ; A cos() cos() PA PQ F 9 A sin() A sin()9A cos() : A sin() sin() ; A cos() sin() PA PQ F ; A cos() A : A cos() 9 A sin() X PQ F
SOLUTION A.6.1 From Table A.1 the location in Cartesian coordinates is x : r cos(), y : r sin(), z : z. Similarly, the vector representation is A A A : (A cos() 9 A sin())x ; (A sin() 9 A cos())y ; A z . PA PA X A V
)
A W
Table A.2 is used to convert from Cartesian or spherical coordinates to cylindrical coordinates. TABLE A.2
Coordinate transformations for cylindrical coordinates
Cylindrical
Cartesian
r A z r A
: : : :
: tan\
z A P AA A X
: : : :
cos()x ; sin()y 9sin()x ; cos()y z (x ; y
y x
z A cos() ; A sin() V W 9A sin() ; A cos() V W A X
Spherical : : : :
sin()r ; cos() Q cos()r 9 sin() Q r sin() Q
: : : : :
r cos() Q A sin() ; A cos() P F AQ A cos() 9 A sin() PQ F
493
A.6 COORDINATE TRANSFORMATIONS TABLE A.3
Coordinate transformations for spherical coordinates
Spherical
Cartesian
r Q r Q
: : : :
: cos\
: tan\
A PQ A F A
sin() cos()x ;sin() sin()y ;cos()z cos() cos()x ;cos() sin()y 9sin()z 9sin()x ; cos()y (x ; y ; z
z
(x ; y ; z
y x
: A sin() cos() ; A sin() sin() V W ; A cos() X : A cos() cos() ; A cos() sin() V W 9 A sin() X : 9A sin() ; A cos() V W
Cylindrical : : : :
sin()r ; cos()z A cos()r 9 sin()z A (r ; z A z : cos\ (r ; z A
: : A sin() ; A cos() PA X : A cos() 9 A sin() P X : A
EXAMPLE A.6.2 Consider a vector A : A r ; A ;A at a point (r , , ) PQ Q F Q in spherical coordinates. Determine its location and representation in cylindrical coordinates. SOLUTION A.6.2 From Table A.2 the location of (r , , ) is r : r sin(), Q A Q :, and z : r cos(). The vector is Q A : (A sin() ; A cos())r ; A ; (A cos() 9 A sin())z . PQ PQ F A F A PA
)
A X
Table A.3 is used to convert from Cartesian or cylindrical coordinates to spherical coordinates.
EXAMPLE A.6.3 Consider a vector A : A x ; A y ; A z at a point (x, y, z) V W X in Cartesian coordinates. Determine its location and representation in spherical coordinates. SOLUTION A.6.3 From Table A.3 the location of (x, y, z) is r : (x;y;z, Q
494
APPENDIX A Vector Analysis
: cos\ z/(x ; y ; z !, and : tan\(y/x). The vector is A : (A sin() cos() ; A sin() sin() ; A cos())r V W X Q A PQ
; (A cos() cos() ; A cos() sin() 9 A sin()) V W X A F
; (9A sin() ; A cos()) . V W
)
A
References 1. Arftgen, G. (1985). Mathematical Methods for Physicists, 3rd ed., New York: Academic Press. 2. Zahn, M. (1987). Electromagnetic Field Theory: A Problem Solving Approach, Malabar, FL: R. E. Krieger Pub. Co. 3. Edminister, J. A. (1979). Theory and Problems of Electromagnetics, Schaum’s Outline Series, New York: McGraw-Hill.
Appendix B: Green’s Functions
In this section we review the basic properties of the free-space Green’s function for Poisson’s equation : f.
(B.1)
This equation plays a key role in both the current and charge models of magnetic materials as discussed in Sections 3.3 and 3.4, respectively. The Green’s function G(x, x) for Poisson’s equation satisfies G(x, x) : .(x 9 x), V where indicates differentiation with respect to the unprimed coordiV nates. The Dirac delta function .(x 9 x) satisfies
.(x 9 x) dv : 1,
\ where the integration is over all space. Physically, .(x 9 x) corresponds to a point source located at the point x. Given G(x, x), we obtain a solution to Eq. (B.1) using the following relation,
(x) :
G(x, x)f(x) dv.
(B.2)
\ In Eq. (B.2) x is an observation point, and x is a source point. 495
496
APPENDIX B Green’s Functions
For 3D problems it is well known that G(x, x) : 9
1 1 . 4 x 9 x
From Eqs. (A.3), (A.15), and (A.27) we obtain explicit formulas for the free space Green’s function in the Cartesian, cylindrical and spherical coordinate systems, respectively. Specifically, we find that 1 1 , 4 ((x 9 x) ; (y 9 y) ; (z 9 z)
(B.3)
1 1 , 4 (r ; r 9 2rr cos( 9 ) ; (z 9 z)
(B.4)
1 1 , 4 (r ; r 9 2rr sin() sin() cos( 9 )
(B.5)
G(x, x) : 9 G(x, x) : 9 and G(x, x) : 9
respectively. These functions play a key role in obtaining the field solutions for free-standing permanent magnet structures. We demonstrated this in various examples in Chapter 4.
Appendix C: Systems of Equations
In this section we review various numerical methods for solving systems of first-order, time-dependent differential equations. Such systems naturally arise as the equations of motion for electromechanical devices (Section 5.6). We briefly discuss the Euler, improved Euler, and fourth-order Runge-Kutta methods. Each of these can be demonstrated using an initial-value problem consisting of a single first-order equation, dy : f(y, t), dt
(C.1)
and an initial condition y(t ) : y . (C.2) We seek a solution to Eqs. (C.1) and (C.2) for a given time interval (t , t ). As a first step, divide the interval (t , t ) into a sequence of discrete values t , t , t , . . . , t , t , . . . , t , as shown in Fig. C.1. Next, L L>
integrate Eq. (C.1) over a typical interval, say (t , t ), L L> RL> dy RL> dt : f(y, t) dt. dt RL RL The left side can be simplified as follows:
y(t ) 9 y(t ) : L> L
RL>
f(y, t) dt.
(C.3)
RL 497
498
APPENDIX C Systems of Equations
Discretization of time interval.
FIGURE C.1
The various solution methods entail different estimates of the integral in Eq. (C.3). The accuracy of a method improves with the accuracy of the estimate.
C.1 EULER’S METHOD Euler’s method is a very basic solution method and is easy to implement. In this method, the area under the curve in Eq. (C.3) is estimated using the value of the function at the left endpoint of the interval (Fig. C.2),
RL>
RL
f(y, t) dt t f(y(t ), t ). L L
(C.4)
We substitute Eq. (C.4) into Eq. (C.3) and obtain y(t ) : y(t ) ; t f(y(t ), t ). L> L L L
(C.5)
This can be rewritten as y(t
) 9 y(t ) L> L : f(y(t ), t ). L L t
(C.6)
By comparing Eqs. (C.1) and (C.6) we see that Euler’s method amounts
499
C.1 EULER’S METHOD
FIGURE C.2
Integral estimate for the Euler method.
to estimating the derivative by a forward difference quotient, dy y(t ) 9 y(t ) L . L> dt t
(C.7)
This is equivalent to keeping only the first term in the Taylor series expansion for y(t ; t), which is of the form y(t ; t) : y(t) ; ;
dy(t) 1 dy(t) t ; t dt 2! dt
1 dy(t) 1 dy(t) t ; t ; · · · . 4! dt 3! dt
(C.8)
It is convenient to rewrite Eq. (C.5) using index notation, y : y ; t f(y , t ). L> L L L
(C.9)
This relation forms the basis for the following solution procedure: 1. Given an initial-value problem (C.1) and (C.2), choose a time increment t and discretize the time interval (t , t ) for which the solution is desired. 2. Start with the initial value point (t , y ) and use Eq. (C.9) to compute the solution point (t , y ). 3. Given (t , y ), use Eq. (C.9) to compute the next solution point (t , y ).
500
APPENDIX C Systems of Equations
FIGURE C3.
Flow chart for implementation of the Euler method.
4. Repeat step 3 using the solution value (t , y ) to compute the L L subsequent solution values (t , y ), for n : 2, 3, 4, . . . . L> L> Continue in this fashion until solution values are obtained for the time interval (t , t ). A flow chart for this procedure is shown in Fig. C.3. Although Euler’s method is easy to implement, it is somewhat limited in terms of its accuracy and stability. Regarding accuracy, the error of this method in any one interval is proportional to t, but the global error is proportional to t. For stability, Euler’s method may
501
C.1 EULER’S METHOD
become unstable if the time constant of the differential equation is negative. For example, consider the equation dy : 9ay, dt
(C.10)
with initial condition y(0) : y 0, and a 0. This equation has an analytical solution y(t) : y e\?R. However, if we apply Eq. (C.5) to (Eq. (C.10) we obtain ) : y(t )(1 9 at). L> L Notice that if at 1, the computed solution is diminishing and positive. However, if at 1, the solution becomes unstable with alternating positive and negative values, and if at 2, the magnitude of the solution increases (in a negative sense) after each step. Thus, the choice of step size t can be crucial to the stability of Euler’s solution procedure. Euler’s method is easily generalized to a system of linear first-order equations as shown in the following example: y(t
EXAMPLE C.1.1 Apply Euler’s method to the equations of motion (5.86) for an electromechanical device. SOLUTION C.1.1 We reduce Eq. (5.86) to a set of algebraic equations using Euler’s method in which derivatives are approximated by the forward difference quotient (C.7). This results in the following system of equations: i(t ; t) : i(t) ;
"(i, x) t V(t) 9 i(t)(R ; R ) 9 x(t) L x
v(t ; t) : v(t) ;
t F(i, x) m
x(t ; t) : x(t) ; t v(t).
(C.11)
Equations (C.11) can be written using index notation,
(i , x ) t L L v V 9 i (R ; R ) 9 i :i ; L> L L L L L x t v : v ; F(i , x ) L> L m L L
x : x ; t v . (C.12) L> L L The solution procedure for Eq. (C.12) is similar to the preceding steps 1 through 4. Specifically, start with the initial conditions [i , v , x ] and use Eq. (C.12) to
502
APPENDIX C Systems of Equations
obtain [i , v , x ], which gives the solution at t : t. Repeat this process to obtain [i , v , x ], [i , v , x ], . . . , [i , v , x ], . . . , until solution values are L L L obtained for the entire interval (t , t ). )
C.2 IMPROVED EULER METHOD The accuracy of Euler’s method can be improved by using a more accurate estimate of the integral in Eq. (C.3). In the improved Euler method the integral is estimated using the trapezoidal rule as shown in Fig. C.4. This gives
RL>
t [f(y(t ), t ) ; f(y(t ), t )]. L> L> L L 2
(C.13)
t ) : y(t ) ; [f(y(t ), t ) ; f(y(t ), t )]. L> L L L L> L> 2
(C.14)
f(y, t) dt
RL We substitute Eq. (C.13) into Eq. (C.3) and obtain y(t
Notice that the unknown y(t ) appears on both sides of Eq. (C.14). L> Thus, it cannot be obtained explicitly. However, Eq. (C.14) can be transformed into an explicit relation for y(t ) by using an estimate on L> the right side. Specifically, y(t ) can be estimated using the Euler L> relation (C.5). Let y (t ) denote the estimate, L> y (t ) : y(t ) ; t f(y(t ), t ). L> L L L
FIGURE C.4
Integral estimate for the improved Euler method.
503
C.3 RUNGE-KUTTA METHODS
Substitute this in the right side of Eq. (C.14) and obtain t y(t ) : y(t ) ; [f(y(t ), t ) ; f(y (t ), t )]. L> L L L L> L> 2
(C.15)
This can be rewritten using index notation as follows: t y : y ; [f(y , t ) ; f(y , t )], L> L L L L> L> 2
(C.16)
where y : y ; t f(y , t ). (C.17) L> L L L Equations (C.16) and (C.17) can be solved iteratively as described in steps 1 through 4 given in the preceding text. A flow chart of the solution procedure is shown in Fig. C.5. The improved Euler method is more accurate and stable than the standard Euler method. Specifically, the error of this method in any one interval is proportional to t and global error is proportional to t.
C.3 RUNGE-KUTTA METHODS Runge-Kutta methods are widely used for integrating first-order equations. The fourth-order Runge-Kutta method is perhaps the most popular. In this method, the estimate of Eq. (C.3) is based on Simpson’s 1/3 rule [1]. This leads to the following set of formulas: k : t f(y , t ) L L
k t k : t f y ; , t ; L 2 L 2 k t k : t f y ; , t ; L 2 L 2 k : t f(y ; k , t ; t) L L y : y ; [k ; 2k ; 2k ; k ]. L> L
(C.18)
A flow chart of the solution procedure is shown in Fig. C.6. The fourth-order Runge-Kutta method is accurate to the fourth-order term in the Taylor expansion. Thus, the error in any interval is proportional to
504
FIGURE C.5
APPENDIX C Systems of Equations
Flow chart for implementation of the improved Euler method.
C.3 RUNGE-KUTTA METHODS
FIGURE C.6
method.
505
Flow chart for implementation of the fourth-order Runge-Kutta
506
APPENDIX C Systems of Equations
t. The following example demonstrates the use of this method for the solution of a first-order system of equations. EXAMPLE C.3.1 Apply the fourth-order Runge-Kutta method to the equations for rotational motion (5.87). SOLUTION C.3.1 The equations governing rotational motion are
di(t) 1 "(i, ) : V(t) 9 i(t)(R ; R ) 9 (t) , L dt
(C.19)
d(t) : (t), dt
(C.20)
and d (t) T(i, ) : . (C.21) J dt K Note that there are three dependent variables i(t), (t), and (t). We need a set of Runge-Kutta coefficients for each variable. We denote these coefficients by (k , k , k , k ), (m , m , m , m ), and (p , p , p , p ), respectively. To implement the Runge-Kutta method, it is convenient to rewrite Eqs. (C.19)—(C.21) as follows: di(t) : f(V(t), i(t), (t), (t)), dt d(t) : g( (t)), dt and d (t) : h(i(t), (t)), dt where f(V(t), i(t), (t), (t)) :
1 "(i(t), (t)) V(t) 9 i(t)(R ; R ) 9 (t) , L
g( (t)) : (t), and h(i(t), (t)) :
T(i(t), (t)) . J K
507
C.3 RUNGE-KUTTA METHODS
The Runge-Kutta coefficients are determined from the following relations: k : t f(V(n), i(n), (n), (n)) m : t g( (n)) p : t h(i(n), (n)), k m p k : t f V(n), i(n) ; , (n) ; , (n) ; 2 2 2
(C.22)
p m : t g (n) ; 2
k m p : t h i(n) ; , (n) ; , 2 2 k m p k : t f V(n), i(n) ; , (n) ; , (n) ; 2 2 2
(C.23)
p m : t g (n) ; 2
k m p : t h i(n) ; , (n) ; , 2 2
(C.24)
and k : t f(V(n), i(n) ; k , (n) ; m , (n) ; p ) m : t g( (n) ; p ) p : t h(i(n) ; k , (n) ; m ). (C.25) Once the coefficients have been computed, they are used to determine the subsequent solution values as follows: i : i ; [k ; 2k ; 2k ; k ] L> L : ; [m ; 2m ; 2m ; m ] L> L : ; [p ; 2p ; 2p ; p ]. (C.26) L> L The solution procedure for Eqs. (C.19)—(C.21) is similar to steps 1 through 4 of Euler’s method. Specifically, start with the initial conditions [i , , ] and use Eqs. (C.22)—(C.25) to determine the coefficients (k , k , k , k ), (m , m , m , m ), and (p , p , p , p ). Once these are known, use Eq. (C.26) to
508
APPENDIX C Systems of Equations
obtain [i , , ], which gives the solution at t : t. Repeat this process to obtain [i , , ], [i , , ], . . . , [i , , ], . . . , until solution values are L L L obtained for the interval (t , t ). ) References 1. Nakamura, S. (1991). Applied Numerical Methods with Software, Englewood Cliffs, NJ: Prentice-Hall.
Appendix D: Units
In this section we discuss three systems of units that are commonly used in magnetism. These are the CGS or Gaussian system, and two SI systems (International System of Units or Système international d’unités). The SI systems are referred to as the Kennelly and Sommerfeld conventions, respectively. The CGS system is based on the fundamental units of centimeter, gram and second, and the SI systems are based on the MKSA units of meter, kilogram, second, and ampere. In this book we use SI system units in the Sommerfeld convention. The basic entities and units in the Sommerfeld and CGS systems are listed in Table D.1. Maxwell’s equations in these two systems are as follows: Maxwell’s Equation SI ;H:J;
CGS (Gaussian) D t
·B:0 B t ·D:
;E:9
4 1 D J; c c t ·B:0 1 B ;E:9 c t · D : 4.
;H:
(D.1)
509
510
APPENDIX D Units TABLE D.1
Fields and units in the SI (Sommerfeld) and CGS systems
Symbol
Description
SI
CGS
H B M
Magnetic field strength Flux density Magnetization Flux
A/m Tesla A/m Weber
Oe Gauss emu/cm Maxwell
The constitutive relations are as follows:
Constitutive Relations SI (Sommerfeld)
CGS (Gaussian)
B : (H ; M) D:+ E;P : 4 ; 10\ 10 + : 4c
B : H ; 4M
(D.2)
D : E ; 4P :1 + : 1.
To convert between the Sommerfeld and CGS systems we use Tables D.2 and D.3. For example, from Table D.1 we find that 1 A/m : 4/10 Oe and that 1 J/m : 40 GOe and so on. Similarly, from Table D.3 we find that 1 Oe : 10/4 A/m and that 1 MGOe : 10/4 kJ/m and so on.
TABLE D.2
Conversion from SI to CGS units
SI H in A/m B in Tesla M in A/m in Webers (BH) in J/m
(BH) in kJ/m
to ; ; ; ; ; ;
4/10 10 10\ 10 40 4/10
CGS : : : : : :
H in Oe B in Gauss M in emu/cm in Maxwells (BH) in GOe
(BH) in MGOe
511
UNITS TABLE D.3
Conversion from CGS to SI units
CGS H in Oe B in Gauss M in emu/cm in Maxwells (BH) in GOe
in MGOe (BH)
to ; 10/4 ; 10\ ; 10 ; 10\ ; 10\/4 ; 10/4
SI : : : : : :
H in A/m B in Tesla M in A/m in Webers (BH) in J/m
(BH) in kJ/m
Other useful relations are 1 T:1
N , A·m
and 1 Wb : 1 T m. The Kennelly and Sommerfeld conventions differ in the definition of magnetization. In the Kennelly convention the constitutive relation is B: H;J (Kennelly convention). Here, J is called the magnetic polarization or intensity of magnetization and is measured in tesla. It is also referred to as the intrinsic flux density and is denoted by B . Notice that J : M. The fundamental element in G magnetism is the magnetic dipole m. In the Sommerfeld convention m is measured in A · m. In the Kennelly and CGS systems it is measured in Wb · m and emu, respectively (1 emu : 4 ; 10\ Wb · m). If a magnetic dipole is subjected to an external B-field, it acquires an energy E : 9m · B,
(D.3)
T : m ; B.
(D.4)
and experiences a torque
In both the Kennelly and CGS systems these are given by E : 9m · H and T : m ; H. In the remainder of this section we give conversion factors for various mechanical units. There are three systems that are commonly used for mechanical analysis. These are the MKS, CGS and British systems. In the MKS system the fundamental units of length, mass, and
512
APPENDIX D Units TABLE D.4
Conversion of mechanical variables from MKS
to CGS units MKS
to ; ; ; ;
mass in kg force in newtons (N) torque in N · m energy in joules
10 10 10 10
CGS : : : :
mass in grams force in dynes torque in dyne · cm energy in ergs
time are the meter (m), kilogram (kg), and second. The unit of force is the newton (N), and the unit of energy is the joule (J). We use the MKS system throughout this text. In the CGS system, the fundamental units are the centimeter (cm), gram (g), and second. In this system the units of force and energy are the dyne and erg, respectively. Last, in the British system the units of mass, length, and time are the slug, foot, and second, respectively. For example, acceleration due to gravity is g : 32 ft/s, whereas in the MKS system g : 9.8 m/s. In the British system the unit of force is the pound (lb), and a commonly used unit of energy is the British thermal unit (Btu). Tables D.4 and D.5 give conversion factors for various mechanical units. Other useful relations are as follows: 1. Length: 1.0 in : 2.54 cm, 1.0 ft : 0.3048 m. 2. Mass density: 1.0 lb/ft : 16.02 kg/m. 3. Torque: 1.0 lb · in:0.113 N · m, 1.0 oz · in:72 gm · cm, 1.0 N · m : 1.019 ; 10 g · cm.
Conversion of mechanical variables from British to MKS units
TABLE D.5
British mass in slugs force in lbs torque in ft · lb energy in Btu
to ; ; ; ;
14.59 4.45 1.3558 1055
MKS : : : :
mass in kg force in newtons torque in N · m energy in joules
Index A Actuators axial-field, 393—401 linear, 413—421 magnetic circuit, 383—392 magnetooptical bias field, 406—413 moving magnet, 449 plunger, 383, 387, 390 resonant, 402—406 rotary bias field, 322 solenoid, 136 Alnico, 48—50 Ampere’s circuital law, 78, 88, 338 Angular momentum, orbital versus spin, 10—11 Anisotropic materials, 5, 77 Anisotropy magnetocrystalline, 27—31 shape, 31—33 Antiferromagnetic material, 6, 7 Atomic magnetic moments, 7—9 multielectron atoms, 14—17 single electron atoms, 9—14 Axial coupling, 286, 288, 297—306 Axial-field actuators, 393—401 Axial-field motors, 421—437 no-load speed, 424 stall-torque, 424 Axially polarized cylindrical shell, 245—253 Axially polarized multipole cylindrical disk magnet, 253—264 B Back emf constant, 368 Barium ferrite, 46 Barkhausen, 36 Bar magnet, 211—217 rectangular, 133—135 Bearing configurations, radial and axial, 273 Bessel function, 176 B-H hysteresis loop, 36—39 Bias field actuators, 406—413 magnetooptical recording and, 323 Bias magnets cylindrical, 218—264 rectangular, 208—218
Biot-Savart law, 104 Bipolar cylindrical magnet, 218—224 Bipolar sphere, 182—185 Bloch walls, 34—35 Bohr magneton, 11 Boltzmann’s constant, 17 Boltzmann statistics, 17 Bonded ferrite magnets, 54 Bonded magnets, 53—55 Boundary conditions, 79—81 Boundary-value problems, 161 Cartesian coordinates, 162—174 cylindrical coordinates, 175—180 magnet between plates, 165—169 spherical coordinates, 180—185 Bound charge, 75 Br, 38 Brillouin function, 19, 22 Brushless dc motors. See Axial-field motors C Can-stack motors, 437—446 Carrier particles, 315 Cartesian coordinates, 469—473 boundary-value problem, 162—174 Laplace’s equation, 162—174 transformations, 491—494 Ceramic magnets, 46—47 CGS/Gaussian unit, 2—5, 509—512 Chain rule, 345 Charge bound, 75 free, 75 Charge density, free, 75 Charge model, 131—134 force, 135—140 torque, 141—143 Circuits current sources, 144—153 equivalent, 157—159 inductance, 151 magnet sources, 153—161 moving, 358—361, 366—376 no flux leakage, 145, 150 stationary, 354—358, 363—366 Coenergy, 379 Coercive force, 31
513
514 Coercivity, 31, 38, 39 intrinsic, 29, 32 Conduction current, 75 Constitutive relations, 75—77 magnetostatic, 99 Contact latching, 268—269 Continuity equation, 75 Convection current, 75 Conversion factors, 3, 509—512 Coordinate transformations, 491—494 Coulomb gauge condition, 86, 103, 127 Coupling, synchronous applications, 286 axial, 286, 288, 297—306 radial, 286, 288—297 Russell-Saunders, 14—15 spin-orbit, 14 Cross product, 471 Curie’s law, 19 Curie temperature, 20 Curl, 472 Current conduction, 75 convection, 75 electrolytic, 75 equivalent, 128 free, 75 line, 104—106 sources, 144—153 surface, 107 Current density displacement, 75 upper limit of, 58 Current model, 126—131 force, 131 torque, 131 Cylindrical coordinates, 473—476 boundary-value problem, 175—180 Laplace’s equation, 175—180 transformations, 491—494 Cylindrical magnets axially polarized multipole disk magnet, 253—264 axially polarized shell, 245—253 axial polarization, 128—130 bipolar, 218—223 radially polarized multipole, 224—245 D Del operator, 343, 472 Demagnetization curve, 44, 66—67 for ceramic, 67 for samarium-cobalt, 67 Demagnetization field, 23—27 Diamagnetic material, 6 Die-upsetting, 53 Differential Maxwell equations, 74, 79, 338
INDEX Dipole-dipole force, 140, 316 Dipole/dipole moment, 3, 108 Energy in external field, 4, 511 magnetic, 140 of sphere in external field, 26 torque in external field, 4, 511 Direct current motors. See Axial-field motors Dirichlet boundary-value problem, 162 Disk magnets, levitation, 273—281 Dislocations, 34 Displacement current density, 75 Divergence, 472 Divergence theorem, 78, 114, 487—489 Domains, 6, 20, 33—36 Domain walls, 34 motion of, 35—36 Dot product, 470 Drag torque, 423 Drive torque, 453 Dyadic, 113 E Earnshaw’s theorem, 273 Eddy currents, 350—353, 411—412 drag force, 352, 353 lift force, 352, 353 Electrical constant, 368 Electrical equations moving coils, 358—361 stationary circuits, 354—358 Electrolytic current, 75 Electromagnets, conventional, 308, 310 Electromechanical devices actuators, axial-field, 393—401 actuators, linear, 413—421 actuators, magnetic circuit, 383—392 actuators, magnetooptical bias field, 406—413 actuators, resonant, 402—406 basics, 335—337 electrical equations, 354—361 electromechanical equations, 363—376 energy analysis, 376—383 examples of, 335 hybrid analytical-FEM analysis, 446—455 mechanical equations, 361—363 micro, 455—465 motors, axial-field, 421—437 motors, stepper, 437—446 quasi-static field theory, 337—353 Electromechanical equations moving coils, 366—376 stationary circuits, 363—366 Electron spin, 10 Electrophotography, 312—318 Electrostatic theory, 89—91
515
INDEX emu, 3, 511 Energy analysis, 376—383 balance equation, 377 co-, 379 contours, 46 density, 117 exchange, 22 magnetostatic, 23, 116—117 point of maximum, 44, 153 product, 157 self-energy equation, 23, 117 Equivalent charge, 132 Equivalent circuit, 157—159 Equivalent currents, 128 Euler’s method, 363, 365—366, 418, 498—503 Exchange energy, 22 Exchange integral, 22 Expectation value, 10 F Faraday’s law of electromagnetic induction, 78, 85, 347 FEA. See Finite element analysis FEL. See Free-electron lasers FEM analysis, hybrid analytical, 446—455 Ferric oxide, 47 Ferrimagnetic material, 6, 7 Ferrites, 46—47 bonded, 54 Ferromagnetic material, 6 Ferromagnetism, 19—22 Finite bar magnet, 211—217 Finite difference method, 200—204 backward, 202 central, 202—203 forward, 202 Finite element analysis (FEA), 60, 190—200 Finite radially polarized multipole cylinder, 235—245 Flux density, 2, 3, 98 line, 62, 108—109 linkage, 119—120, 341 magnetic, 144 remanent, 31 vector field, 78 Force between cylindrical magnets, 139—140 between dipoles, 140 between two wires, 111 charge model, 135—140 coercive, 31 current model, 131 on dipole in external field, 317 Lorentz, 82, 110 magnet above plate, 188, 189
on steady currents, 110—112 on wire above plate, 188 Free charge density, 75 Free current density, 75 Free-electron lasers (FELs), 325—331 Fringe fields, 309 G Galilean transform, 344 Gamma ferric oxide, 47 Gauss’ law, 78, 90, 338 Gears, 282—286 Gradient, 472 Green’s function, free-space, 90, 103, 495—496 H Hall effect, 265, 422 Hard magnetic materials, 44—45 alnico, 48—50 bonded, 53—55 ferrites, 46—47 neodymium-iron-boron, 51—53 samarium-cobalt, 50—51 Helmholtz’s theorem, 489—491 Henry, 118 High field structures, 265—268 Homogeneity, magnetic resonance imaging and, 309 Homogeneous material, 5, 77 Hund’s rules, 15 Hybrid analytical-FEM analysis, 446—455 Hysteresis loop, 36—39 I Images, method of. See Method of images Impulse magnetizers, 58—59 Inclusions, 34 Inductance, 341 external, 119 internal, 119 magnetostatic analysis, 118—126 mutual, 120—121 parallel wires, 122—123 self, 118 solenoid, 123—124, 125—126 toroid, 124—125 wire, 121—122 Induced voltage, 370, 372 Inertial reference frame, 341—342 Infinitely long cylindrical magnet, 224—235 Infinite rectangular magnet, 208—211 Inhomogeneous material, 5, 77 Initial magnetization curve, 37 Integral Maxwell equations, 77—79, 338 Integrals, of vector functions, 479—485 Intrinsic coercivity, 29 shape anisotropy and, 32
516 Iron oxides, 47 Iron-silicon alloys, 39, 43 Irreversible changes, 67, 70 Isotropic materials, 5, 77 K Kennelly units, 2, 509—512 Kerr effect, 320—321 Kirchhoff’s laws current, 146, 338 voltage, 89, 146—147, 339, 341 voltage (moving coil), 358—361 voltage (stationary circuit), 354, 357 L Land g factor, 14 Langevin function, 18 Laplace operator, 472—473 Laplace’s equation Cartesian coordinates, 162—174 cylindrical coordinates, 175—180 spherical coordinates, 180—185 Larmor frequency, 307, 308 Larmor relation, 309 Latching, magnetic, 268—273 Legendre polynomials, 182—185 Legendre’s equation, 181 Lenz’s law, 78 Levitation, passive, 273—281 Linear actuators, 413—421 Linear materials, 5, 77 Line integral, 479—482 Load line, 44—46, 155—156 stability, 66—70 Lodestone, 47 Lorentz force, 82, 110, 362 Lorentz gauge condition, 83 Lorentz transformation, 343—345 M Magic cylinder, 265, 266 Magic sphere, 265—268 Magnequench, 52—53 Magnetic bearing. See Magnetic levitation Magnetic brush technique, 315—318 Magnetic circuit actuators, 383—392 Magnetic circuit analysis, 144—161 current sources, 144—153 equivalent circuits, 157—159 magnet sources, 153—161 Magnetic coupling. See Coupling Magnetic dipole. See Dipole moment Magnetic field bar magnet, 211—217 bipolar sphere, 182—185 current loop, 108 cylindrical magnets. See Cylindrical magnets
INDEX line current, 104—107 magnet between plates, 165, 168 rectangular magnets. See Rectangular magnets solenoid, 100—101 strength, 2, 3, 98 structures of high, 265—268 surface current, 107 toroid, 101—102 wire, 99—100 Magnetic gears, 282—287 Magnetic latching, 268—273 Magnetic levitation, 273—281 Magnetic moments. See Atomic magnetic moments Magnetic quantum number, 10 Magnetic resonance imaging (MRI), 306—307 field strengths, 308 field systems for, 308, 310 fringe fields, 309 homogeneity, 309 hybrid, 310, 312 patient access, 309—310 stability, 309 yoked, 310, 311—312 yokeless, 310—311 Magnetic viscosity, 70 Magnetite, 47 Magnetization process, 55—66 Magnetizing field, 56 Magnetizing fixture design, 58—59 Magnetocrystalline anisotropy, 27—31 Magnetomotive force (mmf), 148 Magnetooptical bias field actuators, 406—413 Magnetooptical (M-O) recording, 318—325 Magnetostatic analysis Biot-Savart law, 104 energy, 116—117 field equations, 97—102 force and torque, 110—112 inductance, 118—126 Maxwell stress tensor, 112—116 vector potential, 102—109 Magnetostatic energy, 23 Magnetostatic theory, 87—89 Magnets equivalent circuit, 157—159 cylindrical. See Cylindrical magnets rectangular. See Rectangular magnets sources, 153—161 Magnets, structures of cylindrical, 218—264 rectangular, 208—218
517
INDEX Materials anisotropic, 5, 77 classification of, 6—7 hard, 44—55 homogeneous, 5, 77 inhomogeneous, 5, 77 isotropic, 5, 77 linear, 5, 77 nonlinear, 5, 77 soft, 39—44 Maxwell, James Clerk, 76 Maxwell’s equations, 74, 338 boundary conditions, 79—81 constitutive relations, 75—77 differential, 74, 79 force and torque, 82 integral, 77—79 Maxwell stress tensor, 112—116 force, 115—116, 362 Mechanical equations, 361—363 Melt spinning, 51, 52 MEMS (micro-electromechanical systems), 455—465 Method of images, 185—190 magnet on plate, 188—189 wire above plate, 188 Micro-electromechanical systems (MEMS), 455—465 Mixed boundary-value problem, 162 MKS system, 368, 509—512 Molecular field, 20 Moment of inertia, 361, 375 Motion, electromechanical equations moving coils, 366—376 stationary circuits, 363—366 Motion, mechanical equations, 361—363 linear, 361 rotational, 361 Motors axial-field, 421—437 stepper (can-stack), 437—446 Moving coils, 358—361, 366—376 Moving reference frames, 341—353 MRI. See Magnetic resonance imaging Multipolar cylinders, radially polarized, 235—245 Multipole disk magnet, axially polarized, 253—264 Mutual inductance, 120—121 N Neodymium-iron-boron, 51—53 Neumann boundary-value problem, 162 Neumann functions, 176 Newton’s second law of motion, 361 Nickel-iron alloys, 39, 43 Noncontact latching, 269
Nonlinear materials, 5, 77 O Ohm’s law, 356 Operating point, 153—161 Orbital angular momentum, 10—11 P Paramagnetic material, 6 Paramagnetism, 17—19 Pauli exclusion principle, 14, 15 Permanent magnet structures. See Magnets Permanent magnets, 44—46 alnico, 48—50 bonded, 53—55 ferrites, 46—47 fields of. See Magnetic field magnetic resonance imaging and, 308, 310 magnetization of. See Magnetization process neodymium-iron-boron, 51—53 samarium-cobalt, 50—51 Permeability absolute, 41 differential, 41 free space, 5, 76 initial, 41 maximum relative, 41 relative, 41 Permeance angular gap, 149 radial gap, 149 rectangular gap, 149 Permittivity, free space, 76 Planck’s constant, 10 Poisson’s equation, 200—201 Potential(s) difference, 90, 340 energy, 23 Maxwell’s equations and, 82—84 scalar, 82 vector, 82, 83, 102—109 Q Quadrupole magnet, 218 Quantum mechanics, 10 Quasi-static field theory, 85—86, 337 moving reference frames, 341—353 stationary reference frame, 338—341 R Radial coupling, 286, 288—297 Radially polarized magnets, 224—245 finite, 235—245 infinitely long, 224—235 Rapid quenching, 51, 52
518 Rare-earth magnets, 58 neodymium-iron-boron, 51—53 samarium-cobalt, 50—51 Rectangular magnets, 208—218 Reference frames moving, 341—353 stationary, 338—341 Relative permeability, 41 Reluctance, 145 torque, 453 Remanence, 31, 38 Remanent flux density, 31 Remanent induction, 38 Resonant actuators, 402—406 Reversible changes, 67 Reversible temperature coefficients, 67 Runge-Kutta method, 363, 418, 503—508 Russell-Saunders coupling, 14—15 S Samarium-cobalt, 50—51 Saturation induction, 38 Scalar-valued functions, 345, 469—473 Schrödinger’s equation, 9 Self-energy equation, 23, 117 Self-inductance, 118 Separation of variables Cartesian coordinates, 162—174 cylindrical coordinates, 175—180 spherical coordinates, 180—185 Shape anisotropy, 31—33 Shorted turn, 418—421 SI (Système international d’unités), 2—5, 509—512 Soft ferrites, 39, 43—44 Soft iron, 39, 42—43 Soft magnetic materials, 39—44 Solenoid B-field, 100—101 inductance, 123—124, 125—126 Sommerfeld units, 2—5, 509—512 Sphere in external field, 26 Spherical coordinates, 476—479 boundary-value problem, 180—185 Laplace’s equation, 180—185 transformations, 491—494 Spin, 10 Spin angular momentum, 10—11 Spin-orbit coupling, 14 Stability, 66—70 magnetic resonance imaging and, 309 Static theory electrostatic, 89—91 magnetostatic, 87—89 Stationary circuits, 354—358, 363—366 Stationary reference frame, 338—341 Stepper motors, 437—446
INDEX Stokes’ theorem, 77, 356, 485—487 Stoner and Wohlfarth model, 28 Strontium ferrite, 46 Structural changes, 67 Superconducting coils, 308, 310 Surface current, 107 Surface integral, 483—485 Susceptibilities absolute, 42 differential, 42 initial, 42 Suspension, 273—281 Synchronous coupling. See Coupling, synchronous T Tesla, 2—3, 510—511 Tesla per amp (TPA) ratio, 60 Toroid B-field, 101—102 inductance, 124—125 Torque, 82, 374 on bipolar cylinder, 142—143 charge model, 141—143 constant, 368 current model, 131 drag, 423 drive, 453 reluctance, 453 on steady currents, 110—112 U Undulator, 326 Units CGS/Gaussian, 2—5 conversion factors, 3 Kennelly, 2 SI (Sommerfeld), 2—5 summary of, 509—512 V Vector field flux, 78 Vector potential, 83, 102—109 line current, 104—106 surface current, 107—108 Vector-valued functions, 74, 345—346, 469— 473 Virtual displacement, 376—377 Volume integral, 483 W Webers, 2 Weiss molecular/mean field theory, 20—22 Wigglers, 326—328 Halbach, 329—330 helical field, 330 hybrid, 330—331