You are on page 1of 634
Acknowledgments In preparing this book, I received contributions, suggestions, encourage- ment, and help from many friends and colleagues. Tam indebted to all of them. I am especially grateful to Giorgio Bertotti, who assisted me in many ways, willingly engaging in many clarifying discussions, and allowing me to benefit from his deep knowledge of electromagnetism and magnetic phenomena. The series editor, Isaak Mayergoyz, fully sup- ported my effort, fostering my confidence in the project and generously handling my outrageously delayed delivery of the manuscript. I would like to acknowledge that this project could only be pursued thanks to the special cooperating milicu, the broad expertise on the physics of magnetic materials, and the array of experimental researches developed by fellow scientists at the Materials Department of IEN. I found advice and support in all of them. I am also indebted to elder scientists in my lab who educated me in the early years of my career. The late Andrea Ferro Milone introduced me to materials science and Piero Mazzetti taught me the basic virtues of the experimental physicist. Aldo Stantero helped me in many ways and on innumerable occasions for more than twenty years. His untimely death was an untold loss to me and to the lab at IEN. Giorgio Bertotti, Vittorio Basso, Carlo Appino, and Alessandro Magni read substantial parts of the manuscript and Oriano Bottauscio provided me with crucial help by expressly performing numerous electromagnetic field computations. The field maps presented in Chapters 4 and 8 are due to him. Vittorio Basso, Cinzia Beatrice, Enzo Ferrara, and Eros Patroi kindly supplied me with their own experimental data and Marco Coisson clarified to me specific aspects of magnetoimpedance measurements. Anna Maria Rietto carried out careful experiments to elucidate a few important details in the magnetic lamination testing with the Epstein test frame and Luciano Rocchino assisted me in assessing the problems related to reference field sources and their traceability to the base units. Sigfrido Leschiutta enhanced my sensitivity to metrological issues and the role of metrology in the physical sciences. I also need to thank all the many colleagues in Europe and elsewhere with whom I shared cooperative research activity and discussions on various scientific xvii xviii Acknowledgments matters. It is finally a pleasure to acknowledge the help provided by Christopher Greenwell and Sharon Brown at Elsevier, who assisted me in the various stages of the book production, and Lucia Bailo, Francesca Fia, and Emanuela Secinaro at the Publication Department of IEN, for their help in literature retrieval. APPENDIX A The SI and the CGS Unit Systems in Magnetism Magnetic measurements are generally traceable to the SI unit system. This means that reference can be made, for any measured quantity, to the standards of the base units maintained by the National Metrological Institutes, a feat accomplished by connecting the primary laboratory with the end user through a chain of comparisons, having stated uncertainties. SI (Sisteme International d’Unités) has been adopted and updated through successive CGPM (Conférence Générale des Poids et Mesures) resolutions, starting with the task assigned by the 9th CGPM in 1948 to the CIPM (Comité International des Poids et Mesures) “to study the establishment of a complete set of rules for units of measurement and to make recommendations on the establishment of a practical system of units of measurement suitable for adoption by all signatories of the Convention du Métre” [A.1]. SI is an absolute system, which means that the corresponding base units are invariant with time and space and all other units can be derived from them through definite mathematical relationships. In particular, a unit {y} of a quantity y is given in this system in terms of the base units {x}, (x2), ..., (x,}, by an expression of the type: (y) = tf) tah), AD with a, B,...,v either positive or negative integers and no numerical factors other than 1. The SI units are then said to form a coherent set of units. SI is based on seven base units: meter (m) for the length, second (s) for the time, kilogram (kg) for the mass, ampere (A) for the electric current, kelvin (K) for the temperature, candela (cd) for the luminous intensity, and the mole (mol) for the amount of substance. The base units relevant for electromagnetic phenomena are m, s, kg, and A. The “vexata quaestio” of unit systems in electromagnetism is rooted in the development of the CGS system in the 19th century, which accompanied over several decades the development of the electrical sciences and their broadening impact on industry and society [A.2]. 613 614 APPENDIX A The SI and the CGS Unit Systems in Magnetism With CGS all electrical and magnetic units are expressed in terms of three base mechanical units: cm, g, and s. Two different unit systems, however, exist, according to whether electrostatic or magnetostatic base equations are used. The electrostatic units (e.s.u.) are defined starting from Coulomb's law, where the interaction force F between two electrical charges (Q;,Qs) at a distance r is expressed by the equation: = pRB FokxS (A2) By taking the adimensional proportionality constant K = 1, the dimen- sions of the electric charge are [M'/7L°/?T~] and the corresponding unit is 1 statcoulomb = 1 dyn'/? cm. With another choice of the primary equation, where electrodynamic forces are described, the electromagnetic units (e.m.u.) are defined. The force per unit length F/L mutually exerted by two infinitely long straight wires carrying the currents i, and i, is F pio ithe ra K oo (A3) where r is the distance between the wires. With the constant k’ = 1, the dimensions of the current are [M'/?L'/?T~] and those of the electric charge, which is the time integral of the current, are [M‘/?L'/?]. It is therefore apparent that the electric charge has different dimensions in the es.u. and the e.m.u. systems, the ratio between them having the dimension of a velocity. This duality is a drawback and, to compound it, the mixed CGS (or Gaussian) system has been introduced, where the e.s.u. units are used for the electric quantities and the e.m.u. units are used for the magnetic quantities. The Gaussian system has been instrumental in the development of the electrical sciences; it has been largely applied in the past literature, and is still applied to some extent in papers and textbooks on magnetism and magnetic materials. But most of its units have inconvenient size with respect to practical engineering needs and objective complications arise from the multiplicity of notations. Conse- quently, since their official adoption by the British Association for the Advancement of Science in 1873, the absolute Gaussian units have been used in association with practical electrical units, such as the volt, the ampere, and the ohm. In 1901, G. Giorgi, by remarking that the watt could equally be the absolute unit of mechanical and electric power, the latter being the product of voltage and current, proposed a system where a fourth independent electrical quantity was associated with the base mechanical units m, kg, and s [A.3]. By abandoning a purely mechanical system, dimensional factors had evidently to be introduced in the laws of force between charges and between currents, but now all the practical APPENDIX A The SI and the CGS Unit Systems in Magnetism 615 electrical units slipped coherently into this system. Giorgi’s proposal won immediate recognition, but it took several decades for general and official acceptance to be achieved, following thorough discussions by Interna- tional Union of Applied and Pure Physics (IUPAP), International Electro- technical Commission (IEC), and the Consultative Committee for Electricity (CCE) of CIPM. In 1946, CIPM approved the adoption of the meter, kilogram, second, ampere system (MKSA) [A.4]. The ampere was thereby defined as “that constant current which, if maintained in two straight parallel conductors of infinite length, of negligible circular cross- section, and placed 1 m apart in vacuum, would produce between these conductors a force equal to 2x 10-7 N of force per meter of length”. MKSA was eventually sanctioned, with the addition of the K and the ed as base units, by the 10th CGPM in 1954 as practical system of units of measurement [A.5]. According to this definition of the ampere, the proportionality constant in Eq. (A.3) has definite dimensions and the value 2 x 10-7 F Posie EOS aa AA Lage 2 The magnetic constant ji (formerly called “magnetic permeability of free space”) has therefore an exact value by definition jg = 4x 10-” N/A?. The unit of electric charge is now 1C=1As and the constant K in Coulomb’s law (A.2) is no more adimensional = _1 Qi tre The electric constant ¢9 is related, according to the solution of Maxwell's equations for the propagation of the electromagnetic field in vacuum, to Ho and the speed of light c: (A.5) = (9p). (A.6) In 1983, the 17th CGPM re-defined the meter, assigning the speed of light the precise value c = 299 792 458 m/s [A.6]. In force of this decision, the electric constant € takes, according to Eq. (A.6), the exact value £9 = 8.854187817 x 10"? F/m. The SI unit system is strongly recommended by the international metrological and standardization organizations and is increasingly applied in magnetism. However, a massive body of literature exists which makes use of CGS and a substantial resilience to SI, dictated either by attachment to routinary use of old units or by sheer distaste of the awkward redundancy of fields in free space generated by SI, is observed in some areas (for example, in the field of permanent magnets). ABLE AL ‘ST units (my kg, 5, A) _ ‘Gaussian units (cm, g 8) Lorenizs equation — ~~ B= qE-+qvxB (N) Maxwell's equations in fre space | vane, Bem dmn OBO vxe=-228 ype i+ aE B, VXB= pa) temas ‘Maxwells equations in the presence of materials WD=p, YKE Magnetic fx density generated by a straight infinitely long wire cnrrying a current i ata distance r 2ixe 2ixr = plete = 2 AF Ginesm), B= GS (in Bo= Bo ze ), Be) = 35a Hin A) @ [Magnetic flac density in the center ofa plane circular coil of rats R carrying a current i =m xR BERG iyoguy, B= AR Gin Bo = BO Boo) = FAR ),BO)= Far (fin AD) Magnetic la density onthe ans of sleoid with NAL turns pr wit length caring a caren _ Ni dei ar Ni 5 a= wl $2 Ginesu), B= 7% Gin ayG) Magnetic field in fee space H= B/uy A/mm) H=B (Ox) 919 ‘wsnauieyy ut swarsis un SOD aM PUES aU, V XIGNaddv Ferri lng ete aed igh candcters sance dcarying he cred Fai F 2G andigineauy Fm 25 (and nA) ye Loma N/m L (i and iy a), joa id iz in A} (dyn/cm) uae ome of ea cl of wea Aa ac of wise m = iAn (Am) i i m= lanGinesu).m= ‘yin in 8} g/08 “ 1 oom, (erg/G) Marti deny ater Brat nM sat oT batt) Domain fd fa phe Hg = —NeM = — 3M (A/mm) Density of ragnetic energy in fice spare Bim B us = my 5q rele) Energy lass per ce per unit volume ina magnetic material Hd3 d/m’) w s $HaB (erg/em") F force; E,slectic field; D, electric displacement vector; B, magnetic ux density; m, magnetic moment; M, magnetic moment per unit velume; H, magnetic field i electric current; j, electric current density; J, magnetic polarization; f, time; g, electric Charge; 2, velocity; speed of lightin fee space; p, volume density of electric charge; electric constant Ho, megnetic constant. usnauteyy ur swoysig mun $99 OWN PUE IS >HL_Y XIN a9 TABLE A2 (Quantity ‘Symbol Sl units Gaussian units Conversion SI~Gaussian L m om M kg 8 T s 5 F N w J w Im? P w ° c v v LV = 3.335641 x 10 statvolt E NC‘ Vm = 3335641 x 10° statvolt cm” i A 1A =2.997925 x 10° statempere m Am JT? 1 Am? = 10° ema © Wo, Vs 1 Wo = 10° maxwell B T IT=10G H Am? LA m? = 1.2566 x 107700 M Am? ergGtem“,emucm™ Am? =10emuem™ Magnetic polarization J T = : Magnetic sasceptibility x 5 : tsi = 4m s19 ‘usspauepy Ur swaysts Yun SOO aM PvE IS OUI Vv XICNBdd¥ APPENDIX A The SI and the CGS Unit Systems in Magnetism. 619 It is therefore required in many instances to pass from one system to another. Tables permitting to do so are found in textbooks, but it seems nevertheless appropriate to summarize here the chief magnetic equations under the Gaussian and SI systems, to suggest some simple translation rules and provide a conversion table for the units of interest in magnetism (Tables A.1 and A.2). Equations in SI can be translated into the corresponding equations in the Gaussian system (and vice versa) by means of a few simple rules [A.7], employing relationships like the following ones: st Pa EG He Dg Ik. PS = S = ige, = TS = iy, 2 is Pg Est He Mo Ds Bo _ Ms _ [4a By = Mg Ho" (A) where the meaning of the symbols is understood from Table A.1. To make, for example, the conversion of an SI equation containing the variables Xs1, ¥s1; «-- Zs, We solve Eq. (A.7) for them, which become then expressed in terms of the corresponding Gaussian variables XG, yg, .-.,2¢ and the magnetic or electric constants. By substituting and making additional use of the relationship e419 = 1/c*, the equivalent Gaussian equation is obtained. An identical procedure in reverse leads to the translation of the Gaussian equations into the SI ones. References A.1. 9th CGPM, “Resolution 6: proposal for establishing a practical system of units of measurement,” Comptes Rendus des Séances de la Conférence Générale des Poids et Mesures, (1948), 64. A2. JJ. Roche, The Mathematics of Measurement (London: The Athlone Press, 1998), 163. A3. G. Giorgi, “Rational units of electromagnetism (in Italian),”” Atti dell’Asso- ciazione Elettrotecnica Italiana, 5 (1901), 402~418. A. CIPM, “Resolution 2: definitions of electric units,” Procts-Verbaux des Séances du Comité International des Poids et Mesures, 20 (1946), 129-137. AS. 10th CGPM, “Resolution 6: practical system of units,” Comptes Rendus des Séances de la Conférence Générale des Poids et Mesures, (1954), 80. A6. 17th CGPM, “Resolution 2: recommended value for the speed of light,” Comptes Rendus des Séances de la Conférence Générale des Poids et Mesures, (1983), 103. AJ. AS. Arrott, “Magnetism in SI units and Gaussian units,” in Ultrathin Magnetic Films (B. Heinrich and J.A.C. Bland, eds., Berlin: Springer, 1994), 7-19. APPENDIX B Physical Constants The values of the physical constants presented here are those recom- mended by Committee on Data for Science and Technology (CODATA), an organization of the International Council of Scientific Unions, which has the duty of evaluating, storing and retrieving data produced in science and technology in a coordinated and critical fashion. The funda- mental physical constants play a fundamental role in science and measurements today, since they are increasingly used in defining and maintaining the SI units. The constants presented here have relevance in magnetism. For a complete set of data, see Ref. [B.1]. Quantity Symbol Value Unit Speed of light in vacuum c 2.99792458 x 10° ms! Magnetic constant Ho 40x 107 NA” Electric constant 0 8.8541878 x10"? Em} Elementary charge e 16021765 x10" = C Avogadro constant Na 9.274009 x 10774 mol"? Boltzmann constant k 1.3806503 x 10° JK“? Planck constant h 6.6261x10% = Js Magnetic flux quantum, h/2e Dy 2.0678336 x 10715 Wb Bohr magneton, efi/2m, Be 9.274009 x 107-24 ites Nuclear magneton, eh/2m,, BN 5.0507832 x 10727 es Electron magnetic moment Me ~9.2847636 x 10-* = JT Electron g-factor Be ~2.0023193 ~ Electron gyromagnetic ratio, 2te/h Ye 1.760859794 x 10 ave Proton magnetic moment Mp 14106066 x 10776 iis Proton gytomagnetic ratio, 211p/f — p 2.675221 x 10° Proton g-factor 8p 5.585694675 - Neutron magnetic moment Bn — 0.9662364 x 10776 tes Neutron gyromagnetic ratio, 2jin/f Yq 18324719 x 10° teas e, electron charge; me, electron mass; imp, proton mass. 621 622 APPENDIX B Physical Constants Reference B.l. PJ. Mohr and BN. Taylor, “CODATA recommended values of the fundamental physical constants: 1998,” J. Phys. Chem. Ref. Data, 28 (1999), 1713-1852. Evaluation of Measuring Uncertainty C.1 TYPE B METHOD OF EVALUATION OF THE UNCERTAINTY 1. A standard resistor in a laboratory is associated with a calibration certificate declaring a value R = 100.0039 0 + 0.0011 ©, with the expanded uncertainty (see Section 10.3) U(R) = ku(R) = 0.0011 0 defined with a confidence level of 95%. The coverage factor k is, for the given confidence level, k = 1.96. The resistor can then be assigned the absolute and relative standard uncertainties u(R) = 0.00056 0 and u(R)/R = 5.6 x 10-6. . A digital voltmeter is used in the 10V range, where the specifications of the manufacturer provide, at a distance of 1 year after the last calibration, an accuracy of 35ppm of reading +7 ppm of range at a temperature of 23+5°C, following a 1h warm-up. By making a series of repeated readings under the same conditions, we determine an average value V = 7.558244 V. It is assumed that the accuracy declaration by the instrument maker identifies an interval +a of equally likely values centered on V. It can alternatively be stated that the correction for the bias AV = 0, being it equally likely to lie in the interval +a. We obtain a = 7.558244 35x 10-6 + 10x7x 1076 = 335 x 10-6 V. From Eq. (10.7), the Type B uncertainty turns out to be u(AV) = a/V3 = 193x 10° V and u(AV)/V = 2.55x 10. C.2 EVALUATION OF COMBINED UNCERTAINTY C.2.1 Example 1 The resistance R= V/I of a component is determined by means of simultaneous voltage and current measurements. Five repeated and simultaneous readings of Vand I are made, as summarized in Table C.1. 623 624 APPENDIX C_ Evaluation of Measuring Uncertainty We focus our attention here on the Type A contributions, assumed in this specific case dominant with respect to the Type B contributions. Two approaches are possible. In the first one (Mode 1), the best estimate of the output quantity is defined as the ratio of the best estimates of the input quantities R= V/I. In the second one (Mode 2), it is defined as R=(V/D, the average of the ratios of the ordered couples (V,1). Following Mode 1, the expression for the combined variance of R is, according to Eq. (10.12), | : 2p) = aed +(&\ zm —of 2\EV wep wR) = /D2(0) + (z) #0 7\(z “VD, Ca) where the covariance u(V/,I) is obtained by application of Eq. (10.19). Following Mode 2, the same quantity is obtained in the ordinary way, according to Eqs. (10.4) and (10.5): (V/D® — VIP Pe eee ~ wR) =u w= CEE) (C2) TABLE C1 i vv) I (mA) Vv/TQ) 1 4.15558 30.8742 134.605 2 4.14524 30.8332 134.441 3 4.15415 30.8348 134.723 4 4.14001 30.9132 133.924 5 4.14918, 30,8944 134.302 Best estimate 7 = 4.14888 1 = 30.8700 (V/1) = 134.399 (arithmetic mean) Experimental 7(7) = 8.40 w(t) = 2.54 w(V/1 = 0.0192. 0? variance «10-6 v2 107" Az Standard u(V) =2.90x103 Vu) =159K 105A W(V/I) = 0.1390 uncertainty i Covariance (7, ]) = -1.99x 10 VA, 1(V,F) = —0.43 correlation coefficient Mode 1 R= 134.399, u2(R) = 0.0192 0?, u_(R) = 0.139 0, u(R)/R = 1.03 x 10> Mode 2 R= 134.399, u2(R) = 0.0192 07, u-(R) = 0.139 0, u_(R)/R = 1.03 x 1073 C2 EVALUATION OF COMBINED UNCERTAINTY 625 It is observed that the two approaches generate the same values for the best estimate of the output quantity and its uncertainty, provided the correlation of the input quantities is considered in Mode 1. C.2.2 Example 2 A resistor is calibrated by making a stable DC current to circulate in a series formed by the resistor under test and a standard resistor and by measuring the voltages across these resistors by means of two high precision calibrated digital voltmeters. Let us indicate with V and V, the voltages developed across the unknown resistor and the standard resistor, respectively. R and R, denote their resistance values. R is deter- mined by comparing the two voltages and is given by the expression: ae R= GR (C3) For the employed voltmeter scales, the 1-year manufacturer's specifica- tions provide at a temperature of 23 + 2°C an accuracy of 12 ppm of reading + 2 ppm of range. The calibration certificate of the standard resistor provides a value R, = 100.0145, an uncertainty of 30 ppm (20 level) at 23 °C and a thermal drift of 5 ppm/K. Vand V, are subjected to 10 repeated simultaneous readings. The arithmetic mean of the ratio V/V, and the associated experimental variance (Type A evaluation) are calculated. For a specific measurement, we obtain, in accordance with Eqs. (10.3)-(10.6), (V/V,) = 1.499939 and ua(V/V,) = 21 x 10° (n = 10). We consequently obtain R = 150.0156 and the uncertainty ua(R) = R,ua(V/V,) = 2.11073 0. In order to obtain the best estimate R of the unknown resistance value, we express it in terms of the significant input quantities: R= g(V/V,),8R1, Ro, Bs, Ry) = (V/V DR, + 8R; + BRp +.8Rs + 8Ry. (C4) In this equation, the corrections 8R;, 8R:, 8R; and 8R, for bias compensation have been added to the estimate generated by the repeated readings (Type A evaluation). 8R; and 8R, are the corrections accounting for the bias in Vand V, readings, respectively. 8R3 is the bias correction associated with the calibration of the standard resistor, whose thermal drift is compensated by 8Ry. Although the measurement might be prepared in such a way that these corrections are zero, the associated uncertainties are not and they must be considered in the calculation (Type B evaluation) of the combined uncertainty u,(R). This can be estimated 626 APPENDIX C_ Evaluation of Measuring Uncertainty by means of Eq. (10.16), since it is fair to assume that all the input quantities are uncorrelated. We therefore write the expression: UR) = WA(R) + (Ry / VP ub (V) + (RV / VP ub Va) + (0/0 X (uB3(Re) + WB 4(Rs)), (C5) where the Type B variances up;, ..., 4p are associated with the corrections 8Ry. In Eq. (C.5) we have assumed, according to Example 1, V/V. The relative combined uncertainty is consequently (C.6) TABLE C2 Source of Distribution Relative Sensitivity Degrees of uncertainty function uncertainty coefficient freedom Digital Rectangular ysi() 1 oo voltmeter (V) “y= 15x10 Digital Rectangular yy,(7) od °° voltmeter (V,) ay = 8x 10% Standard Normal upa(Rs) _ 1510-6 ul oo resistor R (calibration) (Rs) i © Standard Rectangular —ug,(R,) =15x10- resistor Ry (temperature drift) (R,) : 5 Repeatability Normal ua) : ®) “ae = 14x 10 Combined Normal mr) a Yay = 124 relative “EO 77x10 uncertainty Expanded = u® . 7 uncertainty ae C.2 EVALUATION OF COMBINED UNCERTAINTY 627 The previous specifications are used to estimate the terms deriving froma Type B evaluation. The declared accuracy interval of the employed voltmeter is taken as the semi-amplitude a of a rectangular distribution. We obtain a= 38x10~° V for the voltage reading on the unknown resistor (1.5 V read on the 10 V range) and a = 14x 10~° V for the voltage reading on the standard resistor (1V read on the 1V range). The corresponding uncertainties (u(%) = a/V3) are up;(V) = 22x 10° V and Up2(V,) = 8X 10”° V. The uncertainty provided with the standard resistor calibration certificate is associated with a normal distribution, at a 20 level, and is up3(R,) ~ 15x 107 x 100 = 15x 10~* ©. With an allowed fluctuation of +0.5 °C of the standard resistor temperature, the distribu- tion semi-amplitude is a ~ 2.5x 107° x 100 = 2.5x 1074 0 and u4(R,) ~ 1.5107. One can now apply Eq. (C.6) and calculate the relative combined uncertainty. Table C.2 summarizes the whole procedure. The measuring test report will provide the expanded uncertainty U=ku., with the value of the constant k set according to a defined confidence level. The result of the measurement is therefore declared here as R= R+U= 150.0156 +5.4x107°Q. In the present case, a 95% confidence level is assumed, for which k ~ 2. For a discussion on the expanded uncertainty and the degrees of freedom, see Section 10.3. APPENDIX D Specifications of Magnetometers We summarize here the main specifications of commercially available magnetic field measuring devices. These devices are broadly classified according to their working principle and are assigned typical measuring field ranges and uncertainties. The provided figures refer to the sensing system and can be affected to some extent by electronic circuitry and software. Magnetomer type— Field range Frequency Relative working principle range uncertainty Rotating/vibrating coil— 1 nT-10T pe 1073-107? fluxmetric method Search coil—fluxmetric 0.1nT-10-*T 1 Hz~0.5MHz 1073-107? method Hall effect in semi- 10eT-10T = DC—10kHz 107-10" (DE) conductors -? (AC) AMR and GMR 1pT-100mT = DC—1MHz Fluxgate magnetometers 1nT-1 mT. DC—1 kHz Thin-film inductive 100 nT-100 wT DC—10 kHz magnetometers Magnetostriction 1nT-100nT DC—100Hz 10°? Continuous wave NMR 40mT-20T = DC 2x 10-105 magnetometers Free proton precession 20 uT-100 pT = DC 210ee Flowing-water NMR SpT-2T Dc 105 magnetometes Optical pumping 10T-1004T DC 10se SQUID 10pT-ImT 5x10°T/s 1073 629 CHAPTER 1 Basic Phenomenology in Magnetic Materials 1.1 MAGNETIZED MEDIA In September 1820 H. C. Oersted demonstrated that electrical currents and magnets displayed equivalent effects. In a matter of weeks, A. M. Ampére, elaborating on Oersted’s discovery and making his own experiments, boldly interpreted the magnetism of materials as electricity in motion, ie. the result of hidden microscopic currents, circulating around “electrodynamic molecules”. The pedestal of electromagnetism was built in those few weeks, to be crowned in less than 50 years by the towering achievement of Maxwell's equations. Nowadays, we know that these currents exist, but they are quantum-mechanical in nature. They naturally slip into the classical Maxwellian scenario through the concept of permanent magnetic moment and the useful intermixing of classical and quantum concepts in the description of their relationship with the electronic angular momenta. A material sample is fundamentally described, from the viewpoint of magnetic properties, as a collection of magnetic moments, resulting from the motion of the electrons. Classically, orbiting electrons generate microscopic currents and are endowed witha magnetic moment m = —(¢/2m,)-L, ife and m, are charge and mass of the electron, and L is the angular moment. Quantum mechanics makes the view of electronic magnetic moment physically consistent, besides providing the additional basic concept of magnetic moment associated with spin angular momentum. When writing the classical equations of electromagnetism and assigning a meaning to the value of the physical quantities involved, we look at the material as a continuum. This means that all atomic scale intricacies are lost. In particular, the internal currents of quantum- mechanical origin are retained as averages over elementary volumes AV, sufficiently small to be defined as local over the typical scale of the problem, but large enough with respect to the atomic scale. These currents 3 4 CHAPTER 1 Basic Phenomenology in Magnetic Materials (Amperian currents), resulting from electron trajectories at the atomic scale, do not convey any flow of charge across the body. Let us call juy(r) the associated current density. Because of its solenoidal character, jy(r) can be expressed as the curl of another vector function M(t) jm@) = VX M(x). a. Remarkably, it can be demonstrated that the vector function M(r) represents the magnetic moment per unit sample volume [1.1]. This quantity takes the name “magnetization”. If the previous elementary volume contains a certain number of moment carriers, it is M(r) = 2im;/AV, where the summation runs over the moments contained in such avolume. Thanks to Eq. (1.1), we are in a position to describe the magnetic effects ensuing from steady external currents when media are involved. If such currents are made to circulate in the absence of media, the induction vector B (often called “B-field”) is given by the Biot-Savart law, which is expressed in differential form as VXB= aj, a2) where jc is the density of the supplied currents and pg = 477 10-7 N/A? is the magnetic constant (sometimes called permeability of vacuum). Analysis of the Biot-Savart law additionally shows that the B vector is solenoidal, i.e. it obeys the equation V-B = 0. It can be shown that this equation and Eq. (1.2) determine B uniquely for given j.. We recall here that the operative definition of the vector B is provided by Lorentz’s law, which describes the coupling of electrical and magnetic fields with electrical charges. It states that a charge q moving at velocity i is subjected to a force F=q(E+ixB). (1.3) In the presence of magnetic media, both the externally supplied currents and the internal microscopic averaged currents will contribute to the B-field. The two base equations will be VB=0, VXxXB=polje+iw)- (4) When dealing with experiments on magnetic materials and their applica- tions, we endeavor to drive the magnetic state of the material by means of external currents, ie. acting on the quantity j.. We can single out j. in Eq. (1.4), introducing through Eq. (1.1) the magnetization M, a quantity directly accessible to experiments, in place of the awkward internal currents ju. In this way we define the so-called H-field vx (5 -M)=VxH=je as) Ho 1.1 MAGNETIZED MEDIA 5 H is the quantity conventionally defined as the magnetic field as it is the quantity susceptible to direct control by means of the external currents. On the other hand, B appears to be the fundamental field vector because it is characterized by the condition V-B = 0 everywhere, in the free space and inside the matter, and Lorentz’s equation everywhere applies to it. According to Eq. (1.5), the general relationship connecting the vectors B, H, and M is B = gH + oM. (1.6) In the SI unit system the magnetization M is expressed, like the magnetic field, in A/m, putting in evidence the Amperian origin of the magnetic moment. In the absence of media, M = 0 and B = poH, i.e. magnetic field and induction (i.e. H-field and B-field) are equivalent quantities, as they are related by the proportionality constant j1. In many kinds of experiments, we exploit the Faraday-Maxwell law VXE= —dB/dt, where E is the electric field, in order to determine the magnetic behavior of the material. We detect in this case the electromotive force generated in a linked search coil by the time variation of the induction. Normally, we wish to get rid of the term yigH in Eq. (1.6), because we are only interested in the contribution pM deriving from the material. This contribution is called magnetic polarization, J = gM, a quantity having the same dimensions as B (tesla, T) and the same properties as M. We then write B = 4H + J. This general relationship will be specialized to the magnetic properties of the investigated medium by means of some constitutive equation J(H) or B(H). In ferromagnetic materials, these relationships can be very complex and very difficult to predict. In some well-defined instances, it is meaningful to define the relationships B = WH = y1,p19H and M = yH, where pis the permeability, Mt; is the relative permeability, and y is the susceptibility. The quantities j1, and x are related by the equation. Br =1ty. a) Note that, in the old Gaussian system, the base Eq. (1.6) is written as B=H+4cM, (1.8) ie. field, induction, and magnetization have all the same dimensions (though different names, oersted (Oe) for H and gauss (G) for 47M and B). In this book, the SI system will be used throughout and little reference will be made to the increasingly obsolete Gaussian units. A complete set of conversion formulae is nevertheless provided, together with a discussion on their logical foundation, in Appendix A. 6 CHAPTER 1 Basic Phenomenology in Magnetic Materials Ferromagnetic materials are characterized by hysteresis. A residual magnetization M is always left when the external field, associated with the presence of a current density j.,, is completely released having once attained a certain peak value. We can say that some internal mechanism, associated with the nature and structure of the material, preserves a non-zero curl of the microscopic average current jy (Eq. (1.1)). With j. = 0, we obtain from the previous Eqs. (1.4) and (1.5), that induction and H-field are described by the equations VB=0, VXB=poim (9) V-H = —V-M, VxH=0. (1.10) Equation (1.9) simply states that the B-field is now uniquely generated by the internal currents and it preserves its solenoidal character. The two equations (Eq. (1.10)) are instead formally equivalent to the equations for the electrostatic field V-E = p/e) and Vx E= 0, with p the electric charge density and é5 the electric constant. Thus, in the absence of external currents, the field H, whose divergence can by analogy be written as V-H = py, with py = —V-M, can be considered as the gradient of a scalar magnetic potential H = —V,. This potential satisfies the Poisson's equation V’@y = —py. The electrostatic analogy then permits us to introduce, in a purely fictitious way, magnetic charges of volume density Pm acting as sources of the field H, whenever it occurs that V-M #0. Although devoid of physical reality, the concept of magnetic charges is constantly applied in the investigation of magnetic materials and in magnetic measurements because of the simplifications it introduces in the description of many phenomena and in the calculations. It permits one, for example, to derive fields from scalar potential functions, which are solutions of Poisson’s equation __ 1 (MG) 5 Py (x) = x] — dy, (ab thereby applying the conventional methods of electrostatics. Figure 1.1 provides a classical example where the role of magnetic charges can be invoked. It is the case of a cylindrical permanent magnet, where the magnetization M is uniform and axially directed. Since M suffers a discontinuity at the sample ends, the conditions are created for quasi- singular behavior of the divergence V-M. It turns out that the potential function can be written as - 1 ove) 2 Oy) = zl, Fovit 1, (1.12) 1.1 MAGNETIZED MEDIA 7 (@) (b) ©) FIGURE 1.1 Induction B and field H in a cylindrical permanent magnet in the absence of an external applied field (j, = 0). It is assumed that the sample remanent magnetization M is uniform. The induction B= gH + joM is solenoidal (V-B =0) and the field H satisfies the condition VxH = 0. This means that H can be expressed, in formal analogy with the electrostatic field, as the gradient of a scalar potential. In this respect, it is as if fictitious magnetic charges of equal densities and opposite signs were uniformly distributed over the top and bottom surfaces of the cylinder. where the integration is performed over the total area A of the top and bottom surfaces. The quantity oy(r) = n-M(r), where n is the unit vector normal to these surfaces, plays the role of surface magnetic charge density. The correspondingly calculated magnetostatic field H(r) and the induction B(x) = soH(r) + gM are schematically shown in Fig. 1.1b,c. Notice that within the sample, H(x) is directed in such a way as to oppose the magnetization. It is for this reason called a “demagnetizing field”. If the magnetization is not uniform or the material is inhomogeneous, internal demagnetizing fields can also arise. In the free space, B(x) and H(x) (which takes the name stray field) coincide (but for the proportionality factor jig). Note further that the condition V-B = 0 implies that, on traversing the sample surface, the normal component B-n is preserved. The condition VxH =0, however, implies that the same occurs to the tangential component of H. 8 CHAPTER 1 Basic Phenomenology in Magnetic Materials ‘A magnet brought under the permanent condition shown in Fig. 1.1 is endowed with a certain magnetostatic energy content. Part of this energy is contained within the sample and part is associated with the stray field. Under very general terms, we can write the total energy as E= 5H fav, (1.13) where we have defined as H, the demagnetizing field and the integration extends all over the space. This is the energy that must be spent for the formation of the magnetic charges and it can be equivalently written as Ep=— 5H | HeM dv, (1.14) 2 v where integration is made over the sample volume. 1.2 DEMAGNETIZING FIELDS Demagnetizing effects are ubiquitous. Even in accurately closed speci- mens, (e.g. ring samples), one cannot get rid of them completely. This has fundamental consequences from the point of view of magnetic character- ization and it requires measuring strategies aimed at minimizing and/or precisely controlling the demagnetizing fields. We shall discuss and clarify practical methods devised to this purpose, both in soft and hard magnets, in later chapters. In this section, we shall briefly discuss the basic problems connected with the prediction of the demagnetizing fields under different sample geometries. Calculations of demagnetizing fields date back to the 19th century. They were pursued by, among others, Maxwell [1.2], Lord Rayleigh [1.3], and Ewing [14]. One chief problem at that time involved a ship’s magnetism and the correction to be made on the apparent declination of the magnetic compass to determine a ship's position. It was recognized that only in samples shaped as ellipsoids (or spheres) could the demagnetizing field be homogeneous and susceptible of full mathema- tical treatment. The general approach consists in determining the volume and surface charge densities py(!) = —V-M(r) and o4(") = n-M(r) of the uniformly magnetized body and in correspondingly expressing the potential ia oe Put) = aly meet [eee aa 1.2 DEMAGNETIZING FIELDS: from which the demagnetizing field can be derived as the gradient Ha(x) = —Vy(x). With M constant in modulus and direction everywhere inside the spheroidal sample, the volume charge density py(r’) is zero and the surface charge density oy(r) is easily calculated. If M lies along one of the principal axes (a, , c), the corresponding homogeneous demagnetizing field is Na Hy = —NgM= —-—J, (1.16) ld a J, where the proportionality factor Ny is called “demagnetizing coefficient”. For a generic direction, we have Hy = —lINgIM, where IINqil is a tensor having only the diagonal elements different from zero. These are the demagnetizing coefficients along the three principal axes Nga, Nap; Nac- They obey the constraint Naz + Nap + Nac = 1. In the general ellipsoid, ab *c and the demagnetizing factor Ng, is obtained as the integral 0 1 Naa = = 5 [@ FOYE + OP + OC + | dg. (1.17) which can be numerically calculated, together with Ng and Nac = 1— Na — Nap. Results are reported in the literature (see, for instance, Ref. [1.5]). Closed expressions are found for ellipsoids of revolution (Fig. 1.2). In the limiting case of a sphere, we have, for reasons of symmetry, Naa = Ney = Nac = 1/3. For a prolate spheroid, where a = b and the rotational symmetry axis c > a,b, the demagnetizing factors are given, for the defined ratio r = c/a = c/b > 1, by the expressions Ne alge P-7) -1} 1 Naa = Nav = 5 — Nac). For r>>1, the approximation Ng, = In 2r — 1/7 holds. If the same axis ¢ ) NY FIGURE 1.2 Prolate (a = b, c > a,b) and oblate (a = b, c < a,b) spheroids. c is the rotational symmetry axis. The demagnetizing field H, is always uniform in ellipsoids if the magnetization M is uniform, whatever its direction. The demagnetizing coefficients calculated along the three symmetry axes, which completely define the problem, are given by Eqs. (1.18) (prolate spheroid) and (.19) (oblate spheroid). They depend only on the ratio r = c/a = c/b. apparent here that the demagnetizing field does not depend on the sample volume, but only on its geometrical properties (the ratio r). Demagnetizing fields can never be ignored in measurements. To recover the intrinsic magnetic properties of the material under test, a correction is required, where the effective field H = H, — Hy, obtained as the difference between the applied field H, and the demagnetizing field, is calculated. The problem is apparent with bulk soft magnets having a relatively low aspect ratio in the direction of magnetization, but also there is heavy interference by demagnetizing effects with strips and ribbons. In a typical experiment, a 20 um thick, 200 mm long, and 10 mm wide high permeability amorphous ribbon is tested. It is found by hysteresis loop shearing analysis that this sample has what could be deemed a very low demagnetizing factor (Ng ~ 1.3 x 1075), correspond- ing to a value r~ 700 in a prolate ellipsoid. For peak polarization value, J, =0.8T, the demagnetizing field is, according to Eq. (1.16), Hg ~ 8.5 A/m, which is about eight times larger than the effective field H. In hard magnets, it is fortunately possible in principle to perform accurate measurements with open samples also, thanks to the very high fields intrinsically required for their magnetization and demagnetization. However, permanent magnets generally come as bulk specimens 1.2, DEMAGNETIZING FIELDS ue (cylinders, parallelepipeds, spheres), which have high demagnetizing coefficients. It may also happen that some kinds of tapes or thin films preferentially magnetize normal to their plane, thereby approaching the value Ng = 1, and the correction procedure may become very complex. Ellipsoidal test samples are seldom employed. A notable exception is represented by some open sample methods applied in permanent magnet testing (for example, with the vibrating sample magnetometer; see Section 8.2.1), where small spherical specimens (diameter around few mm) are adopted. Use of cylindrical, parallelepipedic, or disk-shaped test specimens is generally the rule, but these geometries engender significant complications in the definition and determination of the demagnetizing coefficient. In fact, even with homogeneous magnetization, the demagne- tizing field is not homogeneous. In addition, it is also dependent on the material permeability. In soft magnets, this brings about notable compli- cations, because permeability is high and, even if we disregard hysteresis, corrections become difficult and inaccurate. Large errors are expected to occur when, as is often the case, H, and Hy have very close values. The problem of predicting demagnetizing fields in non-ellipsoidal bodies can be attacked in principle by computation of the integrals in Eq. (1.15), a relatively complex and time-consuming approach. Three- dimensional and two-dimensional finite element calculations have been developed for this purpose. They are indispensable for treating complex micromagnetic problems and the analysis of specific domain patterns [1.8, 1.9]. For regular cylindrical or parallelepipedic shapes and structurally and magnetically homogeneous materials (no domains and constant susceptibility value y everywhere), calculations of the demagnetizing coefficient have been performed to various degrees of approximation. For uniformly magnetized samples and y = 0, analytical approaches have been carried out and reasonable approximations can often be given, although, in a strict sense, they apply only to diamagnets, paramagnets, and saturated ferromagnets. For example, a uniformly magnetized cylinder has magnetic charge density «= at its top and bottom surfaces. The demagnetizing field at the center of the cylinder is straightforwardly calculated by integrating the field generated by an infinitesimally thick annulus of surface charges and summing up the contributions of the top and bottom ends (details are given in Section 4.4). For a cylinder with height-to-diameter ratio r, such a field turns out to be He NM (a (1.20) vier 12 CHAPTER 1 Basic Phenomenology in Magnetic Materials For r= 1, this provides Ny = 0.293, as compared with Ny = 0.333 in a sphere. The demagnetizing field at the center of a prism with square cross-section and height-to-side length ratio r, uniformly magnetized along the axial direction, is similarly obtained as 2 1 Hg = —NgM = ——arcsin( ——)m. : j= —NaM qaresin( es ;)M a2 We see in Fig. 1.3 how the associated demagnetizing factor compares with that of an inscribed spheroid as a function of the ratio r. In particular, it turns out that, for a cube (r = 1) Ng = 1/3, exactly the same as for the sphere. Magnetic materials are normally characterized using regularly shaped samples, where either the measurement of the magnetic flux upon a well-defined cross-section or the determination of the total magnetic moment of the test specimen is performed and related to the material magnetization. At the same time, the concept of effective field must be given a practical meaning. In connection with these two measuring approaches, one talks of fluxmetric and magnetometric methods. When dealing with non-ellipsoidal open samples, it is therefore expedient to distinguish between fluxmetric (or ballistic) and magnetometric iM & a M i i Lo- ee \ 1 2 a 4 5 r=cla=c/b FIGURE 1.3 Behavior of the demagnetizing factor Ny = Hg/M, with Hg the field at the center, in a uniformly magnetized prismatic sample with square cross- section and y = 0 (dashed line, Eq. (1.21)). Comparison is made with the same quantity calculated for an inscribed spheroid. 1.2 DEMAGNETIZING FIELDS: B demagnetizing factors. Let us consider, as in the earlier example, a cylindrical or prismatic sample. The material is homogeneous, does not show hysteresis, and the susceptibility is isotropic and constant every- where. A qualitative idea of the non-homogeneity of the demagnetizing field in connection with homogeneous magnetization is provided as a sketch in this case by Fig. 1.1. An example of quantitative derivation of the dependence of the axial component of the magnetization and the demagnetizing field for different susceptibility values in a cylindrical sample is illustrated in Fig. 1.4 [1.10]. We thus define the fluxmetric demagnetizing factor as the ratio of the average demagnetizing field to the average magnetization over the midplane perpendicular to the sample axis (z-direction) i) HadA NO = “A ___= ee (1.22) Jymnaa ’ For reasons of symmetry, both demagnetizing field and magnetization at the midplane are directed along the cylinder axis. The magnetometric demagnetizing factor is defined, however, as the ratio of the volume- averaged demagnetizing field to the volume-averaged magnetization (1.23) NY and N{™ depend on the height-to-diameter ratio r and the effective susceptibility x of the material. This is defined as y = (M,)/H and can be expressed in terms of the directly measured apparent susceptibility y, = (M.)/H, and the demagnetizing factor N{? a 1-NP xe Chen et al. have carried out a comprehensive calculation of fluxmetric and magnetometric demagnetizing factors in cylindrical samples for a very large range of aspect ratio r and magnetic susceptibility y values. For the case x = 0, these authors use Brown’s method [1.11], where the cylindrical sample is assimilated to a solenoid bearing the mantle of surface Amperian currents associated, according to Eqs. (1.1) and (1.9), with uniform magnetization. For long cylinders (r > 10) and y # 0, N® and N“™ are analytically calculated by means of a unidimensional model, where the axial magnetization component M, is assumed uniform over each x (1.24) 14 CHAPTER 1 Basic Phenomenology in Magnetic Materials 08 06 M,(2)/M,(0) 04 0.2 (a) zh “——— =104 -0.2 ‘ 1=30 Hage)! Ha 1=108 O02 02 0406) os 410 (b) zh FIGURE 1.4 Behavior of magnetization and demagnetizing fields along the axis of a slender cylindrical sample (height-to-diameter ratio r= h/a = 20). Calcula- tions have been made for different values of the effective susceptibility x. With high x values, the magnetization tends to vanish at the sample ends and the demagnetizing field H,, nearly completely compensates the applied field H, (adapted from Ref. [1.10]. cross-section of the cylinder. The model takes into account the presence of free poles both at the top and bottom ends and on the lateral surface of the sample. Figure 1.5b provides the results obtained with this model for y= 0, y= 100, and y= with r ranging between 10 and 1000. For short cylinders, where the previous assumption of uniform M, is untenable, a two-dimensional model, where the lateral surface of the 1.2 DEMAGNETIZING FIELDS. 15 3 3 8 EH N 3 5 8 £ 5 3 (@) § 2 g 3 2 = e 5 3 . 5 10 100 1000 (b) raha FIGURE 1.5 Magnetometric N{” and fluxmetric N{ demagnetizing factors calculated for cylindrical samples as a function of the height-to-diameter ratio r for different values of the susceptibility (adapted from Ref. [1.10]). 16 CHAPTER 1 Basic Phenomenology in Magnetic Materials cylinder is subdivided into a suitable set of non-overlapping rings of uniform charge density, is numerically implemented. Selected behaviors of Nf’ and N{* obtained with this model for 0.01 <7 = 10 are shown, in comparison with the behavior of Ng in a spheroid, in Fig. 15a. The following general features of the predicted properties of the demagnetizing coefficients should be stressed: (1) For a given x value, both N{? and N{® decrease in slender samples. This is the obvious consequence of the crowding of the free poles at the sample ends. (2) For the same reason, it is always N{ > N{. (3) Whatever the value of 1, N§™ always decreases with increasing y. In contrast N¢ increases with increasing y if r > 1. The opposite occurs if r < 1.4. The ratio N™ /N® increases with increasing 1, for given x, and decreases with increasing y, for given r. The subject of analytical determination of N? and N‘™ in prisms of rectangular cross-section has been treated by Aharoni, who provided expressions for both of them under the assumption of saturated sample (y=) [1.12, 1.13]. The results were applied to experiments on 2.35 um X 95 pum X 250 pm NigoFe29 thin films with easy axis directed along the major side. It was verified that a minimum applied field equal to the calculated fluxmetric demagnetizing field Hy = —N‘?M, (M,, satura- tion magnetization) was required to reverse the magnetization in the central part of the sample. 1.3 MAGNETIZATION PROCESS AND HYSTERESIS We primarily identify the expression “magnetic characterization” with the experimental process by which we determine the constitutive law of a magnetic material. This is the functional dependence M(H) of the magnetization on the effective field or, according to Eq. (1.6), the equivalent laws J(H) and B(H). This book will be chiefly devoted to reviewing and discussing the measuring methods by which such laws can be experimentally derived. Diamagnetic and paramagnetic materials are described by simple single valued relationships between field and magnetization. This by no means implies that such relationships are easily accessible to experiments because the associated susceptibilities are very low and, consequently, it is difficult to discriminate between the response of the sample under test and that of the background. Ferromagnets (and ferrimagnets) display intrinsically large responses. The landmark property of these materials is that they are endowed with a molecular field of quantum-mechanical origin, which implies magnetic ordering, ie. possible extraordinarily high values of the magnetic 1.3 MAGNETIZATION PROCESS AND HYSTERESIS 17 moment per unit volume. However, the phenomenology of the magnetization process in these materials has the traits of complexity, as embodied in the manifold manifestations of hysteresis. This appears as the macroscopic outcome of an intricate combination of microscopic processes, centered on the existence of domains (or, in limiting cases, single-domain particles), which lead to collective rearrangements of the magnetic moments under a changing applied magnetic field. The experimental investigation of magnetic hysteresis and the general testing of materials require that some kind of accessible reference state is defined. Magnetic saturation and demagnetized state are two such reference conditions. All trajectories in the H-M (or H-J) plane converge when reaching the saturated state. On recoiling from it, the system follows a unique trajectory. The demagnetized state acquires a similar distinctive character when the condition H=0 and M=0 is attained. To this end, the sample is either brought beyond the Curie temperature and cooled in a zero-field environment or an alternating field of progressively and finely decreasing peak amplitude is applied, starting from saturation and ending at zero value. The demagnetized state should realize, from the microscopic viewpoint, that special arrangement of the internal magnetic structure corresponding to the condition of absolute minimum of the magnetic free energy. Indeed, an infinite number of trajectories can lead to sample demagnetization without leading to the demagnetized state. If, starting from such a state, the applied field amplitude is increased, a curve in the (H-J) or, equivalently, (H—M) plane is described. It is the initial magnetization curve. In current practice, thermal demagnetization is seldom adopted. When this is the case, it is not expected to lead to the same set of microscopic states obtained with the conventional demagnetization under an alternating field. The curve described after thermal demagne- tization takes the name virgin curve and it can be slightly different from the initial curve. Figure 1.6 provides examples of initial magnetization curves in soft and hard magnets with both the (/-H) and (B-H) representations shown. These curves have been obtained using a closed magnetic circuit, where the applied field H, = H. If open samples are tested or the flux closure is imperfect, H, = H +(Na/po)J and the curves (-H,) and (B-H,) will appear sheared with respect to the intrinsic curves (J-H) and (B-H). We see that in practical soft magnets there is a detectable difference between these two representations only on approaching the saturated state, where the term jigH can be appreciated (Eq, (1.6). In order to bring the material along the magnetization curve, we must spend energy. Let us assume that a field, slowly increasing with time, is applied to a sample forming a closed magnetic circuit 18 CHAPTER 1 Basic Phenomenology in Magnetic Materials BUH) 20F 15 f Ee 2 of é * = ost Non-oriented Fe-Si 0.0 a ost ‘ 1 1 i we tot (@) H(Alm) Ee = Ee 5 Nd-Fe-Dy-Al-B 0 500 1000 1500 (by H(Alm) FIGURE 1.6 Initial magnetization curves in non-oriented Fe-Si laminations and in a Nd—Fe-Dy-AI-B sintered magnet. The soft magnetic laminations have been tested as strips forming a closed magnetic circuit within an Epstein frame. The permanent magnet has been demagnetized and magnetized as a parallelepipedic sample inserted between the pole faces ofan electromagnet. The shaded area provides the energy per unit volume E = [4° HdB to be spent for bringing the material from the demagnetized state to the peak induction value B,. 1.3 MAGNETIZATION PROCESS AND HYSTERESIS a (e.g. a ring specimen) by use of a suitably linked winding supplied with a current i(f). At any instant of time, the supplied voltage is balanced by the resistive voltage drop of the winding Ryi(t) and the induced emf. d@/dt g(t) = Ryilt) + db/dt. (1.25) Starting from the demagnetized state, a certain final state with induction value B, is reached after a time interval fp. The correspondingly supplied energy E = [?uc(thi(t)dt is partly dissipated by Joule heating in the conductor and partly delivered to the magnetic system ty ” BS j Ryi2(Odt + j ° Ny Ait) Bat (1.26) 0 0 dt where N,, is the number of turns of the winding and A is the cross- sectional area of the sample. We are interested in the second term on the right-hand side of this equation. If, to simplify the matter, we consider a ring sample of average circumferential length Im, we can write i(t) = (in/Nw)H() and we find that the energy delivered by the field-supplying external system in bringing the magnet to the final state is ‘Bp u= vf Ho Bat= vf HaB. (1.27) If we refer to the graphical representation of the initial magnetization curve in Fig. 1.6, we conclude that the energy per unit volume to be supplied in order to reach the induction value B, (or, equivalently, the polarization value J,), is given by the area between the B(H) curve and the ordinate axis. Introducing Eq, (1.6) in Eq. (1.27), we obtain H, ty u= vf oHdH + vf Haj, (1.28) and we can distinguish between the energy stored in the magnetic field and the energy stored in the material. Part of the latter is expected to be lost during the process. Except in somewhat unusual cases, the magnetization process is always associated with measurable energy dissipation. This is mirrored in hysteresis, the phenomenon of output lagging behind input. Figure 1.7 provides two examples of hysteresis loops in the previous soft and hard magnets, as obtained by cycling the field between two symmetric peak values +H,. If the integration in 20 CHAPTER 1 Basic Phenomenology in Magnetic Materials 1.5 Je osf 0.0 J(7),B) ost Non-oriented Fe-Si -1000-500 0 500 1000 (a) H(Alm) o Nd-Fe-Dy-AI-B J(T),B(T) ° —— -1500 -1000 -500 oO 500 1000 1500 ©) fe) FIGURE 1.7 Examples of hysteresis loops in soft and hard magnets. In the soft Fe-Si laminations, there is no detectable difference between the B(H) and J(H) curves for magnetizations and fields of technical interest. The difference is instead apparent in permanent magnets. It leads to two different definitions of coercive field: H.y is the ficld required to bring the induction to zero value starting from the saturated state and Hy is required to reduce to zero the polarization J (ie. magnetization M). It is always Hz > Hes (courtesy of E. Patroi). 1.3 MAGNETIZATION PROCESS AND HYSTERESIS 21 Eq. (1.28) is carried out over a full cycle, the energy balance is obtained W= puottadt +f Hay = pray = 4 Hap. (1.29) The quantity W is the energy lost per unit volume in the sample. In fact, the purely reactive term jHdH averages out to zero over a cycle. It then turns out that the area of the B(H) loop is equal to the area of the J(H) loop. Notice in the loops represented in Fig. 1.7 the remanence point, where J, = B,, and the distinction existing between the coercive fields H., and Hz, the latter being the field value where the material is demagnetized (indeed, a demagnetized state far removed from the one previously discussed). The phenomenology of magnetic hysteresis is extremely complex but endowed with a certain mathematical regularity, which has attracted relevant modeling efforts starting from the milestone approach of Preisach [1.14]. A full account of the mathematical aspects of hysteresis is found in Ref. [1.15], while the physics of magnetic hysteresis is treated with deep insight in Ref. [1.16]. We only remark here that, if we limit ourselves to material testing at low field strengths, we can fully describe initial curve and symmetric hysteresis loops by means of a defined function, the Rayleigh Law. Its discovery was a triumph of physical intuition and experimental skill in measurements [1.17], which is paralleled by the physical explanation advanced by Néel [1.18], based on the statistical description of the reversible and irreversible displace- ments of the domain walls. An example of loops determined in the Rayleigh region is reported in Fig. 1.8, where it is observed that the material polarization follows a quadratic dependence on the magnetic field. In particular, any hysteresis loop determined between the peak field values +H, has ascending and descending branches described by the equations b J(H) = (a+ DH, JH # 5 (Hp ~ H), (1.30) where a and b are called reversible and irreversible Rayleigh constants. The tip points of the loops (Hy, Jp) describe the initial magnetization curve (also called normal magnetization curve), and follow the law Jp = aH, + bHR. (31) In the limit of very low fields, the magnetization curve becomes linear (as is apparent in Fig. 1.8) with the constant a proportional to the initial 2 CHAPTER 1 Basic Phenomenology in Magnetic Materials coe] Pure Ni foil (Fp) 0.044 0.02: -0.024 -0.044 -0.06- -300 -200 -100 0 100 200 200 (Aim) FIGURE 18 Experimental hysteresis loops in the Rayleigh region. The ascending and descending branches of the loops and, a fortiori, the initial magnetization curve connecting the tip points of the loops, follow the quadratic law (1.30). The energy per unit volume dissipated upon completing a full cycle is given by the equation W = $bH5, where b is a structure dependent coefficient and Hy is the peak field value (courtesy of C. Beatrice). susceptibility x; i in a= Pay i MoXi- (1.32) Integration of the loop area provides the hysteresis loss per unit volume 4 W = 5 bHp, (1.33) where, as expected, the reversible coefficient a has disappeared. Equivalently, we can express the energy loss for a given peak polarization value J, in the Rayleigh domain as w= wl la? + 4b) al (1.34) Remarkably, Néel’s theory predicts that the quantities aH. and bH./a, where H, is the coercive field, as obtained with a major loop, are independent of the structure of the material. This conclusion derives from 1.3 MAGNETIZATION PROCESS AND HYSTERESIS 23 the assumption of scale invariance of the equations describing the interaction of the Bloch walls with the pinning defects. Our discussion has so far highlighted some essential facts concerning the so-called “quasi-static” magnetic behavior of the materials, which is observed when the rate of change of the magnetization is so low that dynamic viscous-type effects do not interfere with the magnetization process. When this condition is no longer fulfilled, we observe rate- dependent hysteresis effects. In metallic materials, they are almost exclusively due to eddy currents, whose circulation in the sample takes patterns depending on, besides the magnetization rate, the material resistivity, sample geometry, and domain structure. We shall consider dynamic loss behavior in connection with the properties of soft magnetic materials (Chapter 2) and the related measurements (Chapter 7). We only remark here that the fundamental consequence of the viscous effects associated with long-range eddy currents is the increase of the energy dissipated in any cycle, as manifest in the observed broadening of the hysteresis loop with increasing magnetizing frequency (Fig. 1.9). Co7,Fe4B Slip amorphous ribbon 02 O4F 0.0 J (1) 60-40-20 0 20 40 60 FIGURE 1.9 Broadening of the hysteresis loops with increasing magnetizing frequency in a soft magnetic alloy. The measurements refer to an amorphous ribbon (thickness 20 pm), wound into a many layer ring sample and tested under controlled sinusoidal induction waveform. 24 CHAPTER 1 Basic Phenomenology in Magnetic Materials The whole phenomenology can be assessed in most cases by applying the concept of loss separation [1.16]. References 1.1. peed 13. 14, 15. 16. ee 18. 1.9. 1.10. Li. 1.12. ae 114. 1.15. ce 1.17. 1.18, LID. Landau and EM. Lifshitz, Electrodynamics of Continuous Media (Oxford: Pergamon Press, 1989). J.C. Maxwell, A Treatise on Electricity and Magnetism (London: Clarendon, 1891, 3rd edition. Reprinted by Dover, New York, 1954), Vol. 2, p. 59. J.W. Strutt (Lord Rayleigh), “Notes on magnetism. On the energy of magnetized iron,” Philos, Mag., 22 (1886), 175-183. ].A. Ewing, “Experimental researches in magnetism,” Philos. Trans. Roy. Soc. London, 176 (1885), 523-640. A. Hubert and R. Schafer, Magnetic Domains (Berlin: Springer, 1998), p. 184. B.D. Cullity, Introduction to Magnetic Materials (Reading, MA: Addison- Wesley, 1972), p. 58. D. Jiles, Introduction to Magnetism and Magnetic Materials (London: Chapman & Hall, 1991), p. 39. S.W. Yuanand HN. Bertram, “Fast adaptive algorithms for micromagnetics,” IEEE. Trans. Magn, 28 (1992), 2031-2036. D.V. Berkoy, K. Ramstick, and A. Hubert, “Solving micromagnetic problems—towards an optimal numerical method,” Phys. Status Solidi, A-137 (1993), 207-225. DX. Chen, J.A. Brug, and RB. Goldfarb, “Demagnetizing factors for cylinders,” IEEE. Trans. Magn., 27 (1991), 3601-3619. WE. Brown, “Single domain particles: new uses of old theorems,” Am. J. Phys., 28 (1960), 542-551. A. Aharoni, “Demagnetizing factors for rectangular ferromagnetic prisms,” J. Appl. Phys., 83 (1998), 3432-3434. ‘A. Aharoni, L. Pust, and M. Kief, “Comparing theoretical demagnetizing factors with the observed saturation process in rectangular shields,” J. Appl. Phys., 87 (2000), 6564-6566. E. Preisach, “Uber die magnetische Nachwirkung,” Z. Phys., 94 (1935), 277-302. LD. Mayergoyz, Mathematical Models of Hysteresis (New York: Springer Verlag, 1991). G. Bertotti, Hysteresis in Magnetism, (San Diego: Academic Press, 1998). J.W. Strutt (Lord Rayleigh), “On the behaviour of iron and steel under the operation of feeble magnetic forces,” Philos. Mag., 23 (1887), 225-245. L. Néel, “Théorie des lois d’aimantation de Lord Rayleigh. 1 Les déplacements d'une paroi isolée,” Cahiers de Physique, 12 (1942), 1-20. CHAPTER 2 Soft Magnetic Materials A magnetic material is considered “soft” when its coercivity is of the order of, or lower than, the earth’s magnetic field. A soft magnetic material (SMM) can be employed as an efficient flux multiplier in a large variety of devices, including transformers, generators, and motors, to be used in the generation and distribution of electrical energy, and in a wide array of apparatus, from household appliances to scientific equipment. With a market around €6 billion/year, SMMs are today an ever more important industrial product, offering challenging issues in properties understanding, preparation and characterization. SMMs have been at the core of the development of the early industrial applications of electricity. Steel production was sufficiently developed at the turn of the century to satisfy the increasing need of mild steel for the electrical machine cores. In 1900, Hadfield, Barrett, and Brown proved that, by adding around 2% in weight Si to the conventional magnetic steels, one could achieve an increase of permeability and a decrease of energy losses [2.1]. Fe-Si alloys were more expensive and more difficult to produce and gained slow acceptance. In addition, the poor control of the C content was to mask the prospective performance of this product compared with mild steels. It took more than two decades, characterized by a gradual improvement of the metallurgical processes, for Fe-Si to become the material of choice for transformers. An empirical attitude towards research in magnetic materials was prevalent at the time and applications came well before theoretical understanding. This was the case for the Goss process, developed in the early 1930s, by which the first grain-oriented Fe-Si laminations could be produced industrially [2.2]. In the years 1915-1923, Elmen and co-workers at the Bell Telephone Laboratories systematically investigated alloys made of Fe and Ni, discovering the excellent properties of the extra-soft permalloys (78% Ni) [2.3]. Snoek is credited for the successful industrial development of ferrites in the 1940s [2.4], following attempts dating back to the first decade of the century. The discovery in 1967 of the soft magnetic amorphous alloys again occurred nearly by chance [2.5], but it provided a fertile field for technologists and 25 26 CHAPTER 2. Soft Magnetic Materials theorists. The discovery enriched the landscape of applicative magnetic materials, while straining existing theories on magnetic ordering. 2.1 GENERAL PROPERTIES The rough attribution of magnetic softness, based on the value of coercivity, is completed and made useful by the (J,H) hysteresis loop. Figure 2.1 provides a comparison of major DC hysteresis loops in a number of representative soft magnetic alloys. All measurements have been performed under closed-flux conditions. The Ni-Fe (Mumetal type) and the Co-based amorphous alloys reach the highest values of permeability and the lowest coercivities, but their saturation magnetization is somewhat reduced with respect to the Fe-Si and Fe-based amorphous alloys. The ferrites do not display prominent soft DC properties but, being non- conductive, become the best choice at frequencies in the MHz range. The actual and the prospective applications of magnetic materials have thus to be evaluated against a number of parameters, such as initial and peak permeabilities, coercive field, remanence, AC energy losses, squareness ratio, etc., which are the result of both compositional and structural properties, The composition determines the values of the so-called intrinsic magnetic parameters, such as the saturation magnetization, the magnetic anisotropy constants and the magnetostriction constants, which, in turn, affect the magnetization process in a way related to the material structure “F80-100-80 030100 180, S200 H(Alm) (Aim) FIGURE 2.1 Representative DC hysteresis loops in different soft magnetic alloys. (a) Grain-oriented Fe~(3 wt%)Si, non-oriented Fe-(3 wt%)Si, MnZn ferrites. (b) FeygBi3Sig annealed amorphous ribbons (1), Co7FesBisSiz0 as-quenched amor- phous ribbons (2), Fe~Ni (Mumetal type) alloys (3). 2.1 GENERAL PROPERTIES 27 (eg. crystallographic texture, grain size, foreign phases, lattice defects, etc.). By a proper choice of composition and suitable metallurgical and thermal treatments, extra-soft magnets are obtained, where the coercive field and the relative permeability attain values of the order of 0.1 A/m and 10°, respectively. However, it should be stressed that a number of additional properties, such as thermal and structural stability, stress sensitivity of the magnetic parameters, mechanical properties and machinability, and thermal conductivity, have to be considered. The final acceptance of a material in applications will result from a cost-benefit evaluation of all these properties. The magnetization process in an SMM occurs by means of two microscopic mechanisms: motion of the domain walls and uniform rotation of the magnetization inside the magnetic domains. Rotations require high field strengths in the conventional Fe-based crystalline alloys because the Zeeman energy Ej; = —H-J, (with H and J, the applied field and the saturation polarization, respectively) must balance a magneto- crystalline anisotropy energy term Ex roughly of the order of the aniso- tropy constant K;. Given that K; is of the order of few 10 J/m’, fields in the 10° A/m range must be applied to achieve substantial rotations. A soft magnetic behavior can possibly be achieved in these materials only through displacements of the domain walls. Frictional forces, inevitable in real defective materials, resist these displacements. The coercive field measures the typical field strengths at which the domain walls are unpinned from defects and a substantial part of the magnetization is reversed. Energy is lost in this process, and magnetic hysteresis is accordingly observed. The subject of coercivity and hysteresis is classically treated by theorizing the motion of a domain wall, assumed either as a rigid [2.6, 2.7] or a flexible [2.8] object, in a perturbed medium. Basically, it is assumed that the structural perturbation generates a random wall energy profile, whose spatial derivative represents the pressure to be applied by a field in order to achieve wall motion. The hindering effect of the structural defects is chiefly controlled by the value of the anisotropy constant K, implying that the value of coercivity tends to increase with the strength of the anisotropy effects. However, even with the relatively high values attained by Ky, very soft magnetic behavior can be achieved in Fe and Fe-Si when the microstructure is suitably controlled. In practice, this means having the least content of precipitates, voids, dislocations, and point defects, together with large and favorably oriented grains. By having the applied field directed as far as possible alongside one of the (100) easy axes, i.e. in the plane of the main 180° domain walis, one gets an obvious directional advantage for the wall displacements, as remarkably demonstrated by the grain-oriented Fe-Si 28 CHAPTER 2. Soft Magnetic Materials laminations. The role of the microstructural defects is clearly observed in Fe. Here, one can reach coercive fields as low as H.~1A/m upon prolonged purification and annealing treatments [2.9], leading to very low dislocation densities and C and N concentrations (some 10-20 parts in 10° (ppm)). Coercivities of a few hundred A/m can be found instead when these concentrations are in the 100 ppm range [2.7]. C and N are basically insoluble in Fe and tend to form carbides and nitrides, which act as strong pinning centers for the domain walls. With much higher C content (say around 1wt%), graphite precipitates and martensitic domains are additionally formed, and H. can reach values typical of hard magnets (several 10* A/m). Soft and extra-soft magnetic properties are naturally associated with very low values of the magnetic anisotropy (say with K in the range of a few tens J/m’ and less). This is the case, for example, of Fe-Ni alloys, with composition around Fez9—Nigo. Ki is positive in a-Fe (bcc cell) and negative in Ni (fcc cell) and it passes through zero on the high Ni side in the Fe-Ni alloys. Vanishing anisotropy can equally be obtained in amorphous and nanocrystalline alloys because the structural order in these materials is extended over limited distances, from a few atomic spacings to a few nanometers. The characteristic length L controlling the magnetization process, represented by the domain wall thickness, encompasses a large number of the local ordered structural units, so that the magnetocrystalline anisotropy is effectively averaged out. The residual anisotropy is calculated to be [2.10] 4 K= ne : (2.1) where A is the exchange stiffness constant and 8 is the average size of the structural units. By taking 6 = 10~? m, K; = 4.8 x 10 J/m’ (as in Fe) and A~ 10" J/m, we find the negligible value Ky = 5.3 10~° J/m’. With the elimination of the magnetocrystalline effects, other sources of magnetic anisotropy can be brought to light in these materials. For instance, in the highly magnetostrictive Fe-based amorphous alloys, a substantial anisotropy can result from the magnetoelastic coupling between frozen-in or applied stresses and the magnetization. In the typical alloy of composition, FesB,3Sis, the saturation magnetostriction is A, ~ 30 x 10”® and the long-range internal stresses generated by the rapid solidification process are of the order of 50-100 MPa [2.11]. Induced anisotropies K, = 3A,0, in the range of several hundred J/m®, can therefore arise and annealing treatments are required in order to achieve excellent soft magnetic behavior. Co-based amorphous alloys (like 2.1 GENERAL PROPERTIES 29 Cor1FeB,sSijg or CogsFesCryBi5Siz) and Fe—Ni alloys of the permalloy type have vanishing magnetostriction (in the 10~-1077 range) and, lacking also the magnetocrystalline anisotropy, attain the lowest coercivities and record values of permeability. In addition, their proper- ties can be adjusted by inducing calibrated uniaxial anisotropies through annealing treatments under saturating fields. In a saturated Fe—Ni alloy, the magnetization interacts with the Fe-Fe and Ni~Ni atomic pairs in such a way that, if the temperature is sufficiently high, they tend to distribute preferentially along the field direction, although the alloy preserves its character of random solid solution. In the amorphous alloys, anisotropic atomic rearrangements in the local ordered units, with symmetry influenced by the direction of the magnetization, are expected to occur [2.12]. In all cases, dramatic changes of the magnetization curve can be obtained by means of field annealing, as illustrated for a Co-based amorphous alloy by the example shown in Fig. 2.2. Itshould be noted that Amorphous alloy Co,,Fe,B, Sing 08 04 Field annealed, H Field annealed, H,, -20 =10 0 10 20 H (Alm) FIGURE 2.2 DC hysteresis loops in amorphous ribbons of composition Cor Fei. B,sSiio after annealing under a saturating magnetic field. A rectangular loop with H.~ 0.5 A/m and peak relative permeability 1, ~ 10° is obtained after long- itudinal field annealing (Hj, 1000s at 340°C). Transverse field annealing (H., T = 260 °C) leads to linearization of the loops, as shown here after 600 and 3000 s long treatments. The correspondingly induced transverse anisotropies are K,~5}/m’ (2) and K, ~ 25J/m’ (3). 30 CHAPTER 2 Soft Magnetic Materials the hysteresis loops in these amorphous ribbons can be linearized by means of a transverse induced anisotropy around some 10 J/m®. In this case the rotation of magnetization becomes an easy process, leading to quite high initial permeabilities (of the order of some 10°), a welcome property in many applications. SMMs find most applications in magnetic cores of AC apparatus, from 50 Hz to several MHz. The quantity of basic technical interest is in this case the energy loss. Eddy currents are generated by time-varying magnetic flux, leading both to shielding of the core by the associated counterfields (skin effect) and generation of heat by Joule effect. Excluding ferrites, which are basically insulating materials, all soft magnets have to be used in sheet form in order to minimize skin effect and losses. The theoretical assessment of these effects is by no means a simple one because it is seldom possible to treat the material as a continuum, characterized by a given magnetic permeability, and apply to it the Maxwell equations. The magnetic structure is made of domains and domain walls and the distribution of eddy currents can be extraordinarily complex and non- uniform. Williams, Shockley and Kittel were the first to take such a complexity at face value by investigating the dynamic behavior of a 180° Bloch wall in single crystals of Fe—Si. They theorized it through a balance equation involving, on the one hand, the applied magnetic field and, on the other hand, the structural pinning field and the eddy current counterfield (WSK model) [2.13]. Pry and Bean generalized the WSK model to a system of 180° walls, emulating the real domain structure of a grain-oriented Fe~Si lamination [2.14]. The whole problem has in recent times received a complete assessment by Bertotti, who has shown that the complexity of the dynamic magnetization process in real structures can be properly described by means of statistical methods [2.15]. The general phenomenology of losses in magnetic laminations is summarized in Fig, 2.3, illustrating the dependence of the hysteresis loop shape and area (ie. energy loss) on the magnetizing frequency in a grain-oriented Fe-Si alloy. Bertotti’s theory provides a physically based demonstration of the concept of loss separation, as expressed by the equation: W = Wy + Wa t+ Were: (2.2) The energy loss per cycle W is taken, at a given magnetizing frequency f and peak polarization Jp, as the sum of three components, having the following meaning. Wr, called hysteresis loss, is the residual energy dissipated in the limit f— 0. Wa is the loss component calculated by applying the Maxwell equations to the material when it is assumed completely homogeneous from the magnetic viewpoint (absence of domains). By summing up Wp and Wg, one falls short of the measured 2.1 GENERAL PROPERTIES 31 'SEOFe Gms 30 Hy GO Fe-(3mt%)Si : 25Hy 421.927 os 2 z E 00 t = 05 S (00 Hz -1.0 . 5 200 Hz | “60-46-20 “0 "204060 0 80 100 150 200 @ H (Alm) (b) t(H2) FIGURE 2.3 Hysteresis loops (a) and specific energy loss per cycle (b) vs. magnetizing frequency in grain-oriented Fe~Si laminations (thickness 0.29 mm) under sinusoidal time dependence of the polarization (J, = 1.32). The energy loss in (b) is proportional, at a given frequency f, to the area of the corresponding UF hysteresis loop. The loss separation concept formulated in Eq. (2.2) is illustrated in (b). value. The remainder, Wex¢, is called excess loss. The three loss components are associated with different eddy current mechanisms and different space-time scales of the magnetization process. The classical loss W. is a sort of background, always present and independent of any structural feature. For a lamination of thickness, d, conductivity o and density § one finds, under sinusoidal time dependence of the magnetic polarization, the classical loss per unit mass: 2 Wy = SE, oe where dis the density of the material and it is assumed that complete flux penetration in the lamination cross-section occurs. Figure 2.3 demonstrates how an elementary approach to losses based on the classical approxima- tion largely underestimates the measured loss in grain-oriented Jamina- tions at 50-60 Hz, the frequencies at which the large part of electrical energy is generated and produced. In addition, the dependence of W on f is non-linear, in contrast with the prediction on Wa (Eq. (2.3)). The components Wh and We derive from the heterogeneous nature of the magnetic structure of the material and the inherently discrete behavior of the magnetization process. Flux reversal is concentrated at moving domain walls and, even under quasi-static excitation, eddy currents arise because the domain wall displacements, hindered by the pinning centers, occur in i 32, CHAPTER 2 Soft Magnetic Materials jerky fashion (Barkhausen effect). Intense current pulses, with lifetimes around 10~* s, are generated, localized around the jumping wall segments. In this way, the hysteresis loss W,, is generated and, since the time constant of the eddy current pulses is always many orders of magnitude smaller than the typical magnetization period T = 1/f, it is concluded that Wy is independent of frequency. For a given value J, Ws gives a measure of the coercive field. This is totally consistent with the fact that the Barkhausen mechanism is a volume effect, independent, as the coercivity should be, of the material conductivity. The excess loss W,xc is associated with the large- scale motion of the domain walls. The theory shows that, in most SMMs, the following expression holds: Were = kexey of, 24) where kexc is a parameter related to the type of existing domain structure and its relationship with the material structure. Very broadly, it can be stated that the larger is k.,.. the more discrete is the magnetization process. Very large domains are therefore not desirable from this viewpoint and, as discussed below, methods have sometimes to be devised to increase the number of domain walls in the material. In conclusion, if minimization of the AC energy losses is sought, not only the lamination thickness and 10 -— Low-carbon steel 1 ae — Grain-oriented Fe-Si Power loss (W/kg) ot a 4 Amorphous FeBSi 1880 1900 1920 1940 1960 1980 2000 Year FIGURE 2.4 Record loss figures over a century in soft magnetic laminations for transformer cores. 2.2 PURE IRON AND LOW-CARBON STEELS 33 TABLE 2.1 Representative soft magnetic materials and typical values of some basic magnetic parameters Composition Bmax H.(A/m) J; (1) Fe Fe100 3-50x10° 1-100 2.16 NO Fe-Si Fe(>96)-Si(<4) 3-10 10° 30-80 1.98-2.12 GO Fe-Si Fe97-Si3 20-8010? 4-15 2.08 Fe-Si 6.5% Fe93.5-Si6.5 5-30 10? 10-40 1.80 Sintered powders Fe99.5-P0.5 02-2 10° 100-500 1.65-1.95 Permalloy Fel6-Ni79-MoS 5X10" 04 0.80 Permendur Fet9-Co49-V2 2x10? 100 24 Ferrites (Mn,Zn)O-Fe,O3 3x10? 20-80 02-05 Sendust Fe85-Si95-AI5 50x10 5 1.70 Amorphous (Fe based) Fe7sB;3Si9 wo 2 1.56 Amorphous (Co based) Co7;Fe;BisSizo 5x10 05 0.86 Nanocrystalline Fer3 sCu;NbsSii3.5By 10° 05 12 Hmaxr Maximum DC relative permeability; H., coercive field; J., saturation polarization at room temperature. The composition is given in wt%, but for the amorphous and nanocrystalline alloys, where it is expressed in at.%. the material conductivity need to be reduced, as suggested by the classical approach (Eq. (2.3), butalso the microstructure must be controlled in order to minimize both W,, (ie. the coercive field) and Wex<. This emphasizes the role of metallurgical processing, whose continuous refinement over the years has produced increasingly better control of the various structural parameters (e.g. impurities, defects, grain size, and crystallographic texture) and clear progress of the magnetic properties of the materials (see Fig. 2.4). A representative list of different types of SMMs is given in Table 2.1. Most of these materials are produced and traded under defined specifications. The measurement of their properties is therefore subjected to acknowledged standards issued by national and international bodies. The International Electrotechnical Commission (IEC) Standards of the 60404 Series cover specifications and measurements of commercial SMMs. 2.2 PURE IRON AND LOW-CARBON STEELS Tron is referred to as “high purity” when the total concentration of impurities (typically C, N, O, PS, Si, and Al) does not exceed a few hundred ppm. It is otherwise called low-carbon steel or non-alloyed steel. When soluble elements like Si and Al are deliberately introduced, 34 CHAPTER 2. Soft Magnetic Materials typically in the range of a few percent, it is appropriate to speak of “silicon steels”. Very pure iron is rarely used in applications, but the study of its properties is of basic physical interest. The main practical drawbacks of pure Fe are its relatively high electrical conductivity, which makes it unsuitable for AC applications, its poor mechanical properties and its cost. Low-cost low-carbon steels (C < 0.1 wt%) are largely applied in a multitude of small electrical machines (for instance, fractional horse power motors) and devices where efficiency is not of primary concern. Together with the silicon steels, they cover about 80% of the world tonnage of SMMs. More efficient low-carbon steels are increasingly developed today, under the pressure of rising energy costs and environmental concerns. Higher grades are therefore now available, where improved magnetic properties are obtained chiefly by introducing a small amount of Si (<1 wt%) and decreasing the content of impurities (especially sulfides, carbides, and nitrides). A practical method to obtain high-purity iron is to start with commercially pure iron (e.g. of the ARMCO type) and refine it by suitable methods. These include prolonged annealing in pure H, at temperatures not far from the melting point (for instance, 48 h at 1480 °C), zone melting and levitation melting. By means of these methods, some 20-30 ppm maximum total impurity content can be reached, with C and N less than 10 ppm. Record permeabilities p,~ 10° and coercivities H, = 1-2 A/m have been obtained in highly purified iron samples. Some common iron grades are listed in Table 2.2. Low-carbon steels to be used in magnetic cores are generally produced as sheets, through a sequence of hot and cold rolling passes and thermal treatments, as shown schematically in Table 2.3. The specifications for these steels are provided in the IEC Standard 60404- 8-3 [2.16]. In the case of low-cost materials, heat and mechanical treatments are limited to those necessary to reach the final sheet thickness, in the range 0.50-0.85mm. To improve the magnetic performance, TABLE2.2 Typical impurities and their concentrations (wt ppm) in several grades of iron and in low-carbon steel Iron type C N Oo Mae s si cu Ni ARMCO 150 Di) 0 oo 25030 150 Electrolytic 40 i 10s 30 30 40 10 Hy treated 30 10 3080 < 30 Zone refined 7 <10 2.05 95%) is sought in big electrical machines, not only to save energy, but also to avoid overheating and shortened machine lifespan. The development of improved non-oriented alloys is related to the control of a number of structural parameters, namely impurities, grain size, crystallographic texture, surface state, residual and applied stresses. A few tens ppm concentrations of impurities like C, N, S, O tend to increase coercivity and losses (see Figs. 2.5 and 2.9) [2.22]. They can do this directly, by forming precipitates that act as pinning centers for the domain walls, and indirectly, by adversely affecting grain growth and texture. The role of grain size (s) is illustrated in Fig. 2.10, where it is observed that the optimal (s) value is, depending on the composition, around 100-200 pm, where 42 CHAPTER 2 Soft Magnetic Materials the total power loss attains a minimum value [2.23]. This is understood in terms of the opposite dependencies on grain size exhibited by the hysteresis loss component Wy, decreasing as (s)~", with n = 0.5-1, and by the excess loss Wex, approximately increasing as (s)/?. A low impurity content is mandatory in order to achieve this optimal grain size because precipitates tend to hinder grain growth. In addition, some particles, like MnSand AIN, favor the establishment of a detrimental texture, rich in {111} planes. On the other hand, there are soluted impurities, like Sb and Sn, that can induce selective growth of those re-crystallized grains that have orientations close to the ideal random cubic texture {100}(0vw). A similar texture can be approached, in two-stage reduced alloys, by increasing the Alconcentration up to 1.1 wt% [2.22] or even 1.8 wt% [2.24], which permits one to achieve W15/50 ~ 2 W/kg. The troublesome aspect of an increased Al content, besides the cost, is an increased tendency to surface and subsurface oxidation, occurring especially during the wet Hp decarburiza- tion. In this case itis expedient to achieve the desired non-aging properties by decarburization and denitrogenization of the melt through vacuum degassing. The punching operation generates localized internal stresses 6 T 5p Si (1) ™: Power loss (Wikg) eo (4) (8) ap IP dpa 1.5T f=50Hz NO Fe-Si 0 ot 10 100 Average grain size (um) FIGURE 2.10 Power loss vs. grain size in 0.50 mm thick NO Fe-Si laminations. The curves (1)-(5) correspond to different (Si + Al) concentrations: (1) Si 0.01 wt%; (2) Si 0.3 wt%; (3) Si O.8wt% +Al O.2wt%; (4) Si 11wt% + Al 0.2 wt%; (5) Si 3 wt% + Al 1 wt% [2.23]. 23 IRON-SILICON ALLOYS 43 and, consequently, it might affect the loss figure in fully processed materials where, in general, stress relief annealing is not performed. In large machines, however, a much larger effect on magnetic losses is expected to derive from the stresses permanently introduced by stacking and assembling the laminations in the core. 2.3.2 Grain-oriented Fe—Si alloys Fe single crystals exhibit minimum coercivity and maximum permeability when magnetized along one of the (001) axes. This property has fundamental implications on a theoretical level and outstanding practical consequences. In fact, most transformer cores are built today with grain- oriented (GO) Fe-Si laminations, where the crystallites have their [001] easy axis close to the rolling direction (RD) and their (110) plane nearly parallel to the lamination surface (Fig. 2.11). This is the so-called (110)[001] texture or Goss texture, after Goss, the first to develop such materials [2.2]. The remarkable texture of the GO alloys, together with a large grain size (from a few millimeters to a few centimeters) and a small impurity content, leads to coercive fields as low as 4-10 A/m and a maximum 7 (010) FIGURE 2.11 Domains in HGO Fe-Si laminations. They are oriented along the [001] axis and tend to multiply upon scribing the sheet surface. 44 CHAPTER 2 Soft Magnetic Materials permeability around 5 x 10*. These figures differ by about an order of magnitude from those typically found in NO alloys. Single-phase transformer cores can be made either by rolling up a Jong lamination or by stacking and suitably joining separate sheet pieces at the corners, so that the magnetic-flux path is everywhere aligned to RD. Three-phase cores are always of the stacked type and cover the high power range (starting from some 50 kV A). GO laminations are subdivided in two main classes: conventional grain oriented (CGO) and high permeability (HGO) alloys, characterized by a dispersion of the [001] axes of the crystallites around RD of the order of ~7 and ~3°, respectively. Specifications are provided in the IEC Standard 60404-8-7 [2.25]. The CGO materials, although lower performing than the HGO ones, cover about 80% of the market. As illustrated in Table 2.5, commercial products are offered with thicknesses ranging between 0.35 and 0.23 mm and a loss figure Wz 50 = 1.40-0.80 W/kg at 1.7T and 50Hz. More than 1 million ton/year of GO alloys are produced worldwide, with a market value estimated around €1.5 billion. Following the original method of Goss, a number of patented processes for the production of GO sheets have been developed, based on various complex thermomechanical sequences [2.26-2.28]. Table 2.6 offers a schematic illustration of the chief processing methods. In the case of CGO materials, the main preparation steps can be summarized as follows: (1) Melting in the arc furnace, vacuum degassing TABLE 2.5 Typical specifications for conventional and high- permeability grain-oriented Fe-(3 wt%)Si alloys Thickness Specific power losses Polarization (mm) Jp =1.7T, f= 50Hz(W/kg) at 800 A/m (T) Conventional (CGO) 0.35 1.40 1.82 0.30 1.30 1.83 0.27 1.21 1.84 0.23 115 1.84 High permeability (HGO) 0.30 1.06 191 0.27 0.99 1.92 0.23 0.92 1.92 0.23 (scribed) (0.80 1.90 2.3 IRON-SILICON ALLOYS 45 TABLE 2.6 Summary of industrial processing of grain-oriented silicon steel sheets cco HGO-1 HGO-2 HGO-3 Composition (wt%) 3-3.2, Si 29-33, Si 29-3, Si 3.1-3.3, Si 0.04-0.1, Mn 0.03, Al 0.05, Mn 0.02, Mn 0.02, S 0.07, Mn 0.02, Se 0.02, $ 0.03, C 0.03, 5 0.04, Sb 0.001, B Balance, Fe 0.015,N 0.03-0.07, C 0.005, N 0.05-0.07,C Balance, Fe 0.03-0.05, C Balance, Fe Balance, Fe Inhibitors Mn: MnS+AIN — MnSe + Sb B+N+S Melting, vacuum degassing and continuous casting of slabs Re-heating—hot rolling 1320°C 1360 °C 1320°C 1250 °C Annealing 900-1100 °C 1100-1150 °C 900 °C 870-1020°C Cold reduction 70% 87% 60-70% 80% Annealing 800-1000 °C - 800-1000 °C - Cold reduction 55% i 65% 7 Decarburization 800-850 °C (wet Hz atmosphere) MgO coating and coiling Box annealing (secondary recrystallization) 1200°C 1200°C 820-900 °C +1200°C 120°C Phosphate coating and thermal flattening CGO denotes the conventional grain-oriented steels. HGO refers to the high permeability steels, of which three processing methods are presented. CGO and HGO-2 processes adopt a two-stage cold reduction. HGO-1 and HGO-3 reach the final gauge in a single step. The inhibitors are: MnS precipitates (CGO); MnS + AIN particles (HGO-1); MnSe particles + solute Sb (HGO-2); solute B+N+Sin HGO3. 46 CHAPTER 2 Soft Magnetic Materials and continuous casting. Besides Si, ranging in concentration between 2.9 and 3.2 wt%, the following impurities are usually present: Mn (0.04- 0.1 wt%), S (0.02 wt%), C (0.03 wt%). (2) Slab re-heating at 1300-1350 °C, followed by hot rolling to the thickness of 2mm, annealing at a temperature of 900-1100 °C and rapid cooling. Times and temperatures through this stage are finely adjusted in order to achieve a homogeneous distribution of precipitates, namely MnS particles of around 10-20 nm. (3) Two-stage cold rolling (~70% plus ~55%) to the final gauge, with intervening annealing treatment at 800-1000 °C. (4) Decarburizing anneal in wet H; atmosphere at 800-850 °C. Since a huge amount of deformation is accumulated by the previous cold reduction, complete primary re-crystallization takes place at this stage. However, the newly formed grains are strongly inhibited in their growth because of the presence of the finely precipitated MnS impurities. (5) MgO coating, coiling and 48h box annealing at a temperature of 1200 °C. During this final annealing treatment, secondary re-crystallization takes place, where abnormally large and sharply (110)[001] oriented grains grow within the precipitate stabilized primary matrix and eventually cover the whole sheet. This is thought to occur because these grains have a boundary mobility much larger than the great majority of the primary grains, whose prevalent texture is around {111}(110) and {111}(112). It is estimated that one in about 10° primary grains is (110)[001] oriented, which explains the final large secondary-grain size. At the end of box annealing, the precipitates are completely dissolved and harmful effects on the magnetization process are avoided. (6) Phosphate coating and thermal flattening. The HGO laminations are obtained with some variants to the previous sequence, which lead to a sharper Goss texture. Three basic industrial HGO processes are employed today, as summarized in Table 2.6. Ina first process, where a single-stage 87% cold reduction is adopted, the MnS inhibitors are reinforced with AIN precipitates [2.26]. A partial austenitic transformation, during hot rolling and successive annealing at 1100- 1150 °C, is achieved by adjusting the C concentration around 700 ppm. With 3.25 wt% Si and 700 ppm C, the y-phase fraction at 1150 °C amounts to 40-50%, which leads to optimally sized AIN particles. The second method is a two-step cold rolling process which exploits the combined inhibiting action of MnSe precipitates and Sb solute atoms, segregated at grain boundaries [2.27]. In the third process, solute B, N, and S atoms act as inhibitors and a single-stage cold reduction is again performed [2.28]. Commercial HGO laminations come into the thickness range 0.30- 0.23 mm. To achieve lower gauges while maintaining a comparable textural quality is more difficult. It has been shown, however, that high permeability laminations in wide gauge range (0.18-0.50 mm) can be 2.3 IRON-SILICON ALLOYS 47 produced with the process HGO-1 (87% single-stage cold reduction), by forming the AIN inhibitor at the end of the decarburizing stage, through Nz injection (acquired inhibitor method) [2.29]. The beneficial effects on permeability and coercivity produced by an. improved Goss texture are directly related to the morphology of the magnetic domains, which, for a perfect (110)[001] grain, form a bar-like array, with the 180° Bloch walls running along the [001] direction. This ideal structure is approached in the HGO sheets, which exhibit the lowest coercivity and the highest permeability among the Fe-Si alloys, thanks to the combined absence of DW pinning centers and the presence of large well-oriented grains. Figure 2.11 shows a typical domain structure in an HGO lamination. One can notice the presence of spike-shaped supple- mentary domains. These domains form when the [001] axis does not lie in the plane of the lamination. Their role is to create closed paths for the magnetic flux in order to reduce the magnetostatic energy of the system. Ina material with few, wide bar-like domains, like the one in Fig. 2.11, the magnetization reversal, localized at each instant of time at the moving walls, is highly non-homogeneous and the excess loss component Wexe is accordingly large (see Eq. (2.4) and related discussion). In HGO materials with a sharp (110)[001] texture, this feature has detrimental consequences on power losses because the excess loss can be quite high, wiping out the beneficial effects on permeability and DC coercivity introduced by the excellent textural properties of the material. In order to optimize the material behavior in terms of both permeability and losses, it is expedient to apply, both in CGO and HGO laminations, a coating (thickness ~2.5 um) capable of exerting a tensile stress of 2~10 MPa. This stress leads, via magnetoelastic interaction, to partial or complete disappearance of the flux closing domains, which have a high magnetoelastic energy cost. Under these conditions, the system can reduce its magnetostatic energy by reducing the size of the main bar-like domains, leading to a more homogeneous magnetization process and reduced dynamic losses. However, the energy balance is such that stress-induced domain refinement hardly occurs when the angle made by the [001] axis with the plane of the lamination is 8 = 1.5°. For this reason, in the best HGO materials, significant domain multiplication is achieved through a combination of magnetoelastic and magnetostatic effects, by scribing patterns on the lamination surface (Fig. 2.11). An array of scribing techniques (e.g. mechanical scratching, laser irradiation, plasma jet scribing, etch pitting) have been devised and are employed in the production of the highest HGO grades. The joint effects of tensile stress and surface scribing on power losses can be appreciated by experiments carried out in single crystals characterized by different values of the angle 48 CHAPTER 2 Soft Magnetic Materials B [2.30]. As shown in Fig. 2.12, the 50 Hz power loss is drastically reduced by scribing when f < 1°, and further benefit is introduced by the tensile stress. This occurs with little detrimental effects on permeability. For applications at medium frequencies (0.4-10 kHz), thinned GO laminations are commercially available, which are produced by rolling standard laminations to a reduced gauge of 0.15-0.10 mm and carrying out an appropriate heat treatment. As these alloys present inferior texture and permeability, new processes have been devised in the laboratory in order to achieve highly oriented extra-thin materials [2.31]. Excellent performances have been obtained in laminations with thickness ranging between 10 and 100 um by adopting, after cold reduction to the final gauge, a special sequence of annealing treatments. These induce a sharply defined (110)[001] texture through a tertiary re-crystallization process Power loss (Wikg) 0.0 7 ee B (deg) FIGURE 2.12 Power loss Pi7/so measured at 50 Hz and 1.7 T on 0.20 mum thick Fe-(3 wt%)Si single crystal strips, as a function of the misorientation angle 6 made by the [001] direction with respect to the strip plane. Different tensile stresses and scribing conditions are considered: (1) «= 0, before scribing; (2) a= 0, scribed; (3) ¢ = 15 MPa, before scribing; (4) « = 15 MPa, scribed. Tensile stress and scribing of the strip surface both produce a decrease of the power loss, due to a decrease af the excess loss component. Combination of texture perfection with scribing is the route to combined maximum permeabilities and minimum losses. Data taken from Ref. [2.30]. Se 7 2.3 IRON-SILICON ALLOYS 49 [2.32], which occurs because the (110) crystallographic planes are those having minimum energy when exposed to the lamination surface. This principle is also the basis of the preparation of cube-on-face (100)[001] textured Fe-(3 wt%)Si laminations by secondary re-crystallization in a slightly oxidizing atmosphere [2.33]. These materials are, in principle, attractive because the biaxial symmetry introduced by the two easy axes in the lamination plane has potential for applications both in transformers and rotating machines. However, they have failed to attract commercial interest and their production was discontinued many years ago. 2.3.3 Fe-(6.5 wt%)Si, Fe-Al and Fe-Si-Al alloys Fe-(6.5 wt%)Si alloys are a prospective route to low loss materials. When compared with the conventional Fe-(3 wt%)Si alloys, they exhibit a favorable combination of lowered anisotropy (K = 2.1 x 10'J/m’ instead of K=3.6x10*J/m*) and increased resistivity (p=80x10°Qm instead of p = 45 x 10° Q m), which compound with vanishing magne- tostriction (A1p = — 0.5 X 10°, Ar = 2 10 %, see Fig. 2.13) to provide a potentially excellent soft magnetic alloy for applications at power and Coy Cay (wt) FIGURE 2.13 Magnetostriction constants Aigo and Ar» in Fe-Si and Fe-Al alloys and their evolution with the concentrations of Si and Al. The magnetostriction in Fe-Si becomes vanishingly small for cg; ~ 6.5 wt%. 50 CHAPTER 2 Soft Magnetic Materials medium frequencies. Fe—(6.5 wt%)Si alloys cannot be prepared by cold rolling because the heterogeneous formation of ordered FeSi (Bz) and Fei (DOs) phases during cooling makes them hard and brittle. To avoid ordering, cooling rates greater than about 10°°C/s in the temperature interval 800-500 °C are needed, a condition that can be satisfied by rapid quenching from the melt. By planar flow casting (PFC), a method where a molten metal stream is ejected onto a rotating metallic drum (Fig. 2.14), ductile Fe-(6.5 wt%)Si ribbons are obtained, with thickness typically ranging between 30 and 100 um. Once annealed in vacuum at 1100— 1200 °C, they exhibit a columnar grain structure ((s) ~ 100-500 jm) with a prevalent texture (1000vw) (random cube-on-face). The final ribbons may show, at the price of a somewhat reduced ductility with respect to the as-quenched state, a coercivity lower than 10 A/m and a maximum relative permeability larger than 10* and constant up to the kHz region. At 1 kHz and 1T, the power loss can be lower than 15 W/kg in 50 jum thick ribbons, a value that compares favorably with the typical loss figure of thinned GO Fe (3 wt%)Si laminations. Fe~(6.5 wt%)Si alloys can equally be prepared by Si enrichment of standard NO and GO laminations. Classically, Si enrichment is carried out by means of a chemical vapor Planar Flow Casting Chill Block Melt Spinning FIGURE 2.14 Preparation of soft magnetic ribbons by means of rapid quenching, from the melt. The peripheral velocity of the metallic wheel (e.g. copper, iron. steel) typically ranges between 10 and 40 m/s. Casting can be performed either in air, inert gas, or vacuum. 2.4 AMORPHOUS AND NANOCRYSTALLINE ALLOYS 51 deposition (CVD) process, where SiCl, is used as the donor phase, according to the reaction SiCl, + 5Fe + FesSi + 2FeCl,. This process has been industrially applied to NO laminations, prepared with thickness ranging between 0.10 and 0.30 mm [2.34], which exhibit good workability. For peak induction values below 1 T and frequencies higher than a few hundred Hz, the reported loss figures are better than those found in GO laminations of the same thickness. At the same time, the reduced magnetostriction yields a definite decrease of the acoustic noise level under operating conditions. Solute Al and Si atoms affect the physical properties of Fe in a similar way [2.35], but Si is preferred in magnetic alloys because it is less prone than Al to reaction with oxygen and is less expensive. On the other hand, Fe-Al alloys are ductile, even when partial ordering occurs (above 7~ 8 wt% Al concentration). The magnetostriction constant Ajo suffers an approximately fivefold growth on increasing Al up to 10 wt% (from 201076 to ~90 x 10~°), thus following an opposite and unfavorable trend with respect to Si addition (see Fig. 2.13), and eventually drops around zero for Al concentrations ~16 wt%. At the same time, the anisotropy constant follows a monotonic decrease and, depending on the degree of ordering, passes through zero around Al = 11-14 wt%. Two compositions have applicative relevance. The Fe—(13 wt%)Al alloy com- bines high magnetostriction with low anisotropy and is of interest for magnetoelastic transducers. The Fe-(17 wt%)Al alloy is characterized by mechanical hardness and high permeability and is used in magnetic heads. The ternary Fe—(9.6 wt%)Si—(5.5 wt%)Al alloy, known as Sendust, is char- acterized by an extremely soft magnetic behavior because the constants K,, Aroo, and Ay; all approach the zero value simultaneously. The coercive field can reach values around 1-2 A/m and the relative permeability is of the order of 10°. Sendust alloys are extremely brittle and are therefore used in cast form for DC applications and as powder cores in AC devices. 2.4 AMORPHOUS AND NANOCRYSTALLINE ALLOYS Amorphous SMMs can be obtained as thin laminations by means of rapid solidification. Today, the predominantly employed rapid solidification technique is PFC, by which one can prepare ribbons of variable width (up to 100-200 mm) and thickness usually ranging between 10 and 40 pm. In a typical PFC setup (Fig. 2.14), a quartz crucible, which holds the liquefied master alloy, is placed nearly in contact with the surface of a rotating metallic drum, which drags the liquid at a velocity of 10-40 m/s. This ensures a cooling rate of the order of 10°-10° °C/s, sufficient to undercool

You might also like