Download Michael Coey, Stuart S.P. Parkin - Handbook of Magnetism and Magnetic Materials-Springer (2021)

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
Michael Coey
Stuart S. P. Parkin
Editors
Handbook of
Magnetism
and Magnetic
Materials
Handbook of Magnetism and Magnetic
Materials
J. M. D. Coey • Stuart S. P. Parkin
Editors
Handbook of Magnetism
and Magnetic Materials
With 618 Figures and 157 Tables
123
Editors
J. M. D. Coey
School of Physics
Trinity College
Dublin
Ireland
Stuart S. P. Parkin
Max Planck Institute of Microstructure Physics
Halle (Saale)
Germany
ISBN 978-3-030-63208-3
ISBN 978-3-030-63210-6 (eBook)
ISBN 978-3-030-63209-0 (print and electronic bundle)
https://doi.org/10.1007/978-3-030-63210-6
© Springer Nature Switzerland AG 2021
All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction
on microfilms or in any other physical way, and transmission or information storage and retrieval,
electronic adaptation, computer software, or by similar or dissimilar methodology now known or
hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
Magnetism is a natural phenomenon that arouses curiosity in people of all ages.
Electromagnetism, a mainstay of the industrial revolution, supports urban life and
communications everywhere, served by soft magnetic materials that guide and
concentrate magnetic flux. Permanent magnets are now ubiquitous flux generators,
enabling electric mobility, robotics and energy conversion in a range from μW to
MW. A handful of about a dozen optimized bulk functional magnetic materials
address well over 90% of practical applications.
The ability to pattern magnetic thin films has transformed our subject. Progressively scaled to nanometer dimensions, tiny magnetic regions store binary data,
which forms the basis of today’s digital world. Their stray fields are detected using
minute and exquisitely sensitive magnetic field sensors formed from atomically
engineered multi-layer stacks of magnetic thin films that are the first and, still
today, the most important crop product of spin electronics. Spintronics, especially,
concerns the generation, manipulation and control of the spin angular momentum,
which is the source of the electron’s magnetism. Spin-polarized electrical currents
or main pure spin currents with no net charge flow can be used to excite or switch
the direction of magnetization of magnetic nano-elements. This has opened the door
to a range of magnetic devices with properties that go beyond those of charge-based
electronics. There are new prospects for memory, storage and computation that
are fundamentally spin based. The emerging field of chiral spintronics combines
fundamental aspects of chirality, spin and topology.
On a more fundamental level, although the theoretical foundations of magnetism
in relativity and quantum mechanics were established a century ago, the behaviour
of strongly correlated electrons in solids is an unfailing source of surprises
for physicists and chemists, materials scientists and engineers. Model magnetic
materials can be created to exhibit an astonishing range of physical properties,
and increasingly we are learning how to tailor them to suit a particular practical
application or theoretical model.
The shift of emphasis from bulk, functional magnets to thin films has transformed
the range of elements we can use in our materials. Practically, any stable element in
the periodic table can now be pressed into service, because the quantities needed in
a device are so minute. A billion thin film devices each needing a few nanograms
of some new magnetic material consume just a few grams of an unrecoverable
resource.
v
vi
Preface
This handbook aims to offer a broad perspective on the state of the art in
magnetism and magnetic materials. The discovery and dissemination of reliable
knowledge about the natural world is a complex process that depends on interactions
of individuals with shared values and presumptions. Information is the primary
product of their endeavour. It is contained in in papers, patents, reviews, handbooks
monographs and textbooks. This is a perpetual work in progress. Knowledge
percolates through this sequence, taking ever-more digestible and definitive forms
as it is consolidated or eliminated. Now information technology is facilitating this
dynamic. Whereas papers are replaced by more up-to-date papers with new sets
of references to trace their pedigree and textbooks may be updated perhaps after
10 years, handbooks are compendia of information that need updating on a shorter
timescale. This was impractical within the constraints of traditional publication, but
the greater flexibility of electronic publication now opens the possibility for authors
to update their contributions as time passes, and perspectives shift.
The book’s 34 chapters are organized into four parts. After an introduction to the
history and basic concepts in the field, there follow 12 chapters covering the fundamentals of solid state magnetism, and the phenomena related to collective magnetic
order. Eight chapters are then devoted to the main classes of magnetic materials –
elements, metallic compounds, oxides and other nonmetallic compounds, thin films,
nanoparticles and artificially engineered materials. Another six chapters treat the
methods for preparing and characterizing magnetic materials, and the final part is
devoted to some major applications.
No fewer than 85 authors have contributed to this handbook. It has taken
longer than we originally anticipated, and the patience of the early responders is
sincerely appreciated. The format for subsequent updating of the electronic text is
by individual chapter, which will avoid such difficulty in the future.
We are grateful to the staff at Springer, Claus Ascheron for initiating the
project, Werner Skolaut for his patience and encouragement, and Barbara Wolf for
efficiently bringing the handbook to hand.
October 2021
J. M. D. Coey
Stuart S. P. Parkin
Contents
Volume 1
Part I Fundamentals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1
1
History of Magnetism and Basic Concepts . . . . . . . . . . . . . . . . . . . . .
J. M. D. Coey
3
2
Magnetic Exchange Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Ralph Skomski
53
3
Anisotropy and Crystal Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Ralph Skomski, Priyanka Manchanda, and Arti Kashyap
103
4
Electronic Structure: Metals and Insulators . . . . . . . . . . . . . . . . . . . .
Hubert Ebert, Sergiy Mankovsky, and Sebastian Wimmer
187
5
Quantum Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Gabriel Aeppli and Philip Stamp
261
6
Spin Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sergej O. Demokritov and Andrei N. Slavin
281
7
Micromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lukas Exl, Dieter Suess, and Thomas Schrefl
347
8
Magnetic Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Rudolf Schäfer
391
9
Magnetotransport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Michael Ziese
435
10
Magneto-optics and Laser-Induced Dynamics of Metallic Thin
Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Mark L. M. Lalieu and Bert Koopmans
11
Magnetostriction and Magnetoelasticity . . . . . . . . . . . . . . . . . . . . . . .
Dirk Sander
477
549
vii
viii
Contents
12
Magnetoelectrics and Multiferroics . . . . . . . . . . . . . . . . . . . . . . . . . . .
Jia-Mian Hu and Long-Qing Chen
595
13
Magnetism and Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Ilya M. Eremin, Johannes Knolle, and Roderich Moessner
625
Part II
Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
657
14
Magnetism of the Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Plamen Stamenov
659
15
Metallic Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
J. Ping Liu, Matthew Willard, Wei Tang, Ekkes Brück, Frank de
Boer, Enke Liu, Jian Liu, Claudia Felser, Gerhard Fecher, Lukas
Wollmann, Olivier Isnard, Emil Burzo, Sam Liu, J. F. Herbst,
Fengxia Hu, Yao Liu, Jirong Sun, Baogen Shen, and Anne de
Visser
693
16
Metallic Magnetic Thin Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Wu and X.-F. Jin
809
Volume 2
17
Magnetic Oxides and Other Compounds . . . . . . . . . . . . . . . . . . . . . . .
J. M. D. Coey
847
18
Dilute Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Alberta Bonanni, Tomasz Dietl, and Hideo Ohno
923
19
Single-Molecule Magnets and Molecular Quantum Spintronics . . .
Gheorghe Taran, Edgar Bonet, and Wolfgang Wernsdorfer
979
20
Magnetic Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1011
Sara A. Majetich
21
Artificially Engineered Magnetic Materials . . . . . . . . . . . . . . . . . . . . 1047
Christopher H. Marrows
Part III
Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1081
22
Magnetic Fields and Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 1083
Oliver Portugall, Steffen Krämer, and Yurii Skourski
23
Material Preparation and Thin Film Growth . . . . . . . . . . . . . . . . . . . 1153
Amilcar Bedoya-Pinto, Kai Chang, Mahesh G. Samant, and
Stuart S. P. Parkin
24
Magnetic Imaging and Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1203
Robert M. Reeve, Hans-Joachim Elmers, Felix Büttner, and
Mathias Kläui
Contents
ix
25
Magnetic Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1255
Jeffrey W. Lynn and Bernhard Keimer
26
Electron Paramagnetic and Ferromagnetic Resonance . . . . . . . . . . . 1297
David Menard and Robert Barklie
27
Magnetization Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1333
Andrew D. Kent, Hendrik Ohldag, Hermann A. Dürr, and
Jonathan Z. Sun
Part IV
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1367
28
Permanent Magnet Materials and Applications . . . . . . . . . . . . . . . . . 1369
Karl-Hartmut Müller, Simon Sawatzki, Roland Gauß and
Oliver Gutfleisch
29
Soft Magnetic Materials and Applications . . . . . . . . . . . . . . . . . . . . . . 1435
Frédéric Mazaleyrat
30
Magnetocaloric Materials and Applications . . . . . . . . . . . . . . . . . . . . 1489
Karl G. Sandeman and So Takei
31
Magnetic Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1527
Myriam Pannetier-Lecoeur and Claude Fermon
32
Magnetic Memory and Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1553
Wei Han
33
Magnetochemistry and Magnetic Separation . . . . . . . . . . . . . . . . . . . 1593
Peter Dunne
34
Magnetism and Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1633
Nora M. Dempsey
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1679
About the Editors
Michael Coey was born in Belfast in 1945. He studied physics at Cambridge, and then taught English
and physics at the Sainik School, Balachadi (Gujarat).
There he read Allan Morrish’s Physical Principles of
Magnetism from cover to cover (while recovering from
jaundice) before moving to Canada in 1968 to join Morrish’s group at the University of Manitoba for a PhD on
Mõssbauer spectroscopy of iron oxides. He has worked
on magnetism ever since – a life of paid play. After
graduating in 1971, he joined Benoy Chakraverty’s
group at the CNRS in Grenoble as a postdoc with a
letter of appointment signed by Louis Néel. Entering
the CNRS the following year, he worked on the metalinsulator as well as the magnetism of amorphous solids
and natural minerals. In France, he built the network
of collaborators which sustained much of his career.
On a sabbatical with Stefan von Molnar at the IBM
Research Center at Yorktown Heights, he learned about
magneto-transport and the crystal field. Then, in 1979,
he moved to Ireland as a lecturer at Trinity College
Dublin and set about establishing a magnetism research
group in a venerable but woefully underfunded Physics
Department. Luckily, support from the EU substitution programme enabled him to begin research on
melt-spun magnetic glasses. Following the discovery
of Nd2 Fe14 B permanent magnets in 1984, he and
colleagues from Grenoble, Birmingham and Berlin
launched the Concerted European Action on Magnets.
xi
xii
About the Editors
CEAM blossomed into an informal association of 90
academic and industrial research institutes interested in
every aspect of the properties, processing and applications of rare-earth iron permanent magnets. He and
his student Sun Hong discovered the interstitial nitride
magnet Sm2 Fe17 N3 in 1990. The group investigated
other rare-earth intermetallic compounds, as well as
magnetic oxide films produced by pulsed-laser deposition. During this period, he and David Hurley started up
Magnetic Solutions to develop innovative applications
of permanent magnets.
The scientific landscape in Ireland was transformed
by the establishment of Science Foundation Ireland in
2000, given the mission of developing competitive scientific research in Ireland with a budget to match. His
group were able to develop a programme in thin film
magnetism and spin electronics, producing Europe’s
first magnetic tunnel junctions to exhibit 200 % tunnel magnetoresistance. Later they discovered the first
zero-moment ferrimagnetic half-metal and explored the
garden of magneto-electrochemistry. Michael coey was
a promotor of CRANN, Ireland’s nanoscience research
centre, and the Science Gallery, now an international
franchise, was his brainchild. Together with Dominique
Givord, he launched the Joint European Magnetic Symposia (JEMS) and, while chair of C9, the IUPAP Magnetism Committee, inaugurated the Néel medal that is
awarded triennially at the International Conference on
Magnetism. The 2015 JEMS meeting in Dublin saw a
reunion of many of his 60 PhD students, from all over
the world. Together they have published many papers.
Books include Magnetic Glasses, 1984 (with Kishin
Moorjani): Permanent Magnetism, 1999 (with Ralph
Skomski): and Magnetism and Magnetic Materials,
2010. Honours include Fellowship of the Royal Society, International membership of the National Academy
of Sciences, a Fulbright fellowship, a Humboldt Prize,
the Gold Medal of the Royal Irish Academy and the
2019 Born Medal. He has enjoyed visiting professorships at the University of Strasbourg, the National
University of Singapore and Beihang University in
Beijing.
Michael Coey married Wong May, a writer, in 1973;
they have two sons and a grand-daughter.
About the Editors
xiii
Stuart S. P. Parkin is a director of the Max Planck
Institute of Microstructure Physics, Halle, Germany,
and an Alexander von Humboldt Professor, Martin Luther University, Halle-Wittenberg. His research
interests include spintronic materials and devices for
advanced sensor, memory and logic applications, oxide
thin-film heterostructures, topological metals, exotic
superconductors, and cognitive devices. Stuart’s discoveries in spintronics enabled a more than 10,000fold increase in the storage capacity of magnetic disk
drives. For his work that, thereby, enabled the ‘big
data’ world of today. In 2014, he was awarded the
Millennium Technology Award from the Technology
Academy Finland and, most recently, the King Faisal
Prize for Science 2021 for his research into three
distinct classes of spintronic memories. Stuart is a
fellow or member of: The Royal Society, the Royal
Academy of Engineering, the National Academy of
Sciences, the National Academy of Engineering, the
German National Academy of Science – Leopoldina,
The Royal Society of Edinburgh, The Indian Academy
of Sciences, and TWAS – The academy of sciences
for the developing world. Stuart is also a fellow of
the American Physical Society: the Institute of Electrical and Electronics Engineers (IEEE) the Institute
of Physics, London: the American Association for the
Advancement of Science (AAAS); and the Materials
Research Society. Stuart has published more than 600
papers and has more than 121 issued patents. His h
factor is 120. Clarivate Analytics named him a Highly
Cited Researcher in 2018, 2019, 2020 and 2021.
Stuart’s numerous awards include the American
Physical Society International Prize for New Materials (1994); the Europhysics Prize for Outstanding
Achievement in Solid State Physics (1997); the 2009
IUPAP Magnetism Prize and Néel Medal; the 2012
von Hippel Award – Materials Research Society; the
2013 Swan Medal – Institute of Physics; an Alexander
von Humboldt Professorship – International Award for
Research (2014); and ERC Advanced Grant – SORBET (2015). Stuart has been a distinguished visiting
professor at several universities worldwide including:
National University of Singapore; National Taiwan
University; National Yunlin University of Science and
Technology, Taiwan; Eindhoven University of Tech-
xiv
About the Editors
nology, The Netherlands; KAIST, Korea; and University College London. Stuart has been awarded four
honorary doctorates by: RWTH Aachen University
(2007), Eindhoven University of Technology (2008),
The University of Regensburg (2011), and Technische
Universität Kaiserslautern, Germany (2013).
Prior to being appointed to the Max Planck Society,
Stuart had spent a large part of his career with IBM
Research at the San Jose Research Laboratory, which
became the Almaden Research Center when it moved
to a new campus. Stuart was appointed an IBM Fellow,
IBM’s highest technical honour, by IBM’s chairman,
Louis Gerstner in 1999. He received his BA physics
and theoretical physics (1977), an MA, and his PhD
(1980) from the University of Cambridge. He was a student at Trinity College, Cambridge, where he received
an entrance scholarship (1974), a senior scholarship
(1975), a research scholarship (1977) and was elected
a research fellow (1979). In 2014, he became an honorary fellow. Stuart received a Royal Society European
Exchange Fellowship to carry out postdoctoral research
at the Laboratoire de Physique des Solides, Université
Paris-Sud, France, in 1980–1981 and an IBM World
Trade Fellowship to carry out research at IBM in San
Jose.
Contributors
Gabriel Aeppli Physics Department (ETHZ), Institut de Physique (EPFL) and
Photon Science Division (PSI), ETHZ, EPFL and PSI, Zürich, Lausanne and
Villigen, Switzerland
Robert Barklie School of Physics, Trinity College, Dublin, Ireland
Amilcar Bedoya-Pinto Max Planck Institute of Microstructure Physics, Halle
(Saale), Germany
Alberta Bonanni Institut für Halbleiter- und Festkörperphysik, Johannes Kepler
University, Linz, Austria
Edgar Bonet Néel Institute, CNRS, Grenoble, France
Ekkes Brück Delft University of Technology, Delft, The Netherlands
Emil Burzo Babes-Bolyai University, Romania, Cluj-Napoca, Romania
Felix Büttner Helmholtz-Zentrum Berlin für Materialien und Energie, Berlin,
Germany
Kai Chang Beijing Academy of Quantum Information Sciences, Beijing, China
Long-Qing Chen Materials Research Institute, and Department of Materials Science and Engineering, The Pennsylvania State University, University Park, PA, USA
Michael Coey School of Physics, Trinity College, Dublin, Ireland
Frank de Boer University of Amsterdam, Amsterdam, The Netherlands
Anne de Visser Van der Waals-Zeeman Institute, University of Amsterdam,
Amsterdam, The Netherlands
Sergej O. Demokritov Institute for Applied Physics and Center for Nanotechnology, University of Muenster, Muenster, Germany
Nora M. Dempsey Institut Néel, CNRS & Université Grenoble Alpes, Grenoble,
France
xv
xvi
Contributors
Tomasz Dietl International Research Centre MagTop, Institute of Physics, Polish
Academy of Sciences, Warsaw, Poland
WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan
Peter Dunne Institut de Physique et de Chimie des Matériaux de Stasbourg,
Strasbourg, France
Hermann A. Dürr Department of Physics and Astronomy, Uppsala University,
Uppsala, Sweden
Hubert Ebert München, Department Chemie, Ludwig-Maximilians-Universität,
München, Germany
Hans-Joachim Elmers Institute of Physics, Johannes Gutenberg University Mainz,
Mainz, Germany
Ilya M. Eremin Institut für Theoretische Physik III, Ruhr-Universität Bochum,
Bochum, Germany
Lukas Exl University of Vienna Research Platform MMM Mathematics – Magnetism – Materials, University of Vienna, and Wolfgang Pauli Institute, Wien,
Austria
Gerhard Fecher Max-Planck-Institute für Chemische Physik fester Stoffe, Dresden, Germany
Claudia Felser Max-Planck-Institute für Chemische Physik fester Stoffe, Dresden,
Germany
Claude Fermon Service de Physique de l’Etat Condensé, DRF/IRAMIS/SPEC
CNRS UMR 3680 CEA Saclay, Gif sur Yvette, France
Roland Gauß EIT RawMaterials GmbH, Berlin, Germany
Oliver Gutfleisch Technische Universität Darmstadt, Materialwissenschaft, Darmstadt, Germany
Wei Han International Center for Quantum Materials, School of Physics, Peking
University, Beijing, China
J. F. Herbst Research & Development, General Motors R&D Center, Warren,
MI, USA
Fengxia Hu Institute of Physics, Chinese Academy of Sciences, Beijing, China
Jia-Mian Hu Department of Materials Science and Engineering, University of
Wisconsin-Madison, Madison, WI, USA
Olivier Isnard Institute Néel and Université Grenoble Alpes, Grenoble, France
X.-F. Jin Department of Physics and State Key Laboratory of Surface Physics,
Fudan University, Shanghai, People’s Republic of China
Arti Kashyap IIT Mandi, Mandi, HP, India
Contributors
xvii
Bernhard Keimer Max-Planck Institute for Solid State Research,
Germany
Stuttgart,
Andrew D. Kent Center for Quantum Phenomena, Department of Physics, New
York University, New York, NY, USA
Mathias Kläui Institute of Physics, Johannes Gutenberg University Mainz, Mainz,
Germany
Johannes Knolle Blackett Laboratory, Imperial College London, London, UK
Bert Koopmans Department of Applied Physics, Eindhoven University of Technology, Eindhoven, The Netherlands
Steffen Krämer LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA
Toulouse, Univ. Toulouse 3, Grenoble, France
Mark L. M. Lalieu Department of Applied Physics, Eindhoven University of
Technology, Eindhoven, The Netherlands
Enke Liu Institute of Physics, Chinese Academy of Sciences, Beijing, China
J. Ping Liu University of Texas at Arlington, Arlington, TX, USA
Jian Liu Ningbo Institute of Materials Technology and Engineering, Chinese
Academy of Sciences, Ningbo, China
Sam Liu University of Dayton, Dayton, OH, USA
Yao Liu Institute of Physics, Chinese Academy of Sciences, Beijing, China
Jeffrey W. Lynn NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, MD, USA
Sara A. Majetich Physics Department, Carnegie Mellon University, Pittsburgh,
PA, USA
Priyanka Manchanda Howard University, Washington, DC, USA
Sergiy Mankovsky München,
Universität, München, Germany
Department
Chemie,
Ludwig-Maximilians-
Christopher H. Marrows School of Physics and Astronomy, University of Leeds,
Leeds, United Kingdom
Frédéric Mazaleyrat SATIE, CNRS, École Normale Supérieure Paris-Saclay,
Gif-sur-Yvette, France
David Menard Department of Engineering Physics, Polytechnique Montreal, Montréal, QC, Canada
Roderich Moessner Max-Planck Institut für Physik komplexer Systeme, Dresden,
Germany
Karl-Hartmut Müller IFW Dresden, Institute for Metallic Materials, Dresden,
Germany
xviii
Contributors
Hendrik Ohldag Advanced Light Source, Lawrence Berkeley National Laboratory,
Berkeley, CA, USA
Department of Physics, University of California Santa Cruz, Santa Cruz, CA,
USA
Department of Materials Science, Stanford University, Stanford, CA, USA
Hideo Ohno WPI Advanced Institute for Materials Research, Tohoku University,
Sendai, Japan
Laboratory for Nanoelectronics and Spintronics, Research Institute of Electrical
Communication, Tohoku University, Sendai, Japan
Center for Spintronics Integrated System, Tohoku University, Sendai, Japan
Center for Innovative Integrated Electronic Systems, Tohoku University, Sendai,
Japan
Center for Science and Innovation in Spintronics (Core Research Cluster), Tohoku
University, Sendai, Japan
Center for Spintronics Research Network, Tohoku University, Sendai, Japan
Myriam Pannetier-Lecoeur Service de Physique de l’Etat Condensé, DRF/
IRAMIS/SPEC CNRS UMR 3680 CEA Saclay, Gif sur Yvette, France
Stuart S. P. Parkin Max Planck Institute of Microstructure Physics, Halle (Saale),
Germany
Oliver Portugall LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA
Toulouse, Univ. Toulouse 3, Toulouse, France
Robert M. Reeve Institute of Physics, Johannes Gutenberg University Mainz,
Mainz, Germany
Mahesh G. Samant IBM Research, San Jose, CA, USA
Karl G. Sandeman Department of Physics, Brooklyn College of the City
University of New York, Brooklyn, NY, USA
The Physics Program, The Graduate Center, CUNY, New York, NY, USA
Dirk Sander Max Planck Institute of Microstructure Physics, Halle, Germany
Simon Sawatzki Technische
Darmstadt, Germany
Universität
Darmstadt,
Materialwissenschaft,
Vacuumschmelze GmbH & Co.KG, Hanau, Germany
Rudolf Schäfer Institute for Metallic Materials, Leibniz Institute for Solid State
and Materials Research (IFW) Dresden, Dresden, Germany
Institute for Materials Science, Dresden University of Technology, Dresden,
Germany
Contributors
xix
Thomas Schrefl Christian Doppler Laboratory for Magnet Design Through Physics
Informed Machine Learning, Department of Integrated Sensor Systems, Danube
University Krems, Wiener Neustadt, Austria
Baogen Shen Institute of Physics, Chinese Academy of Sciences, Beijing, China
Ralph Skomski University of Nebraska, Lincoln, NE, USA
Yurii Skourski Hochfeld-Magnetlabor Dresden (EMFL-HLD), HelmholtzZentrum Dresden-Rossendorf, Dresden, Germany
Andrei N. Slavin Department of Physics, Oakland University, Rochester, MI, USA
Plamen Stamenov School of Physics and CRANN, Trinity College, University of
Dublin, Dublin, Ireland
Philip Stamp Pacific Institute of Theoretical Physics, University of British
Columbia, Vancouver, BC, Canada
Dieter Suess University of Vienna Research Platform MMM Mathematics – Magnetism – Materials, and Physics of Functional Materials, Faculty of Physics,
University of Vienna,Wien, Austria
Jirong Sun Institute of Physics, Chinese Academy of Sciences, Beijing, China
Jonathan Z. Sun IBM T. J. Watson Research Center, Yorktown Heights, NY, USA
So Takei The Physics Program, The Graduate Center, CUNY, New York, NY, USA
Department of Physics, Queens College of the City University of New York,
Flushing, NY, USA
Wei Tang Materials Science and Engineering, Ames Laboratory, Ames, IA, USA
Gheorghe Taran Physikalisches Institute, KIT, Karlsruhe, Germany
Wolfgang Wernsdorfer Physikalisches Institute, KIT, Karlsruhe, Germany
Matthew Willard Materials Science and Engineering, Case Western Reserve
University, Cleveland, OH, USA
Sebastian Wimmer München,
Universität, München, Germany
Department
Chemie,
Ludwig-Maximilians-
Lukas Wollmann Max-Planck-Institute für Chemische Physik fester Stoffe,
Dresden, Germany
D. Wu National Laboratory of Solid State Microstructures and Department of
Physics, Nanjing University, Nanjing, People’s Republic of China
Michael Ziese Fakultät für Physik und Geowissenschaften, Universität Leipzig,
Leipzig, Germany
Part I
Fundamentals
1
History of Magnetism and Basic Concepts
J. M. D. Coey
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Early History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Compass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Emergence of Modern Science . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Electromagnetic Revolution [9] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetostatics and Classical Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Earth’s Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Properties of Ferromagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetism of the Electron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Demise of Classical Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Phenomenology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Micromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Intermetallic Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Model Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Amorphous Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Fine Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Recording . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Methods of Investigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Materials Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendix: Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4
5
5
7
9
9
16
18
20
20
23
25
27
29
32
34
35
36
38
40
40
41
43
43
48
48
50
J. M. D. Coey ()
School of Physics, Trinity College, Dublin, Ireland
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_1
3
4
J. M. D. Coey
Abstract
Magnetism is a microcosm of the history of science over more than two
millennia. The magnet allows us to manipulate a force field which has catalyzed
an understanding of the natural world that launched three revolutions. First
came the harnessing of the directional nature of the magnetic force in the
compass that led to the exploration of the planet in the fifteenth century.
Second was the discovery of the relation between electricity and magnetism that
sparked the electromagnetic revolution of the nineteenth century. Third is the
big data revolution that is currently redefining human experience while radically
transforming social interactions and redistributing knowledge and power.
The emergence of magnetic science demanded imagination and observational
acuity, which led to the theory of classical electrodynamics. The magnetic field is
associated with electric currents and the angular momentum of charged particles
in special materials. Our current understanding of the magnetism of electrons
in solids is rooted in quantum mechanics and relativity. Yet only since about
1980 has fundamental theory underpinned rational design of new functional
magnetic materials and the conception of new spin electronic devices that can
be reproduced on ever smaller scales, leading most notably to the disruptive,
60-year exponential growth of magnetic information storage. The development
of new magnetic concepts, coupled with novel materials, device and machine
designs has become a rich source of technical innovation.
Introduction
The attraction of ferrous objects to a permanent magnet has been a source of wonder
since the Iron Age. Feeble magnets are widespread in nature in the form of rocks
known as lodestones, which are rich in magnetite, an oxide mineral with ideal
formula Fe3 O4 . Rocky outcrops eventually get magnetized by huge electric currents
when lightning strikes, and these natural magnets were known and studied in ancient
Greece, Egypt, China, and Mesoamerica. Investigations of magnetic phenomena led
to the invention of steel magnets – needles and horseshoes – then electromagnets and
eventually the panoply of hard and soft materials that support the modern magnetics
industry. Magnetism in a rare example of a science with recorded history goes back
well over 2000 years [1, 2].
Theory and practice have been loose partners for most of that time. What people
are able to see and rationalize is inevitably conditioned by a priori philosophical
beliefs about the world. The scientific method of critically interrogating nature
by experimentation and then amassing and exchanging data and ideas among the
community of the curious came to be established only gradually. Mathematics
emerged as the supporting scaffold of natural philosophy in Europe in the seventeenth century, when precisely formulated natural laws and explanations began to
take root. Nevertheless, most of the progress that has been made in magnetism in
the past – from the discovery of horseshoe magnets or electromagnetic induction
1 History of Magnetism and Basic Concepts
5
to the development of Alnico – was based on intuition and experience, rather than
formal theory. That situation is changing.
The discovery of the electron in the closing years of the nineteenth century
impelled the great paradigm shift from classical to modern physics. Magnetism,
however familiar and practically important it had become, was fundamentally
incomprehensible in classical terms. Charged particles were theoretically expected
to exhibit no magnetism of any kind. It took 25 years and the insights of quantum
mechanics and relativity to resolve that conundrum. Magnetism then went on to play
a key role in clarifying basic concepts in condensed matter physics and Earth science
over the course of the twentieth century. Now it is a key player in the transformative
information technology of the twenty-first century.
Early History
Aristotle attributed the first reflections on the nature of magnetic attraction to Thales,
the early Greek philosopher and mathematician who was born in Miletus in Asia
Minor in 624 BC. Thales was an animist who credited the magnet with a soul,
on account of its ability to create movement, by attraction. This curious idea was
to linger until the seventeenth century. The magnet itself is believed to be named
after Magnesia, a city in Lydia in Asia Minor that was a good source of lodestone.
In the fifth century BC, when Empedokles postulated the existence of the four
elements – Earth, water, air, and fire – magnetism was associated with air. Special
effluvia somehow passing through the invisible pores in magnetic material were
invoked to explain the phenomenon, a theory echoed much later by Descartes in
a mechanistic picture that finally laid the magnet’s soul to rest. The Roman poet
Lucretius writing in the first century BC mentions magnetic induction (the ability
of a magnet to induce magnetism in pieces of nonmagnetic iron) and for the first
time notes the ability of magnets not just to attract but also to repel one another.
The Greek approach of developing a philosophical framework into which natural
observations were expected to fit was not conducive to open-minded exploration of
the natural world.
The Compass
The Chinese approach to the magnet was more practical. Their magnetism was
initially linked to practical concerns of geomancy and divination [3]. The art of
adapting the residences of the living and the tombs of the dead to harmonize
with local currents of the cosmic breath demanded knowledge of its direction. A
south-pointer consisting of a carved lodestone spoon that was free to rotate on a
polished baseplate (Fig. 1) was already in use at the time of Lucretius and may have
originated hundreds of years earlier. An important discovery, attributed to Zeng
Gongliang in 1064, was that iron could acquire a thermoremanent magnetization
when rapidly cooled from red heat in the Earth’s magnetic field. A short step
6
J. M. D. Coey
Fig. 1 Magnetic direction finders. (a) Baseplate and lodestone spoon of the south-pointer used in
China from about the first century BC (Needham, courtesy of Cambridge University Press). (b) A
Chinese floating compass from 1044. (c) Fifteenth-century Chinese and (d) Portuguese mariners’
compasses. (Boorstin, courtesy of Editions Robert Laffont)
led to the suspended compass needle, which was described by Shen Kuo around
1088, together with declination, the deviation of the needle from a north-south axis.
Floating compasses had also been developed by this time, often in the form of an
iron fish made to float in a bowl of water.
The compass appeared about a century later in Europe, where it was first
described by Alexander Neckam in 1190. The direction-finding ability of the
magnetic needle or fish was also exploited by Arabs and Persians from the thirteenth
century, both for navigation and to determine the sacred direction of Mecca [4].
Compasses (Fig. 1) were the enabling technology for the great voyages of discovery
of the fifteenth century, bringing the Ming admiral Cheng Ho to the coasts of Africa
in 1433 and Christopher Columbus (who rediscovered declination) to America in
1492, where he landed on the continent where the Olmecs may once have displayed
a knowledge of magnetism in their massive stone carvings of human figures and sea
turtles dating from the second millennium BC.
1 History of Magnetism and Basic Concepts
7
Before long, the landmasses and oceans of our planet were mapped and explored.
According to Francis Bacon, writing in Novum Organum in 1620 [5], the magnetic
compass was one of three things, along with printing and gunpowder had “changed
the whole face and state of things throughout the world.” All three were originally
Chinese inventions. The compass helped to provide us with an image of the planet
we inhabit. This was the first of three occasions when magnetism changed the world.
The Emergence of Modern Science
A landmark in the history of magnetism in Europe was the work of the French
crusader monk Petrus Peregrinus. His tract Epistola de Magnete [6] recounts
experiments with floating pieces of lodestone and carved lodestone spheres called
terella, which he wrote up in Southern Italy during the 1269 siege of Lucera. He
describes how to find the poles of a magnet and relates magnetic attraction to
the celestial sphere. The same origin had long been associated with the magnet’s
directional property in China [3]; we should not forget that before electric light,
people were acutely aware of the stars and scrutinized them keenly. Peregrinus’s
tract included an ingenious proposal for a magnetic perpetual motion device – a
theme that has been embraced by charlatans throughout the ages, right up to the
present day.
Much credit for the inauguration of the experimental method in a recognizably
modern form belongs to William Gilbert. Physician to the English Queen Elizabeth
I, Gilbert personally conducted a series of experiments on terellas, which led him to
proclaim that the Earth itself was a great magnet. The lodestone or steel magnets
aligned themselves not with the celestial sphere, but with the Earth’s poles. He
induced magnetism by cooling iron in the Earth’s field and then destroyed it by
heating or hammering. Gilbert was at pains to debunk the millennial accretion of
superstition that clung to the magnet, confidently advocating in a robust polemical
style reliance on the evidence of one’s own eyes. He described his investigations in
his masterwork De Magnete, published in 1600 [7]. It is arguably the first modern
scientific text.
Subsequent developments were associated with improvements in navigation and
the prestige of the great voyages of discovery. Gilbert’s theories dominated the
seventeenth century up until Edmond Halley’s 1692 shell model for the Earth’s
magnetic structure, which strongly influenced compass technology and navigation.
Naval interests were the principal drivers of magnetic research during this period,
and Halley was sponsored by the British Navy to survey and prepare charts of
the Earth’s magnetic field in the North and South Atlantic oceans (Fig. 2), This
was in the vain hope of addressing the pressing longitude problem, by pinpointing
magnetically the position of a vessel on the Earth’s surface.
The following century was marked by the professionalization of natural philosophy (as physical science was then known in Europe) [8]. Accordingly, the natural
philosopher with his mantle of theory was rewarded with social status, access to
public funding, and credibility beyond that extended to artisans on the one hand and
8
J. M. D. Coey
Fig. 2 A section of Halley’s world chart of magnetic variation published in 1700
quacks on the other, such as the colorful Anton Mesmer, who propagated theories of
animal magnetism in his salon in Paris or James Graham with his royal Patagonian
magnetic bed for nightly rental in a fashionable London townhouse. The English
entrepreneur Gowin Knight, representative of a new breed of natural philosopher,
greatly improved the quality of bar magnets and compasses, coupling scientific
endeavor with manufacturing enterprise and a keen sense of intellectual property.
An outstanding technical breakthrough of the eighteenth century was the 1755
discovery by the Swiss blacksmith Johann Dietrich that the horseshoe was an ideal
1 History of Magnetism and Basic Concepts
9
shape for a steel magnet [1]. His invention, a clever practical solution to the age-old
problem of self-demagnetization in bar magnets, was enthusiastically promoted by
his mentor, the Swiss applied mathematician Daniel Bernoulli, who garnered most
of the credit.
The Electromagnetic Revolution [9]
The late eighteenth century in Europe was a time of great public appetite for lectures
and demonstration of the latest scientific discoveries, not least in electricity and
magnetism. This effervescent age witnessed rapid developments in the harnessing
of electricity, with the 1745 invention of the Leyden jar culminating in Alessandro
Volta’s 1800 invention of the voltaic cell. Analogies between electrostatics and
magnetism were tantalizing, but the link between them proved elusive.
Magnetostatics and Classical Electrodynamics
The torsion balance allowed Charles-Augustin de Coulomb to establish in 1785
the quantitative inverse square laws of attraction and repulsion between electric
charges, as well as similar laws between analogous magnetic charge or poles that
were supposed to be located near the ends of long magnetized steel needles [2]. The
current convention is that the north and south magnetic poles are negatively and
positively charged, respectively. His image was of pairs of positive and negative
electric and magnetic fluids permeating matter, which became charged if one
of them dominated or polarized if they were spatially separated. Unlike their
electric counterparts, the magnetic fluids were not free to flow and could never
be unbalanced in any piece of magnetic material. Coulomb found that the force F
between two magnetic poles separated by a distance r fell away as 1/r2 . Siméon
Denis Poisson then interpreted Coulomb’s results in terms of a scalar potential
ϕm , analogous to the one he used for static electricity, such that the magnetic field
could be written as H(r) = −∇ϕm . In modern terms, ϕm is measured in amperes,
and H in Am−1 . Magnetic charge qm is measured in Am, and the corresponding
potential ϕm = qm /4πr. The magnetic field due to a charge is H(r) = qm r/4πr3 ,
and Coulomb’s inverse square law for the force between two charges separated by
r is F = μ0 qm qm ’r/4πr3 . Here μ0 is the magnetic constant, 4π 10−7 NA−2 , which
appears whenever the magnetic field H interacts with matter. (Other equivalent ways
of writing the units of μ0 are Hm−1 or TmA−1 .)
In Poisson’s opinion, the practice and teaching of mathematics were the purpose
of life. He developed his mathematical theory of magnetostatics from 1824, which
included the equation that bears his name ∇ 2 ϕm = −ρm , where ρm is the density of
magnetic poles. However, the association of H with a scalar potential is only valid
in a steady state and when no electric currents are present. The coulombian picture
of the origin of magnetic fields was dominant in textbooks until about 1960, and it
persists in popular imagery.
10
J. M. D. Coey
A revolutionary breakthrough in the history of magnetism came on 21st April
1820, with the discovery of the long-sought link between electricity and magnetism.
During a public lecture, the Danish scientist Hans Christian Oersted noticed that a
compass needle was deflected as he switched on an electric current in a copper
wire. His report, published in Latin a few months later, triggered an experimental
frenzy. As soon as the news reached Paris, François Arago (who briefly served as
President of France in 1848) immediately performed an experiment that established
that a current-carrying conducting coil behaved like a magnet. A week after Arago’s
report, André-Marie Ampère presented a paper to the French Academy suggesting
that ferromagnetism in a magnetized body was caused by internal currents flowing
perpendicular to the axis of magnetization and that it should therefore be possible
to magnetize steel needles in a solenoid. Together with Arago, he successfully
demonstrated his ideas in November 1820, showing that current loops and coils
were functionally equivalent to magnets, and he subsequently established the
law of attraction or repulsion between current-carrying wires. Ten days later,
the British scientist Humphrey Davy had similar results. The electromagnet was
invented by William Sturgeon in 1825; within 5 years Joseph Henry had used a
powerful electromagnet in the USA for the first electric telegraph. As early as 1822,
Davy’s assistant Michael Faraday produced the first rudimentary electric motor, and
Ampère envisaged the possibility that the currents causing magnetism in solids were
“molecular” rather than macroscopic in nature.
In formal terms, Ampère’s equivalence between a magnet and a current loop of
area A carrying a current I is expressed as
m = IA
(1)
where A is in square meters, I is in amperes, and the magnetic moment m is therefore
in Am2 . Magnetization, defined in a mesoscopic volume V as M = m/V, has units
Am−1 . The direction of m is conventionally related to that of the electric current
by the right-hand rule. At the same time as the experimental work of Ampère
and Arago, Jean-Baptiste Biot and Félix Savart formulated the law expressing the
relation between a current and the field it produces. A current element Iδl generates
a field δH = Iδl × r/4πr3 at a distance r. Integrating around a current loop yields an
expression for the H-field due to the moment m:
H = [3 (m.r) r − m] /4r 3
(2)
The form of the field represented by Eq. (2) and illustrated in Fig. 3 is identical
to that of an electric dipole, so m is often referred to as a magnetic dipole although
we have no evidence for the existence of independent magnetic poles. The dipole
moment is best represented by an arrow in the direction of m, although it is still
commonplace to see the north-seeking and south-seeking poles of a magnet denoted
by the letters N and S. Old habits die hard.
Magnetic moments tend to align with magnetic fields in which they are placed.
The torque on the dipole m is Γ = μ0 m × H, and the corresponding energy of the
1 History of Magnetism and Basic Concepts
11
Fig. 3 Contours of equal magnetic field produced by a magnetic dipole moment m, represented
by the grey arrow
dipole is E = − μ0 m . H. These equations are better written in terms of the more
fundamental magnetic field B, as discussed below; in free space the two are simply
proportional, B = μ0 H, so the torque is
=m×B
(3)
E = −m.B
(4)
and the corresponding energy is
The two rival descriptions of magnetization in solids following from the work of
Coulomb or Ampère, based either on magnetic poles or on electric currents, have
colored thinking about magnetism ever since (Fig. 4). The poles have no precise,
independent physical reality; they are fictitious entities that are a mathematicallyconvenient way to represent the H-field, which is of critical importance in magnetism because it is the local H-field that determines the state of magnetization
of a solid. Currents are closer to reality; electric current loops exist, and they do
act like magnets. Although it is difficult to attribute the intrinsic spin moment of
the electron to a current, the amperian picture of the origin of magnetic fields is
generally adopted in modern textbooks.
Nineteenth-century electromagnetism owed much to the genius of Michael
Faraday. Guided entirely by observation and experiment, with no dependence on
formal theory, he was able to perfect the concept of magnetic field, which he
12
+++++
σ m+
jms
-----
Fig. 4 Alternative
coulombian (left) and
amperian (right) descriptions
of the magnetization of a
uniformly magnetized
cylinder, with a magnetic
dipole moment m in the
direction represented by the
black arrow; σm ± is the
surface magnetic charge
density, jms is the surface
electric current density
J. M. D. Coey
σ m-
described by lines of force [10]. Faraday classified substances in three magnetic
categories. Ferromagnets like iron were spontaneously magnetized and strongly
attracted into a magnetic field; paramagnets were weakly magnetized by a field
and feebly drawn into the regions where the field was strongest; diamagnets, on
the contrary, were weakly magnetized opposite to the field and repelled by it.
Working with an electromagnet, he discovered the law that bears his name and
the phenomenon of electromagnetic induction – that a flow of electricity can be
induced by a changing magnetic field – in 1831. His conviction that a magnetic field
should have some effect on light led to his 1845 discovery of the magneto-optic
Faraday effect – that the plane of polarization of light rotates upon passing through
a transparent medium in a direction parallel to the magnetization of the medium.
The epitome of classical electrodynamics was the set of equations formulated
in 1865 by James Clerk Maxwell, the Scottish theoretician, who had “resolved
to read no mathematics on the subject till he had first read through Faraday’s
‘Experimental Researches in Electricity’.” Maxwell’s magnificent equations formally defined the relationship between electricity, magnetism, and light [11]. As
reformulated by Oliver Heaviside, the equations are a succinct statement of classical
electrodynamics. In the opinion of Richard Feynman, Maxwell’s discovery of the
laws of electrodynamics was the most significant event of the nineteenth century.
The equations in free space are formulated in terms of the fundamental magnetic
and electric fields B and E. Using the international system of SI units adopted in
this Handbook, the equations read:
∇.B = 0
ε0 ∇.E = ρ
(1/μ0 ) ∇ × B = j + ε0 ∂E/∂t
(5)
∇ × E = −∂B/∂t
The first and third equations express the idea that there are no sources of the
magnetic B-field other than time-varying electric fields and electric currents of
1 History of Magnetism and Basic Concepts
13
density j, whereas the second and fourth equations show that the electric field
results from electric charge density ρ and time-varying magnetic fields. Maxwell’s
equations are invariant in a moving frame of reference, although the relative
magnitudes of E and B are altered.
The famous wavelike solutions of these equations in the absence of charges and
currents are electromagnetic waves, which propagate in free space with velocity
c = 1/(ε0 μ0 )1/2 . In SI, the definition of the magnetic constant μ0 is linked to the
fine structure constant. To nine significant figures, it is equal to 4π 10−7 NA−2 . ε0
is then related to the definition of the velocity of light. Heinrich Hertz demonstrated
Maxwell’s electromagnetic waves experimentally in 1888, and he showed that their
behavior was essentially the same as that of light. Hertz could think of no practical
application for his work, yet within a few decades, it had become the basis of radio
broadcasting and wireless communication!
The mechanical effects of electric and magnetic fields were summarized by
Hendrik Lorentz in his expression for the force density FL :
F L = ρE + j × B
(6)
The equivalent expression for the force on a particle of charge q moving with
velocity v is f = q(E + v × B).
Two further fields H and D are introduced in the formulation of Maxwell’s
equations in a material medium to circumvent the inaccessibility of the current and
charge distributions in the medium. We have no direct way of measuring the atomic
charges associated with the polarization of a ferroelectric material or the atomic
currents associated with the magnetization of a ferromagnetic material, so we define
H and D in terms of fields created by the measurable free charges ρ and free currents
j, with dipolar contributions from the magnetization M or polarization P of any
magnetic or dielectric material that may be present. The equations now read:
∇.B = 0
∇.D = ρ
∇ × H = j + ∂D/∂t
(7)
∇ × E = −∂B/∂t
They are further simplified in a static situation when the time derivatives are
zero. The new fields are trivially related to B and E in free space since B = μ0 H and
D = ε0 E, but in a material medium, the H-field is defined in terms of the B-field and
the magnetization M (the magnetic moment per unit volume) as H = B/μ0 – M or
B = μ0 (H + M)
(8)
14
J. M. D. Coey
H
M
B
+++++
–––––
B = P0(H + M)
Fig. 5 B, H, and M for a uniformly magnetized ferromagnetic bar. Eq. (8) is represented by the
vector triangle. The H-field can be regarded as originating from a distribution of positive and
negative magnetic charge (south and north magnetic poles) on opposite faces
Likewise D = ε0 (E + P), where P is the electric polarization. To specify a
situation in magnetostatics or electrostatics, any two of the three magnetic or electric
fields are needed. (Magnetization M and polarization P are regarded as vector
fields.) The defining relation between B, H, and M for a uniformly magnetized
ferromagnetic bar is illustrated in Fig. 5. Note that the B-field is solenoidal – the field
lines are continuous with no sources or sinks; it is divergenceless and can therefore
be expressed as the curl of a vector potential A – whereas the H-field is conservative;
it is irrotational provided j is zero and can be expressed as the gradient of a scalar
potential. Outside the magnet, the H-field is called the stray field, but within the
magnet where it is oppositely oriented to M, the name changes to demagnetizing
field. Boundary conditions that B⊥ and H|| are continuous across an interface in
a steady state (j = 0) follow from the first and third of Maxwell’s equations 7. B
is the fundamental magnetic field, because no elementary magnetic poles exist in
nature (∇. B = 0), but it is the local value of H (and perhaps the sample history)
that determines the magnetic state of a solid, including its micromagnetic domain
structure. The H-field acting in a solid is the sum of the applied field H and the
local demagnetizing field Hd created by the solid body itself.
When describing the stray field outside a distribution of magnetization M(r)
in a solid, the coulombian and amperian descriptions are formally equivalent.
The coulombian expression for the magnetic field is obtained by integrating the
expression for the field due to a distribution of a magnetic charge qm per unit volume
ρm = −∇. M in the bulk, and per unit area σm = M. en at the surface, where en is
the unit vector normal to the surface:
1 History of Magnetism and Basic Concepts
1
H (r) =
4π
−
15
∇ · M r − r
V
|r − r |3
3 M · en r − r d r +
S
|r − r |3
2 d r
(9)
This formula gives H(r) both inside and outside the magnetic material. Outside
B(r) = μ0 H(r).
The amperian expression for the magnetic field produced by a distribution of
currents is based on the Biot-Savart expression for the field due to a current element,
including contributions from the current density jm = ∇ × M in the bulk, and
jms = M × en at the surface:
μ0
B (r) =
4π
∇ × M × r − r
V
|r − r |3
3 (M × en ) × r − r d r +
S
|r − r |3
2 d r
(10)
This formula gives B(r) both inside and outside the magnetic material. The same
result can be obtained by appropriate integration of Eq. 2 over a magnetization
distribution M(r) [12].
For uniformly magnetized ellipsoids, the demagnetizing field Hd is related to the
magnetization by
H d = −N M
(11)
where N is a tensor with unit trace [13]. It reduces to a simple scalar demagnetizing
factor 0 < N < 1 when the magnetization lies along a principal axis of the ellipsoid.
N ≈ 0 for a long needle magnetized along its axis, and N = 1 for a flat plate
magnetized perpendicular to the plane. A sphere has N = 1/3. For any shape
less symmetric than an ellipsoid, the demagnetizing field is nonuniform. There are
useful approximate formulae for square bars and cylinders [14], such as 1/(2n + 1)
√
and 1/[(4n/ π) + 1], respectively, but they should not obscure the fact that the
demagnetizing field in these shapes really is quite nonuniform. Here n is the ratio of
length to diameter. The demagnetizing field is the reason why for centuries magnets
were condemned to take awkward shapes of bars or horseshoes to avoid substantial
self-demagnetization and why the most successful electromagnetic machines of
the nineteenth century were built around electromagnets rather than permanent
magnets. The hardened steel magnets of the day showed little coercivity and were
easily demagnetized. Demagnetizing fields are also the cause of ferromagnetic
domains. The shape constraint on permanent magnets was not lifted until the middle
of the twentieth century. Permanent magnets then came to the fore in the design of
electric motors and magnetic devices. Fig. 6 illustrates a collection of magnets from
the eighteenth, nineteenth, and twentieth centuries.
The imaginative world of Maxwell and his followers in the latter part of the
nineteenth century when the electromagnetic revolution was in full swing was
16
J. M. D. Coey
Fig. 6 Magnets from four centuries; top, seventeenth-century lodestone, nineteenth-century electromagnet; bottom, eighteenth-century horseshoe magnet, twentieth-century alnico and Nd2 Fe14 B
magnets (not to scale)
actually far removed from our own [15]. They envisaged light and other Hertzian
waves as propagating in an all-pervasive aether, which was believed to possess
magical mechanical properties – it had to be a massless incompressible fluid,
transparent and devoid of viscosity, yet millions of times more rigid than steel!
Elaborate mechanical models were envisaged for the waves and fields. In due course
it came to be understood that reality was represented by the abstract mathematics,
which remained after all the mechanical props had been discarded.
The Earth’s Magnetic Field
The Earth’s field was the prime focus of attention of magnetism for over a millennium, especially after it was understood that the magnetic field was of terrestrial
origin. By the beginning of the nineteenth century, the components of the field were
1 History of Magnetism and Basic Concepts
17
being recorded regularly in laboratories across the world. A comparison of the daily
magnetic records at Paris and Kazan, cities lying 4000 km apart, for the same day
in 1825, showed astonishingly similar short-term fluctuations. This inspired Carl
Friedrich Gauss to establish a worldwide network of 50 magnetic observatories,
coordinated from Göttingen, to make meticulous simultaneous measurements of the
Earth’s field, in the hope that if enough high-quality data could be collected, the
mystery of its origin and its fluctuations might be solved. This heroic pioneering
venture in international scientific collaboration amassed stores of data that were
enormous for that time. It inspired Gauss to develop spherical harmonic analysis,
from which he calculated that the leading, dipolar term accounted for about 90%
of the field and that the origin of the stable component was essentially internal.
Edward Sabine later spotted that the intensity of the short-term fluctuations tracked
the 11-year sunspot cycle, which we now know corresponds to reversals of the
solar magnetic field. But in its primary aim, Gauss’s Magnetische Verein must be
counted a failure. No amount of data, however copious and precise, could reveal a
deterministic origin of a phenomenon that was fundamentally chaotic. Piles of data
with no theory or hypothesis through which to view and be tested by them are not
very informative. This lesson was learned slowly.
The pole picture of the Earth’s magnetic field, albeit with poles that needed to
travel tens of kilometers every year to account for the secular variation, yielded
eventually in the academy if not in the popular imagination to one based on
electric currents driven by convection in the Earth’s liquid core. Joseph Larmor,
a dogged believer in the aether, was an early proponent of the geomagnetic dynamo.
He demonstrated the precession of a magnet in a magnetic field at a frequency
fL = γB/2π that bears his name. The precession is analogous to that of a spinning
top in a gravitational field; it is a consequence of the torque on a magnetic moment
expressed by Eq. 3. The constant γ, known as the gyromagnetic ratio, is the ratio of
the magnetic moment to its associated angular momentum. The proportionality of
these two quantities that at first sight appear quite dissimilar, the famous Einstein-de
Haas effect, was eventually demonstrated experimentally in 1915 (Fig. 7).
Fig. 7 The Einstein-de Haas
experiment. The iron rod
suspended from a torsion
fiber twists when a
magnetizing current in the
surrounding solenoid is
reversed, thereby
demonstrating the
relationship between
magnetism and angular
momentum
18
J. M. D. Coey
The Properties of Ferromagnets
If the luminiferous aether was inaccessible to experimental investigation, as the
1887 Michelson-Morley experiment suggested, the same could not be said for
magnetic materials. With its focus on electromagnetism, the nineteenth century
brought a flurry of investigations of the magnetic properties of the ferromagnetic
metals, iron (discovered in the fourth millennium BC), cobalt (discovered in 1735),
and nickel (discovered in 1824) and some of their alloys, which were at the
heart of electromagnetic machines. In 1842 James Joule, a brewer and natural
philosopher, discovered the elongation of an iron bar when it was magnetized
to saturation and demonstrated in a liquid displacement experiment that the net
volume was unchanged in the magnetostrictive process, owing to a compensating
contraction in the perpendicular directions [16]. Magnetostriction is the reason why
transformers hum. Gustav Wiedemann observed that an iron bar twisted slightly
when a current was passed through it in the presence of a magnetic field. Anisotropic
magnetoresistance (AMR) was discovered by William Thomson in 1856; the
resistance of iron or nickel is a few percent higher when measured in the direction
parallel to the magnetization than in the perpendicular direction [17]. The Hall
effect, the appearance of a transverse voltage when a current was passed through
a gold foil subject to a transverse magnetic field was discovered by Edwin Hall in
1879, And the contribution e proportioal to the magnetization of a ferromagnet —
tha anomalous Hall effect — was found shortly afterwards, in iron. John Kerr
showed in 1877 that the rotation of the plane of polarization of electromagnetic
radiation, demonstrated by Faraday for light passing through glass, could also be
measured in reflection from polished ferromagnetic metal surfaces [18].
Gauss’s collaborator Wilhelm Weber, who had constructed the first electromagnetic telegraph in 1833, formally presented the idea that molecules of iron were
capable of movement around their centers, suggesting that they lay in different
directions in an unmagnetized material, but aligned in the same direction in the
presence of an applied magnetic field. This was the origin of the explanation of
hysteresis by James Alfred Ewing, who coined the name for the central phenomenon
of ferromagnetism that he illustrated using a board of small, pivoting magnets [19].
Ewing’s activities as a youthful scottish professor at the University of Tokyo in
the 1890s helped to establish the strong Japanese school of research on magnetic
materials that thrives to the present day.
The hysteresis loop, illustrated in Fig. 8, is the icon of ferromagnetism. Except
in very small particles, a magnetized state is always metastable. The saturated
magnetic state is higher in energy relative to a multidomain state on account
of the
demagnetizing field that creates a positive magnetostatic self-energy -½μ0 Ms .Hd
dV in the fully magnetized state, where the only contribution to the integral comes
from the magnet volume. The hardened steel magnets of the nineteenth century
showed little coercivity, Hc Ms , and could only survive as bars and horseshoes
where the demagnetizing factor N of Eq. 11 was 1. The principal achievement
in technical magnetism in the twentieth century was the mastery of coercivity; this
needed new materials having Hc Ms .
1 History of Magnetism and Basic Concepts
spontaneous magnetization
19
M
remanence
coercivity
virgin curve
initial susceptibility
H
major loop
Fig. 8 The hysteresis loop of magnetization M against magnetic field H for a typical permanent
magnet, showing the initial magnetization curve from the equilibrium multidomain state and the
major loop. Ms is the saturation magnetization, Mr the remanent magnetization at zero field, and
Hc the coercive field required to reduce the magnetization to zero
The astonishing transformation of science and society that began in 1820
deserves the name electromagnetic revolution. By the end of the century, electromagnetic engineering was electrifying the planet, changing fundamentally our
communications and the conditions of human life and leisure. Huge electric
generators, powered by hydro or fossil fuel, connected to complex distribution networks were bringing electric power to masses of homes and factories
across the Earth. Electric light banished the tyranny of night. Electric motors
of all sorts were becoming commonplace, and public transport was transformed.
Telegraph and telephone communication connected people across cities, countries, and continents. Valdemar Poulsen demonstrated magnetic voice recording
in 1898. Much of the progress was achieved by engineers who relied on practical knowledge of electrical circuits and magnetic materials, independently of
the conceptual framework of electrodynamics that had been developed by the
physicists.
The electromagnetic revolution and the subsequent electrification of the planet
were the second occasion when magnetism changed the world. The century closed
with Pierre Curie’s 1895 accurate measurements of the Curie point TC (the critical
temperature above which a material abruptly loses its ferromagnetism) and with the
all-important discovery of the electron. Yet ferromagnetism was hardly understood
at all at a fundamental level at the turn of the century, and it was becoming evident
that classical physics was not up to the task.
20
J. M. D. Coey
Magnetism of the Electron
The discovery of the electron in the closing years of the nineteenth century was a
huge step toward the modern understanding of magnetism. The elementary charged
particle with mass me = 9.109 10−31 kg and charge e = −1.602 10−19 C had been
named by the Irish scientist George Johnstone Stoney in 1891, several years before
Jean Perrin in France actually identified negatively charged particles in a cathode
ray tube and J. J. Thompson in England measured their charge to mass ratio e/me ,
by deflecting the electrons in a magnetic field and making use of Eq. 6. Another
Irish scientist, George Francis FitzGerald, suggested in 1900 that magnetism might
be due to rotational motion of these electrons. They turned out to be not only the
carriers of electric current but also the essential magnetic constituent of atoms and
solids.
The Demise of Classical Physics
At the beginning of the twentieth century, the contradictions inherent in contemporary physics could no longer be ignored, but 25 years were to elapse before they
could be resolved. In that heroic period, classical physics and the lingering wisps
of aether were blown away, and a new paradigm was established, based on the
principles of quantum mechanics and relativity.
Magnetism in particular posed some serious puzzles. In order to account for the
abrupt disappearance of ferromagnetism at the Curie point, Pierre Weiss, who had
developed Ewing’s concept of magnetic domains, postulated in 1907 the existence
of an internal molecular field.
H i = nW M
(12)
proportional to magnetization in order to explain the spontaneous magnetization
within them. His theory of ferromagnetism was based on Paul Langevin’s 1905
explanation of the Curie law susceptibility of an array of disordered classical
magnetic moments.
χ = C/T
(13)
Susceptibility χ can be conveniently defined as the dimensionless ratio M/H,
where H is the applied magnetic field. The expression is modified for a ferromagnet
above its Curie point where it becomes the Curie-Weiss law χ = C/(T – θp ) with
θp ≈ TC . With Eq. (12) and Langevin’s theory of paramagnetism, Weiss invented
the first mean-field theory of a phase transition. For iron, where M = 1.71 MAm−1 ,
the Weiss constant nW is roughly 1000. According to Maxwell’s equation ∇. B = 0,
the component of B normal to the surface of a magnet is continuous, so there should
1 History of Magnetism and Basic Concepts
21
be a stray field of order μ0 Hs ∼ 1000 T in the vicinity of a magnetized iron bar. In
fact, the observed stray fields are a thousand times smaller.
Furthermore if, as Ampère believed, all magnetism was traceable to circulating
electric currents, the magnetization of an iron bar requires an incredible surface
current of 17,100 A for every centimeter of its length. How could such a current
be sustained indefinitely? Why does the iron not melt? What did the sobriquet
molecular really mean? The anomalous Zeeman splitting of spectral lines in a
magnetic field was another mystery. In retrospect, the most startling result was a
theorem proved independently in their theses by Niels Bohr in 1911 and Hendrika
van Leeuwen in 1919. They showed that at any finite temperature and in any
magnetic or electric field, the net magnetization of a collection of classical electrons
vanishes identically. So, in stark contrast with experiment, classical electron physics
was fundamentally incompatible with any kind of magnetism!
By 1930, quantum mechanics and relativity had ridden to the rescue, and a
new understanding of magnetism emerged in terms of the physics of Einstein,
Bohr, Pauli, Dirac, Schrödinger, and Heisenberg. The source of magnetism in
matter was identified with the angular momentum of elementary particles, especially
the electron [20]. The connection between angular momentum and magnetism
had been demonstrated directly on a macroscopic scale in 1915 by the Einsteinde Haas experiment (Fig. 7), where angular recoil of a suspended iron rod was
observed when its magnetization was reversed by an applied field. It turned out
that the perpetual currents in atoms were quantized in stationary states that did not
decay and that the angular momentum of the orbiting electrons was a multiple of
Planck’s constant = 1.055 10−34 Js. Furthermore, the electron itself possessed an
intrinsic angular momentum or spin [20] with eigenvalues of ±½ along the axis of
quantization defined by an external field. Weiss’s molecular field was no magnetic
field at all, but a manifestation of electrostatic coulomb interactions constrained by
Wolfgang Pauli’s exclusion principle, which forbade the occupancy of a quantum
state by two electrons with the same spin.
The intrinsic angular momentum of an electron with two eigenvalues had been
proposed by Pauli in 1924; Samuel Goudsmit and George Uhlenbeck demonstrated
a year later that the spin angular momentum had a value of ½. The Pauli spin
matrices representing the three components of spin angular momentum are
s=
01
0 −i
1 0
,
,
10
i 0
0 −1
/2
(14)
The corresponding electronic magnetic moment was the Bohr magneton,
μB = e/2me
or 9.274 × 10−24 Am2 , twice as large as the moment associated with a unit of
orbital angular momentum in Bohr’s model of the atom. The gyromagnetic ratio
of magnetic moment to angular momentum for the electron spin is γ ≈ e/me ,
so the Larmor precession frequency eB/2πme for the electron is 28 GHzT−1 .
22
J. M. D. Coey
The problem of the electron’s magnetism was finally resolved by Paul Dirac
in 1928 when he succeeded in writing Schrödinger’s equation in relativistically
invariant form, obtaining the non-relativistic electron spin in terms of the 2 × 2
Pauli matrices. Together with Dirac, Werner Heisenberg formulated the exchange
interaction represented by the famous Heisenberg Hamiltonian
H = –2J S i .S j
(15)
to describe the coupling between the vector spins Si and Sj of two nearby manyelectron atoms i and j. The spin vectors S are the spin angular momenta in units of .
The value of the exchange integral J was closely related to Weiss’s molecular field
coefficient nW and depends strongly on interatomic distance. It can be positive, if it
tends to align the two spins parallel (ferromagnetic exchange), or negative if it tends
to align the pair antiparallel (antiferromagnetic exchange). The value of S is obtained
from the first of the three rules, discussed below, that were formulated by Friedrich
Hund around 1927 for finding the ground state of a multi-electron atom. The
exchange interactions among the electrons of the same atom are much stronger than
those between the electrons of adjacent atoms given by Eq. (15). The fundamental
insight that magnetic coupling of electronic spins is governed by electrostatic
coulomb interactions, subject to the symmetry constraints of quantum mechanics,
was the key needed to unlock the mysteries of ferromagnetism. Exchange is
discussed in Chap. 2, “Magnetic Exchange Interactions.”
The magnetic moment of an atom or ion is the sum of two contributions. One
arises from the intrinsic spin angular momentum of the atomic electrons. The other
comes from their quantized orbital angular momentum. The moments associated
with each type of angular momentum have to be summed according to the rules of
quantum mechanics. The moment associated with ½ of spin angular momentum is
practically identical to that associated with of orbital angular momentum, namely,
one Bohr magneton in each case. The quantum theory of magnetism is therefore the
quantum theory of angular momentum. Hund’s rules were an empirical prescription
for determining the total angular momentum of the many-electron ground state of
electrons belonging to the same atom or ion. Firstly, the rule is to maximize the
spin angular momentum S while respecting the Pauli principle that no two electrons
can be in the same quantum state. Secondly, the orbital angular momentum L is
maximized, consistent with the value of S, and thirdly the spin and orbital momenta
are coupled together to form the total angular momentum J = L ± S, according
to whether the electronic shell is more or less than half full. The total magnetic
moment (in units of μB ) is then related to the total angular momentum (in units of
) by a numerical Landé g-factor, which is 1 for a purely orbital moment and 2 for
pure spin.
The spin-orbit coupling, which arises in the atom from motion of the electron
in the electrostatic potential of the charged nucleus and gives rise to Hund’s third
rule, is another key interaction. Of fundamentally relativistic character, it emerges
naturally from Dirac’s relativistic quantum theory of the electron, and it turns out to
be at the root of many of the most interesting phenomena in magnetism, including
1 History of Magnetism and Basic Concepts
23
magneto-optics, magnetocrystalline anisotropy, and the spin Hall effect. The spinorbit interaction for a magnetic ion is represented by the Hamiltonian L.S, where
L is the orbital angular momentum of the many-electron atom in units of and is the atomic spin-orbit coupling constant. Like the exchange constant J , has
dimensions of energy.
Felix Bloch in 1930 described the spin waves that are the quantized elementary
excitations of a ferromagnetic array of atoms whose spins are coupled by Heisenberg exchange. These excitations have an angular frequency ω and a wavevector
k that are related by the dispersion relation ω = Dk2 , where D is the spin wave
stiffness constant. It is proportional to J .
The first quantum theories of magnetism regarded the electrons as localized
on the atoms or ions, but an alternative magnetic band theory of ferromagnetic
metals was developed by John Slater and Edmund Stoner in the 1930s. It accounted
for the non-integral, delocalized spin moments found in Fe, Co, and Ni and their
alloys, although the theory in its original form greatly overestimated the Curie
temperatures. The delocalized, band electron model of Slater and the localized,
atomic electron model of Heisenberg were two distinct paradigms for the theory
of magnetism that persisted until sophisticated computational methods for treating
the many-body interelectronic correlations in the ground state of multi-electron
atoms were devised toward the end of the twentieth century. The differences
between the two approaches are epitomized in the calculation of the paramagnetic
susceptibility. Pauli found a small temperature-independent susceptibility resulting
from Fermi-Dirac statistics for delocalized electrons, whereas Léon Brillouin had
used Boltzmann statistics and the Bohr model to derive the Curie law susceptibility
of an array of atoms with localized electrons.
The sixth Solvay Conference, held in Brussels in October 1930 (Fig. 9), was
devoted to magnetism [21]. It followed four years of brilliant discoveries in
theoretical physics, which set out the modern electronic theory of condensed matter.
Yet the immediate impact on the practical development of functional magnetic
materials was surprisingly slight. Dirac there made the perceptive remark “The
underlying physical laws necessary for the mathematical theory of a large part of
physics and the whole of chemistry are completely known, and the difficulty is only
that the exact application of these laws leads to equations much too complicated to
be soluble.”
Magnetic Phenomenology
In view of the immense computational challenge posed by many-body electron
physics in 1930, a less fundamental theoretical approach was needed. Louis
Néel pursued a phenomenological approach to magnetism with notable success,
oblivious to the triumphs of quantum mechanics. His extension of the Weiss
theory to two equal but oppositely aligned magnetic sublattices led him to the
idea of antiferromagnetism in his 1932 doctoral thesis. This hidden magnetic order
24
J. M. D. Coey
Fig. 9 The 1930 Solvay Conference on Magnetism Back row: Herzen, Henriot, Verschaffelt,
Manneback, Cotton, Errera, Stern, Piccard, Gerlach, Darwin, Dirac, Bauer, Kapitza, Brioullin,
Kramers, Debye, Pauli, Dorfman, van Vleck, Fermi, Heisenberg. Front row: de Donder, Zeeman,
Weiss, Sommerfeld, Curie, Langevin, Einstein, Richardson, Cabrera, Bohr, de Haas
awaited the development of neutron scattering in the 1950s before it could be
directly revealed, initially for MnO. Néel went on to explain the ferrimagnetism
of oxides such as magnetite, Fe3 O4 , the main constituent of lodestone, in terms of
two unequal, antiferromagnetically coupled sublattices. The three most common
types of magnetic order, and their temperature dependences, are illustrated in
Fig. 10.
The spinel (MgAl2 O4 ) structure of magnetite has an A sublattice of 8a sites with
fourfold tetrahedral oxygen coordination and twice as many 16d sites with sixfold
octahedral coordination forming a B sublattice. The spinel structure is illustrated
in Fig. 14 where the 8a sites are at the centers of the blue tetrahedra, which have
oxygen ions at the four corners, and the 16d sites are at the centers of the brown
octahedra, which have six oxygen ions at the corners. The numbers of each type
of site in the unit cell are indicated by the labels. The 16d sites in magnetite
are occupied by a mixture of ferrous Fe2+ and ferric Fe3+ ions with electronic
configurations 3d5 and 3d6 and spin moments of 5 μB and 4 μB , respectively,
whereas the 8a sites are occupied by oppositely aligned Fe3+ ions. This yields a net
spin moment of 4 μB per formula (0.48 MAm−1 ) – a quantitative explanation of the
magnetism of the archetypical magnet in terms of lattice geometry and the simple
rule that each unpaired electron contributes a spin moment of one Bohr magneton.
Néel added two new categories of magnetic substances – antiferromagnets and
ferrimagnets – to Faraday’s original three. Their magnetic ordering temperatures are
known as antiferromagnetic or ferrimagnetic Néel temperatures. The ferrimagnetic
one is also called a Curie point.
1 History of Magnetism and Basic Concepts
25
1/c
1/c
T
T C ,qp
1/c
qp
T
TN
M
M
qp
M
B
B
B A
TC T
T
T fN
B
TN
A
T
A
T fN
T
A
Fig. 10 Schematic temperature dependences of the inverse susceptibility (top) and (sub)lattice
magnetization (bottom) of a ferromagnet (left), an antiferromagnet (center), and a ferrimagnet
(tight)
Micromagnetism
For many practical purposes, it is possible to follow in the footsteps of Néel,
sidestepping the complications engendered by the atomic and electronic basis of
magnetism, and regard magnetization as a continuous vector in a solid continuum
[13], as people have for about 200 years. The iconic hysteresis loop M(H) (Fig. 8)
is the outcome of a metastable structure of domains of uniformly magnetized
ferromagnetic Weiss domains separated by narrow domain walls between domains
magnetized in different directions. The structure depends on the thermal and
magnetic history of a particular sample. Aural evidence for discontinuous jumps
in the size of the domains as the magnetization was saturated was first heard by
Heinrich Barkhausen in 1919 with the help of a pickup coil wound around some
ferromagnetic wires, a rudimentary amplifier, and a loudspeaker. Then in 1931 the
domains were directly visualized by Francis Bitter using a microscope focused on a
polished sample surface and a colloidal suspension of magnetite particles that were
drawn by the stray field to the domain walls. These colloids, known as ferrofluids,
behave like ferromagnetic liquids.
The idea of a domain wall as a region where the magnetization rotates progressively from one direction to the opposite one in planes parallel to the wall
was introduced by Felix Bloch in 1932. His walls create no bulk demagnetizing
26
J. M. D. Coey
Fig. 11 Two types of 180◦ domain walls: a) the Bloch wall and b) the Néel wall
field and cost little magnetostatic energy because ∇. M = 0; the magnetization
in each plane is uniform, and there is no component perpendicular to the planes
(see Eq. 9). The exchange energy cost, written in the continuum approximation as
A(∇M)2 where A ∝ J , is balanced by the anisotropy energy cost associated with the
magnetization in the wall that is misaligned with respect to a magnetic easy axis of
the crystal. Magnetic anisotropy is introduced below, and it is discussed in detail in
Chap. 3, “Anisotropy and Crystal Field.” A Néel domain wall, where the
magnetization rotates in a plane perpendicular to the wall so that ∇. M = 0 in the
bulk, but there is no surface magnetic charge, is higher in energy except in thin films.
The two types of wall are illustrated in Fig. 11.
In principle, the sum of free energy terms associated with exchange, anisotropy,
and magnetostatic interactions, together with the Zeeman energy in an external field,
could be minimized to yield the M(H) loop and the overall domain structure of
any solid. Further terms can be added to take into account the effects of imposed
strain and spontaneous magnetostriction. In practice, however, crystal defects such
as grain boundaries spoil the continuum picture and can exert a crucial influence on
the walls. It is then necessary to resort to models to develop an understanding of
hysteresis.
The basic theory of micromagnetism was developed by William Fuller Brown
in 1940 [13]. The magnetostatic interaction between the magnetic dipoles that
constitute the magnetization is a dominant factor. The dipole fields fall off as 1/r3
(Eq. 2), providing a long-range interaction unlike exchange, which is short-range
because it depends on an overlap between electronic wavefunctions that decays
exponentially with interatomic spacing. This is why weak magnetostatic interactions
that are of order 1 K for a pair of ions are able to compete on a mesoscopic length
scale with the much stronger exchange interactions of electrostatic origin that can
be of order 100 K to control the domain structure of a given ferromagnetic sample.
Magnetocrystalline anisotropy is represented phenomenologically in the theory
by terms in the energy that depend on the orientation of M with respect to the
local crystal axes. The electrostatic interaction of localized atomic electrons with
the potential created by all the other atoms in the crystal is known as the crystal field
1 History of Magnetism and Basic Concepts
27
interaction; the effect of chemical bonding with the ligands of an atom is the ligand
field interaction. The two effects are comparable in magnitude for 3d ions [22].
Magnetocrystalline anisotropy arises from the interplay of the crystal/ligand field
and spin-orbit coupling. The simplest case is for uniaxial (tetragonal, hexagonal,
rhombohedral) crystals, where the leading term in the energy density is of the form
Ea = K1 sin2 θ + . . . ..
(16)
where θ is the angle between M and the symmetry axis. Two opposite easy
directions lie along the crystal axis if the anisotropy constant K1 is positive, but
there are many easy directions lying in an easy plane perpendicular to the crystal
axis (θ = π/2) when K1 is negative. Anisotropy arises also from overall sample
shape, due to the demagnetizing energy ½MHd , which gives another contribution in
sin2 θ that depends on the demagnetizing factor N with
K1 sh =
1
μ0 Ms 2 (1 − 3N )
4
(17)
where Ms is the spontaneous magnetization. There is obviously no shape anisotropy
for a sphere, which has N = 1/3. An expression equivalent to (16) at the atomic scale
is εa = Da sin2 θ , where Da /kB ∼ 1 K. The magnitude of the crystal field energy is
comparable to the magnetostatic energy, but it is much smaller than the exchange
energy in practical magnetic materials. It remains challenging to calculate K1 or Da
precisely in metals.
An instructive paradox arising from Brown’s micromagnetic theory is his result
that the coercivity Hc of a perfect, defect-free ferromagnetic crystal lattice must
exceed the anisotropy field Ha = 2 K1 /μ0 Ms . In practice Hc is rarely as much as a
fifth of Ha . The explanation is that no real lattice is ever free of defects, which act as
sites for the nucleation of reverse domains or as pinning centers for domain walls.
The sequence of metastable states represented on the hysteresis loop is generally
dominated by asperities and lattice defects that are very challenging to characterize
in any real macroscopic sample. Control of these defects in modern permanent
magnets having Hc Ms has been as much a triumph of metallurgical art as physical
theory. Micromagnetism is the subject of Chap. 7, “Micromagnetism.”
Magnetic Materials
The traditional magnetic materials were alloys of the ferromagnetic metals, Fe,
Co, and Ni. The metallurgy and magnetic properties of these alloy systems were
the focus of investigations of technical magnetism in the first half of the twentieth
century, when useful compositions were developed such as Permendur, Fe50 Co50 ,
the alloy with the highest magnetization (1.95 MAm−1 ); Permalloy Fe20 Ni80 , which
has near-zero anisotropy and magnetostriction, together with very high relative
permeability (μr = (1 + χ) ≈ 105 ); and Invar Fe64 Ni36 a composition with near-
28
J. M. D. Coey
Fig. 12 Unit cells of the ferromagnetic elements Fe (body-centered cubic, left), Ni (face-centered
cubic, center), and Co, Gd (hexagonal close-packed, right) [29], with kind permission from
Cambridge University Press
zero thermal expansion around room temperature. The early investigations are well
summarized in Bozorth’s 1950 monograph [23]. The fourth ferromagnetic element
at room temperature is the rare earth gadolinium. The crystal structures of these
elemental ferromagnets are illustrated in Fig. 12.
An important practical advance in the story of permanent magnet development
was the thermal processing of a series of Al-Ni-Co-Fe alloys, the Alnico magnets,
that was initiated in Japan in 1932 by Tokushichi Mishima. Their coercivity relied
on achieving a nanostructure of aligned acicular (needle-like) regions of Co-Fe in
a matrix of nonmagnetic Ni-Al. It was the shape of the ferromagnetic regions that
gave the alloys some built-in magnetic anisotropy (Eq. 17), but it still had to be
supplemented with global shape anisotropy by fabricating the Alnico into a bar or
horseshoe in order to avoid self-demagnetization. The mastery of coercivity that
was acquired over the course of the twentieth century (Fig. 13) was spectacular,
and burgeoning applications in technical magnetism of soft and hard magnetic
materials were the direct consequence. The terms “soft” and “hard” were derived
originally from the magnetic steels that were used in the nineteenth century. The
most useful figure of merit for the hard, permanent magnets is the maximum energy
product |BH|max , equal to twice the energy in the stray field produced by a unit
volume of magnet. The SI unit is kJm−3 . Energy product doubled every 12 years
for most of the twentieth century, thanks to the discovery in the 1960s of rare
earth cobalt intermetallic compounds and the discovery of new rare earth ironbased materials in the 1980s. Comparable progress with decreasing hysteresis losses
in soft, electrical steels continued to the point where they became a negligible
fraction of the resistive losses in the copper windings of electromagnetic energy
converters. Ultrasoft amorphous magnetic glasses were developed in the 1970s.
Applications of soft and hard magnetic materials are discussed in Chaps. 29, “Soft
Magnetic Materials and Applications,” and 28, “Permanent Magnet Materials and
Applications” respectively.
A good working knowledge of the quantum mechanics of multi-electron atoms
and ions had been developed by the middle of the twentieth century, mainly from
1 History of Magnetism and Basic Concepts
29
Nd-Fe-B
Fig. 13 The development of coercivity over the ages and in the twentieth century
observations of optical spectra and the empirical rules formulated by Hund to
specify the ground state L, S, and J multiplet, which is the one of interest for
magnetism. All this led naturally to a focus on the localized electron magnetism
found in the 3d and 4f series of the periodic table. For 3d ions in solids, the
ionic moment is essentially that arising from the unpaired electron spins left after
filling the orbitals according to the Pauli principle and Hund’s first rule. The orbital
moment expected from the second rule is quenched by the crystal field, which
impedes the orbital motion so that it barely contributes to the ionic magnetism.
But the crystal field is weaker for the 4f elements in solids, whether insulating or
metallic, and the magnetism is more atomic-like with spin and orbital contributions
coupled by the spin-orbit interaction according to Hund’s third rule to yield the total
angular momentum J.
Microscopic quantum theory began to play a more important part in magnetic
materials development after the 1970s with the advent of rare earth permanent
magnets SmCo5 and especially Nd2 Fe14 B, when an understanding of the intrinsic,
magnetocrystalline anisotropy in terms of crystal field theory and spin-orbit coupling began at last to make a contribution to the design of new permanent magnet
materials.
Magnetic Oxides
The focus on localized electron magnetism in the 1950s and 1960s led to systematic
investigations of exchange interactions in insulating compounds where the spin
30
J. M. D. Coey
moments of magnetic 3d ions are coupled by indirect overlap of their wavefunctions via an intervening nonmagnetic anion, usually O2− . A systematic empirical
understanding of the dependence of these superexchange interactions on electron
occupancy and bond angle emerged in the work of Junjiro Kanamori and John
Goodenough [24], based on the many new magnetic compounds that were being
fabricated at that time. There is a multitude of solid solutions between end-members,
with extensive opportunities to tune magnetic properties by varying the chemical
compositions of oxide families such as ferrites [25]. Superexchange, like direct
exchange in the ferromagnetic 3d elements, depends on the overlap of wavefunctions of adjacent atoms and decays exponentially with interatomic distance.
The magnetite family of cubic spinel ferrites M2+ Fe3+
2 O4 was the first to be
thoroughly investigated, with M = Mg, Zn, Mn, Fe, 2/3Fe3+ (γFe2 O3 ), Co, or Ni.
Ferrimagnetic Neél temperatures of these ferrites range from 700 to 950 K, although
spinel itself (MgAl2 O4 ) is nonmagnetic. Several of the insulating compounds
with Mn, Ni, and Zn are suitable as soft magnetic materials for audio- or radiofrequency applications. Other important families investigated at that time were
3+
garnets, perovskites, and hexagonal ferrites. The garnet ferrites R3+
3 Fe5 O12 have
a large cubic unit cell containing 160 ions, with ferrimagnetically aligned ferric
iron in both tetrahedral 24d and octahedral 16a sites, and large R3+ ions in
eightfold oxygen coordination in deformed cubal 24c sites. R may be any rare
earth element, including Y, which forms yttrium iron garnet (YIG), Y3 Fe5 O12 , a
superlative microwave material that exhibits ultra-low magnetic losses on account
of its insulating character. The net magnetic moment of YIG is 5μB per formula
unit. Substituting magnetic rare earths in the structure provides an opportunity to
study superexchange between 3d and 4f ions. That interaction is weak, and the
4f ions couple antiparallel to the 24d site iron, but their sublattice magnetization
decays much faster with temperature, giving rise to the possibility of a compensation
temperature, where the net magnetization of the two ferrimagnetic sublattices
crosses zero at a temperature below the ferrimagnetic Neél point. The compensation
temperature of Gd3 Fe5 O12 , for example, is 290 K, whereas its ferrimagnetic Néel
point is at 560 K, a typical value for the whole rare earth iron garnet series.
Another important oxide family, the hexagonal ferrites especially M2 Fe12 O19 ,
where M = Ba2+ or Sr2+ , have uniaxial anisotropy and crystallize in the magnetoplumbite structure. There are four Fe3+ sites in the structure, including a fivefold 2b
site with trigonal symmetry where the threefold axis is parallel to the c-axis of the
hexagonal unit cell. The net ferrimagnetic moment is 20 μB per formula unit, since
eight iron ions belong to one sublattice and four to the other. The large nonmagnetic
M cations occupy sites that would otherwise belong to a hexagonal close-packed
oxygen lattice. The 2b site contributes rather strong uniaxial anisotropy, and the
anisotropy field of 1.4 MAm−1 is more than three times the magnetization (0.38
MAm−1 ), making it possible in the early 1950s to achieve coercivity comparable
to the magnetization and manufacture cheap ceramic magnets in any desired shape,
thereby overcoming the shape barrier that had impeded the development permanent
magnets for a millennium. A million tonnes of these ferrite magnets is sold every
year.
1 History of Magnetism and Basic Concepts
31
The drawback of any oxide magnetic material is that its magnetization is never
more than a third of that of metallic iron. This is unavoidable because most of the
unit cell volume is occupied by large, nonmagnetic O2− anions, with the high-spin
ferric iron Fe3+ or other magnetic ions confined to the interstices in the oxygen
lattice. To make matters worse, a ferrimagnetic structure reduces the magnetization
further. There are relatively few ferromagnetic oxides; CrO2 is one example. It is
not an insulator, but a half metal, with a gap in the minority-spin conduction band.
A search for insulating ferromagnetic oxides in the 1950s led to the investigation of ABO3 compounds with the perovskite structure. Here the magnetic B
cations occupy the 1a octahedral sites, and the nonmagnetic A cations occupy
the 12-coordinated 1b sites in the ideal cubic structure. It proved to be possible
to obtain ferromagnetism provided the A cations are present in two different
valence states. This works best in mixed-valence manganites [26], with composition
2+
3+ (3d4 )
(La3+
0.7 M0.3 )MnO3 where M = Ba, Ca, or Sr. The resulting mixture of Mn
4+
3
and Mn (3d ) on B sites leads to electron hopping with spin memory from
one 3d3 core to another. This is the ferromagnetic double exchange interaction,
envisaged by Clarence Zener in 1951. Similar electron hopping occurs for Fe2+
and Fe3+ in the octahedral sites of magnetite. A consequence is that the oxides,
though ferromagnetic, are no longer insulating, and the Curie temperatures are not
particularly high – they do not exceed 400 K. A notable feature of the mixed-valence
manganites, related to their hopping conduction, is the “colossal magnetoresistance”
observed near the Curie point, where there is a broad maximum in the resistance
that can be suppressed by applying a magnetic field of several tesla. All four
oxide structures are presented in Fig. 14. They illustrate the importance of crystal
chemistry for determining magnetic properties.
Fig. 14 Crystal structures of magnetic oxides: perovskite (top left), spinel (bottom left), garnet
(center), magnetoplumbite (right). The oxygen coordination polyhedral around the magnetic
cations (tetrahedrons, blue, or octahedrons, brown) is illustrated. The spheres are large nonmagnetic cations. Unit cells are outlined in black. Magnetoplumbite is hexagonal, and the others are
cubic [31], with kind permission from APS
32
J. M. D. Coey
Research on localized electron magnetism in oxides and related compounds has
passed through three phases. Beginning with studies of polycrystalline ceramics
from about 1950, single crystals were grown for specific physical investigations
after about 1970, and then in the late 1980s, following the high-temperature
superconductivity boom, came the growth and characterization of ferromagnetic
and ferrimagnetic oxide thin films and first steps toward all-oxide spin electronics. A
similar pattern was followed by sulfides, fluorides, and other magnetic compounds.
All are discussed further in Chap. 17, “Magnetic Oxides and Other Compounds.”
Intermetallic Compounds
A rich class of functional magnetic materials is the intermetallic compounds of
rare earth elements and transition metals. The atomic volume ratio of a 4f to a
3d atom is about three, so the alloys tend to be stoichiometric line compounds
rather than solid solutions. The first of these was SmCo5 , developed for permanent
magnet applications in the USA in the mid-1960s by Karl Strnat. It was followed
by Sm2 Co17 in the early 1970s, and then in 1983 came the announcement of
the independent discovery of the first iron-based rare earth magnet, the ternary
Nd2 Fe14 B, by Masato Sagawa in Japan and John Croat in the USA. This was
a breakthrough because iron is cheaper and more strongly magnetic than cobalt.
Nd2 Fe14 B has since come to dominate the global high-performance magnet market,
with an annual production in excess of 100,000 tonnes. The coercivity needed in
these optimized rare earth permanent magnets is comparable to their magnetization,
and the optimization of the microstructure of a new hard magnetic material to attain
the highest possible energy product, which scales as Ms 2 but can never exceed
¼μ0 Ms 2 , is a long empirical process. It generally takes many years to achieve
a coercivity as high as 20–30% of the anisotropy field [28]. The battle to create
the metastable hysteretic state that permits a permanent magnet to energize the
surrounding space with a large stray field is never easy to win, and each material
requires a different strategy.
The fundamental significance of these intermetallics and related interstitial
compounds such as Sm2 Fe17 N3 that were discovered in the 1990s is that crystal field
theory and quantum mechanics were involved in their design. All have a uniaxial
crystal structure with a single easy axis and strong magnetocrystalline anisotropy.
Such anisotropy is a prerequisite for the substantial coercivity, Hc Ms needed to
overcome the shape barrier and create a magnet with any desired form.
The practical significance of the rare earth permanent magnets has been the
appearance of a wide range of compact, energy-efficient electromagnetic energy
converters that are being used in consumer products, electric vehicles, aeronautics,
robotics, and wind generators.
Besides magnetocrystalline anisotropy, another potentially useful consequence
of the spin-orbit interaction in rare earth intermetallics is their strong magnetostriction. The rare earth elements order magnetically at or below room temperature
so, just as for the permanent magnets, it was necessary to form an intermetallic
1 History of Magnetism and Basic Concepts
33
Fig. 15 Crystal structures of ferromagnetic intermetallic compounds: YFe2 (cubic, left) SmCo5
(hexagonal, top centre), Co2 MnSi (cubic, bottom centre), Nd2 Fe14 B (tetragonal, right). Fe and Co
Mn are the small brown/red, blue, and scarlet spheres. Rare earths are the large spheres. Si and B
are grey and black
compound with iron or cobalt to obtain a functional material with a useful
Curie temperature that should be substantially greater than room temperature to
ensure adequate magnetic stability. A functional magnetostrictive material has to
be magnetically soft, and this was achieved in the RFe2 rare earth Laves phase
compounds by Arthur Clark in 1984, who combined Dy and Tb, which have the
same sign of magnetostriction, but compensating anisotropy of opposite sign, in
the cubic alloy (Tb0.3 Dy0.7 )Fe2 , known as Terfenol-D. Single crystals exhibited
Joulian magnetostriction of up to 2000 parts per million (ppm), a hundred times
greater than Joule had measured 150 years earlier in pure iron [16] (see Chaps.
28, “Permanent Magnet Materials and Applications,” and 11, “Magnetostriction
and Magnetoelasticity”).
Magnetically soft rare earth intermetallics are also of interest as magnetocaloric
materials for solid-state refrigeration when their Curie point is close to room
temperature (see Chap. 30, “Magnetocaloric Materials and Applications”). Some
crystal structures of rare earth intermetallics are shown in Fig. 15.
Among the other intermetallic families, the ordered body-centered cubic Heusler
families of X2 YZ or XYZ alloys are notable in that they include a wide variety
of magnetically ordered compounds, such as the magnetic shape-memory alloy
NiMnSb or the half-metallic ferromagnet Co2 MnSi, which, like CrO2 , has a gap at
the Fermi level for minority-spin electrons. Information on a great many metallic
magnetic materials is collected in Chap. 4, “Electronic Structure: Metals and
Insulators.”
34
J. M. D. Coey
Model Systems
Magnetism has proved to be a fertile proving ground for condensed matter theory.
The first mean-field theory was Weiss’s molecular field of magnetism, later
generalized by Lev Landau in the USSR in 1937. There followed more sophisticated
theories of phase transitions, with magnetism providing much of the data to support
them. The single-ion anisotropy of rare earth ions due to the local crystal field
reduces the effective dimensionality of the magnetic order parameter from three
to two for easy-plane (xy) anisotropy or from three to one for easy-z-axis (Ising)
anisotropy. Magnetically ordered compounds can be synthesized with an effective
spatial dimension of one (chains of magnetic atoms), two (planes of magnetic
atoms), or three (networks of magnetic atoms), as well as ladders and isolated
motifs. Magnetism has provided a treasury of materials that show continuous
phase transitions as a function of temperature or quantum phase transitions at zero
temperature as a function of pressure or magnetic field, as well as topological phases
such as the two-dimensional xy model, investigated by David Thouless, Michael
Kosterlitz, and Duncan Haldane. It is frequently possible to realize magnetic
materials that embody the essential electronic or structural features of the theoretical
models.
An early theoretical milestone was Lars Onsager’s 1944 solution of the twodimensional Ising model, where spins are regarded as one-dimensional scalars
that can take only values of ±1. The behavior of more complex and realistic
systems such as the three-dimensional Heisenberg model near its Curie temperature
was solved numerically using the renormalization group technique developed by
Kenneth Wilson in the 1970s. The ability to tailor model magnetic systems, with
an effective spatial dimension of 1 or 2 due to their structures of chains or planes
of magnetic ions and an effective spin dimension of 1, 2, or 3 determined by
magnetocrystalline anisotropy due to the combination of the crystal/ligand field and
the spin-orbit interaction, was instrumental in laying the foundation of the modern
theory of phase transitions. The theory is based on universality classes where
power-law temperature variations of the order parameter and its thermodynamic
derivatives with respect to temperature or magnetic field in the vicinity of the phase
transition are characterized by numerical critical exponents that depend only by the
dimensionality of the space and the magnetic order parameter.
Another fecund line of enquiry was “Does a single impurity in a metal bear a
magnetic moment?” This was related to Jun Kondo’s formulation of a problem
concerning the scattering of electrons by magnetic impurities in metals and its
eventual solution in 1980. In the presence of antiferromagnetic coupling between
an impurity and the conduction electrons of a metallic host, the combination enters
a nonmagnetic ground state below the Kondo temperature TK . The Kondo effect
is characterized by a minimum in the electrical resistivity. The study of magnetic
impurities in metals focused attention on the relation between magnetism and
electronic transport, which has proved extremely fruitful, leading to several Nobel
Prizes and the emergence in the 1990s of spin electronics.
1 History of Magnetism and Basic Concepts
35
The exchange interaction between two dilute magnetic impurities in a metal
is long-range, decaying as 1/r3 while oscillating in sign between ferrromagnetic
and antiferromagnetic, where r is their separation. The following is the RudermanKittel-Kasuya-Yosida (RKKY) exchange interaction
J (r) = aJsd 2 (sinξ − ξcosξ) /ξ4
(18)
where a is a constant, Jsd is the exchange coupling between the localized impurity
and the conduction electrons, and ξ is twice the product of r and the Fermi
wavevector. It was studied intensively in the 1970s in dilute alloys such as AuFe
or CuMn, known as spin glasses (the host is in bold type, and the impurity in
italics). The impurity in these hosts retains its moment at low temperatures, and
the RKKY exchange coupling J (∇) between a pair of spins is as likely to be
ferromagnetic (positive) as antiferromagnetic (negative). The impurity spins freeze
progressively in random orientations around a temperature Tf that is proportional to
the magnetic concentration. The nature of this transition to the frozen spin glass state
was exhaustively debated. A related issue, the long-range exchange interactions
associated with the ripples of spin polarization created by a magnetic impurity in
a metal, led to an understanding of complex magnetic order in the rare earth metals
( Chap. 14, “Magnetism of the Elements”).
The magnetism of electronic model systems such as a chain of 1s atoms with an
on-site coulomb repulsion U when two electrons occupy the same site, formulated
by John Hubbard in 1963, has proved to be remarkably complex. Control parameters
in the Hubbard model are the band filling and the ratio of U to the bandwidth, and
they lead to insulating and metallic, ferromagnetic, and antiferromagnetic solutions.
Amorphous Magnets
An important question, related to the dilute spin glass problem, was what effect does
atomic disorder have on magnetic order and the magnetic phase transition in magnetically concentrated systems? Here a dichotomy emerges between ferromagnetic
and antiferromagnetic interactions. The answer for materials with ferromagnetic
exchange and a weak local electrostatic (crystal field) interaction is that the atomic
disorder has little effect.
Techniques for rapidly cooling eutectic melts at rates of order 106 Ks−1
developed around 1970 produced a family of useful amorphous ferromagnetic alloys
based on Fe, Co, and Ni, with a minor amount of metalloid such as B, P, or Si. These
metallic glasses, frequently in the form of thin ribbons obtained by melt spinning,
were magnetically soft and proved that ferromagnetic order could exist without a
crystal lattice. There are no crystal axes, and weak local anisotropy due to the local
electrostatic interactions averages out. The magnetic metallic glasses are mechanically strong and have found applications in transformer cores and security tags.
36
J. M. D. Coey
Amorphous materials with antiferromagnetic interactions are qualitatively different. Whenever the superexchange neighbors in oxides or other insulating compounds form odd-membered rings, these interactions are frustrated. No collinear
magnetic configuration is able to satisfy them all. In crystalline antiferromagnets like rocksalt-structure NiO, the partial frustration leads to a reduced Néel
temperature, but in fully frustrated pyrochlore-structure compounds, for example, the Néel point is completely suppressed. In the amorphous state, however,
frustration has a spatially random aspect, and it leads to random spin freezing
with a tendency to antiferromagnetic nearest-neighbor correlations, known as
speromagnetism.
The situation for amorphous rare earth intermetallic alloys, which are best
prepared by prepared by rapid sputtering, is different. There the local anisotropy
at rare earth sites is strong, and does not average out, but it tends to pin the rare
earth moments to randomly oriented easy axes in directions that are roughly parallel
to that of the local magnetization of the 3d ferromagnetic sublattice for the light rare
earths and roughly antiparallel to it for the heavy rare earths. The sign of the 3d-4f
coupling changes in the middle of the series, so that amorphous Gd-Fe alloys, for
example, are ferrimagnetic. (Gd is the case where there are no orbital moment and
no magnetocrystalline anisotropy on account of its half-filled, 4f7 shell.)
Rapid quenching can also be used to produce nanocrystalline material with
isotropic crystallite orientations of nanocrystals embedded in an amorphous matrix.
Certain soft magnetic materials have such a two-phase structure. Nanocrystalline
Nd-Fe-B produced by rapid quenching shows useful coercivity due to domain wall
pinning at the Nd2 Fe14 B nanocrystallite boundaries, but the remanence is only about
half the saturation magnetization on account of the randomly directed easy axes
of the tetragonal crystallites. The magnitude of the anisotropy and the nanoscale
dimension are critical for the averaging that determines the magnetic properties.
Magnetic Fine Particles
An early approach to the difficult problem of calculating hysteresis was to focus on
magnetization reversal in single-domain particles that were too small to benefit from
any reduction in their energy by forming a domain wall. Edmund Stoner and Peter
Wohlfarth proposed an influential model in 1948. The particles were assumed each
to have a single anisotropy axis, and the reverse field parallel to the axis necessary
for magnetic reversal was the anisotropy field Ha = 2Ku /μ0 Ms , potentially a very
large value. There was no coercivity when the field was applied perpendicular to the
axis. Insights arose from the substantial deviation of real systems from the idealized
Stoner-Wohlfarth model.
Meanwhile, the following year Néel, seeking to understand the remanent magnetism and hysteresis of baked clay and igneous rocks, proposed a model of
thermally driven fluctuations of the magnetization of nanometer-sized ferromagnetic
particles of volume V, a phenomenon known as superparamagnetism. The fluctuation time depended exponentially on the ratio of the energy barrier to magnetic
1 History of Magnetism and Basic Concepts
37
reversal reversal ≈ Ku V to the thermal energy kB T. Here Ku is the uniaxial
anisotropy (Eq. 16) of shape or magnetocrystalline origin. The expression for the
time τ that elapses before a magnetic reversal is
τ = τ0 exp (/kB T )
(19)
where the attempt frequency 1/τ0 was taken to be the natural resonance frequency,
∼109 Hz. When the particles are superparamagnetic, the magnetization of particles
smaller than a critical size fluctuates rapidly above a critical blocking temperature.
The magnetization at lower temperatures, or for larger particles, does not fluctuate
on the measurement timescale, and the particles are then said to be blocked. The
blocking criterion for magnetic measurements at room temperature is defined,
somewhat arbitrarily, as /kB T ≈ 25, corresponding to τ ≈ 100 s and ≈ 1 eV
(see Chap. 20, “Magnetic Nanoparticles”). The 10-year stability criterion is
/kB T ≈ 40. Cooling an ensemble of particles through the blocking temperature
Tb = Ku V/25kB in a magnetic field leads to a relatively stable thermoremanent
magnetization. The typical size of iron oxide particles that are superparamagnetic at
room temperature is 10 nm.
The magnetization of baked clay becomes blocked on cooling through Tb in the
Earth’s magnetic field. From the direction of the thermoremanent magnetization
of appropriately dated hearths of pottery kilns, records of the historical secular
variation of the Earth’s field could be established, a topic known as archeomagnetism. Application of the same idea of thermoremanent magnetization to cooling
of igneous rocks in the Earth’s field provided a direct and convincing argument
for geomagnetic reversals and continental drift; rocks cooling at different periods
experienced fields of different polarities (Fig. 16), which followed an irregular
sequence on a much longer timescale than the secular variation. The reversals could
be dated using radioisotope methods on successive lava flows. This gave birth to the
subfield of paleomagnetism and in turn allowed dating of the patterns of remanent
magnetization picked up in oceanographic surveys conducted in the 1960s that
established the reality of seafloor spreading. The theory of global plate tectonics
has had far-reaching consequences for Earth science [29].
Superparamagnetic particles have found other practical uses. Ferrofluids, the
colloidal suspensions of nanoparticles in oil or water with surfactants to inhibit
agglomeration, are just one. They behave like anhysteretic ferromagnetic liquids.
Individual particles or micron-sized polymer beads loaded with many of them
may be functionalized with streptavidin and used as magnetic labels for specific
biotin-tagged biochemical species, enabling them to be detected magnetically and
separated by high-gradient magnetic separation based on the Kelvin force on a
particle with moment m, fK = (m.∇)B. Medical applications of magnetic fine
particles include hyperthermia (targeted heating by exposure to a high-frequency
magnetic field) and use as contrast agents in magnetic resonance imaging. However
the most far-reaching application of magnetic nanoparticles so far has been in
magnetic recording.
38
J. M. D. Coey
Fig. 16 Polarity of the thermoremanent magnetization measured across the floor of the Atlantic
ocean (left). Current polarity is dark; reversed polarity is light. The pattern is symmetrical about the
mid-ocean ridge, where new oceanic crust is being created. Random reversals of the Earth’s field
over the past 5 My, which are dated from other igneous lava flows, determine the chronological
pattern (right) that is used to determine the rate of continental drift, of order centimeters per year.
(McElhinney, Palaeomagnetism and Plate Tectonics [29], courtesy of Cambridge University Press)
Magnetic Recording
Particulate magnetic recording enjoyed a heyday that lasted over half a century,
beginning with analog recording on magnetic tapes in Germany in the 1930s through
digital recording on the hard and floppy discs that were introduced in the 1950s and
1960s, before eventually being superseded by thin-film recording in the late 1980
[27]. Particulate magnetic recording [30] was largely based on acicular particles of
γFe2 O3 often doped with 1–2% Co. Elongated iron particles were also used, and
acicular CrO2 was useful for rapid thermoremanent reproduction of videotapes on
account of its low Curie temperature. Magnetic digital tape recording with hard
ferrite particulate media continues to be used for archival storage.
The trend with magnetic media has always been to cram ever more digital
data onto ever smaller areas. This has been possible because magnetic recording
technology is inherently scaleable since reading is done by sensing the stray
field of a patch of magnetized particles. It follows from Eq. 2 that since the
dipole field decays as 1/r3 and the moment m ∼ Mr3 , the magnitude of B is
unchanged when everything else shrinks by the same scale factor – at least until
the superparamagnetic limit KV/kB T ≈ 40 is reached, at which point the magnetic
records become thermally unstable. To continue the scaling to bit sizes below
1 History of Magnetism and Basic Concepts
39
Fig. 17 Exponential growth of magnetic recording density over 50 years. The lower panel shows
the magnetized magnetic medium with successive generations of read heads based on anisotropic
magnetoresistance (AMR), giant magnetoresistance (GMR), and tunnel magnetoresistance (TMR)
100 nm, granular films of a highly anisotropic tetragonal Fe-Pt alloy are used
to maintain stability of the magnetic records on ever-smaller oriented crystalline
grains. The individual grains are less than 8 nm in diameter. Over the 65-year history
of hard disc magnetic recording, the bit density has increased by eight orders of
magnitude, at ever-decreasing cost (Fig. 17). Copies cost virtually nothing, and the
volume of data stored on hard discs in computers and data centers doubles every
year, so that as much new data is recorded each year as was ever recorded in all
previous years of human history. This data explosion is unprecedented, and the
third magnetic revolution, the big data revolution, is sure to have profound social
and economic consequences. Although flash memory has displaced the magnetic
hard discs from personal computers. The huge data centres, which are the physical
embodiment of the ‘cloud’ where everything we download from the interenet is
stored continue to use hard disc drives.
40
J. M. D. Coey
Methods of Investigation
Magnetism is an experimental science, and progress in understanding and applications is generally contingent on advances in fabrication and measurement technology, whether it was fourteenth-century technology to fabricate a lodestone sphere
or twenty-first-century technology to prepare and pattern a 16-layer thin-film stack
for a magnetic sensor. The current phase of information technology relies largely on
semiconductors to process digital data and on magnets for long-term storage.
For many physical investigations, magnetic materials are needed in special
forms such as single crystals or thin films. Crystal growers have always been
assiduously cultivated by neutron scatterers and other condensed matter physicists.
Only with single crystals can tensor properties such as susceptibility, magnetostriction, and magnetotransport be measured properly. Nanoscale magnetic composites
have extended the range of magnetic properties available in both hard and soft
magnets. After 1970, thin-film growth facilities (sputtering, electron beam evaporation, pulsed laser deposition, molecular beam epitaxy) began to appear in
magnetism laboratories worldwide. Ultra-high vacuum has facilitated the study of
surface magnetism at the atomic level, while some of the motivation to investigate
magneto-optics or magnetoresistance of metallic thin films, especially in thin-film
heterostructures, arose from the prospect of massively improved magnetic data
storage. Experimental methods are discussed in the chapters in Part 3 of this
Handbook.
Materials Preparation
Silicon steel has been produced for electromagnetic applications by hot rolling
since the beginning of the twentieth century. Annual production is now about 15
million tonnes, half of it in China. Permanent magnets, soft ferrites, and specialized
magnetic alloys are produced in annual quantities ranging from upward of a hundred
to a million tonnes. All such bulk applications of magnetism are highly sensitive to
the cost of raw materials. This effectively disqualifies about a third of the elements
in the periodic table and half of the heavy transition elements from consideration
as alloy additives in bulk material. Newer methods such as mechanical alloying
of elemental powders and rapid quenching from the melt by strip casting or melt
spinning have joined the traditional methods of high-temperature furnace synthesis
of bulk magnetic materials.
The transformation of magnetic materials science that has gathered pace since
1970 has been triggered by the ability to prepare new materials for magnetic devices
in thin-film form. The minute quantity of material needed for a magnetic sensor
or memory element, where the layers are tens of nanometers thick, means that
any useful stable element can be considered. Platinum, for example, may sell for
$30,000 per kilogram, yet it is an indispensable constituent of the magnetic medium
in the 400 million hard disc drives shipped each year that sell for about $60 each.
1 History of Magnetism and Basic Concepts
41
Uniform magnetic thin films down to atomic-scale thicknesses are produced in
many laboratories by e-beam evaporation, sputtering, pulsed laser deposition, or
molecular beam epitaxy, and the more complex tools needed to make patterned
multilayer nanometer-scale thin-film stacks are quite widely available in research
centers, as well as in the fabs of the electronics industry, which deliver the hardware
on which the technology for modern life depends.
Experimental Methods
Advances in experimental observation underpin progress in conceptual understanding and technology. The discovery of magnetic resonance, the sharp absorption
of microwave or radiofrequency radiation by Zeeman split levels of the magnetic
moment of an atom or a nucleus in a magnetic field, or the collective precession
of the entire magnetic moment of a solid was a landmark in modern magnetism.
Significant mainly for the insight provided into solids and liquids at an atomic
scale, electron paramagnetic resonance (EPR) was discovered by Yevgeny Zavoisky
in 1944, and Felix Bloch and Edward Purcell established the existence of nuclear
magnetic resonance (NMR) 2 years later. In 1958, Rudolf Mössbauer discovered a
spectroscopic variant making use of low-energy gamma rays emitted by transitions
from the excited states of some stable isotopes of iron (Fe57 ) and certain rare earths
(Eu151 , Dy161 , etc.). All except Zavoisky received a Nobel Prize. The hyperfine
interactions of the multipole moments of the nuclei (electric monopole, magnetic
dipole, nuclear quadrupole) offered a point probe of electric and magnetic fields at
the heart of the atom.
Larmor precession of the total magnetization of a ferromagnet in its internal
field, usually in a resonant microwave cavity, was discussed theoretically by Landau
and Evgeny Lifshitz in 1935, and ferromagnetic resonance (FMR) was confirmed
experimentally 10 years later.
Of the non-resonant experimental probes, magnetic neutron scattering has
probably been the most influential and generally useful. A beam of thermal neutrons
from a nuclear reactor was first exploited for elastic diffraction in the USA in 1951
by Clifford Shull and Ernest Wohlan, who used the magnetic Bragg scattering
to reveal the antiferromagnetic order in MnO. Countless magnetic structures
have been determined since, using the research reactors at Chalk River, Harwell,
Brookhaven, Grenoble, and elsewhere. Magnetic excitations can be characterized by
inelastic scattering of thermal neutrons, with the help of the triple-axis spectrometer
developed in Canada by Bertram Brockhouse at Chalk River in 1956. Complete
spin-wave dispersion relations provide a wealth of information on anisotropy and
exchange. Newer accelerator-based neutron spallation sources at ISIS, Oak Ridge,
and Lund provide intense pulses of neutrons by collision of highly energetic protons
with a target of a heavy metal such as tungsten or mercury. They are most useful
for studying magnetization dynamics. The low neutron scattering and absorption
cross sections of most stable isotopes mean that neutrons can penetrate deeply into
condensed matter.
42
J. M. D. Coey
Besides neutrons, other intense beams of particles or electromagnetic radiation
available at large-scale facilities have proved invaluable for probing magnetism. The
intense, tunable ultraviolet and X-ray radiation from synchrotron sources allows
the measurement of magnetic dichroism from deep atomic levels and permits the
separate determination of spin and orbital contributions to the magnetic moment.
The spectroscopy is element-specific and distinguishes different charge states of the
same element. Spin-sensitive angular-resolved photoelectron spectroscopy makes it
possible to map the spin-resolved electronic band structure. Muon methods are more
specialized; they depend on the Larmor precession of short-lived (2.20 μs) positive
muons when they are implanted into interstitial sites in a solid. Magnetic scattering
methods are discussed in Chap. 25, “Magnetic Scattering.” The specialized
instruments accessible at large-scale facilities supplement the traditional benchtop
measurement capabilities of research laboratories.
Perhaps the most versatile and convenient of these, used to measure the magnetization and susceptibility of small samples, is the vibrating sample magnetometer
invented by Simon Foner in 1956 and now a workhorse in magnetism laboratories
across the world. The sample is vibrated in a uniform magnetic field, produced by
an electromagnet or a superconducting coil, about the center of a set of quadrupole
pickup coils, which provide a signal proportional to the magnetic moment. Since
sample mass rather than sample volume is usually known, it is generally the mass
susceptibility χ m = χ /ρ that is determined.
Superconducting magnets now provide fields of up to 20 tesla or more for
NMR and general laboratory use. The 5–10 T magnets are common, and they
are usually cooled by closed-cycle cryocoolers to avoid wasting helium. Coupled
with superconducting SQUID sensors, ultrasensitive magnetometers capable of
measuring magnetic moments of 10−10 Am2 or less are widely available. (The
moment of a 5 × mm2 ferromagnetic monolayer is of order 10−8 Am2 .)
High magnetic fields, up to 35 T, require expensive special installations with
water-cooled Bitter magnets consuming many megawatts of electrical power. Resistive/superconducting hybrids in Tallahassee, Grenoble and Tsukuba, and Nijmegen
can generate steady fields in excess of 40 T. Higher fields imply short pulses;
the higher the field, the shorter the pulse. Reusable coils generate pulsed fields
approaching 100 T in Los Alamos, Tokyo, Dresden, Wuhan, and Toulouse.
Magnetic domain structures are usually imaged by magneto-optic Kerr
microscopy, magnetic force microscopy, or scanning electron microscopy, although
scanning SQUID and scanning Hall probe methods have also been developed.
The Bitter method with a magnetite colloid continues to be used. All these
methods image the surface or the stray field near the surface. Ultra-fast, picosecond
magnetization dynamics are studied by optical pulse-probe methods based on the
magneto-optic Kerr effect (MOKE). Transmission electron microscopy reveals the
atomic structures of thin films and interfaces with atomic-scale resolution, while
Lorentz microscopy offers magnetic contrast and holographic methods are able to
image domains in three dimensions. Atomic-scale resolution can be achieved by
point-probe methods with magnetic force microscopy or spin-polarized scanning
tunnelling microscopy. The shift of focus in magnetism toward thin films and
1 History of Magnetism and Basic Concepts
43
thin-film devices has been matched by the development of the sensitive analytical
methods needed to characterize them. Hysteresis in thin films is conveniently
measured by MOKE or by anomalous Hall effect (AHE) when the films are
magnetized perpendicular to their plane. Magnetic fields and measurements are
discussed in Chap. 22, “Magnetic Fields and Measurements” and other chapters
in Part 3.
An important consequence of the increasing availability of commercial superconducting magnets from the late 1960s was the development of medical diagnostic
imaging of tissue based on proton relaxation times measured by NMR. Thousands
of these scanners in hospitals across the world provide doctors with images of the
hearts, brains, bones, and every sort of tumor.
Computational Methods
After about 1980, computer simulation began to emerge as a third force, besides
experiment and theory, to gain insight into the physics of correlated electrons in
magnetic systems. Contributions are mainly in two areas. One is calculation of
the electronic structure, magnetic structure, magnetization, Curie temperature, and
crystal structure of metallic alloys and compounds by using the density functional
method. Magnetotransport in thin-film device structures can also be calculated. Here
there is potential to seek and evaluate new magnetic phases in silico, before trying
to make them in the laboratory. This magnetic genome program is in its infancy;
success with magnetic materials to date has been limited, but the prospects are
enticing.
The other area where computation has become a significant source of new insight
is micromagnetic simulation. The domain structure and magnetization dynamics of
magnetic thin-film structures and model heterostructures are intensely studied, both
in industrial and academic laboratories. Simulation overcomes the surface limitation
of experimental domain imaging. Software is generally based on finite element
methods or the Landau-Lifshitz-Gilbert equation for magnetization dynamics.
Spin Electronics
As technology became available in the 1960s and 1970s to prepare high-quality
metallic films with thicknesses in the nanometer range, interest in their magnetostansport properties grew. The terrain was being prepared for the emergence of a
new phase of research that has grown to become the dominant theme in magnetism
today – spin electronics. Spin electronics is the science of electron spin transport in
solids. Many chapters in the Handbook deal with its various aspects.
For a long time, conventional electronics treated electrons simply as elementary Fermi-Dirac particles carrying a charge e, but it ignored their spin angular
momentum ½. At first this was entirely justified; charge is conserved – the electron
has no tendency to flip between states with charge ± e, no matter how strongly
44
J. M. D. Coey
it is scattered. But angular momentum is not conserved, and spin flip scattering
is common in metals. Perhaps one scattering event in 100 changes the electron
spin state, so the spin diffusion length ls should be about ten times the mean free
path λ of the electron in a solid. When electronic device dimensions were many
microns, there was no chance of an electron retaining the memory of any initial spin
polarization it may have had, unless the device itself was ferromagnetic. Anisotropic
magnetoresistance, where the scattering depends slightly on the relative orientation
of the current and magnetization because of spin-orbit coupling, can be regarded
as the archetypical spin electronic process. The relative magnitude of effect in
permalloy, for instance, is only ∼2%, but the alloy is extremely soft, on account
of simultaneously vanishing anisotropy and magnetostriction, so a permalloy strip
with current flowing at 45◦ to the magnetic easy axis along the strip for maximum
sensitivity – which can be achieved by a superposed “barber pole” pattern of highly
conducting gold – makes a simple, miniature sensor for low magnetic fields, with
a reasonable signal-to-noise ratio. AMR sensors replaced inductive sensors in the
heads used to read data from hard discs in 1990, and the annual rate of increase of
storage density improved sharply as a result.
Meanwhile, research activity on thin-film heterostructures where the layer
thickness was comparable to the spin diffusion length began to pick up as more
sophisticated thin-film vacuum deposition tools were developed. Spin diffusion
lengths are 200 nm in Cu, or about ten times the mean free path, as expected, but
they are shorter in the ferromagnetic elements and sharply different for majorityand minority-spin electrons. The mean free path for minority-spin electrons in Co
is only 1 nm. Particularly influential and significant was the work carried out in
1988 in the groups of Peter Grunberg in Germany and Albert Fert in France on
multilayer stacks of ferromagnetic and nonferromagnetic elements that led to the
discovery of giant magnetoresistance (GMR). The effect depended on electrons
retaining some of their spin polarization as they emerged from a ferromagnetic
layer and crossed a nonmagnetic layer before reaching another ferromagnetic layer.
Big changes of resistance were found when the relative alignment of the adjacent
ferromagnetic iron layers in an Fe-Cr multilayer stack was altered from antiparallel
to parallel by applying a magnetic field (Fig. 18). At first, large magnetic fields and
low temperatures were needed to see the resistance changes, but the structure was
soon simplified to a sandwich of just two ferromagnetic layers with a copper spacer
that became known as a spin valve. Spin valves worked at room temperature, and
they were sensitive to the small stray fields produced by recorded magnetic tape or
disc media. In order to make a useful sensor, it was necessary to pin the direction
of magnetization of one of the ferromagnetic layers while leaving the other free to
respond to an in-plane field (Fig. 19).
It was here that the phenomenon of exchange bias came to the rescue. First
discovered in Co/CoO core shell particles by Meiklejohn and Bean in 1956, it was
extended to antiferromagnetic/ferromagnetic thin-film pairs in Néel’s laboratory
in Grenoble in the 1960s. By pinning one ferromagnetic layer with an adjacent
antiferromagnet (initially NiO), a useful GMR sensor could be produced with
a magnetoresistance change of order 10%. Exchange-biased GMR read heads
1 History of Magnetism and Basic Concepts
45
Fig. 18 Original measurement of giant magnetoresistance of a FeCr multilayer stack, where
the iron layers naturally adopt an antiparallel conduction, which can be converted to a parallel
configuration in an applied field [31]
developed by Stuart Parkin and colleagues went into production at IBM in 1998 – a
remarkably rapid transfer from a laboratory discovery to mass production. Exchange
bias was the first practical use of an antiferromagnet. The Nobel Physics Prize was
awarded to Fert and Grunberg for their work in 2007.
Subsequent developments succeeded in eliminating the influence of the stray
field of the pinned layer on the free layer by means of a synthetic antiferromagnet.
This was another sandwich stack, like the slimmed-down spin valve, except the
spacer was not copper, but an element that transferred exchange coupling from one
ferromagnetic layer to the other. Ruthenium proved to be ideal, and a layer just
0.7 nm thick was found to be ideal for antiferromagnetic coupling [32].
GMR’s tenure as read-head technology was to prove as short-lived as that of
AMR. A new pretender with a much larger resistance change was based on the
magnetic tunnel junction (MTJ), a modified spin valve where the nonmagnetic
metal spacer is replaced by a thin layer of nonmagnetic insulator. Electron tunneling
across an atomically thin vacuum barrier had been a striking prediction of quantum
mechanics implicit in the idea of the wavefunction. The thin barrier was at first made
of amorphous alumina, but it was replaced by crystalline MgO after it was found in
2004 that junctions where the MgO barrier acts as a spin filter exhibit tunneling
magnetoresistance (TMR) in excess of 200% [33, 34] (Fig. 19). The adoption of
46
J. M. D. Coey
B
B
I
free
pinned
free
pinned
af
I
ΔR
af
ΔR
MR = (R↑↓−R↑↑)/R↑↑
B
Spin valve sensor
B
Magnetic tunnel junction (MTJ)
Fig. 19 Magnetic bilayer spin-valve stacks used as sensor (left) or as a memory element (right). In
each case, the magnetization lies in-plane, and the lower ferromagnetic reference layer is pinned by
exchange bias with the purple underlying antiferromagnetic layer, while the upper ferromagnetic
free layer changes its orientation in response to the applied magnetic field. The change in stack
resistance is plotted as a function of applied field. The magnetoresistance ratio MR is defined as the
normalized resistance change between parallel and antiparallel orientation of the two ferromagnetic
layers
TMR sensors in read heads in 2005 was accompanied by a change from in-plane to
perpendicular recording on the magnetic medium.
Despite the changing generations of readers, the hard disc writer remained
what is always had been, a miniature electromagnet that delivers sufficient flux
to a patch of magnetic medium to overcome its coercivity and write the record.
The extreme demands of magnetic recording have driven contactless magnetic
sensing to new heights of sensitivity and miniaturization requiring increasingly
hard magnetic media and new ways of writing them. Thin-film GMR and TMR
structures have also taken a new life as magnetic switches for nonvolatile memory
and logic. Most prominent is magnetic random access memory (MRAM), where
huge arrays of memory cells are based on magnetic tunnel junctions. Magnetic
sensing is discussed in Chaps. 31, “Magnetic Sensors,” and 22, “Magnetic
Fields and Measurements.”
Magnetic thin-film technology has now advanced to the point where uniform
layers in synthetic antiferromagnets and magnetic tunnel junctions only a few atoms
thick are routinely deposited on entire 200 or 300 mm silicon wafers. A corollary
of the short spin diffusion length of electrons in metals is the short distance – a
few atomic monolayers – necessary for an electron to acquire spin polarization on
transiting a ferromagnetic layer. Spin-polarized electron currents are central to spin
electronics.
1 History of Magnetism and Basic Concepts
47
The relation between magnetism and the angular momentum of electrons was
unveiled in Larmor precession and the Einstein-de Haas experiment over a hundred
years ago, but only in the present century has it become commonplace to associate
electric currents with short-range flows of angular momentum. A spin-polarized
current carrying its angular momentum into a ferromagnetic thin-film element can
exert torque in two ways. It can create an effective magnetic field, causing Larmor
precession of the magnetization of the element, and it can exert spin transfer torque,
described by John Slonczewski in 1996 that counteracts damping of the precession
and can be used to stabilize high-frequency oscillations or switch the magnetization
without the need for an external magnetic field. Spin torque switching is effective
for elements smaller than 100 nm in size, and unlike switching by current-induced
“Oersted” fields, it is scalable – an essential requirement for electronic devices. Luc
Berger showed that spin torque can also be used to manipulate domain walls.
A recurrent theme in the recent development of magnetism is the role of the spinorbit interaction. It is critically important in thin films [35], being responsible not
only for the Kerr effect, magnetocrystalline anisotropy, and anisotropic magnetoresistance but also for the anomalous Hall effect and the spin Hall effect, whereby
spin-orbit scattering of a current passing through a heavy metal or semiconductor
produces a buildup of electrons with opposite spin on opposite sides of the
conductor. This transverse spin current created by spin-orbit scattering enables the
injection of angular momentum into an adjacent ferromagnetic layer and the change
of its magnetization direction, an effect known as spin-orbit torque. Conversely,
the inverse spin Hall effect is the appearance of a voltage across the heavy metal on
pumping spin-polarized electrons into it from an adjacent ferromagnet, for example,
by exciting ferromagnetic resonance.
The origin of the intrinsic anomalous Hall effect was an open question in
magnetism, for well over a hundred years. A consensus is now building that it is due
to the geometric Berry phase acquired by electrons moving adiabatically through a
magnetic medium. The phase can be acquired from a non-collinear spin structure
in real space or from topological singularities in the band sturcture in reciprocal
space. Circular micromagnetic defects, known as skyrmions are also topologically
protected.
Another manifestation of spin-orbit interaction is the Rashba effect; when an
electric current is confined at an interface or surface, it tends to create a spin
polarization normal to the direction of current flow. One of the most remarkable
surface phenomena, arising from work by Haldane in 1988, is the possibility of
topologically protected spin currents. A special feature of the band structure ensures
that electrons at the surface or edges of some insulators or semiconductors are in
gapless states. Electrons in these states can propagate around the surface without
scattering, and they exhibit a spin order that winds around the surface as the
direction of electron spin is usually locked at right angles to their linear momentum.
Electrons at surfaces and interfaces can behave quite differently from electrons in
the bulk, and interfaces are at the heart of electronic devices. The introduction
of topological concepts into the discussion of spin-polarized electronic transport
48
J. M. D. Coey
and magnetic defects is providing new insight into magnetism at the atomic and
mesoscopic scales.
Conclusion
Magnetism since 1945 has been an area rich in discovery and useful applications,
not least because of the tremendous increase in numbers of scientists and engineers
working in the field. Magnet ownership for citizens of the developed world has
skyrocketed from 1 or 2 magnets in 1945 to 100–200 60 years later or something
of order a trillion if we count the individual magnetic bits on a hard disc in
a desktop computer. Countless citizens throughout the world during this period
already experienced magnetism’s bounty at first hand in the form of a cassette tape
recorder, and nowadays they can access the vast stores of magnetically recorded
information in huge data centers via the Internet using a handheld device.
Magnetism is therefore playing a crucial role in the big data revolution that is
engulfing us, by enabling the permanent data storage, from which we can make
instant copies at practically no cost. It may deliver more nonvolatile computer
memory if MRAM proves to a winning technology and possibly facilitate data
transfer at rates up to the terahertz regime with the help of spin torque oscillators.
There are potential magnetic solutions to the problems of ballooning energy
consumption and the data rate bottleneck. There is potential to implement new
paradigms for computation magnetically. While there is no certainty regarding the
future form of information technology, improved existing solutions often have an
inside track. Magnetism and magnetic materials may be a good bet.
There have been half a dozen paradigm shifts – radical changes in the ways of
seeing and understanding the magnet and its magnetic field – during its 2000-year
encounter with human curiosity. Implications of the big data revolution for human
society are only beginning to come into focus, but they are likely to be as profound as
on the previous two occasions when magnetism changed the world. This Handbook
is a guide to what is going on.
Acknowledgments The author is grateful to Science Foundation Ireland for continued support,
including contracts 10/IN.1/I3006, 13/ERC/I2561 and 16/IA/4534.
Appendix: Units
By the middle of the nineteenth century, it was becoming urgent to devise a standard
set of units for electrical and magnetic quantities in order to exchange precise
quantitative information. The burgeoning telegraph industry, for example, needed
a standard of electrical resistance to control the quality of electrical cables. Separate
electrostatic and electromagnetic unit systems based on the centimeter, the gram and
the second had sprung into existence, and Maxwell and Jenkin proposed combining
them in a coherent set of units in 1863. Their Gaussian cgs system was adopted
1 History of Magnetism and Basic Concepts
49
internationally in 1881. Written in this unit system, Maxwell’s equations relating
electric and magnetic fields contain explicit factors of c, the velocity of light.
Maxwell also introduced the idea of dimensional analysis in terms of the three basic
quantities of mass, length, and time. The magnetic field H and the induction B are
measured, respectively, in the numerically identical but dimensionally different units
of oersted (Oe) and gauss (G).
Another basic unit, this time of electric current, was adopted in the Système
International d’Unités (SI) in 1948. The number of basic units and dimensions in any
system is an arbitrary choice; the SI (International System of Units) uses four insofar
as we are concerned, the meter, kilogram, second, and ampere (or five if we include
the mole). The system has been adopted worldwide for the teaching of science and
engineering at school and universities; it embodies the familiar electrical units of
volt, ampere, and ohm for electrical potential, current, and resistance. Maxwell’s
equations written in terms of two electric and two magnetic fields contain no factors
of c or 4π in this system (Eq. 7), but they inevitably crop up elsewhere. B and H are
obviously different quantities. The magnetic field strength H, like the magnetization
M, has units of Am−1 . The magnetic induction B is measured in tesla (1 T ≡
1 kgs2 A−2 ). Magnetic moments have units of Am2 , clearly indicating the origin
of magnetism in electric currents and the absence of magnetic poles as real physical
entities. The velocity of light is defined to be exactly 299,792,458 ms−1 . The two
constants μ0 and ε0 , the permeability and permittivity of free space, are related by
μ0 ε0 = c2 , where μ0 was 4π 10−7 kgs−2 A−2 according to the original definition of
the ampere. However, in the new version of SI, which avoids the need for a physical
standard kilogram, the equality of μ0 and 4π 10−7 is not absolute, but it is valid to
ten significant figures.
Only two of the three fields B, H, and M are independent (Fig. 4). The relation
between them is Eq. 8, B = μ0 (H + M). This is the Sommerfeld convention for SI.
The alternative Kenelly convention, often favored by electrical engineers, defines
magnetic polarization as J = μ0 M, so that the relation becomes B = μ0 H + J. We
Table 1 Numerical conversion factors between SI and cgs units
Physical quantity
B-field (magnetic flux
density)
H-field (magnetic field
intensity)
Magnetic moment
Magnetization
Specific magnetization
Magnetic energy density
Dimensionless
susceptibility M/H
Symbol
B
SI to cgs conversion
1 tesla = 10 kilogauss
H
1 kAm−1 = 12.57 oersted
m
M
σ
1 Am2 = 1000 emu
1 Am−1 = 12.57 gauss†
1 Am2 kg−1 = 1 emu g−1
(BH)
χ
1 kJm−3 = 0.1257 MGOe
1 (SI) = 1/4π (cgs)
*symbol G; § symbol Oe; † 4πM; Note: 12.57 = 4π; 79.58 = 1000/4π
cgs to SI conversion
1 gauss* = 0.1
millitesla
1 oersted§ = 79.58
Am−1
1 emu = 1 mAm2
1 gauss† = 79.58 Am−1
1 emu g−1 = 1
Am2 kg−1
1 MGOe = 7.96 kJm−3
1 (cgs) = 4π (SI)
50
J. M. D. Coey
follow the Sommerfeld convention in this Handbook. The magnetic field strength H
is not measured in units of Tesla in any generally accepted convention, but it can be
so expressed by multiplying by μ0 .
At the present time, Gaussian cgs units remain in widespread use in research
publications, despite the obvious advantages of SI. The use of the cgs system in
magnetism runs into the difficulty that units of B and H, G and Oe, are dimensionally
different but numerically the same; μ0 = 1, but it normally gets left out of the
equations, which makes it impossible to check whether the dimensions balance.
Table 1 lists the conversion factors and units in the two systems. The cgs equivalent
of Eq. 8 is B = H + 4πM. The cgs unit of charge is defined in such a way that
ε0 = 1/4πc and μ0 = 4π/c so factors of c appear in Maxwell’s equations in place
of the electric and magnetic constants. Convenient numerical conversion factors
between the two systems of units are provided in Table 1.
Theoretical work in magnetism is sometimes presented in a set of units where
c = = kB = 1. This simplifies the equations, but does nothing to facilitate
quantitative comparison with experimental measurements.
References
1. Kloss, A.: Geschichte des Magnetismus. VDE-Verlag, Berlin (1994)
2. Matthis, D.C.: Theory of Magnetism, ch. 1. Harper and Row, New York (1965)
3. Needham, J.: Science and Civilization in China, vol. 4, part 1. Cambridge University Press,
Cambridge (1962)
4. Schmid, P.A.: Two early Arabic sources of the magnetic compass. J. Arabic Islamic Studies. 1,
81–132 (1997)
5. Fowler, T.: Bacon’s Novum Organum. Clarendon Press, Oxford (1878)
6. Pierre Pèlerin de Maricourt: The Letter of Petrus Peregrinus on the Magnet, AD 1292, Trans.
Br. Arnold. McGraw-Hill, New York (1904)
7. Gilbert, W: De Magnete, Trans. P F Mottelay. Dover Publications, New York (1958)
8. Fara, P.: Sympathetic Attractions: Magnetic Practices, Beliefs and Symbolism in EighteenthCentury England. Princeton University Press, Princeton (1996)
9. Mottelay, P.F.: Bibliographical History of Electricity and Magnetism. Arno Press, New York
(1975)
10. Faraday, M.: Experimental Researches in Electricity, volume III. Bernard Quartrich, London
(1855)
11. Maxwell, J.C.: A Treatise on Electricity and Magnetism, two volumes. Clarendon Press,
Oxford (1873) (Reprinted Cambridge University Press, 2010)
12. Bertotti, G.: Hysteresis in Magnetism. Academic Press, New York (1998)
13. Brown, W.F.: Micromagnetics. Interscience, New York (1963)
14. Sato, M., Ishii, Y.: Simple and approximate expressions of demagnetizing factors of uniformly
magnetized rectangular rod and cylinder. J. Appl. Phys. 66, 983–988 (1989)
15. Hunt, B.J.: The Maxwellians. Cornell University Press, New York (1994)
16. Joule, J.P.: On the Effects of Magnetism upon the Dimensions of Iron and Steel Bars.
Philosoph. Mag. Third Series. 76–87, 225–241 (1847)
17. Thomson, W.: On the electrodynamic qualities of metals. Effects of magnetization on the
electric conductivity of nickel and iron. Proc. Roy. Soc. 8, 546–550 (1856)
18. Kerr, J.: On rotation of the plane of the polarization by reflection from the pole of a magnet.
Philosoph. Mag. 3, 321 (1877)
1 History of Magnetism and Basic Concepts
51
19. Ewing, J.A.: Magnetic Induction in Iron and Other Metals, 3rd edn. The Electrician Publishing
Company, London (1900)
20. Tomonaga, S.: The Story of Spin. University of Chicago Press, Chicago (1974)
21. Marage, P., Wallenborn, G. (eds.): Les Conseils Solvay et Les Débuts de la Physique Moderne.
Université Libre de Bruxelles (1995)
22. Ballhausen, C.J.: Introduction to Ligand Field Theory. McGraw Hill, New York (1962)
23. Bozorth, R.M.: Ferromagnetism. McGraw Hill, New York (1950) (reprinted Wiley – IEEE
Press, 1993)
24. Goodenough, J.B.: Magnetism and the Chemical Bond. Interscience, New York (1963)
25. Smit, J., Wijn, H.P.J.: Ferrites; Physical Properties of Ferrrimagnetic Oxides. Philips Technical
Library, Eindhoven (1959)
26. Coey, J.M.D., Viret, M., von Molnar, S.: Mixed valence manganites. Adv. Phys. 48, 167 (1999)
27. Wang, S.X., Taratorin, A.M.: Magnetic Information Storage Technology. Academic Press, San
Diego (1999)
28. Coey, J.M.D. (ed.): Rare-Earth Iron Permanent Magnets. Clarendon Press, Oxford (1996)
29. McElhinney, M.W.: Palaeomagnetism and Plate Tectonics. Cambridge University Press (1973)
30. Daniel, E.D., Mee, C.D., Clark, M.H. (eds.): Magnetic Recording, the First Hundred Years.
IEEE Press, New York (1999)
31. Baibich, M.N., Broto, J.M., Fert, A., Nguyen Van Dau, F., et al.: Giant magnetoresistance of
(001)Fe/(001)Cr magnetic superlattices. Phys. Rev. Lettters. 61, 2472 (1988)
32. Parkin, S.S.P.: Systematic variation on the strength and oscillation period of indirect magnetic
exchange coupling through the 3d, 4d and 5d transition metals. Phys. Rev. B. 67, 3598 (1991)
33. Parkin, S.S.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., et al.: Giant tunneling
magnetoresistance with MgO (100) tunnel barriers. Nat. Mater. 3, 862–867 (2004)
34. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3,
868–871 (2004)
35. Sinova, J., Valenzuela, S.O., Wunderlich, J., Bach, C.H., Wunderlich, J.: Spin hall effects. Rev.
Mod. Phys. 87, 1213 (2015)
Michael Coey received his PhD from the University of Manitoba
in 1971; he has worked at the CNRS, Grenoble, IBM, Yorktown Heights, and, since 1979, Trinity College Dublin. Author
of several books and many papers, his interests include amorphous and disordered magnetic materials, permanent magnetism,
oxides and minerals, d0 magnetism, spin electronics, magnetoelectrochemistry, magnetofluidics, and the history of ideas.
2
Magnetic Exchange Interactions
Ralph Skomski
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Quantum-Mechanical Origin of Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
One-Electron Wave Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Electron-Electron Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Stoner Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Correlations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Heisenberg Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Hubbard Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Specific Exchange Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Intra-Atomic Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Indirect Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Itinerant Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Bethe-Slater Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Metallic Correlations and Kondo Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Exchange and Spin Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Curie Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Order and Noncollinearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin Waves and Anisotropic Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Antiferromagnetic Spin Chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Dimensionality Dependence of Quantum Antiferromagnetism . . . . . . . . . . . . . . . . . . . . . . .
Frustration, Spin Liquids, and Spin Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
54
57
57
59
60
61
63
65
66
67
69
72
75
78
83
83
84
89
92
93
94
95
99
R. Skomski ()
University of Nebraska, Lincoln, NE, USA
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_2
53
54
R. Skomski
Abstract
The electrostatic repulsion between electrons, combined with quantum mechanics and the Pauli principle, yields the atomic-scale exchange interaction. Intraatomic exchange determines the size of the atomic magnetic moments. Interatomic exchange ensures long-range magnetic order and determines the ordering
(Curie or Néel) temperature. It also yields spin waves and the exchange stiffness
responsible for the finite extension of magnetic domains and domain walls.
Intra-atomic exchange determines the size of the atomic magnetic moments.
Positive and negative exchange constants mean parallel (ferromagnetic) and
antiparallel (antiferromagnetic) spin alignments. As a rule, direct exchange
and Coulomb interaction favor ferromagnetic spin structures, whereas interatomic hopping tends to be ferromagnetic and is often the main consideration.
The basic interatomic exchange mechanisms include superexchange, double
exchange, Ruderman-Kittel exchange, and itinerant exchange. Exchange interactions may also be classified according to specific models or phenomena.
Examples are Heisenberg exchange, Stoner exchange, Hubbard interactions,
anisotropic exchange, Dzyaloshinski-Moriya exchange, and antiferromagnetic
spin fluctuations responsible for high-temperature superconductivity. From the
viewpoint of fundamental physics, exchange interactions differ by the role of
electron correlations, the strongly correlated Heisenberg exchange and weakly
correlated itinerant exchange at the opposite ends of the spectrum. Correlations
are also important for the understanding of some exotic exchange phenomena,
such as frustration and quantum spin liquid behavior.
Introduction
Solid-state magnetism is caused by interacting atomic moments or “spins” (Fig. 1).
In the absence of such interactions, the spins would point in random directions, and
the net magnetization would be zero. Ferromagnetic (FM) order requires positive
interactions (a), which favor parallel spin alignment, ↑↑, and yields a nonzero
net magnetization. Antiferromagnetic (AFM) order (b) is caused by negative
interactions and corresponds to antiparallel spin alignment ↑↓ between neighboring
atoms. For reasons discussed below, these interactions are referred to as exchange
interactions. Aside from the interatomic exchange illustrated in Fig. 1, there are
intra-atomic exchange interactions. For example, Fe2+ ions in oxides have six 3d
electrons and the spin structure ↑↑↑↑↑↓, which yields a net atomic moment of
2 μB . Magnetic moments in transition-metal elements and alloys tend to be noninteger, as exemplified by Ni, which has a moment of 0.61 μB per atom. Such
non-integer moments reflect the itinerant Stoner exchange, which contains both
inter- and intra-atomic contributions.
By about 1920, it had become clear that magnetostatic interactions cannot
explain ferromagnetism at and above room temperature. Weiss’ mean-field theory assumes that a molecular field stabilizes ferromagnetic order. However, the
2 Magnetic Exchange Interactions
55
Fig. 1 Interatomic exchange and magnetic order: (a) antiferromagnetism (AFM) and (b) ferromagnetism (FM)
molecular fields required in the theory (several 100 teslas) are much higher than
typical magnetostatic interaction fields, which are only of the order of 1 tesla.
Equating the thermal energy kB T with the Zeeman energy μo μB H yields the
conversion μB /kB = 0.672 K/T, meaning that low-temperature thermal excitations
of about 1 kelvin destroy any magnetic order caused by magnetic fields of about
1 T. The weakness of Zeeman and other magnetic interactions reflects the relativistic
character of magnetism: The ratio of magnetic and electrostatic interactions is of the
order of 1/α 2 , where α = 1/137 is Sommerfeld’s fine structure constant.
Aware of the smallness of purely magnetic interactions, Werner Heisenberg
concluded in 1928 that ferromagnetic order must be of electrostatic origin, realized
on a quantum-mechanical level [1]. He found that the Coulomb repulsion between
electrons
U (r 1 , r 2 ) =
e2
1
4πεo | r 1 − r 2 |
(1)
in combination with the Pauli principle yields a strong effective field consistent
with experiment. The Pauli principle forbids the occupancy of an orbital by two
electrons of parallel spin. In real space, it yields an exchange hole, that is, electrons
with parallel spins (↑↑) stay away from each other, while electrons with antiparallel
spin (↑↓) can come arbitrarily close, which carries a Coulomb-energy penalty. In a
nutshell, this is the origin of ferromagnetic exchange. A different consideration is
that even electrons of antiparallel spin (↑↓) avoid each other to some extent due to
their Coulomb repulsion, which is known as the correlation hole. The correlation
hole weakens the trend toward ferromagnetism.
Consider two electrons 1 and 2 in two atomic orbitals L (for left) and R (for
right). The real-space part of the wave function is
56
R. Skomski
1
± (r 1 , r 2 ) = √ (φL (r 1 )φR (r2 ) ± φR (r 1 )φL (r 2 ))
2
(2)
where the upper and lower signs correspond to ↑↓ and ↑↑, respectively.
Using Eq. (2) to evaluate the Coulomb interaction EC = Ψ ± ∗ U Ψ ± dr1 dr2
yields an energy splitting of ±JD . The integral
±JD =
φL∗ (r 1 )φR∗ (r 2 )U (r 1 , r 2 )φR (r 1 )φL (r 2 )dV1 dV2
(3)
is referred to as the exchange integral or direct exchange. If JD is positive, then
the energy of the FM state is lower than that of the AFM state, favoring ↑↑
alignment. Since direct exchange JD is of electrostatic origin, it has the right order
of magnitude to explain ferromagnetism.
However, equating the net exchange J with JD has a number of flaws. First,
JD is the electrostatic self-interaction energy of a fictitious charge distribution
ρ F (r) = – e φ ∗ L (r) φ R (r) and therefore always positive. This is at odds with experiment, because antiferromagnetism is well established in many materials. Second,
Eq. (2) means that the left (L) and right (R) atoms harbor exactly one electron each
[1, 2]. This approximation, known as the Heitler-London approximation in chemistry, amounts to ignoring the ionic configurations φ L (r1 ) φ L (r2 ) and φ R (r1 ) φ R (r2 ).
In fact, some iconicity is expected on physical grounds, because electrons hop
between atoms and therefore temporarily create ionic configurations.
Third, the
wave functions φ L (r) and φ R (r) exhibit some overlap So = φ L (r) φ R (r) dr. This
overlap is responsible for interatomic hopping, affecting the one-electron levels and
reducing the net exchange. Furthermore, overlap corrections diverge with increasing
number N of electrons involved, which is known as non-orthogonality catastrophe.
Sections “One-Electron Wave Functions” and “Electron-Electron Interactions”
solve the overlap problem by using Wannier-type orthogonalized orbitals.
It is nontrivial to predict magnetism from the chemical composition. For
example, MnBi, ZrZn2 , and CrBr3 are all ferromagnetic but do not contain any
ferromagnetic element. Section “Specific Exchange Mechanisms” describes a number of important exchange mechanisms in metals and insulators. It is important to
distinguish between intra-atomic exchange, which is responsible for the formation
of atomic magnetic moments, and interatomic exchange, which determines the type
of magnetic order and the ordering temperature. Examples of magnetic order are the
FM and AFM structures of Fig. 1, but there also exist noncollinear spin structures,
caused, for example, by competing exchange or Dzyaloshinski-Moriya interactions
(Sections “Curie Temperature” and “Magnetic Order and Noncollinearity”).
Beyond determining magnetic order, exchange is important in micromagnetism,
where the exchange stiffness affects the sizes of magnetic domains and domain
walls (Section “Spin Waves and Anisotropic Exchange”). Exchange is also involved
in various “exotic” magnetic systems, such as frustrated spin structures and
quantum-spin liquids (Sections “Curie Temperature” and “Magnetic Order and
Noncollinearity”).
2 Magnetic Exchange Interactions
57
Quantum-Mechanical Origin of Exchange
Exchange reflects the interplay between independent-electron level splittings (T ),
Coulomb repulsion (U ), and the direct exchange integral (JD ). This section
elaborates the fundamentals of this relationship, starting from one-electron wave
functions (Section “One-Electron Wave Functions”), introducing wave functions of
interacting electrons (Section “Electron-Electron Interactions”), and discussing a
number of fundamental limits and models (Sections “Stoner Limit”, “Correlations”,
“Heisenberg Model”, and “Hubbard Model”).
One-Electron Wave Functions
Well-separated atoms are described by atomic wave functions of on-site energy Eat ,
but in molecules and solids, the wave functions of neighboring atoms overlap and
yield interatomic hybridization. This section focuses on two s electrons in atomic
dimers, such as H2 and hypothetical Li2 (Fig. 2). Most features of this model can be
generalized to solids, although solids exhibit additional many-electron effects. The
one-electron Hamiltonian corresponding to Fig. 2 is
H.1 (r) =
2 2
∇ + VL (r) + VR (r)
2me
(4)
where L and R stand for right and left, respectively, and VL/R (r) = Vo (| r – RL/R | )
are the atomic potentials. In the case of hydrogen-like atoms of radius Rat = ao /Z,
the atomic ground state has the eigenfunction φ(r)∼ exp (–r/Rat ) and the energy
Eat = Z2 e2 /8πεo ao . The vicinity of the second atom means that the atomic wave
functions φ L (r) = |Lo > and φ R (r) = |Ro > overlap, which is described by the overlap
Fig. 2 Symmetric and antisymmetric wave functions: (a) atomic wave function, (b) Wannier
function, (c) antibonding state |σ *>, and (d) bonding state |σ >
58
R. Skomski
integral
So =< Lo VR Ro >
(5)
The overlap causes interatomic hopping and yields a level splitting into bonding and
antibonding states. The respective energies are Eo ± T , where T is the hopping
integral. The hybridized eigenfunctions (Fig. 2(c-d)) are |σ > ∼ |Lo > + |Ro >
(bonding) and |σ ∗ > ∼ |Lo > – |Ro > (antibonding), both having So -dependent
normalizations. The label σ refers to the ss σ -bond between the two s orbitals and
leads to T < 0 in this specific model. The on-site energy Eo differs from the atomic
energy Eat by the crystal-field energy ECF ≈ <Lo |VR |Lo >.
Traditional and Modern Analyses The determination and interpretation of the
hopping integral T require care, because hopping affects the net exchange J and
may change its sign. T is often approximated by
To = <Lo |VR |Ro > =
φL∗ (r) VR (r) φR (r) dr
(6)
but this interpretation is qualitative only, because the overlap correction to T is of
the same order of magnitude as To itself. Equation (6) therefore conflates the related
phenomena of hopping and wave-function overlap. This distinction is related to the
abovementioned non-orthogonality catastrophe.
Orthogonality problems are avoided by the use of orthogonalized atomic wave
functions or Wannier functions [3]. These functions are similar to atomic wave functions but contain some admixture of neighboring orbitals to ensure orthogonality. In
the present model [4]
1 1 |L> = √ |σ > + |σ ∗ > and |R> = √ |σ > − |σ ∗ >
2
2
(7)
Figure 2(b) shows one of these Wannier functions. The two wave functions (c-d)
correspond to rudimentary wave vectors k = 0 (bonding) and k = π/a (antibonding).
Solids are very similar in this regard, except that k varies continuously (band
structure).
It is instructive to discuss the parameters involved for large interatomic distances R [2]. In this extreme-tight-binding limit, So = ½ (R/Rat )2 exp(–R/Rat ),
ECF = 2(Rat /R)Eat , and T = So ECF . Since ECF decreases only slowly, scaling
as 1/R, the asymptotic behavior of T is governed by the exponential decay of So .
In terms of the Wannier functions |L > and |R>, the one-electron Hamiltonian of
Eq. (4 ) assumes the very simple matrix structure
H1 =
Eo T
T Eo
(8)
2 Magnetic Exchange Interactions
59
The diagonalization of this Hamiltonian is trivial, reproducing E± = Eo ± T
and yielding
1
1
|σ > = √ (|L> + |R>) and |σ ∗ > = √ (|L> − |R>)
2
2
(9)
Equations (8, 9) remove the overlap integral from explicit consideration and
constitute a great scientific and practical simplification.
Electron-Electron Interactions
Ferromagnetism is caused by electron-electron interactions. Addition of the
Coulomb energy H12 = U (r 1 , r 2 ) to Eq. (4) yields
H (r 1 , r 2 ) = H1 (r 1 ) + H1 (r 2 ) + U (r 1 , r 2 )
(10)
To diagonalize this Hamiltonian, it is convenient to use two-electron wave functions Ψ i constructed from Wannier functions, namely, Ψ 1 = |LL>, Ψ 2 = |LR>,
Ψ 3 = |RL>, and Ψ 4 = |RR>. Since <L|R > = 0, these functions are all orthogonal,
and Eq. (10) becomes
⎛
U
⎜T
H= ⎜
⎝T
JD
T
0
JD
T
T
JD
0
T
⎞
JD
T ⎟
⎟
T ⎠
U
(11)
where a physically unimportant zero-point energy has been ignored. The Coulomb
parameter U is the extra energy required to put a second electron onto a given atom
(R or L), essentially
U=
e2
4πεo
n(r 1 ) n(r 2 )
dr 1 dr 2
| r1 − r2 |
(12)
where n(r) = nL/R (r). Unlike JD , which decrease exponentially with interatomic
distance, U is an atomic parameter and more or less independent of crystal structure.
Both U and T tend to be large, several eV, whereas JD is rather small, typically of
the order of 0.1 eV. This indicates that JD is not the only or even the most important
contribution to interatomic exchange. For example, the exchange in the H2 molecule
is antiferromagnetic, in spite of JD being positive.
Equation (11) can be diagonalized analytically. There are two low-lying states
1
1
|↑↑> = √ |LR> − √ |RL>
2
2
(13)
60
R. Skomski
cos χ
sin χ
|↑↓> = √ (|LR> + |RL>) + √ (|LL> + |RR>)
2
2
(14)
where tan (2χ ) = –4T /U [4]. Equation (14) is a superposition of two Slater
determinants, described by the mixing angle χ . The corresponding energy levels
are
E↑↑ = –JD
E↑↓
U
= +D −
2
(15)
U2
4
4T 2 +
(16)
Defining an effective exchange as J = E↑↑ –E↑↑ /2 yields
U
J = JD + −
4
T2+
U2
16
(17)
This equation shows that interatomic hopping (T ) reduces the net exchange
interaction. The effect comes from the admixture of |LL> and |RR> to |LR> + |RL>
(Eq. (14)) which is ignored in Eq. (2).
Stoner Limit
In the metallic limit of strong interatomic hopping (T
J =
U
+ JD − |T |
4
U ), Eq. (17) becomes
(18)
This equation predicts ferromagnetism for sufficiently small hopping T and roughly
corresponds to the Stoner theory [5] of itinerant transition-metal magnets (Section “Itinerant Exchange”). Since U
JD , the driving force behind Stoner
ferromagnetism is the Coulomb integral U , not the direct exchange JD [6]. The
interatomic hopping competes against the electron-electron interactions described
by the Stoner parameter I = U /4 + JD , whereas a refined calculation for transition
metals yields I = U /5 + 1.2 Jat [7]. Here Jat is the intra-atomic exchange, which
merges with the interatomic exchange in the itinerant limit.
Equation (18) yields a very simple and scientifically successful explanation,
namely, that ferromagnetism occurs when the one-electron level splitting, ±|T | in
the model of Section “Electron-Electron Interactions”, is sufficiently small compared to the nearly crystal-independent Coulomb parameter U . Figure 3 illustrates
the physics behind this mechanism. The Coulomb repulsion U favors the FM configuration, but the FM alignment carries a one-electron energy penalty. More precisely,
2 Magnetic Exchange Interactions
61
Fig. 3 Origin of magnetism in the independent-electron picture. The one-electron level splitting
into bonding (σ ) and antibonding (σ *) states favors ↑↓ spin pairs, whereas the Coulomb repulsion
between the two |σ > electrons yields ↑↑ coupling so long as the Coulomb energy is larger
than the one-electron level splitting. The independent-electron nature of this picture is seen from
two features. First, the electrons occupy one-electron levels (σ and σ *). Second, the Coulomb
interaction can be interpreted as an effective field (Stoner exchange field)
the “one-electron” contributions of this section are actually independent-electron
contributions treated on a quantum-mechanical mean-field level [8], because level
splittings such as ±|T | depend on all other electrons in the system.
An alternative view on the Stoner limit is that antisymmetrized wave functions
|Ψ > diagonalize the leading one-electron part (T -part) of Eq. (11) and can therefore
be used to evaluate electron-electron interactions (U and JD ) in lowest-order
perturbation theory. The antisymmetric wave functions have the character of Slater
determinants if the spin is included. For example, Fig. 3 corresponds to
|FM > = σ (r1 ) σ ∗ (r2 ) − σ ∗ (r1 ) σ (r2 ) ↑ (1) ↑(2)
(19)
|AFM > = σ (r1 ) σ (r2 ) (↑(1) ↑(2) −↑ (1) ↑(2))
(20)
and
In this method, known as the independent-electron or quantum-mechanical meanfield approximation in solid-state physics and the molecular-orbital (MO) method in
chemistry, individual electrons move in an effective potential or “mean field” Veff (r)
created by all electrons in the system (Section “Itinerant Exchange”).
Correlations
The quantum-mechanical mean-field approach, which is the rationale behind the
local-density approximation to density-functional theory (LSDA DFT), has been
highly successful in magnetism, but some red flags indicate the need for a more
thorough analysis. For example, the mean-field result of Eq. (18) leads to the
prediction of positive (FM) exchange for JD = 0 so long as |T | < 14 U . In fact,
putting JD = 0 in the exact result of Eq. (17) shows that the exchange is always
62
R. Skomski
negative for JD = 0. Ferromagnetic coupling requires
|T | <
JD
U
+ JD
2
(21)
which is qualitatively different from |T | < 14 U .
The limitations of the quantum-mechanical mean-field approach are defined by
the treatment of the correlation hole. The correlation energy is defined [9] as the
difference between the correct many-electron energy and the corresponding oneelectron (independent-electron) energy obtained from a single Slater determinant
(Hartree-Fock determinant).
Consider |Ψ AFM > of Eq. (20), whose real-space part has the structure
|σ σ > =
1
(|LL> + |LR> + |RL> + |RR>)
2
(22)
This wave function has an ionic character of 50%, that is, the electrostatically
unfavorable configurations |LL> and |RR> provide half the weight. Since the
electrons equally occupy all two-electron states, Eq. (22) lacks a correlation hole.
The Coulomb penalty associated with the unfavorable ionic contribution leads to an
overestimation of the AFM energy and therefore to an overestimation of the trend
toward ferromagnetism.
In reality, electron correlations lead to a partial suppression of the |LL> and
|RR > occupancies, described by the mixing angle χ in Eq. (14). The Heisenberg
limit, Ψ + in Eq. (2) and χ = 0 in Eq. (14), has |Ψ AFM > ∼ |LR> + |RL>, which
corresponds to an ionic character of 0% and to a fully developed correlation
hole. The Heisenberg model is said to be overcorrelated, as opposed to the
undercorrelated independent-electron approach. An interesting approach is the use
of Coulson-Fischer wave functions, that is, of Slater determinants constructed not
from |L> and |R> but from combinations such as |L> + λ |R>, where λ ≈ |T |/U for
small hopping [10]. This unrestricted Hartree-Fock approximation contains a part
of the correlations at the expense of symmetry breaking in the Hamiltonian [9]. The
approximation is sufficient to reproduce the correct AFM wave function, Eq. (14),
for the H2 model of Eq. (11), but this finding cannot be generalized to arbitrary
many-electron systems. Near the equilibrium H-H bond length of about 0.74 Å, the
electrons are delocalized, described by Eq. (22) and λ = 1, but above 1.20 Å, the
electrons localize very rapidly and λ approaches zero.
Correlations primarily affect AFM spin configurations [6]. For example, the FM
wave function |σ σ ∗ > – |σ ∗ σ > = |LR> – |RL>, Eq. (13), is independent of T and
U and therefore unaffected by correlations. The reason for the absence of ionic
configurations in Eq. (13) is the Pauli principle, which creates the exchange hole
and forbids |LL > and |RR > occupancies with parallel spin. Correlations effects are
most important in half-filled bands, where ferromagnetism means that all bonding
and antibonding real-space orbitals are occupied by ↑ electrons and the net energy
2 Magnetic Exchange Interactions
63
gain due to interatomic hybridization is zero. Electrons (or holes) added to halffilled bands do not suffer from this constraint and make ferromagnetism easier to
achieve.
Solid-state correlations are multifaceted and yield many more or less closely
related magnetic phenomena, such as spin-charge separation (Section “Antiferromagnetic Spin Chains”), wave-function entanglement, and the fractional quantumHall effect (FQHE). The determination of correlations is demanding even- or
medium-sized molecules or clusters, because the number of configurations to be
considered increases exponentially with system size. For example, the complete
description of a single CH4 molecule (10 electrons) requires the diagonalization of
a matrix containing 43,758 × 43,758 determinants [9]. Some methods to describe
correlations [9–13] are microstate approaches, such as those in this chapter, selfenergy methods, the evaluation of matrix elements between Slater determinants
(known as the configuration interactions, CI), dynamical mean-field theory (DMFT)
[14, 15], and the Bethe ansatz [16, 17]. Unlike LSDA+U, the DMFT is a
true correlation approach, because the electrons keep their individuality and the
mean-field character refers to the spatial aspect of the correlations only. Some
other correlation approaches, such as the Hubbard model, are briefly discussed in
Section “Hubbard Model”.
Heisenberg Model
In the strongly correlated Heisenberg limit (U
J = JD −
T ), Eq. (17) becomes
2T2
U
(24)
Putting U = ∞ yields J = JD and reproduces the naïve Heisenberg result of
Eq. (3). Expressions very similar to Eq. (24) can be derived for solids [3], but the
method is cumbersome and the resulting picture not very transparent. It is often
better to eliminate hopping terms and to consider spin Hamiltonians.
To replace the real-space wave functions (R and L) by spins (↑ and ↓), one
considers the full wave functions in the Heisenberg limit, namely, the AFM singlet
|AFM> =
1
(|LR> + |RL>) (|↑↓> − |↓↑>)
2
(25a)
and a FM triplet
1
|FM ↑↑> = √ (|LR> − |RL>) |↑↑>
2
|FM0> =
1
(|LR> − |RL>) (|↑↓> + |↓↑>)
2
(25b)
(25c)
64
R. Skomski
1
|FM ↓↓> = √ (|LR> − |RL>) |↓↓>
2
(25d)
The triplet (25b-d) reflects Sz = (−1, 0, +1) for S = 2 · ½ = 1 and is split by an
external magnetic field (Zeeman interaction).
In the Heisenberg model, one considers the spin part and implicitly understands
that the spins are located on neighboring atoms. The model involves spin operators
S = 12 σ, where σ is the vector formed by the Pauli matrices. The mathematical
direct-product identity
⎛
1
⎜0
σx ⊗ σx + σy ⊗ σy + σz ⊗ σz = ⎜
⎝0
0
0
1
1
0
0
1
−1
0
⎞
0
0⎟
⎟
0⎠
1
(26)
reproduces the eigenfunctions and the singlet-tripletsplitting of Eq. (25), so that the Heisenberg Hamiltonian can be written as H=–2 J S x ⊗ S x +S y ⊗ S y +S z ⊗ S z ,
in vector notation, H = –2 J S 1 · S 2 . An alternative approach is to apply angularmomentum algebra to S = S 1 + S 2 , using S 2 = S 1 2 + 2 S 1 · S 2 + S 2 2 and
S 1 2 = S 1 2 = 3/4, and exploiting that S 2 = S(S + 1) is equal to 2 (S = 1,
↑↑), and S 2 = 0 (S = 0, ↑↓). Considering atomic spins of arbitrary size S ≥ 1/2,
performing a lattice summation over all spin pairs (compare Fig. 1), and including
an external magnetic field, the Heisenberg Hamiltonian becomes
H = −2
i>j
Jij S i · S j − g μo μB
i
S1 · H i
(27)
where the Jij are often treated as parameters. Solutions of the Heisenberg model
will be discussed in Sections “Spin Waves and Anisotropic Exchange”, “Antiferromagnetic Spin Chains”, and “Dimensionality Dependence of Quantum Antiferromagnetism”.
Some definitions of J involve a factor of 2, depending on whether the summation
is over all atoms (subscript ij) or only over pairs of atoms (subscript i > j). Even
opposite signs are sometimes chosen, using J > 0 and J < 0 for AFM and FM
interactions, respectively. The most common definition of J , used in Eq. (27), is
actually an exchange per electron, not per atom. The AFM-FM energy difference per
pair of atoms, E(Sz = 0) – E(Sz = 2S), is equal to 4 S (S + 1/2) J and diverges in
the classical limit (S = ∞). The divergence is removed by introducing renormalized
atomic exchange constants Jat = 2S 2 J or Jat = 2S (S + 1). In the classical limit
(S = ∞), Jat = Jat and H = −Jat s 1 · s 2 , where the unit vector s = S/S = M/Ms
describes the local magnetization direction. The classical energy splitting between
the ↑↑ and ↑↓ states, namely, ±Jat , is formally the same as that for S = 1/2, ±J .
2
Biquadratic exchange, H = –B S · S , as well as other higher-order
terms, may arise for several reasons, for example, in T /U expansions of the full
Hamiltonian [3]. In the case of spin 1/2 interactions, they do not yield new physics,
2 Magnetic Exchange Interactions
2
because S · S
= 3/16 −
nonzero for S ≥ 1.
65
1
2
S · S , but biquadratic exchange effects are
Hubbard Model
Completely ignoring the small direct exchange JD in equations such as (11, 12, 13,
14, 15, 16, 17, 18) leads to the Hubbard model. Generalized to solids, the Hubbard
Hamiltonian is
+
+
H = i,j Tij ĉi↑
ĉj↑ + ĉi↓
ĉ↓ + U i n̂i↑ n̂i↓
(28)
where n̂ = ĉ+ ĉ [18, 19]. In the Hubbard model, correlation effects are described by
the Coulomb interaction U . Equation (21) indicates that the Hubbard model does not
predict ferromagnetism in half-filled bands, but this argument cannot be generalized
to arbitrary bands and band fillings.
The bare Coulomb interaction is very high, about 20 eV for the iron-series
elements, but this value is reduced to about 4 eV due to intra-atomic correlations and
screening by conduction electrons. The screening (Sections “Itinerant Exchange”
and “Metallic Correlations and Kondo Effect”) depends on the crystal structure,
and eg orbitals tend to have slightly higher U values than t2g orbitals, so that U
varies somewhat for a given element. Table I shows typical U values for the three
transition-metal series [20]. Note that the effects of U are complemented by the
moderately strong intra-atomic exchange Jat , also listed in Table I. Approximate
values for U in some main-group elements are 8.0 eV (C), 3.1 eV (Ga), and 4.2 eV
(As). In rare earths, U is equal to and best obtained from the spectroscopic SlaterCondon parameter F0 . It is of the order 10 eV and somewhat increases with number
of 4f electrons.
The Hubbard U yields a number of correlation effects. One of them is the
suppression of metallic conductivity for large values of U (Mott localization),
which reflects the splitting of metallic bands into upper and lower Hubbard bands
with opposite spin directions. The effect is very similar to the Coulson-Fischer
Table 1 Typical values of
screened Coulomb integrals
(U ) and screened intra-atomic
exchange (Jat )(Jat )
nv
3
4
5
6
7
8
9
10
11
Sc
V
Ti
Cr
Mn
Fe
Co
Ni
Cu
U
Jat
2.4
3.1
3.3
4.5
4.5
3.9
4.4
4.0
5.7
0.4
0.5
0.6
0.7
0.7
0.7
0.8
0.8
0.8
Y
Zr
Nb
Mo
Tc
Ru
Rh
Pd
Ag
U
Jat
1.7
2.4
2.7
3.7
3.9
4.2
4.0
3.8
4.8
0.3
0.4
0.5
0.5
0.6
0.6
0.6
0.6
0.6
Lu
Hf
Ta
W
Re
Os
Ir
Pt
Au
U
Jat
1.5
2.0
2.4
3.5
3.7
4.1
3.8
3.6
4.0
0.3
0.3
0.4
0.5
0.5
0.5
0.5
0.5
0.6
66
R. Skomski
Fig. 4 Hubbard interpretation of band gaps: (a) Mott-Hubbard insulator, (b) charge-transfer
insulator, (c) simple interpretation of Hubbard-Mott transition, and (d) refined Hubbard transition
involving a correlated metal phase known as the Brinkman-Rice (BR) phase
electron localization in the H2 molecule (Section “Correlations”). Some oxides are
antiferromagnetic Mott-Hubbard insulators, Fig. 4a, but many are charge-transfer
insulators [21], where the 2p-3d gap Δ is smaller than the Hubbard gap U (Fig. 4b)
and the transition to metallic behavior involves hopping between cation 3d and anion
2p states. The trend toward charge transfer behavior increases from early to late
transition metals and from oxides to halides.
In spite of the simplicity of Eq. (28), there have been no exact solutions for
the Hubbard model so far, except for a few special cases. Even the well-known
Hubbard band splitting (Fig. 4c) is a simplification. A detailed analysis, using
Gutzwiller wave functions [19] and dynamical mean-field theory (DMFT) [14],
yields a correlated-metal or Brinkman-Rice phase [22] with metallic quasiparticles
in the middle of the Hubbard gap (Fig. 4d). This quasiparticle peak is analogous to
impurity peaks near band edges, for example, in the gaps of semiconductors. The
difference is that the disorder responsible for the peak is not caused by impurity
atoms but by correlated electrons (and holes) randomly occupying lattice sites.
Specific Exchange Mechanisms
The involvement of Coulomb integral (U ), and exchange integral (JD ), and oneelectron level splitting (T ) is a common feature of exchange interactions, but the
interplay between these quantities varies greatly among magnetic solids.
2 Magnetic Exchange Interactions
67
Intra-Atomic Exchange
Atomic wave functions inside a given atom are orthogonal, so that the ferromagnetic exchange is not weakened by one-electron level splittings involving
hopping between different orbitals (T = 0). On the other hand, one-electron energy
differences between shells and subshells are typically large, several eV. In terms of
Fig. 3, these energy differences provide a forbiddingly one-electron level splitting.
Ferromagnetic intra-atomic exchange is therefore almost exclusively limited to
the nearly degenerate electrons in the partially filled inner subshells of transitionmetal atoms, namely, 3d, 4d, and 5d electrons in the iron, palladium, and platinum
series, respectively, 4f electrons in rare-earth (lanthanide) atoms, and 5f electrons
in actinides.
Hund’s Rules Intra-atomic exchange and spin-orbit coupling give rise to the
hierarchy of three Hund’s rules [23]. The rules, which are empirical but have a
sound physical basis, determine the magnetic ground state of atoms or ions. Hund’s
first rule reflects intra-atomic exchange and states that the total spin S is maximized
so long as the Pauli principle is not violated. The number of one-electron orbitals
per subshell is 2 l + 1, which yields 5 orbitals per d-shell and 7 orbitals per f-shell.
In the first half of each series, all spins are ↑, and for half-filled shells, the total spin
moment is therefore 5 μB (d-shells) and 7 μB (f -shells). Additional electrons are ↓
due to the Pauli principle. For example, Co2+ has a 3d7 electron configuration and
the spin structure 3d (↑↑↑↑↑↓↓).
Quantum states characterized by quantum numbers L and S form terms denoted
by 2S + 1 L. For example, the term symbol 2 F means L = 3 and S = ½. The next
consideration is Hund’s second rule, which states that the orbital angular moment
L is maximized, subject to the value of S. The vector model usually employed
in magnetism assumes L-S (Russell-Saunders) coupling, where the total orbital
moment L = i Li and the total spin moment S = i S i combine to yield the total
moment J = L + S. The operators obey angular-momentum quantum mechanics,
for example, S 2 = S (S + 1), L2 = L (L + 1), and J2 = J (J + 1). The opposite
limit of j-j coupling, where the spin-orbit interaction dismantles the total ionic spin
and orbital moments, is important only for the ground state of very heavy elements
(Z > 75) and for excited states of most elements [24], which are usually of no
concern in magnetism.
Spin-orbit coupling causes the terms to split into multiplets, which are denoted
by 2S + 1 LJ, and obey |L – S| ≤ J ≤ |L + S|. Hund’s third rule describes how spin (S)
and orbital moment (L) couple to yield the total angular momentum (J): Less than
half-filled subshells have J = |S – L|, whereas more than half-filled shells exhibit
J = |S + L|. This rule explains, for example, the large atomic magnetic moments of
the heavy rare earths, such as 10 μB per atom in Dy3+ and Ho3+ .
68
R. Skomski
Consider, for simplicity, the Hund’s-rules ground state of the p2 configuration,
realized, for example, in free carbon atoms. There are six one-electron states (px↑ ,
py↑ , pz↑ , px↓ , py↓ , pz↓ ), but the Pauli principle reduces the 6 × 6 = 36 two-electron
microstates to 15 Slater determinants. For example, |↑ ◦ ↓> ∼ |x↑ (r1 )y↓ (r2 ) –
y↓ (r1 )x↑ (r2 )> means that the px (Lz = +1) and py (Lz = −1) orbitals are both
occupied by ↑ electrons, while the pz orbital (Lz = 0) is empty. The 15 microstates
of the p2 configuration form three terms: 1 S (L = 0, S = 0), 1 D (L = 2, S = 0), and
3 P (L = 1, S = 1). Hund’s first rule uniquely establishes the ground-state term 3 P,
because the other two terms have zero spin. The term contains (2 L + 1) (2S + 1) = 9
Slater determinants, for example, |↑ ↑ ◦>, where Lz = 1 and Sz = 1. Hund’s third
rule predicts J = L – S = 0, corresponding to a nonmagnetic ground state. The p2
ground-state wave function is a superposition of three Slater determinants described
by Clebsch-Gordan coefficients C(L, Lz , S, Sz |J, Jz ) [2, 25]. Explicitly
1
1
1
|ψ> = √ |↑ ◦ ↓> + √ |↓ ◦ ↑> − √ |◦ ↑↓ ◦>
3
3
3
(29)
The involvement of two or more Slater determinants indicates that correlations are
not necessarily be important even in seemingly simple systems.
Hund’s rules are obeyed fairly accurately by rare-earth ions in metallic and
nonmetallic environments. For example, the ground-state multiplets of rare-earth
ions obey J = |L ± S|, whereas excited multiplets have relatively high energies, with
notable exceptions of Eu3+ and Sm3+ , where the splitting is only about 0.1 eV [26].
One reason for the applicability of the rules is that the 4f -shell radii of about 0.5 Å
are much smaller than the atomic radii of about 1.8 Å. This enhances the spin-orbit
coupling and reduces the interaction with surrounding atoms. By contrast, Hund’s
rules are often violated in 3d, 4d, and 5d transition metals, where orbital moments
are quenched.
Moment Projections and Quenching Exchange interactions are between spins S,
not between total moments J=L+S, which makes it necessary to project the total
moment onto the spin moment. Similarly, the Zeeman interaction with an external
field involves L + 2 S, not J = L + S and S. In the Zeeman case, projection of
L + 2 S onto J yields the symbolic replacement L + 2 S → g and the moment
m = g J. The g-factor is obtained by using (L + 2 S) · J = g J2 and
J (J + 1) = L (L + 1) + 2L · S + S (S + 1)
(30)
The result of the calculation is
g=
3 1 S (S + 1) − L (L + 1)
+
2 2
J (J + 1)
(31)
which yields g = 1 – S/(J + 1) and g = 1 + S/J for the first and second
halves of the lanthanide series, respectively. To account for spin-only character of
2 Magnetic Exchange Interactions
69
interatomic exchange, the atomic projection S · J = (g − 1) J2 must be used.
The corresponding de Gennes factor G = (g – 1)2 J(J + 1) is important for
the finite-temperature behavior of rare-earth magnets, where it controls the Curie
temperature. One implication of Eq. (31) is that the vectors L, S, and J are not
necessarily (anti)parallel but described by the vector model of angular momenta
[24]. A good example is Sm3+ , which has antiparallel spin and orbital moments
L = 5 and S = 5/2, respectively, so that L – 2S could naïvely be expected to yield
a zero magnetic moment. In fact, g = 2/7 and m = 0.71 μB , which corresponds to
angles of 22◦ between L and J and of 44◦ between S and J.
Hund’s rules are often violated in metallic and nonmetallic transition-metal
magnets. The d orbitals of iron-, palladium-, and platinum-series atoms are fairly
extended, so that interactions with neighboring atoms outweigh Hund’s rules
considerations. The rules regarding L are affected most, because the orbital moment
is normally quenched, L ≈ 0. For example, bcc iron has a magnetization of about
2.2 μB , but only about 5% is of orbital origin. The reason is that orbital moments
require an orbital motion of the electrons, but this motion is disrupted by the crystal
field introduced in Section “One-Electron Wave Functions”. Note that L = 0 means
J = S and, according to Eq. (31), g = 2.
High-spin Low-spin Transitions Crystal-field interactions cause the five 3d levels of
transition-metal ions to split. In magnets with cubic crystal structure, this splitting
is of the eg -t2g type: The |z2 > and |x2 – y2 > orbitals form the eg dublet, whereas
the |xy>, |xz>, and |yz> orbitals form the t2g triplet. The crystal-field interaction
yields a moment reduction if the splitting is larger than the combined effect of U
and JD . Such transitions are known as high-spin low-spin transitions. For example,
in octahedral environments, the energy of the t2g triplet is lower than that of the
eg dublet. Co2+ has seven 3d electrons, which translate into the spin configuration
t2g (↑↑↑↓↓) eg (↑↑) and a moment of 3 μB . However, in the limit of large crystalfield splitting, one of the two eg↑ electrons “falls down” in the sense of Fig. 3 and
occupies the empty t2g↓ orbital, yielding the spin configuration t2g (↑↑↑↓↓↓) eg (↑)
and a moment of 1 μB . Examples are the Co2+ complexes [Co(H2 O)6 ]2+ (high
spin) and [Co(CN)6 ]4− (low spin).
Indirect Exchange
The model of Section “Electron-Electron Interactions” describes the so-called direct
exchange between nearest neighbors, where the hopping integral T competes
against U and JD . Exchange in solids is often indirect, mediated by conduction
electrons or by intermediate atoms, such as oxygen.
Superexchange Transition-metal oxides frequently exhibit exchange bonds of type
Mm+ -O2− -Mm+ , where Mm+ is a transition-metal cation. This type of exchange
is known as superexchange and also realized in magnetic halides such as MnF2 .
70
R. Skomski
The net exchange is tedious to calculate [27], but a transparent physical picture
emerges if one assumes that U and JD compete against T and that the outcome of
this competition is largely determined by T , similar to Eq. (24). For one 3d level
per transition-metal atom (M) and one oxygen 2p level (O), Eq. (8) becomes
⎛
⎞
EM Tpd(L) 0
H = ⎝ Tpd(L) EO Tpd(R) ⎠
0 Tpd(R) EM
(32)
Here, EM and EO are the atomic on-site energies, and Tpd(R/L) describes the hopping
between M and O atoms. When Tpd(R) = Tpd(L) , a unitary transformation using
⎛
√1
2
0
√1
2
⎞
⎟
⎜
=⎝ 0 1 0 ⎠
√1 0 √1
2
(33)
2
partially diagonalizes the Hamiltonian and yields
⎛
√EM
Q+ HQ = ⎝ 2Tpd
0
√
⎞
2Tpd 0
EO
0 ⎠
0 EM
(34)
The transformation couples the wave functions of the two transition-metal atoms.
One of the coupled M levels is nonbonding (bottom-right matrix element), whereas
the other one (top left) hybridizes with the oxygen, thereby creating a level splitting
between the two coupled M orbitals. In the Heisenberg limit, Tpd is small, and the
diagonalization of Eq. (34) yields the transition-metal level splitting ±Teff , where
Teff = Tpd 2 /|EM –EO |.
Substitution of Teff into Eq. (24) yields the effective transition-metal exchange
Jeff = JD −
2 Tpd 4
U (EM − EO )2
(35)
Since JD is small, the hopping normally wins, and the exchange in most oxides is
therefore antiferromagnetic. However, the non-s character of the 2p and 3d orbitals
causes Tpd to depend on the type of d orbital (eg or t2g ) and on the bond angle.
Figure 5 compares a 180◦ bond (a) with a 90◦ bond (b). In (a), the two p-d bonds
differ by the sign of the involved 2p wave-function lobe, but Tpd(R) = −Tpd(L)
leaves Eq. (35) unchanged. In (b), Tpd(R) = 0 by symmetry, because the hopping
contributions of the two oxygen lobes (+ and –) cancel each other. This implies
Tpd(R) = 0 in Eq. (32), and the two transition-metal orbitals are no longer coupled
(Teff = 0).
2 Magnetic Exchange Interactions
71
Fig. 5 Overlap and
exchange: (a) nonzero
overlap (180◦ bond) and zero
overlap (90◦ bond). In (a), the
hopping integral is nonzero,
corresponding to
antiferromagnetic indirect
exchange, but in (b), the
hopping integral is zero by
symmetry
The above analysis is the basis for the Goodenough-Kanamori-Anderson rule
[27, 28], which
states that exchange in oxides is antiferromagnetic for bond angles
θ B > 90◦ Teff 2 > 0 but ferromagnetic for bond angles of θ B = 90◦ (Teff = 0).
Examples of the former are rock salt, spinel, and wurtzite oxides, where the
predominant bond angles are 180◦ , 125◦ , and 109◦ , respectively [27]. Ferromagnetic
exchange dominates in CrO2 [27], where the Cr4+ ions yield a net moment of 2 μB
per formula unit.
Ruderman-kittel Exchange The exchange interaction of localized magnetic
moments in metals is mediated by conduction electrons, which is known as the
Ruderman-Kittel-Kasuya-Yosida or RKKY mechanism. Electrons localized at Ri
and conduction electrons of wave vector k undergo a strong intra-atomic s-d
exchange –Jsd S k · S i δ (r–R i ), so that the localized electrons perturb
the sea
of conduction electrons. The perturbed wave functions are ψ k (r) = k ck exp (i
k · r) dk, where the integration is limited to wave vectors |k| < kF . The wavevector cutoff affects the real-space resolution of the response ψ(r) and means
that details smaller than about 1/kF , such as
δ(r – Ri ), cannot be resolved. As
a consequence, the electron density n(r)∼
ψ k (r) ψ k (r) dk contains a wavelike oscillatory contribution. The oscillations are spin-dependent and yield the
oscillatory RKKY exchange
72
R. Skomski
J (R) = Jo
2kF R cos (2kF R) − sin (2kF R)
(2kF R)4
(36)
between localized moments at Ri and Rj = Ri + R. In metals, kF ∼ n1/3 is large, and
the oscillation period does not exceed a few Å. In dilute magnetic semiconductors
(DMS), n can be made small by adjusting doping level and/or temperature, and the
RKKY interaction is then a nanoscale effect.
Equation (36) describes exchange interactions mediated by free electrons, but
the underlying perturbation theory can also be used to treat arbitrary independentelectron systems, such as tight-binding electrons in metals [29] and DMS exchange
mediated by shallow nonmagnetic donors (or acceptors) [30]. At finite temperatures,
the thermal smearing of the Fermi surface yields an exponential decay of the
oscillations, with a decay length proportional to kF /T.
Double Exchange Intra-atomic exchange favors parallel spin alignment, and electrons retain a “spin memory” while hopping between atoms. This process translates
into a ferromagnetic exchange contribution first recognized by Zener [28, 31].
Double exchange occurs in mixed valence oxides, such as Fe3 O4 . This oxide
contains Fe3+ and Fe2+ ions on B-sites. The latter can be considered as Fe3+ ions
plus an extra electron that can hop more or less freely between the d5 ion cores.
The double-exchange mechanism is important in magnetoresistive perovskites
(manganites). The parent compound, LaMnO3 , contains Mn3+ ions only and is
an antiferromagnetic insulator. Partially replacing La3+ by Sr2+ creates a charge
imbalance that is compensated by the formation of Mn4+ ions. In both Mn3+ and
Mn4+ , the low-lying t2g triplets are occupied by three well-localized 3d electrons,
but in Mn3+ , there is an additional eg electron that yields ferromagnetic double
exchange and metallic conductivity.
Itinerant Exchange
The magnetism of 3d, 4d, and 5d elements and alloys is fairly well described by the
independent-electron approximation, which corresponds to the use of a single big
Slater determinant. The electrons move in the solid, and the corresponding hopping
competes against the electrostatic electron-electron interaction.
The simplest approach is to replace the crystal potential V(r) by a chargeneutralizing homogeneous background V(r) = const. (jellium model). The only
free parameter describing the corresponding homogeneous but not necessarily free
electron gas is the electron density n. It is convenient to parameterize n in terms
of the average inter-electronic distance rs = (3/4πn)1/3 and to relate rs to the freeelectron Fermi wave vector kF = (9π/4)1/3 /rs . Typical values of kF ao are 0.34 (Cs),
0.72 (Cu), and 1.03 (Be) [8]. The inverse magnetic susceptibility of the jellium
is [32]
2 Magnetic Exchange Interactions
73
χp
π
1
=1−
+ 2
(0.507 ln (kF ao ) − 0.162)
χ
kF ao
kF ao 2
(37)
where χ p = (α/2π)2 ao kF is the susceptibility of the non-interacting electron gas
(Pauli susceptibility). The onset of ferromagnetism corresponds to χ = ∞, that is,
to 1/χ = 0.
Equation (37) includes the key distinction between kinetic energy (hopping),
scaling as 1/rs 2 ∼ kF 2 , and Coulomb interaction, scaling as 1/rs ∼ kF . The Pauli
susceptibility reflects the kinetic energy, whereas –π/kF ao is the independentelectron Coulomb correction, which corresponds to Bloch’s early theory of itinerant
exchange [8, 33]. As the electron gas gets less dense and kF becomes smaller, the
π/kF ao term in Eq. (37) predicts ferromagnetism for kF ao < 1/π, which is close
to the value for alkali metals such as Cs. Experimentally, the alkali metals are not
particularly close to ferromagnetism, which is caused by d and f electrons, not by a
homogeneous electron gas.
In fact, the last term in Eq. (37), which scales as 1/kF 2 and reflects the
so-called random-phase approximation (RPA), negates the Bloch prediction of
ferromagnetism – χ (kF ao ) never reaches zero in Eq. (37). The physics behind the
RPA is that the charge of any individual electron is screened by the other electrons
in the metal, which amounts to a reduction of the net Coulomb repulsion from 1/rs
to an exponentially decaying interaction. In other words, the screening electrons
form a quasi-particle cloud around the electron and renormalize the Coulomb
interaction.
The Stoner theory replaces Eq. (37) by the semiphenomenological expression
χp
= 1 − I D(EF )
χ
(38)
where the Stoner parameter I ∼ 1 eV [34] describes the electron-electron interaction
(Section “Stoner Limit”). Equation (38) predicts ferromagnetism for high densities
of states (DOS), when the paramagnetic state becomes unstable and the magnets
satisfy the Stoner criterion (EF ) > 1/I. The DOS of d electrons is much higher than
that of the jellium electrons implied in Eq. (37), which explains the occurrence
of ferromagnetism in transition metals. Alternatively, since the DOS (density of
states) is inversely proportional to the bandwidth W ∼ |T |, ferromagnetism
occurs in narrow bands. This finding is in agreement with the general analysis of
Section “Antiferromagnetic Spin Chains”.
Band Structure and Magnetism The hopping aspect of magnetism is determined by
the band structure and by the metallic density of states (DOS). Both are obtained
from the eigenvalues and eigenfunctions of Hamiltonians of the type
H=−
2 2
∇ +
2me
j
Vo (r − r j )
(39)
74
R. Skomski
where the lattice-periodic potential depends, in general, on the electron distribution.
The eigenfunctions of Eq. (39) are Bloch states ψ(r) = exp (ik · r) u(r) and electron
densities n(r) = u*(r)u(r). Equation (39) describes delocalized electrons whose
electrical conductivity is infinite due to the absence of scattering matrix elements.
This includes the tight-binding limit of well-separated atoms, where the hopping
integrals decrease exponentially with interatomic distance, T ∼ exp (–R/Ro ), but
the conductivity remains infinite even for large R [8]. At zero temperature, the
magnets are well described by these Bloch-periodic wave functions. This includes
the explanation of non-integer moments, which are caused by the smearing of oneelectron wave functions and spin densities over many lattice sites.
Inhomogeneous Magnetization States Wave-function and magnetization inhomogeneities may have several reasons. Wave-function localization requires the breaking of structural periodicity due to disorder (Anderson localization) or finite
temperature. Near Tc , atomic-scale itinerant moments behave like Heisenberg spin
vectors (“spin fluctuations”) of random orientation but well-conserved magnitude,
the latter involving some short-range order. Experimentally, this localization manifests itself as a characteristic specific-heat contribution [9]. This spin-fluctuation
picture is realized both in strong ferromagnets (e.g., Co), where the ↑ band is filled,
and in weak ferromagnets such as Fe, where the ↑ band is only partially filled.
Deviations from wave-function periodicity also occur due to electron correlations
(Mott localization, Section “Hubbard Model”), competing exchange in perfectly
periodic lattices (Section “Magnetic Order and Noncollinearity”), and surface
effects.
Very weak itinerant ferromagnets (VWIFs), such as ZrZn2 (Tc = 17 K), barely
satisfy the Stoner criterion, and their behavior is qualitatively different from that
of strong and weak ferromagnets [35, 36]. Thermal excitations act as local spin
perturbations that can be described by the wave-vector-dependent susceptibility
χ (k) [3]. For VWIFs, a good approximation is
χ=
χo
|I D (EF ) − 1 + f (k)|
(40)
and f (k) = a2 k2 . Here χ o is the interaction-free susceptibility, approximately equal
to the Pauli susceptibility χ p of Eq. (37), and a is an effective interatomic distance.
Inverse Fourier transform of Eq. (40) yields |M(r)| ∼ exp.(−r/ξ ), where r is the
distance from the perturbation and ξ = a/|1–I D (EF )|1/2 . In VWIFs, I D ≈ 1,
so that ξ is large by atomic standards and blurs the distinction between intra- and
interatomic exchange. The Stoner transition, I D = 1, yields ξ = ∞ and corresponds
to Bloch-periodic wave functions. A rough Curie temperature approximate is [37].
Tc 2
TS
2
+
Tc
=1
TJ
(41)
2 Magnetic Exchange Interactions
75
This equation interpolates between the Heisenberg limit TJ (spin rotations) and the
Stoner limit Ts (moment reduction).
Strongly exchange-enhanced Pauli paramagnets, such as Pt, are close to the
onset of ferromagnetism and have I – 1/(EF ) 0. Magnetic impurities create spinpolarized clouds of radius ξ in these materials. The corresponding radial dependence
M(r) of the magnetization combines a pre-asymptotic exponential decay (r ξ )
with RKKY oscillations for large distances (r ξ ). The exponential decay length
ξ is described by Eq. (40), in close analogy to VWIFs. For example, magnetic
surfaces of Co2 Si nanoparticles spin-polarize the interior of the particles with a
penetration depth ξ [38]. Spin polarized clouds in strongly exchange-enhanced
Pauli paramagnets are also known as a paramagnons [3]. Left to themselves,
these quasiparticles slowly decay, and by considering the time dependence of the
fluctuations, f (k) → f (k, ω) in Eq. (4), one can show that the relaxation time diverges
at the phase transition (critical slowing down).
Bethe-Slater Curve
It is of practical importance to have some guidance concerning the strength and
sign of the exchange in a given metallic magnet. An early attempt was the
semiphenomenological Bethe-Slater-Néelcurve [39], which plots the net exchange
or the ordering temperature as a function of the interatomic distance or number
of electrons. There are many versions of this curve, and Fig. 6 shows one of
them. The curve predicts antiferromagnetism for small interatomic distances,
ferromagnetism for intermediate distances, and the absence of magnetic order in
the limit of very large distances. Experiment, the results of Section “Stoner Limit”,
and detailed calculations [40] grant some credibility to the approach, but the curve
has nevertheless severe flaws [27, 41].
Equation (38) shows that the onset of ferromagnetism is predominantly determined the density of states (EF ) at the Fermi level. This density somewhat increases
Fig. 6 Early version of the
Bethe-Slater-Néel curve
[27, 39]
76
R. Skomski
with interatomic distance, but a more important consideration is the position of
the Fermi level relative to the big peaks in the DOS. These peaks tend to vary
substantially among materials with similar chemical composition but different
crystal structures. For example, many transition-metal-rich intermetallic alloys have
interatomic distances of about 2.5 Å but show big differences in spin structures and
magnetic ordering temperatures.
A specific example is the distinction between fcc and bcc Fe structures. First, the
interatomic distance R = 2Rat in fcc iron, 2.53 Å, is actually a little bit larger than
that in bcc Fe, 1.48 Å, so that Fig. 6 cannot explain the ferromagnetism of bcc Fe.
Second, the plot ignores that bcc and fcc iron have very different crystal structures.
One difference is the number of nearest neighbors, namely, 8 in the bcc structure and
12 in the fcc structure. The bandwidth increases with the number z of neighbors, so
the ferromagnetism tends to be more difficult to create in dense-packed structures
(z = 12 . . . 14) and easier to create at surfaces (z = 4 . . . 6).
However, the number of neighbors is not the main consideration, because fcc Ni
and fcc Co have 12 nearest neighbors but are both ferromagnetic. More important is
the location of the big peaks in the density of states. For nearly half-filled d-shells
(Cr, Mn), one wants to have the peaks somewhere in the middle of the band, whereas
for nearly filled d-shells (Co, Ni), main peaks near the upper band edge are preferred.
An accurate determination of the peak positions
can only be done numerically, but
the moments theorem [42], dealing with µm = Em (E) dE, provides some guidance
[41]. The respective zeroth, first, and second moments describe the total number of
states, the band’s center of gravity, and the bandwidth, all unimportant in the present
context.
The third moment, μ3 , parameterizes the asymmetry of the DOS, that is, whether
the main peaks of the DOS are in the middle of the band (μ3 = 0) or close to
the upper band edge (μ3 < 0). It can be shown [42] that μ3 reflects the absence
or presence of equilateral nearest-neighbor triangles in the structure, the former
yielding centered main peaks and the latter creating main peaks near the upper
band edge. Figure 7 provides a very simple example of this relationship. Equilateral
nearest-neighbor triangles are present in the fcc structure but not in the bcc structure,
which corresponds to bcc ferromagnetism in the middle of the series and fcc
(or hcp) ferromagnetism for Co and Ni. Fe is intermediate, but bcc Fe becomes
ferromagnetic more easily than fcc Fe.
Manganese Isolated manganese atoms have half-filled 3d shells and a magnetic
moment of 5 μB per atom, which corresponds to a magnetization of approximately
5 T in dense-packed Mn structures. If this magnetization could be realized in a
ferromagnetic material, it would revolutionize technology, particularly since Mn is a
relatively inexpensive element. However, most Mn-based permanent magnets, such
as MnAl, MnBi, and Mn2 Ga, exhibit rather modest magnetizations of the order of
0.5 T [43]. The main reason for the low magnetization of Mn magnets is the halffilled character of the Mn bands.
2 Magnetic Exchange Interactions
77
Fig. 7 Crystallographic
motifs and density of states:
(a) square and (b) equilateral
triangle. The density of states
is largest in the middle (a)
and near the top of the level
distribution (b). The atomic
orbitals (red) are of the
s-type, but in [42], it can be
seen that 3d electrons behave
similarly, and the present
figure can be generalized to
three-dimensional lattices
Fig. 8 Exchange interactions
in hypothetical simple-cubic
Mn [45]
Magnetizations as high as μo Ms = 3.2 T (3.25 μB per atom) have been reported
in thin-film Fe9 Co62 Mn29 deposited on MgO [44], where DFT calculations predict
2.90 μB per atom [45]. A traditional interpretation in terms of Fig. 6 is that
dilution by Fe and Co atoms enhances the average distance between Mn atoms. The
tetragonal structure of the Fe-Co-Mn alloy is loosely related to that of L21 -ordered
Mn2 YZ Heusler alloys, where the Mn atoms occupy a simple-cubic sublattice and
exhibit ferromagnetic exchange [46]. DFT calculations (Fig. 8) actually indicate
that the Mn-Mn exchange never becomes ferromagnetic. Furthermore, the example
of L10 -ordered MnAl shows that large Mn-Mn distances are not necessary for
78
R. Skomski
ferromagnetic exchange: The dense-packed Mn sheets in the (001) planes of MnAl,
which form a square lattice, exhibit a strong FM intra-layer exchange J [47]. This
underlines the crucial role of atomic neighborhoods.
Metallic Correlations and Kondo Effect
The situation in 3d metals is intermediate between the uncorrelated itinerant limit
(U /W = 0) and the strongly correlated Heisenberg limit, with U /W ratios of the
order of 0.5 [9]. For example, electron-electron interactions cause a bare electron
to become surrounded or “dressed” by other electrons, forming a quasiparticle of
finite lifetime, because electrons constantly enter and leave the dressing cloud. The
corresponding relaxation time τ ≈ / Im (Σ), where Σ is the self-energy, decreases
with increasing interaction strength. For metallic electrons of energy Ek , the lifetime
is approximately EF 3 /V2 (Ek – EF )2 [48], meaning that weak interactions and
vicinity to the Fermi surface yield well-defined and slowly decaying quasiparticles
which constitute a Fermi liquid.
As pointed out in Section “Correlations”, the independent-electron approximation involves a single Stater determinant and does not account for correlation
effects. The treatment of correlations requires several Slater determinants, such
as the two determinants of the model of Section “Electron-Electron Interactions” and the three determinants forming the ground state of the p2 configuration (Section “Intra-Atomic Exchange”). An example of correlated manyelectron states is the Gutzwiller wave function |> = exp –η i n̂i↑ n̂i↓ |o >,
where the parameter η depends on U /W and the exponential term has the
effect of creating new Slater determinants from |Ψ o > [9, 19]. The Gutzwiller
method can be interpreted as a many-electron extension of the Coulson-Fischer
approach.
It is sometimes claimed or implied that density-functional theory becomes exact
if one goes beyond the local-density approximation and that LSDA+U approaches
account for correlations. This argumentation is questionable for several reasons.
First, density-functional theory provides the correct ground-state energy [12, 49]
if the density functional is known, but the exchange interaction is an energy
difference between the ferromagnetic and other spin configurations (AFM, PM)
and therefore involves excited states. Second, the density functional is not known
very well. The local-density approximation uses a potential inspired by and well
adapted to nearly homogeneous dense electron gases. The eigenfunctions used in
LSDA, known as Kohn-Sham (KS) orbitals, are pseudo-wave functions without a
well-defined quantum-mechanical meaning. They serve to determine the density
functional [49] and lack, for example, Gutzwiller-type projection features. The local
character of the LSDA, which can be improved by gradient corrections [50], is not
essential in this regard: Hartree-Fock theory involves a single Slater determinant
but is highly nonlocal [8]. Other density functionals, such as the Runge-Zwicknagel
functional for highly correlated electrons in dimers [51] and the density functional
2 Magnetic Exchange Interactions
79
for Bethe-type crystal-field interactions of rare-earth 4f electrons [52], bear little or
no resemblance to the LSDA.
The underlying problem is that the density functional is a generating functional
very similar to partition function Z and free energy F= – kB T ln Z in equilibrium
thermodynamics [52, 53]. The generating functionals correspond to Legendre
transformations, and in thermodynamics, the transformations are realized through
the term – T S, where S is the entropy. Once Z is determined by the summation
or integration over all microstates, such as the atomic positions ri in a liquid,
temperature-dependent physical properties are obtained in a straightforward way
from F(T). The theory is exact in principle, but the predictions depend on the
accuracy of the partition function. One example is that low- and high-temperature
expansions have different domains of applicability. Another example is the statistical mean-field approximation, including Oguchi-type nonlocal corrections [54],
which are unable to describe critical fluctuations. In density-functional
theory, the
Legendre transformation is realized through the integral –
V(r) n(r) dr [53].
The density functional is obtained by eliminating the microstate information in
Ψ ( . . . , ri , . . . , rj , . . . ) and yields the ground state for each lattice potential V(r).
This lattice potential is the DFT equivalent of the temperature in thermodynamics,
and the accuracy of the predictions depends on the quality of the generating
functional.
The density functionals used in LSDA are not calculated but obtained through
intelligent and experimentally supported guesswork. An exception is the weakly
correlated limit (U ≈ 0), where the KS orbitals become quantum-mechanical
wave functions with well-defined physical meaning. There are two reasons for
the great success of the LSDA, and its extensions have two main reasons. First,
transition metals are only weakly correlated and therefore amenable to ad hoc
improvements using “second-principle” approaches (materials-specific choices of
methods and parameters). Second, the KS Slater determinants used in LSDA are of
the unrestricted Hartree-Fock type (Section “Correlations”), which are constructed
from wave functions having symmetries lower than that of the Hamiltonian
[9, 10]. Unrestricted HF determinants can be expanded in terms of “regular” Slater
determinants and therefore contain some correlations [9].
The “Plus U” Method The LSDA+U method modifies the KS one-electron
potential by a potential that depends on the electron’s atomic orbital i, essentially
[55].
1
Vi (r) = VLSDA (r) + U
− ni
2
(42)
A crude approximation is U ∼ U . The presence of U suppresses ↑↓ occupancies
in highly correlated 3d and 4f orbitals. The LSDA+U can be used, for example,
to adjust the charge state of magnetic ions (configurations) to their experimental
values. Such adjustments are sometimes necessary, because there is only one
80
R. Skomski
Fig. 9 LSDA+U for bcc Fe: (a) magnetic moment, (b) weak ferromagnetism and (c) strong
ferromagnetism. The direct exchange and double-counting corrections are ignored in this figure
KS determinant available to account for the uncorrelated subsystem (one Slater
determinant) and for the ion’s intra-atomic couplings (several Slater determinants).
Strictly speaking, U is a well-defined first-principle quantity [55], not a fitting
parameter that can be chosen to obtain a desired computational result. Figure 9
illustrates this point for the magnetic moment of bcc Fe, calculated using the VASP
code for with U varying from 0 to 6. The moment m per Fe atom (a) exhibits an
increase from 2.21 μB to 3.07 μB , the experimental value being about 2.22 μB . Near
U = 0.9 eV (dashed vertical line), the slope dm/dU changes from about 0.4 μB /eV to
0.1 μB /eV, caused by the unphysical transition from weak to strong ferromagnetism
(b-c).
2 Magnetic Exchange Interactions
81
Fig. 10 Model describing quantum-spin-liquid corrections in solids. The quantum-mechanical
mean-field (MF) approximation self-consistently treats an independent electron in a sea of
surrounding electrons (gray) and corresponds to one Slater determinant. Atomic Heisenberg spins
having S = 1/2 yield 2z + 1 Slater determinants
Noncollinear Density-functional Theory The Heisenberg model is based on quantum rotations of atomic spins of fixed magnitude S 2 = S (S + 1). This is a rough
approximation for transition metals, where electrons are delocalized (itinerant)
and atomic moments are often non-integer. However, rotations of electron spins
(S = 1/2), which are realized through Pauli matrices and yield spin-wave functions
such as ψ(θ ) = (cos½θ , sin½θ ), can be implemented in the LSDA and used to
describe noncollinear spin states, including antiferromagnets [56]. This approach
corrects, for example, much of the great overestimation of the Curie temperature in
the Stoner theory.
The spin-wave functions ψ(θ ) are of the quantum-mechanical mean-field type,
weakly correlated, and not eigenstates of the Heisenberg Hamiltonian. Figure 10
illustrates the many-electron aspect of the approximation. The model treats one
↓ electron in a sea of ↑ electrons. The left part of the figure corresponds to the
quantum-mechanical mean-field approximation, where electrons interact with an
effective medium. In a slightly more realistic picture, the interaction with z nearest
neighbors is individualized through exchange bonds, as shown for z = 3 and
z = 5.
The model, which assumes S = 1/2 Heisenberg spins and nearest-neighbor
exchange J< 0, is exactly solvable. The ground state has one ↑ electron and
z ↓ electrons, which leads to the involvement of (z + 1) Slater determinants. The
admixture of these determinants describes whether the ↓ electron stays in its original
central place (Néel state) or “leaks” into the crystalline environment [57]. The calculation shows that the reversed spin occupies neighboring atoms (dark gray) with a
combined weight of 50%, thereby affecting net exchange and ordering temperature.
Sections “Antiferromagnetic Spin Chains” and “Dimensionality Dependence of
Quantum Antiferromagnetism” considers the lattice aspect of this spin leakage.
Exchange in the Kondo Model The Kondo effect, characterized by a resistance
minimum, is a correlation effect caused by the exchange interaction of localized
82
R. Skomski
Table 2 Kondo
temperatures (in kelvin) for
some transition-metal
impurities in nonmagnetic
hosts (gray column) [59, 65]
Rh
Pd
Pt
Cu
Ag
Au
Zn
Al
Cr
–
100
200
1.0
0.2
0.01
3
1200
Mn
10
0.01
0.1
0.01
0.04
0.01
1.0
530
Fe
50
0.02
0.3
22
3
0.3
90
5000
Co
1000
0.1
1
2000
–
200
–
–
impurity spins with conduction electrons [9, 58]. Below the Kondo temperature
TK , the interaction couples the conduction electrons to the impurity spins, which
enhances the electrical resistivity. Some Kondo temperatures for Cr, Mn, Fe and Co
in various matrices are shown in Table 2.
The simplest Kondo model is of the Anderson-impurity type, where a single
conduction electron, described by a delocalized orbital |c>, interacts with a localized
state |f > [9]. The Coulomb U is negligibly small for the delocalized orbital |c> but
large for the localized orbital |f >. Furthermore, the on-site energy of the localized
electron (bound state) is lower than that of the delocalized electron by E. In terms
of the wave functions |ff >, |fc>, |cf>, and |cc>, the Hamiltonian is
⎛
U − E
⎜
T
H=⎜
⎝
T
0
T
0
0
T
⎞
T 0
0 T ⎟
⎟
0 T ⎠
T E
(43)
Since U
T , the |ff > state (energy U − E) does not play any role in the groundstate determination. In the absence of hybridization (T = 0), the ground state would
be degenerate, |f c> ± |c f >, both states having the energy E = 0 and containing
one localized and one delocalized electron. The first excited antiferromagnetic state,
|cc> = |c↑ c↓ >, has the energy E = E, meaning that the localized electron becomes
a conduction electron.
The hopping integral T does not affect the ferromagnetic state |f c>−|c f >,
because a localized ↑ electron cannot hop into a delocalized orbital that already
contains a ↑ electron. However, the localized ↑ electron can hop into a delocalized
orbital containing an electron of opposite spin, which lowers the energy of the
antiferromagnetic state. The corresponding singlet (↑↓) ground state has an energy
of –2T 2 /E = –2JK , roughly translating into a Kondo-temperature of TK =
2T 2 /kB E. Above TK , the |↑↓> and |↑↑> states are populated with approximately
equal probability, the two electrons effectively decouple, and the resistivity drops.
In reality, there are many conduction electrons, so an integration over all k-states
is necessary [58]. The main contribution comes from electrons near the Fermi level,
2 Magnetic Exchange Interactions
83
which form a Kondo screening cloud of size ξ proportional to 1/TK and yield a
Kondo temperature TK = (W/kB ) exp (−1/ [2 JK D(EF )]). Due to its exponential
dependence on JK and (EF ), TK varies greatly among systems [59]. Table II shows
some examples. TK is lowest for impurities in the middle of the 3d series and largest
for nearly filled or nearly empty 3d shells, as exemplified by TK = 5000 K for
Ni in Cu. The dependence JK ∼ 1/S indicates that Kondo exchange become
less effective in the classical limit. Heavy-fermion compounds, such as UPt3 and
CeAl2 , can be considered as Kondo lattices where conduction electrons interact
with localized 4f or 5f electrons [9].
Exchange and Spin Structure
Exchange affects spin structure and magnetic order in many ways. It determines
the ordering temperature, gives rise to a variety of collinear and noncollinear
spin structures, and influences micromagnetism through the exchange stiffness
A. Exchange phenomena include quantum-spin-liquid behavior, high-temperature
superconductivity, and Dzyaloshinski-Moriya interactions.
Curie Temperature
In spite of its simplicity, the Heisenberg model (27) is very difficult to solve,
especially in two and three dimensions. A great simplification is obtained by using
the identity
S i · S J = S i · <S j > + <S i > · S j + Cij + co
(44)
and neglecting
the thermodynamic correlation term Cij = (S i − <S i >) ·
S j − <S j > and the constant co = <S i > · <S j >. The latter is physically
unimportant, because it does not affect the thermodynamic averaging. The former
is important only in the immediate vicinity of the Curie temperature, where it
describes critical fluctuations [60, 61]. Substituting Eq. (44) into Eq. (27) and
assuming z nearest neighbors of spin moments S yield the factorized single-spin
Hamiltonian
H = −2 z J S · <S> − 2μo μB S · H
(45)
This equation amounts to the introduction of a mean field μo μB H =2 z J <S> and
maps the complicated Curie-temperature problem onto the much simpler problem
of a spin S in a magnetic field. This approximation (45) is the thermodynamic meanfield approximation, which must be distinguished from the quantum-mechanical
mean-field approximation used to treat electron-electron interactions.
The partition function belonging to Eq. (45) is a sum over the 2 S + 1 Zeeman
levels Sz . The field dependence of <S> has the form of a Brillouin function (BS ),
84
R. Skomski
and self-consistently evaluating <S> yields the Curie temperature
Tc =
2 S(S + 1)
zJ
3 kB
(46)
A generalization of this equation to two or more sublattices will be discussed
in Section “Dimensionality Dependence of Quantum Antiferromagnetism”. This
generalization includes the Néel temperature of antiferromagnets.
The spin excitations leading to Eq. (46) consist in the switching of individual
spins S i . The corresponding energies are rather high, with temperature equivalents close to Tc . At low temperatures, the mean-field approximation predicts
exponentially small deviations from the zero-temperature magnetization M(0),
which is at odds with experiment. In fact, the low-temperature behavior M(0) –
M(T) of Heisenberg magnets is governed by low-lying excitations (spin waves)
(Section “Spin Waves and Anisotropic Exchange”) and described by Bloch’s law,
M(0) – M(T) ∼ T3/2 , in three dimensions.
Magnetic Order and Noncollinearity
Depending on the sign of the interatomic exchange, there are several types of magnetic order. Figure 11 shows some examples. Often there are two or more sublattices
[4, 54], and the division into sublattices can be of structural or magnetic origin.
Ferrimagnetism (FiM) normally reflects chemically different sublattices, such as
Fe and Dy sublattices in Dy2 Fe14 B. Antiferromagnetism (AFM) is also caused by
negative interatomic exchange constants, but the different sublattices are chemically
and crystallographically equivalent. For example, CoO crystallizes in the rock-salt
structure, but the Co forms two sublattices of equal and opposite magnetization.
Ferromagnetism is frequently encountered in metals (Fe, Co, Ni) and alloys (PtCo,
SmCo5 , Nd2 Fe14 B), the latter having different ferromagnetic sublattices. CrO2 is a
ferromagnet, but most oxides and halides are antiferromagnetic (MnO, NiO, MnF2 )
or ferrimagnetic (Fe3 O4 , BaFe12 O19 ).
Many oxides of chemical composition MFe2 O4 crystallize in the spinel structure, which contains one cation per formula unit on tetrahedral sites [...] (M2+ ,
sublattice A) and two cations per formula unit on octahedral sites {...} (Fe3+ ,
sublattice B). The exchange between the A and B sublattices is negative, which
yields a ferrimagnetic spin structure. The cation distribution over the A and B sites
depends on both chemical composition and magnet processing. For example, Fe3 O4
crystallizes in the so-called “inverse” spinel structure [Fe3+ ] {Fe2+ Fe3+ }(O2− )4
[65]. The total magnetization, measured in μB per formula unit, is therefore
[−5] + {5 + 4} = 4.
In the classical limit, the mean-field Curie temperature is given by the lowest
eigenvalue of the N × N matrix in the equation
kB T <si > =
j Jij <sj >
(47)
2 Magnetic Exchange Interactions
85
Fig. 11 Spin structures (schematic): ferromagnets (FM), antiferromagnet (AFM), ferrimagnet
(FI), Pauli paramagnet (PM), and noncollinear spin structure (NC)
This matrix equation is easily generalized to quantum-mechanical case, by carefully
counting neighbors and using the appropriate de Gennes factors and Brillouin
functions [54, 62]. The number N of sublattices is equal to the number of
nonequivalent atoms. In disordered solids, all atom are nonequivalent and N → ∞.
For two sublattices A and B, Eq. (47) becomes
3 kB T <sA > =
j JAA <sB > + JAB <sB >
(48a)
3 kB T <sB > =
j JBA <sB > + JBB <sB >
(48b)
Here JAA/BB and JAB/BA are the classical intra- and intersublattice exchange
constants, respectively, and the factor 3 reflects the classical limit of the Brillouin
functions. The solution of Eq. (48) is
Tc =
1
6 kB
(JAA + JBB ) ±
(JAA − JBB )2 + 4 JAB JBA
(49)
The two sublattices often have different numbers of atoms, so that JAB = JAB
in general, but the two intersublattice exchange constants enter Eq. (49) in the
form of the product JAB JBA , and it is sufficient to consider J ∗ = (JAB JBA )1/2 .
For one-sublattice ferromagnets, where JBB = J ∗ = 0, Tc is equal to JAA /3kB .
86
R. Skomski
Various scenarios exist for two-sublattice magnets. In the simplest AFM case, the
two intrasublattice exchange interactions JAA = JBB = 0 and J ∗ <0, yielding the
Néel temperature TN = –Tc = |J ∗ |/3kB .
Metallic Sublattices Sublattice effect also occur in metals. In rare-earth transitionmetal (RE-TM) magnets, the RE-TM exchange (spin-spin) coupling is AFM for
the light rare earths and FM for the heavy rare earths. The orbital moment of the
TM sublattice is negligible, but inside each rare-earth atom, L and S are antiparallel
for light RE and parallel for heavy RE. This yields the spin structures [S↑ ]TM [S↓
L↑ ]RE for the light and [S↑ ]TM [S↓ L↓ ]RE for the heavy rare earths. The large but
opposite moment of the heavy rare-earth atoms yields a zero net magnetization
in some transition-metal-rich alloys. This spin state is referred to as compensated
ferrimagnetism (CFiM) and normally occurs at some compensation temperature
T0 , because different sublattices tend to have different temperature dependences
of magnetization [54, 63]. This is one of the features that distinguish CFiM from
AFM. Compensation occurs quite frequently in ferrimagnets, including oxides such
as rare-earth garnets R3 Fe5 O12 .
A rule of thumb for the exchange in transition-metal alloys is the switch rule:
The exchange is negative for interactions between late and early transition-metal
atoms but positive otherwise. For example, Co and Pt are both late transition-metal
elements, so the Co and Pt moments in CoPt are parallel. While the switch rule
includes alloys containing heavy transition-metal atoms (3d-4d and 3d-5d alloys),
it is not very reliable for elements in the middle of the series [64]. It also describes
impurities in host lattices and RE-TM intermetallic compounds, because rare earths
count as early transition metals, with one 5d electron contributing to the exchange
[65].
With the exception of very weak itinerant ferromagnets, intersublattice interactions in metals are well described by the Heisenberg model, and equations
like (49) provide good estimates of the ordering temperature [62]. For example,
transition-metal-rich rare-earth permanent magnets have JTT
J ∗ ≈ JRT and
JRR ≈ 0, so that the rare-earth contribution
to the Curie temperature is given by
Tc ≈ (JTT /3kB ) 1 + J ∗2 /JTT 2 . As a function of the number of 4f electrons, it
peaks in the middle of the lanthanide series, because J ∗2 involves the de Gennes
factor (Section “Intra-Atomic Exchange”)
Noncollinear Spin Structures There is a rich variety of noncollinear spin structures.
Spin glasses are disordered materials whose local magnetization is frozen below
some spin-glass transition temperature Tf [66, 67]. The definition of Tf , the spin state
below and above Tf , the nature of the transition, and the microscopic description are
nontrivial, but there is normally a distribution of exchange interactions Jij , caused,
for example, by RKKY interactions between localized moments in a metallic host.
In the simplest case, Jij = ±Jo for the interaction with z neighbors. For z → ∞,
the eigenvalue distribution of the random matrix Jij obeys Wigner’s semicircle law
√
[66], and the corresponding mean-field estimate is Tf = z Jo /kB . The situation is
2 Magnetic Exchange Interactions
87
further complicated by different types of disorder that can occur. Chemical disorder
means atomic substitutions with little or no changes in atomic positions. Bond and
topological
disorders involve substantial changes in atomic positions and in Jij =
|r i − r j | , but in the latter case, there is no continuous transformation connecting
the ordered and disordered lattices [67].
Helimagnetism arises when competing exchange interactions between nearest
and next-nearest neighbors yield spin spirals of wave vector k. Such structures
are realized, for example, in the heavy rare-earth elements, where k || ez [68].
Consideration of a-b planes labeled by magnetization angles θ n yields the classical
Heisenberg energy
E = –J
n
cos (θn+1 –θn ) –J
n cos (θn+2 –θn )
(50)
where J and J are the exchange interaction between neighboring layers, respectively. The energy (50) is minimized by the ansatz θ n + 1 = θ n + δ, where δ ∼ 1/k
is the magnetization rotation between subsequent layers:
J + 4J cos δ sin δ = 0
(51)
Aside from including FM (δ = 0) and AFM (δ = π) states,
has
this equation
noncollinear or helimagnetic (0 < δ < π) solutions, δ = arccos −J /4J .
Noncollinear spins structures are very common for elements in the middle
of the iron series, notably Cr and Mn, where the exchange contains competing
antiferromagnetic interactions. Elemental Cr forms a spin-density wave where the
AFM sublattice magnetization exhibits a real-space oscillation with a periodicity of
about 6 nm [59].
Dzyaloshinski-Moriya Interactions Noncollinearity may also arise from relativistic
Dzyaloshinski-Moriya (DM) interactions [69–71], which occur in structures with
violated or “broken” inversion symmetry. Examples are MnSi [72], α-Fe2 O3
(hematite) [65], and structurally disordered magnets such as spin glasses [66], as
well as in artificial magnetic nanostructures [73]. DM interactions are described by
the Hamiltonian HDM = – i>j Dij · Si × Sj , wherethe direction
of the DM vector
Dij = −Dj i is given by Dij ∼ n (r i − r n ) × r j − r n . In this expression, i
and j denote the two DM-interacting spins, and rn is the position of a magnetic or
nonmagnetic neighbor (Fig. 12). Physically, d electrons hop from atom i to atom
n and then to atom j. Unless rn is located on the line connecting ri and rj (and the
cross product determining D is zero), the hopping sequence involves a change of
direction at rn , which creates a partial orbit around rn and some spin-orbit coupling
that affects the spins i and j. The DM interaction changes the spin projections onto
the plane created by the vectors ri – rn and rj – rn : It tries to make Si and Sj parallel
to ri – rn and rj – rn , respectively.
88
R. Skomski
Fig. 12 Dzyaloshinski-Moriya Interactions in (a–b) crystals and (b–c) thin films. The red atoms
are magnetic, whereas the blue and white atoms are nonmagnetic but have weak (white) and strong
(blue) spin-orbit coupling
Since DM interactions are caused by spin-orbit coupling, they are a weak
relativistic effect, comparable to micromagnetic dipolar interactions and to magnetocrystalline anisotropy. They compete against the dominant Heisenberg exchange
and create canting angles of the order of 1◦ in typical magnetic materials [74].
By contrast, noncollinearities due to competing ferromagnetic or antiferromagnetic
exchange (Eq. (51)) can assume any value between 0◦ and 180◦ . However, DM
canting angles substantially larger than 1◦ are possible in materials with weak
Heisenberg exchange (low Tc ).
DM effects are strongly point-group-dependent, and the absence of inversion
symmetry is a necessary but not sufficient condition [75]. For example, inverse cubic
Heusler alloys have zero net DM interactions in spite of their noncentrosymmetric
point group Td . Figure 12(a–b) illustrates DM interactions in an orthorhombic bulk
2 Magnetic Exchange Interactions
89
crystal without inversion symmetry (point group C2v ). The fictitious crystal has
an equiatomic MT composition, where M is a magnetic or nonmagnetic metallic
element and T is a transition-metal element. The structure yields a spin spiral in the
x-z plane, that is, perpendicular to the net DM vector. B20-ordered cubic crystals
such as MnSi (point group T) are unique in the sense that their space group (P21 3)
is achiral due to the 180◦ character of the 21 screw axis but becomes chiral through
the incorporation of a chiral MnSi motifs.
Figure 12(c–d) shows the effect of Dzyaloshinski-Moriya interactions in thin
films with perpendicular anisotropy and fourfold (C4v ) or sixfold (C6v ) symmetry
(side view). When a patch of magnetic material is deposited on a material with
strong spin-orbit coupling, for example, Co on Pt, the modified spin structure is
reminiscent of a hedgehog. Such DM interactions are of interest in the context of
magnetic skyrmions. For example, bubble domains in thin films have a nonzero
skyrmion number and therefore yield a topological Hall effect (THE) [76], but DM
interactions change the spin structure of the bubble and the THE, thereby adding
new physics.
Spin Waves and Anisotropic Exchange
The low-lying excitations in Heisenberg magnets are of the spin-wave or magnon
type. Spin waves are of interest in experimental and theoretical physics and also
important in applied physics (microwave resonance, exchange stiffness). Chapter
SPW is devoted to spin waves, and in this chapter, the focus is on exchange in spin1/2 Heisenberg magnets, where quantum effects are most pronounced. To solve the
ferromagnetic Heisenberg model, it is convenient to rewrite the exchange term in
Eq. (27) as
S i · S i+1 =
1 + −
+ Si Si+1 + Si− Si+1
+ Sz,i Sz,i+1
2
(52)
The Sz operators measure the spin projections, Sz |↑>= + 12 |↑> and Sz |↓>=
− 12 |↓>, but leave the wave function unchanged. The spin-flip operators S + and S −
rise and lower the spin by one unit, respectively: S + |↓>=|↑> and S − |↑>=|↓>.
Since the S = 1/2 Heisenberg model has only two spin states, S + |↑>= 0 and
S − |↓>= 0, or symbolically S + S + = 0 and S − S − = 0. The products of
S + and S − in Eq. (52) have the effect of interchanging spins of opposite sign:
|↑↓ > becomes |↓↑ > and vice versa.
The ferromagnetic state, symbolically |0 > = |↑↑↑↑↑↑↑↑...>, is an eigenstate
of the Hamiltonian, because each of the spin-flip terms contains an S + operator
that creates a zero. One might naively expect that a single switched spin creates an
excited eigenstate, for example, |i > = |↑↑↑↓↑↑↑↑...>, where Ri is the position
of the flipped spin. However, the spin-flip operators move the flipped spin and
thereby create wave functions |i + 1 > and |i–1>. The low-lying eigenstates
of the ferromagnetic chain are actually plane-wave superpositions of single-spin
90
R. Skomski
Fig. 13 Spin wave (schematic)
flips, |ψ k > = exp.(ik·Rj ) |j>. These wave-like excited states are the spin waves
or magnons. Each magnon corresponds to one switched spin, but the reversal is
delocalized rather than confined to a single atom (Fig. 13). The corresponding
excitation energy is E = 4(1 – cos(k a)). For arbitrary crystals and spins [63]
E(k) = 2S
jJ
R j 1– cos k · R j
(53)
where Rj is the distance between the exchange-interacting atoms.
Of particular interest is the long-wavelength limit, where the dispersion relation
(53) becomes quadratic. The three monatomic cubic lattices (sc, bcc, fcc) have [63]
E(k) = 2SJ a 2 k 2
(54)
This equation cannot be generalized to more complicated cubic crystals, because
E(k) is governed by the interatomic distance Rj , not by the lattice constant a,
which can be very large due to superlattice formation. Application of Eq. (53) to
crystals without second-order structural anisotropy and z nearest neighbors (distance
R) yields E(k) = 2zSJ (1– sin(kR)/kR), which has the long-wavelength limit
E(k) =
z
SJ R 2 k 2
3
(55)
For sc, bcc, and fcc lattices, Eq. (55) is equivalent to Eq. (54). In good approximation, it can also be applied to slightly noncubic structures. For example, elemental
cobalt has R = 2RCo , where RCo = 1.25 Å nm is the atomic radius of fcc and hcp Co.
Strongly anisotropic structures, such as multilayers, require an explicit evaluation of
Eq. (53).
It is common to write this relation as E = D k2 , where D is the spin-wave
stiffness. In micromagnetism, it is convenient to write the exchange energy as
E =
A [∇s]2 dV
where A is the exchange stiffness. Comparison of Eqs. (54) and (56) yields
(56)
2 Magnetic Exchange Interactions
91
Table 3 Spin-wave stiffness
D and exchange stiffness A
for some materials [78]
Material
Fe
Co
Ni
Ni80 Fe20
Co2 MnSn
Fe3 O4
A = 2 c S2
J
a
D
meV/nm2
2.8
4.5
4.0
2.5
2.0
5.0
A
pJ/m
20
28
8
10
6
7
(57)
where c is the number of atoms per unit cell (c = 1 for sc, c = 2 for bcc, c = 4 for
fcc). Similar to Eq. (54), Eq. (57) cannot be used for arbitrary crystals, whose lattice
constants can be very large, and for dilute magnets [90]. In terms of the interatomic
distance, the rule of thumb is A ≈ zS 2 J /5R. Values of spin wave stiffness D and
exchange stiffness A for some common magnets are given in Table 3.
Anisotropic Exchange Anisotropic exchange is a vague term, used for a variety of
physically very different phenomena, sometimes even for the Dzyaloshinski-Moriya
interactions.
Spin waves are affected by magnetocrystalline anisotropy, especially in noncubic
magnets. The anisotropy adds a spin-wave gap Eg = E(k = 0) to Eq. (54) and
also affects the exchange stiffness. For example, in uniaxial (tetragonal, hexagonal,
trigonal) magnets, one needs to distinguish A|| (along the c-axis) and A⊥ (in the ab-plane). The difference is particularly large in multilayers, where the intra-layer
exchange (A⊥ ) is often much stronger than the interlayer exchange (A|| ). The
Heisenberg interaction behind this type of anisotropic exchange remains isotropic,
as in Eq. (27), and the difference
between
A|| and A⊥ is caused by the nonrelativistic
bond anisotropy, Jij = J R i − R j [79].
Very different physics are involved in the so-called anisotropic Heisenberg
model, which derives from the (isotropic) Heisenberg model by the replacement
J ŝ · ŝ → J
ŝx · ŝx + ŝy · ŝy + Jz ŝz · ŝ z
(58)
The exchange anisotropy J = Jz –J /Jz , which has the same relativistic
origin as magnetocrystalline anisotropy and the Dzyaloshinski-Moriya interaction
(Section “Magnetic Order and Noncollinearity”),
is normally
very small compared
to the average or “isotropic” exchange Jo = 2J + Jz /3. However, J becomes
non-negligible when Jo is very small, for example, in some compounds with low
Curie temperature [80].
92
R. Skomski
The XY and Ising models are obtained by putting Jz = 0 and J = 0 in Eq.
(58), respectively. In classical statistical mechanics, these models have the spin
dimensions n = 2 (XY) and n = 1 (Ising), as compared to n = 3 (Heisenberg model),
n = ∞ (spherical model), and n = 0 [4, 60]. The spin dimension has a profound
effect on the onset of ferromagnetism in D-dimensional crystals. Ising ferromagnets
have Tc = 0 in one real-space dimension (D = 1) but Tc > 0 for D ≥ 2. Heisenberg
magnets have Tc = 0 in one and two real-space dimensions but Tc > 0 for D ≥ 3. For
all ferromagnetic spin dimensionalities, statistical mean-field theory is qualitatively
correct in D > 4 real-space dimensions, with logarithmic corrections in D = 4. For
the geometrical meaning of D = 4, see Figs. 7.9, 7.10 in Ref. [4].
Two-dimensional magnets (D = 2) are particularly intriguing. The Heisenberg
model (n = 2) predicts Tc = 0, but adding an arbitrarily small amount of uniaxial
anisotropy to the Heisenberg model yields Tc > 0 [81]. This feature has recently
received renewed attention in the context of the two-dimensional van der Waals
(VdW) magnetism. The two-dimensional XY model (n = 2, D = 2) yields a
Thouless-Kosterlitz transition with a power-law decay of spin-spin correlations but
no long-range magnetic order.
There are actually two types of Ising models, characterized by similar Hamiltonians, J = 0 in Eq. (58), but different Hilbert spaces. Ising’s original model is de
facto a classical Heisenberg model with infinite magnetic anisotropy [82, 83], which
leads to two spin orientations, ↑ and ↓. The quantum-mechanical Ising model, also
known as the “spin-1/2 Ising model in a transverse field” [84–86], is physically very
different. For example, it allows states with <sx > = 0 and <sy > = 0, whereas the
idea of the (original) Ising model is to suppress such states, <sx > = <sy > = 0.
Exchange anisotropy or exchange bias in thin films means that a pinning layer
yields a horizontal hysteresis-loop shift in a free layer. The bias is realized through
FM or AFM exchange at the interface between the pinning and soft layers, but its
ultimate origin is the magnetocrystalline anisotropy of the pinning layers, which
is often an antiferromagnet. The situation is physically similar to the horizontal
and vertical hysteresis-loop shifts sometimes observed in hard-soft composites,
which are inner-loop effects. Micromagnetically, the exchange-energy density is not
confined to the atomic-scale interface but extends into the pinning and free layers,
so that the net interlayer exchange energy per film area is generally very different
from the atomic-scale interlayer exchange [73].
Experimental Methods
There are many methods to investigate exchange, directly and indirectly, some of
which are briefly mentioned here. Magnetic measurements are used to determine
Curie temperatures Tc ∼ J , from which exchange constants can be deduced. The
low-field magnetization of antiferromagnets is zero, but high fields tilt the AFM
sublattices and yield a small magnetization M(H ) ∼ H /J .
2 Magnetic Exchange Interactions
93
The exchange may also be deduced from the low-temperature M(T) curves,
because the Bloch law involves the exchange stiffness. Magnetic force and, to
a much lesser extent, anomalous magneto-optic microscopies used to investigate
magnetic domain structures, which contain implicit information about the exchange
stiffness. A direct method to probe exchange is magnetic resonance.
Neutron diffraction and, to a much lesser extent, X-ray diffraction (XRD) are
important methods to probe spin structure. The magnetic XRD signal is much
weaker than the neutron-diffraction signal. Interatomic exchange can be probed
by a variety of methods, such as X-ray magnetic dichroism, which also allows
a distinction of L and S contributions to the atomic moments. Electron-transport
measurements are frequently used to gauge and confirm exchange effects.
Antiferromagnetic Spin Chains
Spin waves are particularly intriguing in antiferromagnets, whose low-lying states
correspond to the highest excited states in the ferromagnetic case [8, 16, 17]. By
analogy with FM ground state, |↑↑↑↑↑↑↑↑...>, one could intuitively assume that
the AFM ground state is a superposition of the two quasi-classical Néel states
|AFM (1)>=| ↑↓↑↓↑↓↑↓ · · · >
(59)
|AFM (2)>=| ↑↓↑↓↑↓↑↓ · · · >
(60)
and
However, the spin-flip terms in Eq. (52) do not transform Eqs. (59) and (60) into
each other but create pairs of parallel spins (spinons), for example
S4 + S5− |↑↓↑↓↑↓↑↓ · · · >=|↑↓↑↑↓↓↑↓ · · · >
(61)
Using the Néel states to evaluate Eq. (27) yields an AFM ground-state energy of
−0.5 J per atom, compared to the exact Bethe result of −2 J (ln 2 − 1/4) =
−0.886 J [8]. Systems with such complicated ground states are also known as
quantum spin liquids (QSL). The underlying physics is very similar to the spin
mixing discussed below Fig. 10 but now involves an infinite number of spins.
The derivation of the Heisenberg model, Section “Hubbard Model”, was based
on the neglect of interatomic hopping. Strictly speaking, this is meaningful only
when U is large and the band is half-filled. In more- or less-than-half-filled bands,
a fraction of the electrons can move almost freely. Such magnets have both charge
and spin degrees of freedom, and the corresponding extension of the Heisenberg
chain is known as the Tomonaga-Luttinger model or Luttinger liquid [87, 88]. A
typical wave function is |↑↓↑↑↓◦↑↓>, which contains one hole. The model has a
number of interesting features. For example, spin and charge excitations move with
different velocities, the former being slower, because spin excitations have lower
94
R. Skomski
energies δE = ω than charge excitations. This is an example of a correlation effect
known as spin-charge separation [88]. By contrast, in the itinerant limit, charge and
spin degrees are closely linked. Spin-charge separation is important in the Kondo
mechanism (where low-energy spin flips determine the resistivity) and in hightemperature superconductivity (Section “Dimensionality Dependence of Quantum
Antiferromagnetism”).
The electron distribution n(E) of a Luttinger liquid is very different from a Fermi
liquid [88]. Weak correlations in metallic magnets create particle-hole quasiparticles
but leave the Fermi surface otherwise intact. Strong correlations, as in the Luttinger
liquid, completely destroy the Fermi surface, and n(E) becomes a smooth function.
Dimensionality Dependence of Quantum Antiferromagnetism
The Luttinger liquid is a typical one-dimensional effect: Arbitrarily small perturbations of structural, thermal, or quantum-mechanical origin destroy long-range
magnetic order. Quantum-spin-liquid effect in higher dimensions are generally
less pervasive but not necessarily unimportant. In antiferromagnets, it is possible
to redefine the operators Si ± in Eq. (52) by reversing the spin in each second
atom, which assimilates the AFM problem to the FM problem and allows the
consideration of spin waves. However, this procedure creates terms of the type
Si + Sj + and Si − Sj − , where the atoms i and j belong to different sublattices. These
terms go beyond straightforward spin-wave theory, which exclusively involves
Si + Sj − and Si − Sj + . The additional terms yield a quantum-mechanical mixture
of the two sublattices and require an additional diagonalization procedure known as
Bogoliubov transformation [3]. The sublattice admixture reduces both the energy
of the AFM state and the sublattice magnetizations, the latter meaning that the
↑ sublattice acquires some ↓ character and vice versa.
The ground-state energy is discussed most conveniently by starting from two
interacting spins having S = 1/2, as described by Eq. (26). For a given AFM
exchange J < 0, the energies of the FM and AFM states scale as S2 = 1/4 and
–S(S + 1) = −3/4, respectively. More generally, the AFM energy is proportional
to –S(S + δ), where δ describes the intersublattice admixture and 0 < δ < 1. Spinwave theory yields δ = 0.363 for the linear chain, δ = 0.158 for the square lattice,
and δ = 0.097 for the simple-cubic lattice. The one-dimensional value is close to
the exact result δ = 0.386 for S = 1/2. A rough estimate for the relative reduction
of the sublattice magnetization in hypercubic magnets (z = 2d) is 0.15/S(d − 1),
corresponding to sublattice magnetizations of 0%, 70%, and 85% in one-, two-, and
three-dimensional magnets having S = 1/2.
The complete magnetization
collapse
in one dimension is caused by the involve
ment of the integral k−1 dk ∼ kd – 1 dk, which exhibits an infrared (small-k)
divergence in one dimension. The same integral is behind Bloch’s law and the
Wagner-Mermin theorem, and in all cases, the divergence indicates that fluctuations
destroy long-range magnetic order in one dimension. However, the underlying
physics is different: The fluctuations considered in Bloch’s law and in the Wagner-
2 Magnetic Exchange Interactions
95
Mermin theorem are of thermodynamic origin, whereas the present ones are
zero-temperature quantum fluctuations. These fluctuations, which are largest for
S = 1/2, are a correlation effect and therefore difficult to treat in density-functional
calculations.
One example is high-temperature superconductivity (HTSC) in La2-x Srx CuO4 ,
which involves 3d9 states in the Cu-O planes of the oxides [9] and where spin-1/2
quantum fluctuations trigger the formation of Cooper pairs. The parent compound
La2 CuO4 is a strongly correlated antiferromagnetic semiconductor, but Sr doping
drives the system toward a phase transition. The denominator in Eq. (40) becomes
small, meaning that spin fluctuations (antiferromagnetic paramagnons) are strongly
enhanced by Sr doping. Furthermore, both spin-charge separation [9] and critical
slowing down cause the spin fluctuations to evolve very sluggishly, so that they can
play the role of phonons in BCS superconductors.
Frustration, Spin Liquids, and Spin Ice
A number of exotic topics in physics are more or less closely related to exchange
interactions. This subsection discusses both classical and quantum-mechanical
implications of frustration, as well as some related micromagnetic questions.
Frustration Ring configurations with antiferromagnetic interactions and odd numbers N of atoms offer intriguing physics. Let us start with theclassical exchange
energy between atoms at Ri and Rj , which is equal to −Jij cos φi − φj . Consider
an equilateral triangle (N = 3) with antiferromagnetic nearest-neighbor interactions
(Fig. 14). If the exchange was ferromagnetic, then φ i = 0 (or φ i = const.) would
simultaneously minimize the energy of all bonds and yield a ground-state energy
of −3J . The corresponding antiferromagnetic solution, of energy −3 |J |, does not
exist, because three antiferromagnetic bonds cannot be simultaneously realized in
a triangle (Fig. 14a). This is referred to as magnetic frustration. Fig. 14b shows
that bond angles of 120◦ , rather than 180◦ , may be realized for all spins, and the
corresponding ground-state energy is −1.5 |J |, somewhat lower than the energy
− |J | of (a).
The classical frustration problem of Fig. 14 is elegantly summarized by a
construction known as Frost’s circle. The approach was developed to describe
the hopping of p electrons in cyclic molecules [89] but can also be applied to
s-state electrons such as those in Fig. 7 and to interatomic exchange, because the
involvement of τ ij and Jij is mathematically equivalent. For a ring of N atoms with
nearest-neighbor exchange J , the energy eigenvalues per atom are
En = –2 J cos (2 π n/N )
(62)
where n = 0, ..., N–1, N. These energies can be arranged on a circle, as exemplified
by the example N = 5 and ferromagnetic coupling (J > 0) in Fig. 14(c). The
96
R. Skomski
Fig. 14 Classical frustration in rings of N atoms: (a) frustrated state (N = 3), (b) ground state
(N = 3), and (c) graphical solution (Frost’s cycle) for N = 5
FM ground state has n = 0 and the energy −2J . However, the figure shows
that energy eigenvalues are not necessarily symmetric with respect to changing
the sign of J . For odd values of N, the AFM ground state is double degenerate
and characterized by nearest-neighbor spin angles (1–1/N) 180◦ , as opposed to the
180◦ expected for ideal antiferromagnetism. The incomplete antiparallelity leads to
a ground-state energy 2J cos (π/N), higher than that of an ideal antiferromagnet
(2J ). This analysis shows exchange in antiferromagnets is different from exchange
in ferromagnets, even in the classical limit.
The quantum-mechanical ground state of the AFM spin-1/2 Heisenberg triangle
is obtained by using Eq. (52) to evaluate the matrix elements between the states
|↑↑↓>, |↑↓↑>, and |↓↑↑>. It is sufficient to consider Sz = 1/2, since none of
the terms in Eq. (52) changes the total spin projection Sz . For example, spin
configurations such as |↑↑↑ > (Sz = 3/2) and |↑↑↓ > (Sz = 1/2) do not mix.
Furthermore, Sz = − 1/2 is equivalent to Sz = +1/2 and does not need separate
consideration. For the three states with Sz = 1/2, the Hamiltonian is
⎛
⎞
⎛
⎞
100
111
3
H = J ⎝0 1 0⎠ − J ⎝1 1 1⎠
2
001
111
(63)
The diagonalization of this matrix is trivial and yields one FM eigenstate (1, 1, 1)
of energy – 3J /2 > 0 and two AFM eigenstates of energy 3J /2 < 0, for example,
|Ψ 1 > = (2, −1, −1) and |Ψ 2 > = (0, 1, −1). Explicitly
1
|1 >= √ (2| ↑↑↓ >−| ↑↓↑ >−| ↓↑↑ >)
6
and
(64a)
2 Magnetic Exchange Interactions
1
|2 > = √ (| ↑↓↑ >−| ↓↑↑ >)
2
97
(64b)
Since the two AFM states are degenerate, |Ψ > = c1 |Ψ 1 > + c2 |Ψ 2 > is also an
eigenstate; complex numbers c1/2 mean a spin component in the y-direction.
Alternatively, Eq. (64b) is a product of the type (|↑↓>− |↓↑>)⊗|↑> and contains a
maximally entangled AFM singlet |↑↓>− |↓↑>. According to Eq. (63), the corners
of the triangle are equivalent, so that the singlet may be placed on any of the three
bonds.
The spin-1/2 Heisenberg square has the spin projections Sz = (0, ±1, ±2). In the
antiferromagnetic ground state (Sz = 0), there are six spin configurations, namely,
the two Néel states |↑↓↑↓> and |↓↑↓↑>, as well as six non-Néel states with pairs
of parallel spins, such as |↑↑↓↓>. Classically, non-Néel states are not expected to
appear in the ground state, but the diagonalization of the corresponding 6 × 6 matrix
yields an AFM ground-state singlet with a strong admixture of non-Néel character,
namely
1
1
1
|>= | ↑↓↑↓ >+ | ↓↑↓↑ >− √ (| ↑↑↓↓ >+ . . . )
2
2
8
(65)
The total weight of the non-Néel configurations in the ground state is 50%. This
ground state is rather complicated but, unlike Eq. (64), not frustrated.
Quantum-spin Liquids The quantum-mechanical behavior of the structures of
Fig. 14 is liquid-like, similar to the Luttinger liquid of Sect. “Antiferromagnetic Spin
Chains”. Two-dimensional magnets are particularly interesting. Figure 15 shows 2D
lattices with the triangular structural elements required for AFM frustration. The
wave functions are complicated but contain AFM singlets, and there are many ways
of arranging these singlets on a lattice (a). Some materials investigated as QSL
materials, such as ZnCu3 (OH)6 Cl2 (herbertsmithite), form Kagome lattices (b).
Triangular and Kagome lattices exhibit similar frustration behaviors, but Kagome
lattices differ by having low-lying excitations [91].
Frustration is also important in spin-ice materials, such as pyrochlore-ordered
Dy2 Ti2 O7 . The pyrochlore structure consists of tetrahedra whose corner atoms are
magnetic. The strong crystal field forces the moments to lie on lines from the corners
to the centers of the tetrahedra, but to fix the spin direction (inward or outward),
one needs an additional criterion known as the “two-in, two-out” rule. There are
many two-in two-out configurations, which creates a spin-ice situation reminiscent
of Fig. 15a.
There are excited spin-ice states where all spins point inward (four in) or outward
(four out), which yields an accumulation of magnetic charge (south poles or north
poles) in the middle of the tetrahedron. Such accumulations are also known as
magnetic monopoles, but this characterization is misleading. Magnetic monopoles
are high-energy elementary particles having B·dA = 0. Their existence cannot
98
R. Skomski
Fig. 15 Frustrated lattices in two dimensions: (a) triangular lattice and (b) Kagome lattice
Fig. 16 Static magnetic field
sources: (a) magnetic dipole
and (b) “magnetic monopole”
be ruled out, but they have never yet been observed in the universe. Solid-state
“monopoles”
are no monopoles, because they are formed from dipoles and do not
violate B·dA = 0. For illustrative purposes, the magnetic charges near the ends of
a long bar magnet may be regarded monopoles, but these are not real monopoles but
merely the ends of long dipoles. Figure 16 compares a magnetic dipole (a) with a
putative magnetic monopole (b). The configuration (b) may be created in the form of
a magnetic-dipole layer or by some radial magnetization distribution in a magnetic
material.
In any case, it requires compensating south-pole charges inside the sphere,
so that B · dA = 0. As a consequence, the magnetic field is actually zero outside
the sphere, so that the finite-length arrows in (b) are without physical basis.
In the context of new materials, it is important to keep in mind that exchange
interactions are described in terms of Hamiltonians. The equation of motion of any
system is more fundamentally governed by the Lagrangian and its time integral, the
action. The difference can be ignored in flat spaces but is important in curved and
periodic spaces, where it corresponds to the Berry phase. The contributions of the
phase are basically of a ‘zero-Hamiltonian’ type and ignored in this chapter.
2 Magnetic Exchange Interactions
99
Acknowledgments This chapter has benefited from help in details by P. Manchanda and R.
Pathak and from discussions with B. Balamurugan, C. Binek, X. Hong, Y. Idzerda, A. Kashyap,
P. S. Kumar, D. Paudyal, T. Schrefl, D. J. Sellmyer, and A. Ullah. The underlying research
in Nebraska has been supported by DOE BES (DE-FG02-04ER46152) NSF EQUATE (OIA2044049), the NU Collaborative Initiative, HCC and NCMN.
References
1. Heisenberg, W.: Zur Theorie des Ferromagnetismus. Z. Phys. 49, 619–636 (1928)
2. Pauling, L., Wilson, E.B.: Introduction to Quantum Mechanics. McGraw-Hill, New York
(1935)
3. Jones, W., March, N.H.: Theoretical Solid State Physics I. Wiley & Sons, London (1973)
4. Skomski, R.: Simple Models of Magnetism. University Press, Oxford (2008)
5. Stoner, E.C.: Collective electron ferromagnetism. Proc. Roy. Soc. A. 165, 372–414 (1938)
6. Slater, J.C.: Ferromagnetism in the band theory. Rev. Mod. Phys. 25, 199–210 (1953)
7. Stollhoff, G., Oleś, A.M., Heine, V.: Stoner exchange interaction in transition metals. Phys.
Rev. B. 41, 7028–7041 (1990)
8. Ashcroft, N.W., Mermin, N.D.: Solid State Physics. Saunders, Philadelphia (1976)
9. Fulde, P.: Electron Correlations in Molecules and Solids. Springer, Berlin (1991)
10. Coulson, C.A., Fischer, I.: Notes on the molecular orbital treatment of the hydrogen molecule.
Philos. Mag. 40, 386–393 (1949)
11. Senatore, G., March, N.H.: Recent progress in the field of electron correlation. Rev. Mod. Phys.
66, 445–479 (1994)
12. Avella, A., Manchini, F. (eds.): Strongly Correlated Systems: Theoretical Methods. Springer,
Berlin (2012)
13. Avella, A., Manchini, F. (eds.): Strongly Correlated Systems: Numerical Methods. Springer,
Berlin (2013)
14. Metzner, W., Vollhardt, D.: Correlated lattice fermions in d = ∞ dimensions. Phys. Rev. Lett.
62, 324–327 (1989)
15. V. I. Anisimov, A. I. Poteryaev, M. A. Korotin, A. O. Anokhin, and G Kotliar, First-principles
calculations of the electronic structure and spectra of strongly correlated systems: dynamical
mean-field theory“, J. Phys.: Condens. Matter 9, 7359–7367 (1997)
16. Karbach, M., Müller, G.: Introduction to the Bethe ansatz I. Comput. Phys. 11, 36–44 (1997)
17. Karbach, M., Hu, K., Müller, G.: Introduction to the Bethe ansatz II. Comput. Phys. 12, 565–
573 (1998)
18. Hubbard, J.: Electron correlations in narrow energy bands. Proc. R. Soc. London Ser. A. 276,
238–257 (1963)
19. Gutzwiller, M.C.: Effect of correlation on the ferromagnetism of transition metals. Phys. Rev.
Lett. 10, 159–162 (1963)
20. Şaşıoğlu, E., Friedrich, C., Blügel, S.: Effective Coulomb interaction in transition metals from
constrained random-phase approximation. Phys. Rev. B. 83., 121101R-1-4 (2011)
21. Zaanen, J., Sawatzky, G.A., Allen, J.W.: Band gaps and electronic structure of transition-metal
compounds. Phys. Rev. Lett. 55, 418–421 (1985)
22. Brinkman, W.F., Rice, T.M.: Application of Gutzwiller’s Variational method to the metalinsulator transition. Phys. Rev. B. 2, 4302–4304 (1970)
23. Hund, F.: Atomtheoretische Deutung des Magnetismus der seltenen Erden. Z. Phys. 33, 855–
859 (1925)
24. Herzberg, G.: Atomic Spectra and Atomic Structure. Dover, New York (1944)
25. Condon, E.U., Shortley, G.H.: The Theory of Atomic Spectra. University Press, Cambridge
(1959)
26. Taylor, K.N.R., Darby, M.I.: Physics of Rare Earth Solids. Chapman and Hall, London
(1972)
100
R. Skomski
27. Goodenough, J.B.: Magnetism and the Chemical Bond. Wiley, New York (1963)
28. Anderson, P.W.: Exchange in insulators: Superexchange, direct exchange, and double
exchange. In: Rado, G.T., Suhl, H. (eds.) Magnetism I, pp. 25–85. Academic Press, New York
(1963)
29. Mattis, D.C.: Theory of Magnetism. Harper and Row, New York (1965)
30. Skomski, R., Zhou, J., Zhang, J., Sellmyer, D.J.: Indirect exchange in dilute magnetic
semiconductors. J. Appl. Phys. 99., 08D504-1-3 (2006)
31. Zener, C.: Interaction between the d-shells in the transition metals. II. Ferromagnetic compounds of manganese with perovskite structure. Phys. Rev. 82, 403–405 (1951)
32. Brueckner, K.A., Sawada, K.: Magnetic susceptibility of an electron gas at high density. Phys.
Rev. 112, 328 (1958)
33. Bloch, F.: Bemerkung zur Elektronentheorie des Ferromagnetismus und der elektrischen
Leitfähigkeit. Z. Physik. 57, 545–555 (1929)
34. Janak, J.F.: Uniform susceptibilities of metallic elements. Phys. Rev. B. 16, 255–262 (1977)
35. Wohlfarth, E.P.: Very weak itinerant Ferromagnets; application to ZrZn2 . J. Appl. Phys. 39,
1061–1066 (1968)
36. Mohn, P.: Magnetism in the solid state. Springer, Berlin (2003)
37. Mohn, P., Wohlfarth, E.P.: The curie temperature of the ferromagnetic transition metals and
their compounds. J. Phys. F. 17, 2421–2430 (1987)
38. Balasubramanian, B., Manchanda, P., Skomski, R., Mukherjee, P., Das, B., George, T.A.,
Hadjipanayis, G.C., Sellmyer, D.J.: Unusual spin correlations in a nanomagnet. Appl. Phys.
Lett. 106., 242401-1-5 (2015)
39. Sommerfeld, A., Bethe, H.A.: Elektronentheorie der Metalle.”, Ferromagnetism. In: Flügge, S.
(ed.) Handbuch der Physik, Vol. 24/II, pp. 334–620. Springer, Berlin (1933)
40. Cardias, R., Szilva, A., Bergman, A., Di Marco, I., Katsnelson, M.I., Lichtenstein, A.I.,
Nordström, L., Klautau, A.B., Eriksson, O., Kvashnin, Y.O.: The Bethe-Slater Curve Revisited;
New Insights from Electronic Structure Theory. Sci. Rep. 7, 4058-1-11 (2017)
41. Skomski, R., Coey, J.M.D.: Permanent magnetism. Institute of Physics, Bristol (1999)
42. Sutton, A.P.: Electronic structure of materials. Oxford University Press (1993)
43. Coey, J.M.D.: New permanent magnets; manganese compounds. J. Phys.: Condens. Matter. 26,
064211-1-6 (2014)
44. Snow, R.J., Bhatkar, H., N’Diaye, A.T., Arenholz, E., Idzerda, Y.U.: Large moments in bcc
Fex Coy Mnz ternary alloy thin films. Appl. Phys. Lett. 112, 072403 (2018)
45. Kashyap, A., Pathak, R., Sellmyer, D.J., Skomski, R.: Theory of Mn-based high-magnetization
alloys. IEEE Trans. Magn. 54, 2102106-1-6 (2018)
46. Wollmann, L., Chadov, S., Kübler, J., Felser, C.: Magnetism in cubic manganese-rich Heusler
compounds. Phys. Rev. B. 90, 214420-1-12 (2014)
47. Skomski, R., Manchanda, P., Kumar, P., Balamurugan, B., Kashyap, A., Sellmyer, D.J.:
Predicting the future of permanent-magnet materials. IEEE Trans. Magn. 49, 3215–3220
(2013)
48. Anderson, P.W.: Concepts in Solids: Lectures on the Theory of Solids. W. A, Benjamin, New
York (1965)
49. Kohn, W.: Nobel lecture: electronic structure of matter—wave functions and density functionals. Rev. Mod. Phys. 71, 1253–1266 (1999)
50. Perdew, J.P., Burke, K., Ernzerhof, M.: Generalized gradient approximation made simple. Phys.
Rev. Lett. 77, 3865–3868 (1996)
51. Runge, E., Zwicknagl, G.: Electronic structure calculations and strong correlations: a model
study. Ann. Physik. 5, 333–354 (1996)
52. Skomski, R., Manchanda, P., Kashyap, A.: Correlations in rare-earth transition-metal permanent magnets. J. Appl. Phys. 117, 17C740-1-4 (2015)
53. Lieb, E.H.: Density functionals for coulomb systems. Int. J. Quantum Chem. 24, 243–277
(1983)
54. Smart, J.S.: Effective Field Theories of Magnetism. Saunders, Philadelphia (1966)
2 Magnetic Exchange Interactions
101
55. Anisimov, V.I., Aryasetiawan, F., Lichtenstein, A.I.: First-principles calculations of the electronic structure and spectra of strongly correlated systems: the LDA+ U method. J. Phys.:
Condens. Matter. 9, 767–808 (1997)
56. Liechtenstein, A.I., Katsnelson, M.I., Antropov, V.P., Gubanov, V.A.: Local spin density
functional approach to the theory of exchange interactions in ferromagnetic metals and alloys.
J. Magn. Magn. Mater. 67, 65–74 (1987)
57. Skomski, R., Balamurugan, B., Manchanda, P., Chipara, M., Sellmyer, D.J.: Size dependence
of nanoparticle magnetization. IEEE Trans. Magn. 53, 1–7 (2017)
58. Kondo, J.: Resistance Minimum in Dilute magnetic Alloys. Progr. Theor. Phys. 32, 37–49
(1964)
59. Wijn, H.P.J. (ed.): Magnetic properties of metals: d-elements, alloys, and compounds. Springer,
Berlin (1991)
60. Yeomans, J.M.: Statistical mechanics of phase transitions. University Press, Oxford (1992)
61. Wilson, K.G.: The renormalization group and critical phenomena. Rev. Mod. Phys. 55, 583–
600 (1983)
62. Duc, N.H., Hien, T.D., Givord, D., Franse, J.J.M., de Boer, F.R.: Exchange interactions in rareearth transition-metal compounds. J. Magn. Magn. Mater. 124, 305–311 (1993)
63. Kittel, C.: Introduction to Solid-State Physics. Wiley, New York (1986)
64. Kumar, P., Kashyap, A., Balamurugan, B., Shield, J.E., Sellmyer, D.J., Skomski, R.: Permanent
magnetism of intermetallic compounds between light and heavy transition-metal elements.
J. Phys.: Condens. Matter. 26, 064209-1-8 (2014)
65. Coey, J.M.: Magnetism and magnetic materials. University Press, Cambridge (2010)
66. Fischer, K.-H., Hertz, A.J.: Spin glasses. University Press, Cambridge (1991)
67. Moorjani, K., Coey, J.M.D.: Magnetic glasses. Elsevier, Amsterdam (1984)
68. Koehler, W.C.: Magnetic properties of rare-earth metals and alloys. J. Appl. Phys. 36, 1078–
1087 (1965)
69. Stevens, K.W.H.: A note on exchange interactions. Rev. Mod. Phys. 25, 166 (1953)
70. Dzyaloshinsky, I.: A thermodynamic theory of ’weak’ ferromagnetism of antiferromagnetics.
J. Phys. Chem. Solids. 4, 241–255 (1958)
71. Moriya, T.: Anisotropic Superexchange interaction and weak ferromagnetism. Phys. Rev. 120,
91–98 (1960)
72. Bak, P., Jensen, H.H.: Theory of helical magnetic structures and phase transitions in MnSi and
FeGe. J. Phys. C. 13, L881–L885 (1980)
73. Skomski, R.: Nanomagnetics. J. Phys.: Condens. Matter. 15, R841–R896 (2003)
74. R. Skomski, J. Honolka, S. Bornemann, H. Ebert, and A. Enders, “Dzyaloshinski–Moriya
micromagnetics of magnetic surface alloys”, J. Appl. Phys. 105, 07D533-1-3 (2009)
75. Ullah, A., Balamurugan, B., Zhang, W., Valloppilly, S., Li, X.-Z., Pahari, R., Yue,
L.-P., Sokolov, A., Sellmyer, D.J., Skomski, R.: Crystal structure and Dzyaloshinski–Moriya
micromagnetics. IEEE Trans. Magn. 55, 7100305-1-5 (2019)
76. Seki, S., Mochizuki, M.: Skyrmions in Magnetic Materials. Springer International, Cham
(2016)
77. Balasubramanian, B., Manchanda, P., Pahari, R., Chen, Z., Zhang, W., Valloppilly, S.R., Li,
X., Sarella, A., Yue, L., Ullah, A., Dev, P., Muller, D.A., Skomski, R., Hadjipanayis, G.C.,
Sellmyer, D.J.: Chiral Magnetism and High-Temperature Skyrmions in B20-Ordered Co-Si.
Phys. Rev. Lett. 124, 057201-1-6 (2020)
78. Dubowik, J., Gościańska, I.: Micromagnetic approach to exchange Bias. Acta Phys. Pol A.
127, 147–152 (2015)
79. Skomski, R., Kashyap, A., Zhou, J., Sellmyer, D.J.: Anisotropic Exchange. J. Appl. Phys. 97,
10B302-1-3 (2005)
80. de Jongh, L.J., Miedema, A.R.: Experiments on simple magnetic model systems. Adv. Phys.
23, 1–260 (1974)
81. Bander, M., Mills, D.L.: Ferromagnetism of ultrathin films. Phys. Rev. B. 38, 12015–12018
(1988)
102
R. Skomski
82. Ising, E.: Beitrag zur Theorie des Ferromagnetismus. Z. Phys. 31, 253–258 (1925)
83. Binek, C.: Ising-Type Antiferromagnets: Model Systems in Statistical Physics and in the
Magnetism of Exchange Bias. Springer, Berlin (2003)
84. Sachdev, S.: Quantum Phase Transitions. University Press, Cambridge (1999)
85. Skomski, R.: On the Ising character of the quantum-phase transition in LiHoF4 . AIP Adv. 6,
055704-1-5 (2016)
86. Stinchcombe, R.B.: Ising model in a transverse field. I. Basic theory. J. Phys. C. 6, 2459–2483
(1973)
87. Luttinger, J.M.: An exactly soluble model of a many-fermion system. J. of Math. Phys. 4,
1154–1162 (1963)
88. Schofield, A.J.: Non-Fermi liquids. Contemporary Phys. 40, 95–115 (1999)
89. Frost, A.A., Musulin, B.: A mnemonic device for molecular orbital energies. J. Chem. Phys.
21, 572–573 (1953)
90. Balasubramanian, B., Manchanda, P., Pahari, R., Chen, Z., Zhang, W., Valloppilly, S.R., Li,
X., Sarella, A., Yue, L., Ullah, A., Dev, P., Muller, D.A., Skomski, R., Hadjipanayis, G.C.,
Sellmyer, D.J.: Chiral Magnetism and High-Temperature Skyrmions in B20-Ordered Co-Si.
Phys. Rev. Lett. 124, 057201-1-6 (2020)
91. Mendels, P., Bert, F.: Quantum kagome frustrated antiferromagnets: One route to quantum spin
liquids. C. R. Phys. 17, 455–470 (2016)
Ralph Skomski received his PhD from Technische Universität
Dresden in 1991. He worked as a postdoc at Trinity College,
Dublin, and at the Max-Planck-Institute in Halle, before moving
to the University of Nebraska, Lincoln, where he is presently a
Full Research Professor. He is an analytical theorist with primary
research interests in magnetism, nanomaterials, and quantum
mechanics.
3
Anisotropy and Crystal Field
Ralph Skomski, Priyanka Manchanda, and Arti Kashyap
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Phenomenology of Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lowest-Order Anisotropies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Anisotropy and Crystal Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Tetragonal, Hexagonal, and Trigonal Anisotropies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Higher-Order Anisotropy Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Anisotropy Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Crystal-Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
One-Electron Crystal-Field Splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Crystal-Field Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Many-Electron Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin-Orbit Coupling and Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Rare-Earth Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Rare-Earth Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Operator Equivalents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Single-Ion Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Temperature Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Transition-Metal Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin-Orbit Matrix Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Crystal Fields and Band Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Itinerant Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
104
107
108
109
112
114
115
116
118
122
123
127
132
133
136
139
141
145
147
149
150
151
R. Skomski ()
University of Nebraska, Lincoln, NE, USA
e-mail: [email protected]
P. Manchanda
Howard University, Washington, DC, USA
A. Kashyap
IIT Mandi, Mandi, HP, India
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_3
103
104
R. Skomski et al.
First-Principle Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Other Anisotropy Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetostatic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Néel’s Pair-Interaction Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Two-Ion Anisotropies of Electronic Origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Dzyaloshinski-Moriya Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Antiferromagnetic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetoelastic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Low-Dimensional and Nanoscale Anisotropies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Surface Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Random Anisotropy in Nanoparticles, Amorphous, and Granular Magnets . . . . . . . . . . . .
Giant Anisotropy in Low-Dimensional Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendix A: Spherical Harmonics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendix B: Point Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendix C: Hydrogen-Like Atomic 3d Wave Functions . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
154
156
159
159
161
162
164
165
166
167
167
169
171
174
174
176
177
179
Abstract
Magnetic anisotropy, imposed through crystal-field and magnetostatic interactions, is one of the most iconic, scientifically interesting, and practically
important properties of condensed matter. This article starts with the phenomenology of anisotropy, distinguishing between crystals of cubic, tetragonal,
hexagonal, trigonal, and lower symmetries and between anisotropy contributions
of second and higher orders. The atomic origin of magnetocrystalline anisotropy
is discussed for several classes of materials, ranging from insulating oxides and
rare-earth compounds to iron-series itinerant magnets. A key consideration is the
crystal-field interaction of magnetic atoms, which determines, for example, the
rare-earth single-ion anisotropy of today’s top-performing permanent magnets.
The transmission between crystal field and anisotropy is realized by spin-orbit
coupling. An important crystal-field effect is the suppression of the orbital
moment by the crystal-field, which is known as quenching and has a Janus-head
effect on anisotropy: the crystal field is necessary to create magnetocrystalline
anisotropy, but it also limits the anisotropy in many systems. Finally, we
discuss some other anisotropy mechanisms, such as shape, magnetoelastic, and
exchange anisotropies, and outline how anisotropy is realized in some exemplary
compounds and nanostructures.
Introduction
Magnetic anisotropy means that the energy of a magnetic body depends on the
direction of the magnetization with respect to its shape or crystal axes. It is a quantity
of great importance in technology. For example, it is crucial for a material’s ability to
serve as a soft or hard magnet; governs many aspects of data storage and processing,
3 Anisotropy and Crystal Field
105
such as the areal density in the magnetic recording; and affects the behavior of
microwave and magnetic-cooling materials.
In the simplest case of uniaxial anisotropy, the energy depends on the polar angle
θ but not on the azimuthal angle φ of the magnetization direction:
Ea = V K1 sin2 θ + K2 sin4 θ + K3 sin6 θ
(1)
Here the Kn are the n-th anisotropy constants and V is the crystal volume. The first
anisotropy constant K1 is often the leading consideration. Ignoring K2 and K3 , the
anisotropy energy is equal to K1 V sin2 θ , and two cases need to be distinguished.
K1 > 0 yields energy minima at θ = 0 and θ = 180◦ , that is, the preferential
magnetization direction is along the z-axis (easy-axis anisotropy). When K1 < 0,
the energy is minimized for θ = 90◦ (easy-plane anisotropy).
The magnitudes of the room-temperature anisotropy constants K1 vary from less
than 5 kJ/m3 in very soft magnets to more than 17 MJ/m3 in SmCo5 . A variety of
rare-earth-free transition-metal alloys have anisotropies between 0.5 and 2.0 MJ/m3 .
YCo5 , where the Y is magnetically inert, has K1 = 5.0 MJ/m3 .
This chapter deals with the phenomenological description and physical origin of
anisotropy. A key question is how magnetic anisotropy depends on crystal structure
and chemical composition. The main contribution to the anisotropy energy of
most materials is magnetocrystalline anisotropy (MCA), which involves spin-orbit
coupling, a relativistic interaction [1]. This mechanism involves two steps. First, the
electrons that carry the magnetic moment interact with the lattice, via electrostatic
crystal field and exchange interactions. Second, the spin-orbit coupling (SOC)
ensures that the spin magnetization actually takes its orientation from the lattice.
In the absence of spin-orbit coupling, anisotropic arrangements of atoms do not
introduce magnetocrystalline
anisotropy. A good example is the Heisenberg model,
H = –ij J R i –R j S i · S j , which is magnetically isotropic even if the exchangebond distribution (Ri – Rj ) is highly anisotropic, for example, in thin films and
nanowires.
Magnetocrystalline anisotropy is not the only contribution. Magnetostatic dipolar
interactions are important in some nanostructured materials and also in materials
where the magnetocrystalline anisotropy is zero by coincidence. Shape anisotropy
(Sect. “Néel’s Pair-Interaction Model”) is a dipole contribution of importance
in some permanent magnets (alnicos), and in magnets with a noncubic crystal
structure, there is also a small dipole contribution to the MCA. The latter is
particularly important in Gd-containing magnets, because Gd3+ ions do not exhibit
anisotropic crystal-field interactions (Sect. “Crystal-Field Theory”) but has a large
dipole moment (S = 7/2).
Magnetic anisotropy is most widely encountered in ferro- and ferrimagnets, but it
is also present in antiferromagnets (Sect. “Magnetoelastic Anisotropy”), disordered
magnets (Sect. “Random Anisotropy in Nanoparticles, Amorphous, and Granular
Magnets”), paramagnets, and diamagnets. An example of an anisotropic diamagnet
is graphite, where the magnitude of the susceptibility is 40 times higher along the
hexagonal c-axis than in the basal plane [2], due to the high mobility of the electrons
in the graphene-like carbon sheets that make up the graphite structure.
106
R. Skomski et al.
Anisotropies in bulk materials and this films are closely related to the orbital
moment L (see below) and therefore to the Bohr-van Leeuwen theorem, which
suggests that a magnetic field acting on electrons does not change the magnetization
of solids. In fact, the field leaves the energy of the electrons unchanged, because the
Lorentz force is perpendicular to the velocity, but nevertheless changes the orbital
moment. Note that the definition and physical interpretation of orbital moments
in solids is a rather recent development, associated with the discoveries of Berry
phase and bulk-boundary correspondence. Berry-phase effects are importantant
in curved and periodic spaces and can be considered as geometrical constantenergy phenomena. Furthermore, magnetization processes reflect the rotation rather
than creation of atomic moments. Neither the concept of spin roatation nor the
Berry phase where known in the early 20th century, when the theorem was
formulated.
Depending on the relative strength of the spin-orbit coupling compared to the
interatomic interactions (crystal field, exchange, hopping), there are two important
limits. The magnetocrystalline anisotropy of high-performance permanent magnets,
such as SmCo5 and Nd2 Fe14 B, is largely provided by the rare-earth 4f electrons
[3, 4]. These electrons are close to the nucleus (R ≈ 0.5 Å), which means that they
exhibit a strong spin-orbit coupling (of the order of 200 meV) but do not exhibit
much interaction with the crystalline environment (of the order of 10 meV). The
orbits of the 4f electrons, as well as the charge distribution n4f (r), are determined
by Hund’s rules. Specifically, Hund’s first rule, which has its origin in the Pauli
principle, states that the total spin S is maximized. The remaining degeneracy with
respect to total orbital moment L is removed by Hund’s second rule, which means
that the orbital moment L is maximized due to intra-atomic exchange. Finally,
Hund’s third rule describes how the spin-orbit interaction rigidly couples orbitalmoment vector L to the spin vector S.
Due to Hund’s third rule, a change in the magnetization angles θ and φ
yields a rigid rotation of the charge distribution n4f (r). Figure 1 explains how this
Fig. 1 Basic mechanism of magnetocrystalline anisotropy, illustrated by an Sm3+ ion (blue) in a
tetragonal environment (yellow). The Sm spin (arrow) is rigidly coupled to the prolate Hund’s-rules
4f charge cloud of the Sm, and the anisotropy energy is equal to the electrostatic interaction energy
between the Sm3+ electrons and the electrostatic crystal field. Due to electrostatic repulsion, the
energy of (a) is lower than (b), and the anisotropy is easy-axis (K1 > 0). If the prolate Sm3+ ion was
replaced by an oblate Nd3+ ion, the situation would be reversed to easy-plane anisotropy (K1 < 0)
3 Anisotropy and Crystal Field
107
rotation translates into magnetocrystalline anisotropy. The electric field created by
neighboring atoms in the crystal (yellow) is weaker than the spin-orbit coupling and
has little effect on the 4f electronic structure, but the electrostatic interaction of the
4f shell with the crystal causes the energy to depend on θ and φ. Atomic crystal-field
charges are normally negative, which amounts to a repulsive interaction between
4f electrons and neighboring atoms. Figure 1 shows an Sm3+ ion with a prolate
charge distribution. For the tetragonal crystal environment shown in the figure, the
electrostatic energy of (a) is lower than (b), corresponding to easy-axis anisotropy
(K1 > 0).
In the opposite limit of the Fe-series transition-metal magnets, the spin-orbit
interaction is much weaker than the interatomic interactions involving 3d electrons.
The 3d orbits are therefore determined by the crystal field and interatomic hopping,
and the spin-orbit coupling is a small perturbation unless the unperturbed energy
levels are accidentally degenerate. In other words, crystal-field interactions determine rare-earth anisotropy but are merely an important starting consideration for
the determination of iron-series transition-metal anisotropies.
This chapter starts with the phenomenology of magnetic anisotropy (Sect. “Phenomenology of Anisotropy”), followed by analyses of crystal-field interactions (Sect. “Crystal-Field Theory”), rare-earth anisotropy (Sect. “Rare-Earth
Anisotropy”), and transition-metal anisotropy (Sect. “Other Anisotropy Mechanisms”). The last section deals with some special topics,
Phenomenology of Anisotropy
Magnetocrystalline anisotropy is usually parameterized in terms of anisotropy
constants, as in Eq. (1). The definition of these constants is somewhat arbitrary,
tailored towards experimental and micromagnetic convenience, and there is some
variation in notation. Another approach is to use anisotropy coefficients of order
2n, obtained by expanding the magnetic energy in spherical harmonics (Appendix
A). For example, the anisotropy coefficients κ 2 0 , κ 4 0 , κ 6 0 roughly correspond
to the uniaxial anisotropy constants K1 , K2 , and K3 , respectively. Anisotropy
coefficients are difficult to access by direct magnetic measurements, but they form
an orthonormal system of functions that do not mix crystal-field contributions of
different orders.
Anisotropy energies per atom, typically 1 meV or less, are much smaller than
the energies responsible for moment formation (about 1000 meV). This means that
the spontaneous magnetization Ms = |M| is essentially fixed and that anisotropy
energies can be expressed in terms of the magnetization angles φ and θ . It is
convenient to choose a coordinate frame where
M = Ms sin θ cos φ ex + sin θ sin φ ey + cos θ ez
(2)
The respective directions x, y, and z with unit vectors ex , ey , and ez often correspond
to the crystallographic a-, b-, and c-axes in crystals of high symmetry.
108
R. Skomski et al.
Lowest-Order Anisotropies
The first- (or second-)order anisotropy constant K1 often provides a good description
of the anisotropy of magnetic substances where higher-order anisotropy constants
are negligible. There are, however, many exceptions to this rule. For example, the
relative magnitudes of the different Kn depend on crystal structure and temperature,
and the φ-dependence of the anisotropy cannot be neglected in many cases. On
the other hand, simplifications arise because many anisotropy constants are zero
by crystal symmetry. Figure 2 shows the corresponding hierarchy: the number of
anisotropy constants decreases in the direction of the arrows.
Fig. 2 Relations between crystal systems; the arrows point in directions of increasing symmetry
and decreasing numbers of anisotropy constants. The arrows also suggest how degeneracies may
arise. For instance, stretching a cubic crystal so that c > a = b creates a tetragonal crystal, whereas
stretching a hexagonal crystal so that a > b creates an orthorhombic crystal
3 Anisotropy and Crystal Field
109
Let us start with the lowest-order anisotropy. Expansion of the magnetic energy
in spherical harmonics shows that there are five second-order terms Y2 m , corresponding to five anisotropy coefficients κ 2 m , namely, κ 2 −2 , κ 2 −1 , κ 2 0 , κ 2 1 , and
κ 2 2 . However, three Euler angles are necessary to fix the anisotropy axes ex , ey ,
and ez relative to the crystal axes, so there remain only two independent anisotropy
constants. Explicitly
Ea (θ, φ)
2
= K1 sin2 θ + K1 sin θ cos 2φ
V
(3)
where K1 is the lowest-order (or second-order) in-plane anisotropy constant; K1
and K1 are generally of comparable magnitude. Equation (3) can be used for any
crystal, but by symmetry, K1 = 0 for trigonal (rhombohedral), hexagonal, and
tetragonal crystals. Only triclinic monoclinic and orthorhombic crystals (red unit
cells in Fig. 2) have K1 = 0. Examples of such low-symmetry compounds are the
monoclinic 3:29 intermetallics [5, 6], such as Nd3 (Fe1-x Tix )29 .
Cubic symmetry is not compatible with these second-order anisotropy contributions, which means that K1 = 0 and K1 = 0. For example, the replacement of the
tetragonal environment in Fig. 1 by a cubic environment would mean that the x-,
y-, and z-directions are all equivalent, which can only be achieved if K1 = K1 = 0
in Eq. (3). However, there exists a differently defined fourth-order “cubic” K1 c that
reproduces Eq. (3) for small angles θ . To describe cubic anisotropy, one must add
spherical harmonics so that the sum does not violate cubic symmetry. Based on
Table 16, there are only two independent terms that satisfy this condition:
K c
Ea
K1 c 2
= 4 x 2 y 2 + y 2 z2 + z2 x 2 + 6 x 2 y 2 z2
V
r
r
(4)
where x/r = cosα x , y/r = cosα y , and z/r = cosα z are the direction cosines of the
magnetization. From the power-law behavior of r in Eq. (4), we see that K1 c and
K1 c are fourth- and sixth-order anisotropy constants, respectively. K1 c > 0 favors
the alignment of the magnetization along the [001] cube edges, which is called irontype anisotropy, while K1 c < 0 corresponds to an alignment along the [98] cube
diagonals and is referred to as nickel-type anisotropy . The subscript “c” is often
omitted, but when uniaxial and cubic anisotropies need to be distinguished, then it
is better to use Ku and/or K1 c to distinguish the respective anisotropy constants.
Anisotropy and Crystal Structure
The number of anisotropy constants rapidly increases with increasing order and
decreasing symmetry . By definition, Ea (−M) = Ea (M), so that we need evenorder spherical harmonics only (gray rows in Table 16). The maximum number of
anisotropy constants is therefore 5 (up to second order), 14 (up to the fourth order),
and 27 (up to the sixth order). Since the anisotropy axes do not necessarily corre-
110
R. Skomski et al.
spond to the crystallographic axes, three of these anisotropy constants effectively
function as Euler angles to fix the orientation of the anisotropy axes. The anisotropy
axes are known for most crystals of interest in magnetism, which reduces the number
of anisotropy constants to 2 (up to second order), 11 (up to the fourth order), and 24
(up to the sixth order). Anisotropy constants of the eighth- and higher-order occur
in itinerant magnets, for example, (Sect. “Transition-Metal Anisotropy”), but they
are usually very small and rarely considered.
The Euler angles must be considered in crystals with low symmetry. Table 1 lists
the crystal systems, point groups, and space groups for some magnetic substances.
Appendix B gives a complete list of all 32 points and 230 space groups. Triclinic
crystals always need three Euler angles to relate the crystallographic a-, b-, and
c-axes to the magnetic x-, y-, and z-axes. In all other noncubic crystals, the caxis is parallel to the z-axis, and one needs at most one Euler angle φ o . This
angle corresponds to a rotation of the crystal around the c-axis, and the identity
cos(β – β o ) = cos (β) cos (β o ) + sin (β) sin (β o ) can be used to get rid of
one in-plane anisotropy constant at the expense of introducing a generally unknown
rotation angle.
As a macroscopic property, magnetic anisotropy is determined by the point
group of the crystal. Most magnetic substances with orthorhombic, tetragonal,
rhombohedral (trigonal), or hexagonal structures belong to the cyclic (C) or dihedral
(D) Schönflies groups. The groups Cn have a single n-fold rotation axis (c-axis),
whereas Cnh and Cnv also have one horizontal and n/2 vertical mirror planes,
respectively. The dihedral groups Dn have an n-fold rotation axis and n/2 additional
twofold rotation axes perpendicular to the c-axis. The cubic crystal system contains
tetrahedral (T) and octahedral (O) Schönflies groups.
Most noncubic crystal structures of interest in magnetism belong to the highly
symmetric Schönflies point groups Cnv , Dn , Dnh , and Dnd , which have φ o = 0. This
includes hexagonal (6 mm, 622, 6/mmm, 62m), trigonal (3 m, 32, 3m), tetragonal
(4 mm, 422, 4/mmm, 42m), and orthorhombic (2 mm, 222, 2/mmm) crystals.
Nonzero values of φ o need to be considered in crystals with point groups Cn ,
Cnh , and Sn . This is the case for all triclinic and monoclinic crystals and for some
hexagonal (6, 6/m, 6), trigonal (3, 3), and tetragonal (4, 4/m, 4) point groups. An
example is the monoclinic 3:29 structure [5], which has the point group C2h .
To elaborate on the role of the point groups, it is instructive to compare Cn and
Cnh with Cnv . Figure 3 shows a top view of a fictitious tetragonal crystal. The
fourfold symmetry axis is clearly visible, and since the horizontal mirror plane
is in the plane of the paper, the figure describes both Cn and Cnh . Some of the
nonmagnetic atoms act as “ligands” (red crosses) and create a crystal field that acts
on the rare-earth ions (blue) and establishes local easy axes (dashed line). The local
3 Anisotropy and Crystal Field
111
Table 1 Crystal systems, point groups, space groups, Strukturbericht notation, and prototype
structures for some compounds of interest in magnetism. Not all the examples are ferromagnetic
Crystal system
Monoclinic
Monoclinic
Orthorhombic
Point group
C2h (2/m)
C2h (2/m)
D2h (mmm)
Space group
C2/m
C2/c
Pnma
Tetragonal
Tetragonal
Tetragonal
D4h (4/mmm)
D4h (4/mmm)
D4h (4/mmm)
P4/mmm
P42 /mnm
I4/mmm
Trigonal
Trigonal
D3 (32)
D3d (3m)
P32 12
R3m
Trigonal
Hexagonal
Hexagonal
D3d (3m)
C6v (6 mm)
D6h (6/mmm)
R3c
P63 mc
P63 /mmc
Hexagonal
Cubic
Cubic
Cubic
Cubic
Cubic
D6h (6/mmm)
T (23)
Th (m3)
Td (43m)
Td (43m)
Oh (m3m)
P6/mmm
P21 3
Pa3
F43m
I43m
Fm3m
Cubic
Cubic
Oh (m3m)
Oh (m3m)
Im3m
Pm3m
Cubic
Cubic
Oh (m3m)
Oh (m3m)
Pn3m
Fd3m
Cubic
Oh (m3m)
Ia3d
Examples
D015 (AlCl3 ): DyCl3
B26 (tenorite): CuO
C37 (Co2 Si): Co2 Si; D011 (cementite):
Fe3 C; goethite: α-FeO(OH); orthorhombic
perovskite: SrRuO3
L10 (CuAu): PtCo, FePd, FePt, MnAl, FeNi
Rutile (C4): CrO2 , MnF2 , TiO2 ; Nd2 Fe14 B
ThMn12 (D2b ): Sm(Fe11 Ti); Al3 Ti (D022 ):
Al3 Dy, GaMn3
D04 (CrCl3 ): CrCl3 (P31 12)
C19 (α-Sm): Sm, NbS2 ; Th2 Zn17 : Sm2 Co17 ,
Sm2 Fe17 N3
D51 (corundum): α-Fe2 O3
B4 (wurtzite): MnSe
A3 (hcp): Co, Gd, Dy; B81 (NiAs): MnBi,
FeS; C7 (MoS2 ): TaFe2 ; C14 (MnZn2
hexagonal laves phase): TaFe2 , Fe2 Mo; C36
(MgNi2 hexagonal laves): ScFe2 ; D019
(Ni3 Sn): Co3 Pt*; PbFe12 O19
(magnetoplumbite): BaFe12 O19 , SrFe12 O19 ;
Th2 Ni17 : Y2 Fe17
D2d (CaCu5 ): SmCo5
FeSi (B20): MnSi, CoSi, CoGe
Pyrite (C2): FeS2
C1b (half-Heusler): MnNiSb
A12 (α-Mn): Mn
A1 (fcc): Ni; B1 (NaCl): CoO, NiO, EuO,
US; D03 (AlFe3 ): Fe3 Si; L21 (cubic
Heusler): AlCu2 Mn; D8a (Th6 Mn23 ):
Dy6 Fe23
A2 (bcc): Fe, Cr
B2 (CsCl): NiAl, FeCo, AlCo, B3
(zincblende): CuCl, MnS, GaAs; E21 (cubic
perovskite): BaTiO3 ; L12 (AuCu3 ): Fe3 Pt
C3 (cuprite): CuO2
C15 (cubic Laves phase): SmFe2 , TbFe2 ,
UFe2 , ZrZn2 ; H11 (spinel): Fe3 O4
Fe3 Al2 Si3 O12 (garnet): Y3 Fe5 O12 ,
Gd3 Fe5 O19
112
R. Skomski et al.
Fig. 3 Top view on a unit
cell of a tetragonal crystal
with C4 or C4h symmetry.
Nonmagnetic ligands (red
crosses) create a crystal field
of low symmetry and local
easy axes (dashed lines) that
are unrelated to the crystal
axes (gray lines). For clarity,
the figure shows only some of
the atoms in the unit cell
crystal field may have very low site symmetry, with local easy axes unrelated to the
a- and b-axes (gray). However, since the local easy axes obey the fourfold rotation
symmetry, the sum of all local anisotropy contributions is fourth-order, and there
is no second-order in-plane contribution. The in-plane anisotropy is of the type
cos(4φ – 4φ o ), and in the present example, the angle φ o ≈ 40◦ is equal to the angle
between the dashed local easy axes and the crystal axes. Going from C4(h) to C4v
introduces vertical mirror planes. These two planes ensure that each local easy axis
of angle φ o has a counterpart with –φ o , so that the net anisotropy directions are now
parallel to the crystal axes.
The picture outlined in Fig. 2 and Table 1 focuses on crystallographic point
groups. A more general approach would be to consider magnetic point groups, as
exemplified by the noncubic (tetragonal) electronic structure of magnets having a
cubic crystal structure and a layered antiferromagnetic spin structure. However, the
layers can lie in any of the equivalent cubic lattice planes, so the magnetic anisotropy
remains cubic.
Tetragonal, Hexagonal, and Trigonal Anisotropies
The anisotropy constants belonging to a given point group can be derived by
applying the symmetry elements of the group to the expansion of the magnetic
energy in terms of spherical harmonics. For example, the fourfold rotation symmetry
of tetragonal magnets, Fig. 3, is compatible with cos4φ terms but not with cos2φ
3 Anisotropy and Crystal Field
113
or cos6φ terms. Up to the sixth order, the anisotropy of magnets with a tetragonal
crystal structure is described by
Ea
= K1 sin2 θ + K2 sin4 θ + K2 sin4 θ cos 4φ + K3 sin6 θ + K3 sin6 θ cos 4φ (5)
V
Without further modification, this equation can be used for the space groups C4v ,
D4 , D4h , and D2d . In particular, all tetragonal compounds listed in Table 1 belong
to the highly symmetric point group D4h . The point groups C4 , C4h , and S4 have
a lower symmetry and require consideration of fourth-order angular shifts φ o . The
anisotropy of orthorhombic crystals differs from Eq. (5) by additional second-order
terms, similar to the K2 term in Eq. (3).
The corresponding anisotropy energy expression for trigonal symmetry is
Ea
2
4
3
V = K1 sin θ + K2 sin θ + K2 sin θ cos θ cos (3 φ)
6
6
+ K3 sin φ + K3 sin θ cos (6φ) + K3 sin3 θ cos3 θ cos (3
φ)
(6)
Without modification, this equation can be used for the trigonal point groups
C3v , D3 , and D3d as well as for hexagonal crystals, which can be considered as
degenerate trigonal crystals (Fig. 2). Hexagonal point symmetry is ensured by
putting K2 = K3 = 0, and no further modification is necessary for the point groups
C6v , D6 , D6h , and D3h . The lower-symmetry point groups C3 , C3h , and S3 (trigonal)
and C6 , C6h , C3h , and S6 (hexagonal) require the consideration of an angular
shift φ o in each φ-dependent term. Note that the relationship between trigonal,
rhombohedral, and hexagonal crystals is complicated. The term rhombohedral
denotes the translational symmetry (Bravais lattice), whereas the closely related
term trigonal refers to the point symmetry. Some trigonal crystals have hexagonal
rather than rhombohedral translation symmetry. The trigonal space groups whose
names begin with P (for primitive) are hexagonal, whereas those starting with R
are rhombohedral. For example, Table 1 shows that α-Sm and Sm2 Co17 belong to
the space group R3m and are both trigonal and rhombohedral. Translationally, the
difference between hexagonal (P) and rhombohedral (R) is similar to the difference
between primitive (P) and body-centered (I) cubic crystals, the rhombohedral cell
having two extra lattice points.
Equations (5) and (6) also describe cubic crystals, which can be considered as
degenerate tetragonal or trigonal crystals (Fig. 2). Stretching a cubic crystal along
the [001]-axis yields a tetragonal crystal, whereas stretching it along the [111] cube
diagonal yields a rhombohedral crystal. The tetragonal symmetry axis (θ = 0) is
therefore parallel to the cubic [001] direction, and the anisotropy constants obey
K1 = K1 c , K2 = − 7K1 c /8 + K2 c /8, K2 = – K1 c /8 – K2 c /8, K3 = – K2 c /8, and K3 = K2 c /8. In the trigonal case, θ = 0 refers to the [111] direction, and the √
in lowestorder anisotropy constants are K1 = − 3K1 c /2, K2 = 7K1 c /12, and K2 = 2K1 c /3.
114
R. Skomski et al.
Higher-Order Anisotropy Effects
Equations (5) and (6) are relatively easy to use in experimental magnetism and
theoretical micromagnetism. However, unlike spherical harmonics, the energy terms
in these equations are nonorthogonal and mix anisotropy contributions of different
orders 2n. For example, uniaxial anisotropy, Eq. (1), has the following presentation
in terms of spherical harmonics:
Ea
V
0 0 = κ22 3cos2 θ − 1 + κ84 35 cos4 θ − 30cos2 θ + 3
0 6
231 cos6 θ − 315 cos4 θ − 105 cos2 θ − 5
+ κ16
(7)
Comparison of Eqs. (1) and (7) shows that K1 contains not only second-order
(κ 2 0 ) but also fourth-order (κ 4 0 ) and sixth-order (κ 6 0 ) contributions. In more detail,
K1 = – 3κ 2 0 /2 – 5κ 4 0 – 21κ 6 0 /2, K2 = 35κ 4 0 /8 + 189κ 6 0 /8, and K3 = −231κ 6 0 /16.
In many cases, the only important anisotropy contribution is K1 = −3κ 2 0 /2,
but in some cases κ 2 0 = 0 and K1 are dominated by fourth-order terms. An
important example is Nd2 Fe14 B (tetragonal) in a narrow temperature range below
room temperature, where κ 4 0 causes the sign of K1 to change (Fig. 14(d) in
Sect. “Temperature Dependence” and Ref. 7).
Anisotropy contributions of the same order tend to have similar magnitudes,
which is important for understanding experimental data. For example, the two uniaxial anisotropy constants K1 and K2 provide a consistent fourth-order description
of hexagonal crystals but not of tetragonal crystals, because the non-uniaxial K2 term in Eq. (5) is also of the fourth order. For second-order uniaxial anisotropies,
see Sect. “Lowest-Order Anisotropies”.
Higher-order anisotropy constants may have drastic effects if K1 ≈ 0 by
coincidence, for example, due to competing sublattice contributions. For example,
uniaxial anisotropy with K1 < 0 and K2 > − K1 /2 yields easy-cone magnetism,
where the negative K1 makes the c-axis an unstable magnetization direction but the
positive K2 prevents the magnetization from reaching the basal plane (a-b-plane). In
this regime, the preferred magnetization direction lies on a cone around the c-axis,
described by the angle θ c = arcsin (|K1 |/2K2 ). The temperature dependences of
K1 and K2 are generally very different; K2 usually negligible at high temperatures.
As a consequence, the preferential magnetization direction may change as a
function of temperature, which is known as a spin-reorientation transition. A similar
film thickness-dependent transition is observed in films where surface and bulk
anisotropy contributions compete.
The ratio K1 /μo Ms has the dimension of a magnetic field, which makes it
possible to compare anisotropies with applied magnetic fields and coercivities. It is
customary to define the corresponding anisotropy field of K1 -only uniaxial magnets
as
Ha =
2K1
μo Ms
(8)
3 Anisotropy and Crystal Field
Table 2 First- and
second-order anisotropy
constants at room temperature
[9–11]
115
Substance
Fe
Ni
Co
Fe3 O4
Nd2 Fe14 B
Sm2 Fe17 N3
Sm2 Fe17 C3
YCo5
Y2 Co17
Tm2 Co17
Sm2 Fe14 B
K1 (MJ/m3 )
0.048
−0.005
0.53
−0.011
4.9
8.6
7.4
5.8
4.0
1.6
−12.0
K2 (MJ/m3 )
0.015
0.005
0
0.028
0.65
1.46
0.74
−0.3
0.3
0.2
−0.29
Structure
Cubic
Cubic
Hexagonal
Cubic
Tetragonal
Rhombohedral
Rhombohedral
Hexagonal
Hexagonal
Hexagonal
Tetragonal
The anisotropy field is defined in a formal way and does not actually exist inside a
magnet; it is equal to the external field that creates a certain effect on the magnet.
Subject to shape anisotropy corrections (Sect. “Magnetostatic Anisotropy”), the
anisotropy field establishes an upper limit to the coercivity Hc . In practice, Hc Ha ,
which is known as Brown’s paradox. An approximate relation is Hc = α Ha , where
α 1 is the Kronmüller factor [8, 9].
The inclusion of higher-order anisotropies gives rise to different nonequivalent
anisotropy field definitions. For example, using Eq. (1) and comparing the energies
for θ = 0 and θ = 90◦ lead to Ha = 2(K1 + K2 + K3 )/µo Ms . The initial slope of the
perpendicular magnetization curves yields the same Ha , whereas the nucleation field
of uniaxial magnets is not affected by K2 and K3 , so that Eq. (8) remains valid for
uniaxial magnets of arbitrary order. In cubic magnets, the anisotropy fields for irontype anisotropy (K1 > 0) are described by Eq. (8), whereas nickel-type anisotropy
(K1 < 0) yields Ha = − 4 K1 /3μo Ms (Table 2).
Anisotropy Measurements
Sucksmith-Thompson method. The experimental determination of magnetic
anisotropy is easiest if single crystals or c-axis-aligned single-crystalline powders or
thin films are available. The Sucksmith-Thompson method uses a magnetic field H
perpendicular to the c-axis and measures the magnetization M in the field direction
[12]. Starting from Eq. (1), ignoring K3 and adding the Zeeman energy yield the
energy density η(M/Ms ):
η = K1
M2
Ms 2
+ K2
M4
Ms 4
− μo MH
(9)
where M = Ms sinθ . Minimizing the energy, ∂η/∂M = 0, and dividing the result by
μo M yields
116
R. Skomski et al.
2 K1
H
4 K2
=
+
M2
M
μo Ms 2
μo Ms 4
(10)
Plotting H/M as a function of M2 yields K1 and K2 from the intercept and slope of
the straight line, respectively.
Approach to saturation. Samples are often polycrystalline. In the ideal case
of noninteracting grains with second-order uniaxial anisotropy, the corresponding
random-anisotropy problem can be solved explicitly. The approach to saturation
obeys
M(H ) = Ms
Ha 2
1−
15H 2
(11)
In practice, this method requires the fitting of the three parameters: Ms , the
sought-for Ha = 2 K1 /μo Ms , and a high-field susceptibility that must be used
to ensure that ∂M/∂H = 0 for H = ∞. Note that Eq. (11) does not predict
the sign of K1 , because both easy-axis and easy-plane ensembles yield the same
asymptotic behavior. Note that Eq. (11) is essentially a random-anisotropy relation
(Sect. “Random Anisotropy in Nanoparticles, Amorphous, and Granular Magnets”).
Torque magnetometry. A single-crystalline magnetic sample experiences a
mechanical torque –∂Ea /∂α, where α is a magnetization angle relative to the crystal
axes. The angle α is varied with the help of a rotating magnetic field, and the torque
is monitored as a function of the field direction, for example, by measuring the
twisting angle of a filament to which the sample is attached. The interpretation
of the torque depends on the crystalline orientation of the sample, but if the
torque axis is parallel to a magnetocrystalline symmetry axis, the corresponding
anisotropy constants are readily obtained as Fourier components of the torque
curves [13].
Magnetic circular dichroism. Single-ion anisotropy is closely related to the
orbital moment and approximately proportional to the latter in iron-series transitionmetal magnets (Sect. “Perturbation Theory”). A direct way to probe orbital (and
spin) moments on an atomic scale is X-ray magnetic circular dichroism (XMCD).
Circular dichroism means that circularly polarized photons pass through the sample
and that the absorption is different for left- and right-polarized light [14–16]. This is
because the orbital moment reflects atomic-scale circular currents that interact with
light. Furthermore, due to spin-orbit coupling, the light also interacts with spin, so
that XMCD can also be used to simultaneously measure the spin moment.
Crystal-Field Theory
Electrons in solids occupy states reminiscent of atomic orbitals, even in metals.
This applies, in particular, to the partially filled inner shells of transition-metal
elements, such as the iron-series 3d shells and rare-earth 4f shells. The electrons in
3 Anisotropy and Crystal Field
117
Fig. 4 Angular dependence of 3d wave orbitals: (a) real eigenfunctions and (b) top view on a
mixture of states constructed from m > ∼ exp (ιmφ) with m = ±2. Red and yellow areas in
(a) indicate regions of positive and negative wave functions ψ, respectively, and the darkness
in (b) indicates the electron density ψ*ψ. The wave functions shown in this figure are all
eigenfunctions of the free atoms, but in solids (b), the crystal field, symbolized by ligands (black
dots), favors real wave functions (top), whereas spin-orbit coupling favors complex wave functions
| ± m > (bottom). Details of this “quenching” behavior will be discussed in Sect. “Spin-Orbit
Coupling and Quenching”
the inner shells, which often carry a magnetic moment, interact with the crystalline
environment. The crystal-field (CF) interaction of the Sm3+ ion in Fig. 4 is
one example, but a similar picture is realized in 3d ions, especially in oxides.
Itinerant magnets, such as 3d metals, require additional considerations, because their
electronic structure is largely determined by interatomic hopping (band formation).
Crystal-field theory had its origin in the study of transition-metal complexes
in the last decade of the nineteenth century [17]. An example was the distinction between violet and green [Co(NH3 )6 ]3+ Cl3 3− , which indicates energy-level
differences of stereochemical origin. The quantitative crystal-field theory dates
back to Bethe [18], who treated the atoms as electrostatic point charges. Since
then, the crystal-field theory has been extended to include quantum mechanical bonding effects in a generalization are known as ligand-field (LF) theory
[19]. As emphasized by Ballhausen [17], the latter is quantitatively superior
to Bethe’s CF theory but leaves the main conclusions of the latter unchanged.
In practice, the terms are often used interchangeably: the atoms surrounding a
magnetic ion are called ligands in both complexes and solids, and the term ligand
field is sometimes used. Physically, both electrostatic and hybridization effects
contribute to the crystal field (ligand field), even in oxides. The focus of this
section is on the traditional electrostatic crystal-field theory, but some interatomic
118
R. Skomski et al.
hybridization effects will be discussed in the context of itinerant anisotropy
(Sect. “Transition-Metal Anisotropy”).
One-Electron Crystal-Field Splitting
The wave functions and charge distributions of the electrons are obtained from the
Schrödinger equation. Hydrogen-like 3d wave functions are listed in Appendix C.
The angular parts of the wave functions follow from the spherical character of the
intra-atomic potential and are the same for Fe-series 3d, Pd-series 4d, and Pt-series
5d electrons. However, the radial parts differ for the three series, and they also
depend on non-hydrogen-like details of the atomic potentials. Figure 4 shows the
angular distribution of the five 3d orbitals ψ μ (r). In a free atom, the five orbitals
are degenerate, but in solids and molecules, they undergo crystal-field interactions
described by the Hamiltonian:
HCF =
V (r) n(r) dV
(12)
where V (r) is the crystal or ligand-field potential and n(r) = ψ ∗ (r)ψ(r) refers to
the d or f orbital(s) in question.
To understand crystal-field effects, it is necessary to consider the shape of the
orbitals. Atomic wave functions and charge distributions such as those shown in
Figs. 4 and 1, respectively, have characteristic prolate, spherical, or oblate shapes.
The larger the magnitude of the quantum number m = lz , the more oblate or flatter
the orbitals, as we can in Fig. 4. This is because large orbital moments, m = ± 2
in Fig. 4 , correspond to a pronounced circular electron motion in the plane perpendicular to the quantization axis (z-axis). By contrast, the prolate |z2 > orbital, which
has m = 0 has its electron cloud close to the z-axis. In a crystalline environment, the
different orbital shapes correspond to different electrostatic interactions. Crystalfield charges are negative [20], so that the interaction between the 3d or 4f electronic
charge clouds and those of the surrounding atoms is repulsive. As a consequence,
the prolate |z2 > orbital prefers to point in interstitial directions between the atomic
neighbors, rather than towards them. The opposite is true for the oblate orbitals with
m = ± 2.
The electrostatic repulsion between the 3d electrons and those of the neighboring
atoms removes the degeneracy of the five 3d levels and yields the famous eg -t2g
splitting in an environment with cubic symmetry. Figure 5(a) shows the |z2 > orbitals
in a cubal environment where the central atom is coordinated by 8 neighbors.
The |z2 > orbital points in an electrostatically favorable direction and has a very
low energy. The charge distribution of the |x2 -y2 > orbital also points in directions
away from the neighboring atoms or ligands, and it can be shown that the
|x2 -y2 > and |z2 > have the same energy, forming a so-called eg doublet. The |xy>,
|xz>, |yz> orbitals are equivalent by symmetry and form a t2g triplet. The charge
3 Anisotropy and Crystal Field
119
Fig. 5 A 3d orbital (z2 ) in
some crystalline
environments: (a) cubal, (b)
octahedral, (c) tetrahedral,
and (d) tetragonally distorted
cubal. Note that (a), (b), and
(c) have cubic symmetry,
whereas (d) is tetragonal
distributions of the triplet orbitals are closer to the ligands, so that the triplet energy
is higher than the doublet energy.
The opposite splitting is realized in an octahedral environment, Fig. 5(b), where
the central atoms are coordinated by six ligand atoms. In this environment, the |x2 y2 > and |z2 > orbitals point directly towards the neighboring atoms, whereas the |xy>,
|xz>, |yz> orbitals point in interstitial directions. The tetrahedral environment (c)
has no inversion symmetry but is otherwise very similar to the cubal environment.
Basically, the cubic e-t2 crystal-field splitting is reduced by a factor 2, because
there are only four neighbors. Symmetries lower than cubic partially or completely
remove the eg and t2g degeneracies. Figure 5(d) illustrates this for a tetragonally
distorted cubal environment. Compared to (a), the ligands move towards the basal
plane, which lowers the energy of the |z2 > orbital but raises that of the |x2 y2 > orbital. As a consequence, these states no longer form a doublet. Similarly,
the |xz> and |yz> orbitals become somewhat more favorable compared to the
|xy > orbital, because their charge distribution has a substantial out-of-plane
component. This splits the t2g triplet, but |xz> and |yz> remain degenerate, because
the x and y directions are equivalent in a tetragonal crystal.
Figure 6 summarizes the eg -t2g splitting and the evolution of the levels due
to a symmetry-breaking tetragonal distortion. It is important to note that halffilled (and full) 3d shells have spherical charge distributions and do not interact
with anisotropic crystal fields. Equivalently, the CF interaction leaves the center
of gravity of the 3d levels unchanged. This can be used to gain some quantitative
information about the level splitting. For example, the eg -t2g splitting, also known
120
R. Skomski et al.
Fig. 6 Crystal-field splitting of 3delectrons in cubic and tetragonal environments
as 10Dq, consists of an energy shift of +6Dq for the doublet and a –4Dq shift for
the triplet.
Table 3 lists the crystal-field splittings for the most symmetric point groups in
each crystal system and for axial symmetry. The levels are described by Mullikan
symmetry labels, using t and e for triplets and doublets, respectively [24]. Singlets
are denoted by a or b, depending on whether the reference axis is an n-fold rotation
axis (a) or not (b). The subscripts 1, 2, and 3 indicate C2 symmetries around crystal
axes, and primes ( ) and double-primes ( ) refer to horizontal mirror symmetry
and antisymmetry, respectively. The subscript g (German gerade “even”) denotes
inversion symmetry, which exists for the cubal coordination, Fig. 5(a), but not for
the otherwise very similar tetrahedral coordination, Fig. 5(c). However, the inversion
symmetry of the 3d wave functions means that there are no levels with subscript u
(German ungerade “odd”), so that no confusion arises by dropping the subscript
g [25]. In each crystal system, the complexity of the symmetry labels decreases
with decreasing symmetry. For example, the eg -t2g splitting is limited to the highly
symmetric point group Oh : the respective cubic compounds FeS2 (space group
Th ), MnNiSb (space group Td ), and FeSi (space group T) have eg -tg , e-t2 , and
e-t splittings.
3 Anisotropy and Crystal Field
121
Table 3 Crystal-field splittings of 3d electrons. The colors indicate the crystal-field multiplet
structure: one doublet and one triplet (red), one singlet and two doublets (yellow), three singlets
and one doublet (green), and five singlets (blue). The listed point groups are the most symmetric
ones in each crystal system – less symmetric point groups yield modified symbols, such as missing
subscripts g. In linear molecules (point groups D∞h and C∞v ), the multiplets a1 , e1 , and e2 are
also known as + , , and , respectively
122
R. Skomski et al.
Crystal-Field Expansion
It is convenient to expand the crystal-field potential V (r) into spherical harmonics
Yl m (θ , φ). The corresponding expansion coefficients Al m are known as crystalfield parameters and play an important role in crystal-field theory and magnetism.
Treating the ligands (i = 1 ... N) as electrostatic point charges [18] located at Ri
yields the crystal-field potential energy:
V (r) = −
e
4πεo
N
i=1
qi
| Ri − r |
(13)
This sum is easily converted into a sum of spherical harmonics by exploiting the
identity:
4 π rl
l (2 l + 1) R l+1
1
=
|R−r |
|m|<l
Y1 m∗ (Θ, Φ) Y1 m (θ, φ)
(14)
so long as R > r. Strictly speaking, the l-summation extends from zero to infinity,
but the symmetry of n(r) in Eq. (12) means that the only relevant terms are l = 2, 4
(d-orbitals) and l = 2, 4, 6 (f -orbitals).
Inserting Eq. (14) into Eq. (13) and summing over all ligands leads to the
cancellation of Yl m (θ , φ) terms that are incompatible with the symmetry of the
crystal. For example, cubic crystals have
V (r) = 20A4 0 x 4 + y 4 + z4 − 3r 4 /5
(15a)
where the dimensionless crystal-field parameter 4πεo R5 A4 0 /qe is equal to −7/16,
7/18, and 7/36 for the octahedral, cubal, and tetrahedral ligands of Fig. 5, respectively. The r4 term in Eq. (15a) is isotropic and not necessary for the description
of magnetic anisotropy, but it ensures that the center of gravity of the energy is
conserved during crystal-field splitting. Since x2 + y2 + z2 = r2 , Eq. (15a) is
equivalent to
V (r) = −40A4 0 x 2 y 2 + y 2 z2 + z2 x 2 − r 4 /5
(15b)
and to any linear combination of Eqs. (15a) and (15b). The structure of this equation
mirrors that of Eq. (4) for the anisotropy of cubic magnets. A third version of Eq.
(15) will be discussed in the context of operator equivalents.
Uniaxial crystal fields are described by
V (r) = A2 0 3 z2 − r 2 + A4 0 35 z4 − 35 z2r 2 + 3 r 4
+A6 0 231 z6 − 315 z4 r 2 + 105 z2 r 4 − 5 r 4
(16)
3 Anisotropy and Crystal Field
Table 4 Crystal-field
parameters for some noncubic
rare-earth transition-metal
intermetallics [10]
123
Compound
R2 Fe14 B
R2 Fe17
R2 Fe17 N3
A2 0
K/ao 2
300
34
−358
A4 0
K/ao 4
−13
−3
−39
From Eq. (14) we see that the small parameter in the ligand-field expansion is
r/R, that is, the ratio of d-shell radius to interatomic distance. For this reason,
A4 0 is typically smaller than A2 0 by a factor of order (r/R)2 , or about one order
of magnitude. Exceptions are, for example, weakly distorted cubic structures.
Another way of interpreting crystal fields is to expand V (r) into a Taylor series
with respect to x, y, and z. The nonzero expansion coefficients are the crystal-field
parameters Al m , where l denotes the l-th spatial derivative of V (r). In particular,
A2 0 ∼ ∂ 2 V (r) /∂z2 or, in terms of the electric field, A2 0 ∼ ∂Ez /∂z. This means
that A2 0 is essentially an electric field gradient .
The point-charge model accurately describes the symmetry of the crystal field
[20] and yields semiquantitatively correct numerical predictions for a variety of
systems. It was originally developed for insulators but also approximates rare-earth
ions in metals where the electrostatic interaction is screened by conduction electrons
[21]. This surprisingly broad applicability has its origin in the superposition principle of crystal-field interactions, which states that the effects of different ligand atoms
are additive in very good approximation [20]. Experimentally, crystal-field effects
are measured most directly by spectroscopy, for example, optical spectroscopy or
inelastic neutron scattering, but there are also indirect measurements, such as rareearth anisotropy measurements (Table 4).
Many-Electron Ions
A fixed number n of inner-shell electrons of an ion is called a configuration, such
as 3dn and 4fn . In practice, the configuration corresponds to the ions’ charge state.
All rare-earth elements form tripositive ions, R3+ , as exemplified by Sm3+ (4f5 )
and Dy3+ (4f9 ). Some form R2+ shells such as europium in EuO or R4+ in mixedvalence and heavy-fermion compounds such as CeAl3 [22, 23]. Transition-metal
ions show a greater variety, most commonly T2+ , T3+ , and T4+ , where the ionic
charge is determined by chemical considerations. For example, Fe3 O4 contains both
Fe2+ (3d6 ) and Fe3+ (3d5 ) ions to charge-compensate the O2− anions.
The n electrons are distributed over the available 2 × (2 l + 1) one-electron states
and labeled by sz = ±1/2 and lz = −l, ..., l – 1, l. The relationship between these
electrons is largely governed by the Pauli principle, by Hund’s-rules for electronelectron interactions, and by spin-orbit coupling. The Pauli principle means that
each real-space d or f orbital can accommodate at most one↑ and one↓ electron.
Subject to the Pauli principle, there are several ways to place n electrons onto the 10
124
R. Skomski et al.
one-electron 3d levels, each combination corresponding to a many-electron state.
These can be divided into terms characterized by well-defined total spin S = i
si and orbital quantum numbers L = i li (i = 1 ... n), with each term containing
(2 S + 1) (2 L + 1) states. The terms are usually denoted by 2S + 1 L, where 2S + 1
is the spin multiplicity and L is denoted by as S (L = 0), P (L = 1), D (L = 2), F
(L = 3), G (L = 4), H (L = 5), and I (L = 6). More generally, it is common to use
capital letters for ionic properties, and S, P, D, F are analogous to one-electron states
s, p, d, and f.
An example is the 3d2 configuration, realized, for example, in Ti2+ . The first
electron can occupy any of the 2 × 5 states, leaving nine states for the second
electron. This yields 90/2 = 45 permutations, each corresponding to a two-electron
state. The highest L is achieved by placing two electrons in the lz = 2 state, (↑↓,
−, −, −, −). This yields S = 0 and L = 4, that is, a1 G term containing 9 states.
The wave function (↑, ↑, −, −, −) has S = 1 and L = 3 and therefore belongs to
a3 F term, which contains 21 states. The other 3d2 terms are 1 D (5 states) and 3 P (9
states), and 1 S (1 state). Similar term analyses can be made for all configurations
[17, 24, 26] but will not be discussed here, because in magnetism our main interest
is the ground-state term. A trivial case is 3d1 , which corresponds to a single
term 2 D.
As far as symmetry is concerned, the crystal-field splittings of ions are equal to
those of the one-electron states [17]. For example, the octahedral splitting eg -t2g for
a single d electron corresponds to Eg -T2g in D ions. Table 5 shows basic the CF
splittings of many-electron terms in cubic, tetragonal, and trigonal environments.
The subscript-free symmetry labels A (singlet), B (singlet), E (doublet), and T
(triplet) are of the lowest-symmetry type, and the numbers indicate two or more
distinct levels. Note that most point groups have subscripts (1, 2, g, u) that are
important in spectroscopy but not for the explanation of magnetic anisotropy.
Without interactions, the terms of a configuration would be degenerate. In reality,
the degeneracy is removed by the electron-electron interaction:
1
U=
4πεo
ρ (r) ρ r dV dV | r − r |
(17)
Table 5 Basic branching table for crystal-field splittings of many-electron ions. Both groundstate and excited terms are included, and the table is not restricted to d electrons. For example, the
free-ion triplet of a single p electron (P) remains unaffected by a cubic crystal field but exhibits a
singlet-doublet splitting in tetragonal and trigonal crystals
Term
S
P
D
F
G
H
Cubic CF
A
T
E+T
A+2T
A+E+2T
E+3T
Tetragonal CF
A
A+E
A + 2B + E
A + 2B + 2E
3A + 2B + 2E
3A + 2B + 3E
Trigonal CF
A
A+E
A + 2E
3A + 2E
3A + 3E
3A + 4E
3 Anisotropy and Crystal Field
125
where ρ(r) is the electron charge density. The corresponding term splittings are
large, 1.8 eV for Co2+ , and often dominate the behavior of the ion. The term
energies E(L, S) can be calculated in a straightforward way, by applying the lowestorder perturbation theory to Eq. (17), E(L, S)=< (L,S) | U | (L, S) >[17, 27].
However, the ground-state term is more easily obtained from Hund’s rules. The
first rule states that the total spin S= i si is maximized. In the above 3d2 example,
there are two terms with maximum S, namely, 3 F and 3 P, both having S=1. Hund’s
second rule acts as a tiebreaker, by favoring large L= i li . Since F and P mean L=3
and L=1, respectively, 3 F is the ground-state term of the 3d2 configuration. Table 6
shows some basic properties of 3d ions; 4f ions will be discussed in the context of
rare-earth anisotropy (Sect. Crystal-Field Theory).
Hund’s first rule yields another simplification: in the ground-state term of the
3d5 configuration, there are five ↑ electrons which occupy the five available orbitals,
lz = −2, ... +1, +2. This yields L = lz = 0, meaning that empty, half-filled, and
completely filled 3d shells are all S-type ions. This principle carries over to magnetic
anisotropy: from Table 5 we see that S states do not undergo crystal-field splitting
but remain in their highly symmetric degenerate A state. The corresponding charge
distribution is spherical, and the ion does not contribute to the magnetocrystalline
anisotropy (except via admixture with a higher excited state). Figure 7 shows the
level splittings of the ground-state terms of the 3d ions in an octahedral crystal field.
Note the half-shell symmetry of the splittings: aside from the sign, there are only
two nontrivial cases, namely, one electron or hole (d1 , d4 , d6 , d9 ) and two electrons
or holes (d2 , d3 , d7 , d8 ).
The crystal-field interaction is normally weaker than the intra-atomic exchange.
However, very strong crystal fields may negate Hund’s rules and cause a transition to
a low-spin state. For example, octahedrally coordinated Fe2+ has the configuration
3d6 , and, according to Fig. 7, a T2g ground state, that is, t2g (↑↑↑↓)-eg (↑↑). The
two ↑ electrons in the eg -doublet experience a competition between electronelectron interaction, which favors parallel spin alignment, and the CF, which
favors t2g occupancy. In very strong crystal fields, the electronic structure becomes
t2g (↑↑↑↓↓↓)-eg (empty), and the ion loses its magnetic moment. This is an example
of a high-spin low-spin transition. Aside from d6 , the three ions d4 , d5 , and d7
Table 6 Electronic
configurations of 3d ions. The
listed terms are the
ground-state terms
Ion
3d1
3d2
3d3
3d4
3d5
3d6
3d7
3d8
3d9
Example
Ti3+ , V4+
Ti2+ , V3+
V2+ , Cr3+
Cr2+ , Mn3+
Mn2+ , Fe3+
Fe2+ , Co3+
Co2+ , Ni3+
Ni2+ , Pd2+
Cu2+
Term
2D
3F
4F
5D
6S
5D
4F
3F
2D
L
2
3
3
2
0
2
3
3
2
S
1/2
1
3/2
2
5/2
2
3/2
1
1/2
126
R. Skomski et al.
Fig. 7 Crystal-field splittings
of the ground-state terms of
3d ions in a weak octahedral
crystal field. The energy unit
Dq is one tenth of the eg -t2g
splitting
undergo a high spin low spin in strong octahedral crystal fields, leading to spin
moments of 2 μB (d4 ) and 1 μB (d5 , d7 ).
It is instructive to plot the term energies as a function of the crystal field,
using the eg -t2g splitting 10Dq to quantify the crystal field in an Orgel diagram.
An extension of the Orgel diagram is the Tanabe-Sugano diagram , where both
the crystal field (Dq) and the term energies are divided by the Racah parameter
B [24]. This parameter links Hund’s second rule, namely, the maximization
of L, to the underlying intra-atomic electron-electron interaction and satisfies
E(3 P) – E(3 F) = 15B. The ground-state energy is used as the energy zero, which
helps to visualize transitions. Figure 8 shows a Tanabe-Sugano diagram where the
ground-state term changes from high spin 5 T2g (blue line) to low spin 1 A1g (red
line).
Any splitting ± E of a degenerate state lowers the energy by about E if the
level is only partially occupied. For example, the tetragonal lattice distortion of
Fig. 6 means that the eg doublet splits into a low-lying a1g state and a b1g state
and a single electron in the eg doublet moves to the a1g level. The resulting crystalfield energy gain competes against the mechanical energy necessary to tetragonally
distort the crystal. However, the former is linear in strain ε, whereas the latter is
quadratic, so that the CF should always create a small distortion. This is known as
the Jahn-Teller effect.
3 Anisotropy and Crystal Field
127
Fig. 8 Tanabe-Sugano diagram for a 3d6 ion [24]. The energy unit Dq is one tenth of the eg -t2g
splitting and B is the Racah parameter [28]. The vertical line indicates a transition from a high-spin
state (blue) to a low-spin state (red)
Spin-Orbit Coupling and Quenching
The interatomic interactions (U ) remove the degeneracy between different terms
and create ions with well-defined L and S. However, L and S do not interact
and can point in any direction. In reality, they are subject to relativistic spinorbit coupling, which causes the terms to split into multiplets of well-defined total
angular momentum J, denoted by 2S + 1 LJ . Figure 9 illustrates the origin of spinorbit coupling: the orbital motion of the electron (L) creates a magnetic field that
couples to the electron’s own spin (S). This coupling is important for both isotropic
magnetism (moment formation, and exchange) and magnetic anisotropy. The key
role of spin-orbit coupling in the explanation of magnetic anisotropy was first
recognized and exploited by Bloch and Gentile in 1931 [1].
The quasiclassical model of Fig. 9 correctly reproduces the order of magnitude
of the spin-orbit coupling, aside from a factor 1/2 (Thomas correction). The spinorbit coupling may be derived directly from the relativistic Dirac wave equation.
The coupling is a fourth-order term in the Pauli expansion of the relativistic energy,
similar to the v4 term in the equation:
me c 2
1+
v2
1
1
= me c 2 + me v 2 − me v 4
2
2
8
c
(18)
128
R. Skomski et al.
Fig. 9 Spin-orbit coupling in a free ion (schematic). The orbiting spin acts like a current loop and
creates a magnetic field that acts on the spin. The nucleus does not actively participate in the spinorbit coupling but merely serves to curve the trajectory of the electron: a circular racetrack would
do equally well
Electromagnetic effects are added by including scalar and vector potentials
[10, 29]. The result of the calculation is the SOC energy [29]:
Hso =
3
s · ∇V × k
2
2
2me c
(19)
This equation shows that the spin-orbit coupling favors a spin direction perpendicular to both potential gradient and direction of motion. For example, electrons
in thin films experience a Rashba effect [30], and there is a small interstitial
contribution to the magnetocrystalline anisotropy [31]. The Rashba effect means
that electrons of wave vector k move in the film plane and experience a potential
gradient perpendicular to the film, which naturally occurs due to broken inversion
symmetry at thin-film surfaces and interfaces. According to Eq. (19), the spin then
prefers to lie in the plane, in one direction perpendicular to k. In the opposite inplane spin direction, the energy is enhanced, which is referred to as Rashba splitting
of the electron levels.
The potential gradient is most pronounced near the atomic nuclei, and for
hydrogen-like 1/r potentials
Hso =
Ze2
2
l·s
2
2
2me c 4πεo r 3
(20)
Using Appendix C, we can evaluate the average <1/r3 > and obtain Hso = ξ l · s.
Here ξ is the spin-orbit coupling constant:
ξ=
2
Z 4 e2
1
2
3
2
2me c ao 4πεo n3 l l + 1/ (l + 1)
2
(21)
3 Anisotropy and Crystal Field
129
It is instructive to discuss relativistic phenomena in terms of Sommerfeld’s finestructure constant, α = e2 /4πεo c ≈ 1/137. Electrons in atoms and solids have
velocities of the order of v = αc, so that from Eq. (18):
me c 2 1 +
v2
1
1
= me c 2 + me α 2 c 2 − me α 4 c 4
2
8
c2
(22)
Similarly, Eq. (21) becomes
ξ=
1
me 4 4 2
Z α c +
1
2
3
n l l + /2 (l + 1)
(23)
This equation captures the relativistic nature of spin-orbit coupling and magnetic
anisotropy. In terms of powers of α, ξ is a small relativistic correction, similar to
the v4 term in Eqs. (18) and (22), but Z, which is largest for inner-shell electrons in
heavy elements, greatly enhances the effect in partially filled shells. Tables 7 and 8
show values of spin-orbit coupling constants ξ for 3d, 4d, 5d, 4f, and 5f elements,
obtained from Hartree-Fock calculations [32, 33]. A comparison of experimental
data and theoretical predictions indicates that these tables have an accuracy of the
order of 10% [32–34]. Furthermore, ξ somewhat increases with ionicity [34]: going
from T2+ to T3+ and T+ , respectively, changes the SOC constant of late 3d elements
by about ±10%.
There are two limits for many-electron spin-orbit coupling. Russell-Saunders
coupling means that the ion has well-defined values of L = i li and S = i si .
They are good quantum numbers, and the SOC is a weak perturbation. This limit
is realized when the intra-atomic interactions are stronger than ξ . In the opposite
limit of j-j coupling, the i-th electron first experiences a one-electron SOC so that
J = i (li + si ). Most solid-state magnetism involves Russell-Saunders coupling, but
j-j coupling is important in two limits: (a) low-lying levels of very heavy elements,
such actinides, and (b) excited levels of most elements, except very light ones. In (a),
the j-j coupling is imposed by the large λ in heavy atoms, whereas in (b), it reflects
the increased electron separation in excited states. In many cases, Russell-Saunders
Table 7 Spin-orbit coupling
constants for electrons in the
partially filled dipositive 3d,
4d, and 5d transition-metal
ions. ξ of the inner 1s, 2s, and
2p electrons in heavy
elements is much stronger
than the values in this table,
but closed shells do not
exhibit a net spin-orbit
coupling
d1
d2
d3
d4
d5
d6
d7
d8
d9
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
ξ (meV)
10
15
22
31
41
53
68
86
106
Y
Zr
Nb
Mo
Tc
Ru
Rh
Pd
Ag
ξ (meV)
32
48
65
84
106
129
156
186
221
La
Hf
Ta
W
Re
Os
Ir
Pt
Au
ξ (meV)
69
196
244
302
360
419
485
556
633
130
Table 8 Spin-orbit coupling
constants ξ for tripositive 4f
and 5f transition-metal ions
[32, 33]
R. Skomski et al.
f1
f2
f3
f4
f5
f6
f7
f8
f9
f10
f11
f12
f13
f14
Ce
Pr
Nd
Pm
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu
ξ (meV)
85
102
120
139
159
182
205
230
257
286
318
352
422
143
Th
Pa
U
Np
Pu
Am
Cm
Bk
Cf
Es
Fm
Md
No
Lw
ξ (meV)
197
234
271
308
348
390
432
478
525
576
629
686
–
–
coupling yields the correct multiplet structure but j-j coupling causes quantitative
deviations in the level spacing. For example, the j-j coupling effect on the 3 P2 -3 P0
splitting is negligible in C, about 20% in Si, and dominates in Ge, Sn, and Pb [26].
Magnetic anisotropy reflects low-lying excitations, and Russell-Saunders coupling
therefore applies to both transition metals and rare earths.
The Russell-Saunders coupling establishes the vector model, where J = L + S.
Using the Hund’s-rules ground-state terms to evaluate i ξ i li · si yields the ionic
spin-orbit coupling Λ L·S, where Λ = ± ξ /2S for less and more than half-filled
shells, respectively [17]. The change of sign at half filling yields Hund’s third rule:
for the early elements in each series, J = L – S, and for the late elements, J = L + S.
Each multiplet has 2 J + 1 Zeeman-like intramultiplet levels, Jz = − J, ...,
(J – 1), J, and the degeneracy of these levels is removed by a magnetic field or by the
crystal field. Due to the g-factor of the electron, a magnetic field couples to (L + 2S)
rather than to (L + S). This makes it necessary to project L + 2S onto J, so that
(L + 2S) · J = g J2 . The Landé g-factor of the ion g = 1 for pure orbital magnetism
(L = J) and g = 2 for pure spin magnetism (L = 0). For arbitrary L and S, less and
more than half-filled shells exhibit g = 1 – S/(J + 1) and g = 1 + S/J, respectively.
The exchange between magnetic ions involves spin only, which mandates the use of
the projection S · J = (g − 1) J2 and yields de Gennes factor G = (g – 1)2 J(J + 1).
This makes it possible to write the exchange interaction as J = G Jo , where Jo is a
J-independent Heisenberg exchange constant.
Spectroscopic and magnetic measurements indicate that Hund’s rules are well
satisfied in rare-earth ions (Sect. “Crystal-Field Theory”), but iron-series transitionmetal ions systematically violate them, especially the third rule. For example,
g ≈ 2 for iron-series atoms in almost all metallic and nonmetallic crystalline
environments. In other words, the magnetic moments of Fe, Co, and Ni originate
3 Anisotropy and Crystal Field
131
Fig. 10 Quenching of the 3d
orbital moment (schematic).
The crystal field creates an
energy landscape that inhibits
the circular orbital motion of
the electron and leads to the
charge density of Fig. 4(b)
nearly exclusively from the spin of the 3d electrons, and the atoms look as if L = 0.
For example, iron has a magnetization of about 2.2 μB , but only about 5% of this
moment is of orbital origin. This effect is known as orbital-moment quenching.
Quenching was first recognized explicitly by van Vleck in 1937 [35].
Figure 10 illustrates the physics behind this effect, namely, the disruption of
the electron’s orbital motion by the crystal field. Mathematically, the difference
between quenched and unquenched wave functions is that between real and complex
spherical harmonics (Appendix A). Consider the two states |x2 –y2 >∼cos(4φ) and
|xy>∼ sin(4φ), which are shown in the top row of Fig. 7. Using lz = −i∂/∂φ
to calculate <lz > = − i ψ* lz ψ dφ yields <lz > = 0; it is completely
quenched. Pictorially, the electron “oscillates” in the valleys between the CF
potential mountains, as indicated by the dashed line in Fig. 10, and these oscillations
yield no net orbital motion. The respective electron densities <ψ|ψ> for |x2 –y2 > and
|xy>, namely, ρ = 1 + cos(8φ) and ρ = 1–2cos(8φ), exhibit complementary minima
and maxima, and the positioning of mountains decide which of the two densities
yields the lower energy.
Rather than asking why the orbital is quenched, we should therefore ask how
an orbital moment arises in a solid. Unquenched orbitals are described by wave
functions of the type exp (±imφ) = cos(mφ) ± i sin(mφ), or |±2>= |x2 –y2 >± |xy>.
These functions describe an uninhibited orbital motion and yield <lz > = ± 2 in units
of . However, the corresponding electron charge cloud is ringlike, ρ = const., so
that the electron occupies both valley and energetically costly hill regions, rather
than being confined to valleys. The competition between spin-orbit coupling (SOC)
and crystal field (CF) decides whether the orbital moment is quenched. In the 4f
case, the SOC is large, and the orbital motion of the electrons remains essentially
unquenched by the CF, as in Fig. 9. The opposite is true for 3d electrons, where the
SOC is a small perturbation to the CF, leading to nearly complete quenching.
132
R. Skomski et al.
Rare-Earth Anisotropy
The magnetocrystalline anisotropy of permanent-magnet materials, such as
Nd2 Fe14 B and SmCo5 , largely originates from the rare-earth sublattice. K1
values are 4.9 MJ/m3 and 17.0 MJ/m3 , respectively. By comparison, bcc iron has
K1 = 0.05 MJ/m3 [10]. The 4f wave functions are nearly unquenched, so that the
magnetocrystalline anisotropy energy is equal to the crystal-field energy, as in Fig. 1.
The basic physical picture of this single-ion anisotropy is clear, but a few questions
remain; it is necessary to determine the shape of the rare-earth 4f shells or ions and
to quantify the relationship between crystal-field interaction and anisotropy. Another
question is the temperature dependence. Anisotropy energies per ion correspond to
very low temperatures, at most a few kelvins, so the observation of anisotropy at
and above room temperature must be explained (Table 9).
Table 9 Anisotropy, magnetization, and Curie temperature of some rare-earth transition-metal
intermetallics [9, 10, 37]
Substance
YCo5
SmCo5
NdCo5
Y2 Fe14 B
Pr2 Fe14 B
Nd2 Fe14 B
Sm2 Fe14 B
Dy2 Fe14 B
Er2 Fe14 B
Y(Co11 Ti)
Sm(Fe11 Ti)
Y(Fe11 Ti)
Y2 Co17
Nd2 Co17
Sm2 Co17
Dy2 Co17
Er2 Co17
Y2 Fe17
Y2 Fe17 N3
Sm2 Fe17
Sm2 Fe17 N3
TbFe2
aK c
1
for TbFe2
K1 (RT)a
MJ/m3
5.2
17.2
0.7
1.06
5.6
4.9
−12.0
4.5
−0.03
−0.47
4.9
0.89
−0.34
−1.1
3.3
−2.6
0.72
−0.4
−1.1
−0.8
8.9
0.013
μo Ms (RT)
T
1.06
1.07
1.23
1.36
1.41
1.61
1.49
0.67
0.95
0.93
1.14
1.12
1.25
1.39
1.20
0.68
0.91
0.84
1.46
1.17
1.54
0.84
Tc
K
987
1003
910
571
565
585
618
593
557
940
584
524
1167
1150
1190
1152
1186
320
694
389
749
730
Structure
Hexagonal (CaCu5 )
Hexagonal (CaCu5 )
Hexagonal (CaCu5 )
Tetragonal (Nd2 Fe14 B)
Tetragonal (Nd2 Fe14 B)
Tetragonal (Nd2 Fe14 B)
Tetragonal (Nd2 Fe14 B)
Tetragonal (Nd2 Fe14 B)
Tetragonal (Nd2 Fe14 B)
Tetragonal (ThMn12 )
Tetragonal (ThMn12 )
Tetragonal (ThMn12 )
Hexagonal (Th2 Ni17 )
Rhombohedral (Th2 Zn17 )
Rhombohedral (Th2 Zn17 )
Hexagonal (Th2 Ni17 )
Hexagonal (Th2 Ni17 )
Hexagonal (Th2 Ni17 )
Hexagonal (Th2 Ni17 )
Rhombohedral (Th2 Zn17 )
Rhombohedral (Th2 Zn17 )
Cubic (laves)
3 Anisotropy and Crystal Field
133
Rare-Earth Ions
Rare-earth atoms tend to form tripositve ions in both metals and insulators. Since
spin-orbit coupling is very strong for inner-shell electrons in heavy elements, the
4f electrons experience a rigid coupling of their spin and orbital moments, with
unquenched orbitals and Hund’s-rules spin-orbit coupling. Magnetic anisotropy
is an intramultiplet effect, involving the 2 J + 1 magnetic quantum states Jz
of the ground-state multiplet. Excited multiplets have relatively high energies,
with the notable exceptions of Eu3+ and Sm3+ [33]. In the former, this energy
is only about 40 meV, but the ground-state moment of Eu2+ is zero, and the
element often adopts a Eu2+ configuration with half-filled shell and zero anisotropy.
Otherwise, the Eu3+ ion shows strong van Vleck susceptibility: the Eu3+ moment
is zero in its J = 0 ground-state multiplet, where the contributions from S = 3
and L=3 cancel, but the first-excited multiplet (7 F1 , J = 1) is only 330 K
above the 7 F0 ground-state multiplet. In the case of Sm3+ , the splitting between
the ground-state multiplet (6 H5/2 ) and the first-excited multiplet (6 H7/2 ) is about
100 meV (∼1000 K) [33], so that interatomic interactions and thermal excitations
yield some admixture of 6 H7/2 character (J-mixing). The focus of this section
is on ground-state multiplets, with a brief discussion of the excited Sm multiplet.
To determine the crystal-field energy, it is first necessary to specify the shape of
the 4f shells. Why is the Sm3+ ion in Fig. 1 prolate rather than oblate? Interchanging
oblate and prolate shapes changes the sign of K1 and has far-reaching implications.
A tentative answer is provided by the angular dependence of the (real) one-electron
4f wave functions, which are shown in Fig. 11. States with m = ±3, favored by
Hund’s second rule, are prolate, whereas the m = 0 state is oblate. The strong spinorbit coupling then creates axially symmetric superpositions exp.(±mφ) from states
with equal |m|, and Hund’s rules determine how the one-electron orbitals combine
to yield many-electron orbitals.
Like any electric charge distribution, the many-electron 4f shell can be expanded
in spherical harmonics. This multipole expansion provides a successively improved
description of angular features. In the zeroth order, the 4f shell is approximated by
a sphere of charge Q = Qo and does not support any anisotropy. The first-order
corresponds to an electric dipole moment Q1 , which is absent by wave-function
symmetry. The lowest-order electric moment is the second-order quadrupole
moment Q2 , which describes the prolaticity of a charge distribution. Table 10
lists some Hund’s-rules ground-state properties of the tripositive rare-earth ions,
including Q2 .
There is a systematic dependence of the ground-state ionic shape on the number
of 4f electrons. Gd3+ has a half-filled shell and a spherical charge distribution
because Hund’s rules mandate seven ↑ electrons having l = 3, 2, 1, 0, −1, −2,
−3, so that L = i li = 0 (S-state ion) and Q2 = 0. The other elements follow a
quarter-shell rule: the first and third quarters of the series have oblate ions, and the
second and fourth quarters have prolate ions. This rule is a consequence of particle-
134
R. Skomski et al.
Fig. 11 Angular dependence
of 4f wave functions. Red and
yellow areas indicate regions
of positive and negative wave
functions, respectively. As in
Fig. 4, the wave functions
shown here are the real ones,
and m refers to the wave
functions |m > ∼ exp.(imφ)
from which these wave
functions are constructed
hole symmetry in each half shell: 6 electrons are equivalent to a half-filled shell (7
electrons) with one hole. By Hund’s rules, the first electron(s) in a shell have a large
|m| and are oblate (Fig. 11), corresponding to a negative Q2 . Removing an electron
with a large |m| from a half-filled shell yields one oblate hole, which is the same as
a prolate electron distribution with a positive Q2 .
Table 10 is limited to the quadrupole moment Q2 . Higher-order multipole
moments provide a refined description of the angular dependence of the rare-earth 4f
electron cloud. The third-order octupole and fifth-order triakontadipole moments are
zero by symmetry, but the fourth-order hexadecapole moment (16-pole, Q4 ) and the
sixth-order hexacontatetrapole (64-pole, Q6 ) are generally nonzero. Figure 12 shows
the zoology of the angular dependence of the 4f charge distributions up to the fourth
order. For Hund’s-rules ions, the number of animals is limited by the symmetry of
the wave functions (Figs. 4 and 11), namely, n ≤ 4 for 3d ions and to n ≤ 6 for
4f ions [38]. Furthermore, since the rare-earth 4f electrons are unquenched, the 4f
charge distribution shows axial symmetry, and there are no multipole contributions
Ql m with m = 0. The anisotropy corresponding to the unquenched quadupole
3 Anisotropy and Crystal Field
135
Table 10 Hund’s-rules ground states of 4f ions. The orbitals listed from left to right, lz = 3, 2, 1,
0, −1,–2, −3.
moment of rare-earh ions can be very high, up to a few K per atom in temperature
units [36]. This temperature scale needs to be distinguished from that governing
the temperature dependence of anisotropy constants, which involves interatomic
exchange (Sect. 4.4.4).
136
R. Skomski et al.
Fig. 12 Cartoon illustrating the electrostatic R3+ multipole moments up to the fourth order (Q0 ,
Q2 , and Q4 ). The 4f charge distributions n(r) derive from Figs. 4 and 11 and are both axially and
inversion symmetric
Operator Equivalents
The next step is to quantitatively determine the interaction V (r) n (r) dV
(Sect. “One-Electron Crystal-Field Splitting”) between the crystal field and the
4f charge distribution. This can be done explicitly, in a straightforward but
cumbersome way, but a more elegant method is to use operator equivalents. Both
approaches assume that the crystal field, V (r) or Al m , and the 4f charge distribution,
n(r) or Qn , are known.
The straightforward method is best explained by considering
the lowest-order
uniaxial limit, where Eq. (16) reduces to V (r) = A2 0 3z2 –r 2 . Substituting this
expression into Eq. (12) yields
HCF = A02
3z2 − r 2 n(r)dV
(24)
3 Anisotropy and Crystal Field
137
By definition, the integral in this equation is equal to Q2 , so that HCF = A2 0 Q2 .
Equation (24) is exact and easily generalized to other Al m , but the problem remains
to determine Q2 as a function of the ion’s electronic properties and magnetization
angles. For example, the rare-earth crystal field is normally far too weak to affect
the term and multiplet structures, but it usually affects the intramultiplet structure.
These energy values can all be obtained by specifying n(r), but this is a very tedious
method.
A much more elegant approach is the use of Stevens operator equivalents Ol m .
The idea is to replace the real-space coordinates (x/r, y/r, z/r) in expressions such as
Eqs. (24) by the vector operator (Jx , Jy , Jz ), using J± = Jx ± iJy and identities such
as J2 = J(J + 1). The lowest-order noncubic operator equivalents are
O2 0 (J ) = 3 Jz 2 –J (J + 1)
(25)
corresponding to 3z2 – r2 and
O2 2 (J ) =
1 2
J+ + J− 2
2
(26)
corresponding to x2 – y2 = ½(x + iy)2 + ½(x – iy)2 . The derivation of higher-order
operator equivalents [33, 38] is straightforward but tedious. For example, the fourthorder cubic crystal-field expression Eq. (15a) consists of the term
1
1
20 x 4 + y 4 + z4 − 3r 4 /5 = 35z4 − 30 z2 r 2 + 3r 4 + 5 (x + iy)4 + (x–iy)4
2
2
(27)
2
and corresponds to O4 0 + 5 O4 4 . Here O4 0 = 35 Jz 4 −
30J (J + 1) Jz +
1
4
2
2
4
4
2
25Jz − 6J (J + 1) + 3J (J + 1) and O4 = 2 J+ + J– . The operators have
been tabulated in Refs. 33 and especially 38. It is also possible to define operator
equivalents Ol m (L, Lz ) and related spin Hamiltonians Hspin (S, Sz ) for 3d ions
(Sect. “Transition-Metal Anisotropy”), but the underlying physics is different from
the presently considered rare-earth limit, because L and S are only weakly coupled
(quenching).
The occurrence of Jz and of the ladder operators J± greatly simplifies the
calculation of matrix elements of magnetic ions in a crystal field or exchange field.
For Sm3+ , J = 5/2 yields Jz = ±5/2, ±3/2, and ± 1/2, corresponding to O2 0 = 10,
O2 0 = −2, and O2 0 = −8. The magnitude of the splitting is determined by A2 0 and
by the radial part of n(r), but the evaluation of the Ol m is sufficient to determine the
relative energies, namely, 5:–1:–4 in the present example.
The multipole moments are straightforward linear functions of the operator
equivalents:
Ql = θl < r l >4f Ol 0
(28)
138
R. Skomski et al.
Here the Stevens coefficients θ 2 = α J , θ 4 = β J , and θ 6 = γ J are rare-earth specific
constants that describe how Hund’s rules affect the shape of the R3+ ions [38]. For
example, Sm3+ has α J = 13/32 ·5·7, β J = 2·13/33 ·5·7·11, and γ J = 0. There is no
sixth-order crystal-field interaction for Sm3+ (γ J = 0), because the ground-state
multiplet has J = 5/2 < n/2. However, as mentioned in Sect. “Rare-Earth Ions”,
Sm3+ exhibits a rather unusual low-lying excited multiplet, which has J = 7/2 and
may give a small nonzero γ J contribution due to thermal or quantum mechanical
admixture.
Rare-earth ions in magnetically ordered compounds experience an interatomic
exchange field HJ , so that the rare-earth Hamiltonian becomes [39]
H = l,m Bl m Ol m (J, Jz ) + 2 μo (g–1) J · H J + g μo J · H
(29)
Here Bm n = θ n < rn >4f Al m and g J·H describes the comparatively weak Zeeman
interaction and HJ is the exchange field. The quantities L, S, and λ enter this
equation only indirectly, via Hund’s rules and J = L ± S. However, O l m contains
intramultiplet excitations (−J < Jz < J), and the raising and lowering operators
J± in Eq. (26) indicate that off-diagonal crystal fields, such as A2 2 , can change
Jz . To exactly diagonalize Eq. (29), it is necessary to include matrix elements
<Jz | Ol m (J ) | Jz >, where Jz = J’z . These matrix elements are known [38] but
complicate the calculations and the evaluation of the results.
Major simplifications arise if the term involving the exchange energy is much
larger than the CF interaction. This is approximately the case in rare-earth transitionmetal (RE-TM) intermetallics such as Nd2 Fe14 B [39, 40], where the exchange field
is roughly proportional to the RE-TM intersublattice exchange JRT . This strong
exchange field stabilizes states with Jz = ±J, where the sign determines the net
magnetization but does not affect the anisotropy. Intramultiplet excitations, caused
by the operators J± , are effectively suppressed, and only the Ql = θl < r l >4f Ol 0
terms remain to be considered. Furthermore, putting Jz = ±J drastically simplifies
the operator equivalents:
O2 0 = 2 J · (J − 1/2)
(30)
O4 0 = 8 J · (J − 1/2) · (J –1) · (J –3/2)
(31)
O6 0 = 16 J (J − 1/2) · (J –1) · (J –3/2) · (J − 2) · (J − 5/2)
(32)
The corresponding 4f charge distributions are axially symmetric around the
quantization axis (z-axis), and their multiple moments are given by Eq. (28).
Table 11 lists multipole moments derived from Eqs. (30)–(32).
3 Anisotropy and Crystal Field
Table 11 Rare-earth
multipole moments
Ql = θl < r l > Ol 0 for
Jz = J, measured in ml .
ao = 0.529 Å is the Bohr
radius
139
Element
4f1
Ce3+
4f2
Pr3+
3
4f
Nd3+
5
4f
Sm3+
7
4f
Gd3+
8
4f
Tb3+
9
4f
Dy3+
10
4f
Ho3+
11
4f
Er3+
12
4f
Tm3+
13
4f
Yb3+
Q2 /ao 2
−0.748
−0.713
−0.258
0.398
0
−0.548
−0.521
−0.199
0.190
0.454
0.435
Q4 /ao 4
1.51
−2.12
−1.28
0.34
0
1.20
−1.46
−1.00
0.92
1.14
−0.79
Q6 /ao 4
0
5.89
−8.63
0
0
−1.28
5.64
−10.0
8.98
−4.50
0.73
Single-Ion Anisotropy
The anisotropy constants are extracted by rotating the magnetization, that is, by
rotating the 4f charge distribution and calculating the energy. It is convenient to
choose a coordinate frame where J is fixed, that is, to actually rotate the crystal field
around the rare-earth ions. This can be done for each ligand separately, because
crystal fields obey the superposition principle. It starts conveniently from an axial
coordination, R || ez , and the corresponding crystal fields A 2 , A 4 , and A 6 are
referred to as intrinsic crystal fields [20]. In the point-charge model, A2 (R) =
–eq/4πεo R 3 . Due to the axial symmetry of the 4f charge distribution, the rotation
of R into the correct direction relative to the 4f moment involves a polar angle Θ.
For example
A2 0 = A2
1
3cos2 − 1
2
(33)
describes the rotation of a single ligand. By adding the contributions from all
ligands, one can create any crystal field and any relative orientation between crystal
and magnetic moment. This approach is not limited to uniaxial anisotropy. Equation
(16) is uniaxial, but it contains a z4 term, and by rotating different charges onto
the x- and y-axes, one can create crystal fields of the type x4 + y4 + z4 , which are
cubic. Figure 13 illustrates the rotation of the crystal around the rare-earth ion for a
fourth-order anisotropy contribution. Note that none of the rare-earth ions in Fig. 12
has the ghost shape, but quadrupole moments (Q2 ) do not interact with crystal fields
having fourfold symmetry, so that Fig. 13 actually applies to the UFOs (Ce, Tb) and
to the digesting snakes (Sm, Er, Tb).
Since crystal rotations, for example, Θ = 45◦ in Fig. 13(c), and magnetization
rotations are equivalent, Eq. (33) also describes the energy as a function of the
magnetization angle, that is, the anisotropy energy per rare-earth atom. Explicitly
140
R. Skomski et al.
Fig. 13 Cartoon-like “shaking-ghost” interpretation of fourth-order rare-earth anisotropies. Since
the head, feet, and hands of the ghost are made from negatively charged 4f electrons, electrostatics
favors (a) over (b) and (c). The latter two have the same crystal-field energy, but (c) is easier to
calculate, because it leaves the axis of quantization (arrow) unchanged
Ea =
1
Q2 A2 0 3 cos2 θ − 1
2
(34)
Comparison with Eq. (1) yields
K1 = −
3
A2 0 Q2
2VR
(35)
where VR is the crystal volume per rare-earth atom. This equation resolves the rareearth anisotropy problem by separating the properties of the 4f shell, described by
Q2 , from the crystal environment, described by A2 0 .
Crystal-field parameters such as A2 0 describe the surroundings of the rare-earth
ion and therefore change little across an isotructural series of compounds with
different rare earths. Examples are A2 0 values of 300 K/ao 2 for R2 Fe14 B, 34 K/ao 2
for R2 Fe17 , and – 358 K/ao 2 for R2 Fe17 N3 . In a given crystalline environment, the
sign of the rare-earth anisotropy depends on whether the ion is prolate or oblate.
A positive K1 is obtained by using oblate ions, such as Nd3+ , on sites where the
crystal-field parameter A2 0 is positive, and prolate ions, such as Sm3+ , in crystalline
environments where A2 0 is negative. This explains the use of neodymium in hard
R2 Fe14 B and RT12 N alloys, whereas samarium is preferred in RCo5 , R2 Fe17 N3 ,
and RT12 intermetallics. The rare-earth ions responsible for the anisotropy must
be magnetic, whereas both magnetic and nonmagnetic ligand atoms contribute to
the crystal field. An interesting example is interstitial nitrogen in Sm2 Fe17 , which
changes the anisotropy from easy-plane to easy-axis [41].
Using volume VR per rare-earth ion as a unit volume, the uniaxial anisotropy
constants are
3
21
K1 = − A2 0 Q2 − 5 A4 0 Q4 −
A6 0 Q6
2
2
(36)
3 Anisotropy and Crystal Field
K2 =
141
35 0
189
A4 Q4 +
A6 0 Q6
8
8
K3 = −
231 0
A6 Q6
16
(37)
(38)
Tetragonal magnets also have
K2 =
1
5
A4 4 Q4 + A6 4 Q6
8
8
K3 = −
11
A6 4 Q6
16
(39)
(40)
whereas hexagonal magnets exhibit only one in-plane term
K3 = −
1
A6 6 Q6
16
(41)
Cubic anisotropy can be considered as a special limiting case of tetragonal
anisotropy. Using Eqs. (36)–(40) and dropping terms absent incompatible with cubic
symmetry yields
K1 c = −5 A4 0 Q4 −
K2 c =
21 0
A6 Q6
2
231 0
A6 Q6
2
(42)
(43)
A striking feature in the last two equations is the absence of independent inplane crystal-field parameters, such as A4 4 . While a separate consideration of O4 4 ,
as contrasted to O4 4 ∼ Q4 , is not necessary for rare earths due to the axial symmetry
of the 4f charge clouds, the non-uniaxial CF parameters are not independent but
obey A4 4 = 5A4 0 and A6 4 = − 21A6 0 in cubic symmetry.
Temperature Dependence
Magnetic anisotropy exhibits a temperature dependence that is usually much more
pronounced than that of the spontaneous magnetization. It vanishes at the Curie
point. Figure 14 shows schematic temperature dependences of the anisotropy
constants for some classes of magnetic materials. Anisotropy energies per atom
intrinsically correspond to rather low temperatures, of order 1 K for. Magnetic
anisotropy at or above room temperature therefore requires the help of an interatomic exchange field Hex , which stabilizes the directions of the atomic moments
against thermal fluctuations.
142
R. Skomski et al.
Fig. 14 Temperature dependence of anisotropy (schematic): (a) basic dependence in elemental
magnets, (b) bcc Fe, (c) RCo5 alloys, and (d) Nd2 Fe14 B. The curves in (a) are schematic and less
smooth in practice [70], which reflects subtleties in the electronic structure
Typical rare-earth transition-metal (RE-TM) intermetallics exhibit a strong rareearth anisotropy contribution, and for TM-rich intermetallics, this contribution
dominates below and somewhat above room temperature. For example, the lowtemperature anisotropy constants K1 are 26 MJ/m3 for SmCo5 and 6.5 MJ/m3 for
Sm2 Co17 , as compared to room-temperature values of 17 MJ/m3 and 4.2 MJ/m3 .
The exchange field necessary to realize the RE anisotropy contribution is largely
provided by the rare-earth transition-metal (RE-TM) intersublattice exchange JRT ,
rather than the weaker rare-earth rare-earth (RE-RE) exchange [42].
The RE-TM interaction is proportional to J·Hex , that is, the rare-earth ions
behave like paramagnetic ions in an exchange field Hex ∼ JRT MT created by
3 Anisotropy and Crystal Field
143
and proportional to the transition-metal sublattice magnetization MT . Depending
on the sign of Hex , the RE-TM exchange favors Jz = ±J, and the corresponding low-temperature anisotropy is described by Ol m (J, Jz ) =Ol m (J, ±J ), as in
Eqs. 30–32. However, thermal excitation leads to the population of intermediate intramultiplet levels with |Jz | < J. The randomization becomes important
above some temperature T ∗ ∼ JRT /kB , which is typically of order 100–200 K,
Fig. 14(c–d). Below T*, |Jz | ≈ J, and the anisotropy is only slightly reduced.
Above T*, the rare-earth anisotropy contribution is strongly reduced. In the hightemperature limit, kB T JRT , all Jz levels are equally populated and the rare-earth
anisotropy vanishes, because m Ol m (J, m) = 0. The orientations of the 4f charge
clouds are thermally randomized and the net shape of the charge clouds becomes
spherical.
To quantify the temperature dependence, one must evaluate the thermal averages
< Ol m >th . At low temperatures, the quantization of Jz plays a role. The exchange
splitting between Jz = ±J and ± (J – 1) is of order JRT , so that the anisotropy
remains constant or “plateau-like” for T T*, Fig. 14(c). Above T*, the discrete
level splitting is less important and Jz can be considered as a continuous quantity.
This means that Jz = J cosθ and HRT = –JRT cos (θ ), and the operator equivalents
entering the
anisotropy
expression simplify to Legendre polynomials, for example,
O2 0 ∼ 12 3cos2 θ − 1 = P2 (cos θ ). The thermal averages
π
m
< cos θ >= N
exp
0
JRT
kB T
cos θ cosm θ sin θ dθ
(44)
are readily evaluated by a high-temperature expansion of the exponential function
and yield the rare-earth anisotropy [43].
K1 (T ) = K1 (0)
JRT 2
15 kB T 2
(45)
For anisotropies of arbitrary order m, it can be shown that Km ∼ (JRT /T )2m .
Equation (44) can also be used as a classical estimation for iron-series elements
and for the TM anisotropy contribution in RE-TM intermetallics. However, in
this case, J is not an independent interaction parameter (JRT ) but determined
by the Curie temperature, JTT ≈ kB Tc , and the high-temperature limit of Eq.
(45) is no longer meaningful. For small θ , Eq. (44) leads to <cosm θ > =
1−mk B T /JTT . The exponent m = 1 yields the magnetization, whereas values m>1
are necessary to determine the anisotropy, which is proportional to <Pm > =
1–m (m + 1) kB T /2JTT . These relations correspond to the famous Akulov-Callen
m(m + 1)/2 power laws [44–46]:
Km/2 (T )
=
Km/2 (0)
Ms (T )
Ms (0)
m(m+1)/2
(46)
144
R. Skomski et al.
Table 12 First and second-order anisotropy constants at low temperatures (LT) and at room
temperature
Element
Fe
Co
Ni
Nd2 Fe14 B
Pr2 Fe14 B
Sm2 Fe17 N3
LT
K1 (MJ/m3 )
0.052
0.7
−0.012
−18
24
12
Table 13 Transition-metal
and rare-earth contributions
to the room-temperature
magnetocrystalline anisotropy
[10]. All values are in MJ/m3
RT
K2
(MJ/m3 )
−0.018
0.18
0.03
48
−7
3
Structure Refs.
K1 (MJ/m3 )
0.048
0.41
−0.005
4.3
5.6
8.6
Compound
Nd2 Fe14 B
Sm(Fe11 Ti)
Sm2 Fe17 N3
Sm2 Co17
SmCo5
K2 (MJ/m3 )
−0.015
0.15
−0.002
0.65
≈0
1.9
K1
4.9
4.8
8.6
3.3
17.0
K1T
1.1
0.9
−1.3
−0.4
6.5
bcc
fcc
hcp
tetr.
tetr.
rhomb.
K1R
3.8
3.9
9.9
3.7
10.5
[47]
[47]
[47]
[48]
[48]
[48]
Symmetry
Tetragonal
Tetragonal
Rhombohedral
Rhombohedral
Hexagonal
In other words, 2nd-, 4th-, and sixth-order anisotropy contributions are proportional to the third, tenth and 21st powers of the magnetization, respectively. Equation
(46), which is valid up to about 0.65 Tc for Fe, means that higher-order anisotropy
contributions rapidly decrease with increasing temperature. A crude approximation,
based on Ms ∼ (1 – T/Tc )1/3 and used in Fig. 14(a), yields the linear dependence
K1 (T) ≈ K1 (0) (1 – T/Tc ) for the first anisotropy constant K1 of uniaxial magnets
(Table 12).
In summary, the temperature dependence of the anisotropy is a very complex phenomenon. Each crystallographically nonequivalent site generally yields
a different anisotropy contribution with a different temperature dependence, and
the distinction is most pronounced between rare-earth (4f ) and transition-metal
(3d) sites. As a rule of thumb, the RE or TM contributions dominate at low
or high temperatures [40, 49], and their respective temperature dependences are
approximately given by Eqs. (44 and 45) and Eq. (46). In the latter case, K1 ∼ Ms 3
(uniaxial magnets) and K1 ∼ Ms 10 (cubic magnets). Actinide (5f ) anisotropy is
limited by the interatomic exchange, although the spin-orbit coupling is very large,
and its temperature dependence follows that of the magnetization, K1 ∼ Ms [50].
The anisotropy of 3d–5d (and 3d–4d) intermetallics, such as tetragonal PtCo, largely
originates from the heavy transition-metal atoms, but this anisotropy is realized via
spin polarization by the 3d sublattice, roughly corresponding to K1 ∼ Ms 2 [51, 52].
The same dependence is obtained for the two-ion (magnetostatic) contribution to the
magnetocrystalline anisotropy, Sect. “Two-Ion Anisotropies of Electronic Origin”,
because the magnetostatic energy scales as Ms 2 (Table 13).
3 Anisotropy and Crystal Field
145
Transition-Metal Anisotropy
Typical second- and fourth-order iron-series transition-metal anisotropies are
1 MJ/m3 and 0.01 MJ/m3 , respectively, with large variations across individual alloys
and oxides (Tables 14 and 15). The anisotropy constants are often quoted in meV
or μeV per atom, especially in the computational literature dealing with metallic
magnets. A rule-of-thumb conversion for dense-packed iron-series transition-metal
magnets is 1 meV = 14.4 MJ/m3 . In alloys, the anisotropy must be multiplied by
the volume fraction f of the transition metals. For example, the transition-metal
contribution to the anisotropy of transition-metal-rich rare-earth intermetallics
corresponds to f ≈ 0.7, because about 30% of the crystal volume is occupied
by the rare-earth atoms.
The magnetic anisotropy 3d magnets is largely dominated by the degree of
quenching (Sect. “Spin-Orbit Coupling and Quenching”). For oxides, the degree
of quenching was implicitly considered by Bloch and Gentile [1], whereas Brooks
(1940) explicitly considered quenching in itinerant iron-series magnets [53]. An
explanation of quenching in itinerant magnets is provided by the model Hamiltonian:
Table 14 Anisotropy, magnetization, and Curie temperature of some oxides [9–11, 37, 63]
K1
(RT)
MJ/m3
α-Fe2 O3
−0.007
γ-Fe2 O3
−0.0046
Fe3 O4
−0.011
MnFe2 O4
−0.003
CoFe2 O4
0.270
NiFe2 O4
−0.007
CuFe2 O4
−0.0060
MgFe2 O4
−0.0039
BaFe12 O19
0.330
SrFe12 O19
0.35
PbFe12 O19
0.22
BaZnFe17 O27
0.021
Y3 Fe5 O12
−0.0007
Sm3 Fe5 O12
−0.0025
Dy3 Fe5 O12
−0.0005
CrO2
0.025
NiMnO3
−0.26
(La0.7 Sr0.3 )MnO3 −0.002
Sr2 FeMoO6
0.028
Substance
μo Ms
(RT)
T
0.003
0.47
0.60
0.52
0.50
0.34
0.17
0.14
0.48
0.46
0.40
0.48
0.16
0.17
0.05
0.56
0.13
0.55
0.25
Tc
K
960
863
858
573
793
858
728
713
723
733
724
703
560
578
563
390
437
370
425
Structure
Rhombohedral (Al2 O3 )
Cubic (disordered spinel)
Cubic (ferrite)
Cubic (ferrite)
Cubic (ferrite)
Cubic (ferrite)
Cubic (ferrite)
Cubic (ferrite)
Hexagonal (M ferrite)
Hexagonal (M ferrite)
Hexagonal (M ferrite)
Hexagonal (W ferrite)
Cubic (garnet)
Cubic (garnet)
Cubic (garnet)
Tetragonal (rutile)
Hexagonal (FeTiO3 )
Rhombohedral (perovskite)
Orthorhombic
146
R. Skomski et al.
Table 15 Anisotropy, magnetization, and Curie temperature of some transition-metal structures.
PT indicates a structural change near or below the Curie temperature
Fe
Co (α)
K1 (RT)
MJ/m3
0.048
0.53
μo Ms (RT)
T
2.15
1.76
Tc
K
1043
1360
Co (β)
Ni
Fe0.96 C0.04
Fe4 N
−0.05
−0.0048
−0.2
−0.029
1.8
0.62
2.0
1.8
1388
631
(PT)
767
Fe16 N2
Fe3 B
Fe23 B6
1.6
−0.32
0.01
2.7
1.61
1.70
(PT)
791
698
Fe0.65 Co0.35
FeNi
0.018
1.3
2.43
1.60
1210
(PT)
Fe0.20 Ni0.80
FePd
−0.002
1.8
1.02
1.37
843
760
FePt
6.6
1.43
750
CoPt
4.9
1.00
840
Co3 Pta
MnAl
2.1
1.7
1.38
0.62
1000
650
MnBi
1.2
0.78
630
Mn2 Ga
2.35
0.59
(PT)
Mn3 Ga
1.0
0.23
(PT)
Mn3 Ge
0.91
0.09
(PT)
NiMnSb
−6.3
1.10
698
Fe7 S8
0.320
0.19
598
Substance
a Extrapolation
Structure
Refs.
Cubic (bcc)
Hexagonal
(hcp)
Cubic (fcc)
Cubic (fcc)
Tetragonal
Cubic
(modified fcc)
Tetragonal
Tetragonal
Cubic
(C6 Cr23 )
Cubic (bcc)
Tetragonal
(L10 )
Cubic (fcc)
Tetragonal
(L10 )
Tetragonal
(L10 )
Tetragonal
(L10 )
Hexagonal
Tetragonal
(L10 )
Hexagonal
(NiAs)
Tetragonal
(D022 )
Tetragonal
(D022 )
Tetragonal
(D022 )
Cubic
(half-Heusler)
Monoclinic
[68]
[68]
to fully ordered Co3 Pt has been suggested to yield 3.1
H=
E1 (k) 0
0 E2 (k)
+λ
0 i
−i 0
MJ/m3
[69]
[68]
[70]
[37]
[71]
[72]
[73]
[37]
[68]
[37]
[74]
[74]
[74]
[75]
[74]
[9]
[76]
[76]
[76]
[37]
[37]
[67]
(47)
where E1 (k) and E2 (k) are two 3d subbands connected by a spin-orbit matrix
elements ±iλ. The spin-orbit term favors a nonzero net orbital moment, as required
3 Anisotropy and Crystal Field
147
for magnetic anisotropy, but λ ≈ 50 meV is usually much smaller than |E1 (k) –
E2 (k)|, the latter being comparable to the bandwidth W of several eV. However,
even for |E1 (k) – E2 (k)| = W, perturbation theory leads to a small orbital moment
and some residual anisotropy. Furthermore, accidental degeneracies E1 (k) = E2 (k)
yield the eigenvalues ±λ and completely unquenched orbitals. The corresponding
anisotropy energy, about 50 meV per atom, is then huge compared to typical ironseries anisotropies of 0.1 meV, or about 1 MJ/m3 .
The practical challenge is to add the spin-orbit couplings of all atoms (index i):
Hso = i λi l i · s i
(48)
to the isotropic Hamiltonian and to determine the anisotropy contributions from all
bands and k-vectors. To quantitatively determine the anisotropy, this procedure must
be performed for different spin direction s, s || ez and s || ex .
Perturbation Theory
The simplest approach to 3d anisotropy is the perturbation theory as originally
developed by Bloch and Gentile [1] and later popularized by van Vleck [35] and
Bruno [54]. The idea is to consider the Hamiltonian H = Ho + Hso , where Ho (l i )
is the nonrelativistic isotropic part and to consider Hso as a small perturbation. In
the independent-electron approximation, the lowest-order correction proportional to
ξ i = λ is obtained by using the perturbed wave functions |μ k σ >, where μ is a 3d
subband index and the index σ = {↑, ↓} labels the spin direction. Lowest-order
perturbation theory, linear in λ, uses completely quenched orbitals, <li > = 0, and
therefore <li ·si > = < li > ·si = 0.
The next term is quadratic in λ. For a single electron of wave function |μ k σ >,
the corresponding anisotropy energy is
Ek = λ2
μ,σ k
<μk σ |l · s|μ k σ ><μ k σ |l · s|μkσ >
Eμ k σ −Eμkσ
(49)
The total second-order anisotropy energy is obtained by summation over all
electrons. Since the SOC leaves the centers of gravity of the one-electron energies
unchanged, there is no net anisotropy contribution from level pairs |μ k σ > and |μ
k σ > when both levels are occupied (o) or unoccupied (u). The summation (or
integration) is therefore limited to |μkσ >= |o>and |μ k σ > = |u>:
E = −λ
o,u
of
<o | l · s | u> <u | l · s | o>
Eu −Eo
(50)
The numerical determination of the anisotropy constants requires the evaluation
E for several spin directions s = ½(σ x ex + σ y ey + σ z ez ), where σ x , σ y ,
148
R. Skomski et al.
and σ z are Pauli’s spin matrices. Equation (50) is sometimes reformulated in form
of a statement that the anisotropy energy is proportional to the quantum average
of angular orbital moment. However, this equivalence is limited to small orbital
moments [55] – rare-earth orbital moments are fixed by Hund’s rules (Fig. 13) and
do not change as a function of magnetization direction.
The spin summation is greatly simplified by the factorization of the unperturbed
wave functions, |μ k σ >= |μ k>|σ >, but the k-space summations can only
be performed numerically for most systems. The factorization into |μk> and
|σ > makes it possible to formally perform a summation over |μ k>, |μ k >, and
|σ > only, leaving the spin s unaveraged. This leads to a spin Hamiltonian of the
general many-electron type:
Hspin = −λ2 S · K · S
(51)
where K is a 3 × 3 real-space anisotropy matrix [56, 57]. For uniaxial anisotropy,
Eq. (51) reduces to the anisotropy term:
Hspin = D Sz 2 –S (S + 1) /3
(52)
This expression, which mirrors other second-order anisotropy expressions, is
not restricted to magnetocrystalline anisotropy but can also be used for dipolar
anisotropy (see Sect. “Magnetostatic Anisotropy”). It is most useful for 3d ions,
where D is often considered an adjustable parameter. As a rough approximation, Eq.
(52) can also be used for metallic Fe and Co (S ≈ 1). It cannot be used to describe the
anisotropy of independent conduction electrons (S = ½) nor for Ni (S ≈ ½), because
S = ½ yields Sz 2 – S(S + 1)/3 = 0 for Sz = ±½. It is, however, possible to consider
classical averages over a number of electrons, which yields Hspin = D cos2 θ –1/3
and K1 = −D.
Generalizing the perturbation expansion to arbitrary orders n yields anisotropy
constants of the order:
Kn/2 ∼
λn
Vo (Eo − Eu )n−1
(53)
where Vo is the crystal volume per transition-metal atom. This important relation,
known as spin-orbit scaling, was first deduced for lowest-order cubic anisotropy,
where n = 4 and K1 c ∼ K2 ) [1]. In this case, the anisotropy constant scales as
λ4 /A3 , where A is the energy-level splitting in the absence of spin-orbit coupling
(crystal-field splitting or bandwidth). This scaling behavior explains the low cubic
anisotropy of bcc iron (0.05 MJ/m3 ) and Ni (−0.005 MJ/m3 ), as compared to that
of hexagonal Co (0.5 MJ/m3 ) and YCo5 (5 MJ/m3 ).
Equation (53) provides a semiquantitative understanding of transition-metal
anisotropies. In metallic systems, Eu – Eo ∼ W, where the bandwidth W is about
5 eV for iron-series (3d) magnets and somewhat larger for palladium -series (4d),
3 Anisotropy and Crystal Field
149
platinum-series (5d), and actinide (5f ) magnets. The spin-orbit coupling rapidly
increases as the atoms get heavier (Tables 7 and 8), so that heavy transitionmetal elements are able to support very high anisotropies so long as the induced
magnetic moments on the heavy atoms are appreciable. In particular, FePt magnets
are important in magnetic recording [58], but both the low Curie temperature and
the low intrinsic magnetic moment per heavy transition-metal atom make it very
difficult to exploit the high anisotropy of very heavy atoms, up to 1000 MJ/m3 for
actinide compounds such as uranium sulfide [59].
As outlined in Eqs. (49 and 50), quantitative anisotropy calculations require a
summation of all occupied and unoccupied states. This summation involves matrix
elements <o|l·s|u>, which couple wave functions of equal |Lz |, namely, Lz = ±1 and
Lz = ±2, where the quantization axis (z-axis) is parallel to the spin direction (see
below). These matrix elements affect the sign and magnitude of the anisotropy but
do not change its order of magnitude, because they are of order unity. The order of
magnitude of the anisotropy is given by the spin-orbit coupling, which is essentially
fixed for a given element (Tables 7 and 8) and by the denominator Eo – Eu , which
requires a detailed discussion.
Spin-Orbit Matrix Elements
In Eq. (50), the itinerant wave functions |o > and |u > are of the Bloch type and can
therefore be expanded into atomic wave functions. Including spin, there are 10 3d
orbitals per atom, which yield 100 matrix elements <l·s > for each spin direction.
However, the number of independent matrix elements is drastically reduced by
symmetry. First, for the highly symmetric point groups Cnv , Dn , Dnh , and Dnd
(Sect. “Anisotropy and Crystal Structure”), only three spin and orbital-moment
directions need to be considered, namely, x, y, and z. Second, the matrix elements
between ↑↑ and ↓↓ pairs are the same, whereas those for ↑↓ and ↓↑ are equal
and opposite in sign. Third, many of the remaining matrix elements are zero by
symmetry [60].
Explicit matrix elements are obtained by applying equations such as
lˆ z = i(y∂/∂x – x∂/∂y) or lˆ z = −i∂/∂φ to the real or quenched 3d wave functions
of Fig. 4. For example, |xy>∼ sin(4φ) and |x2 –y2 >∼ cos(4φ) yield:
<xy | l̂z |x 2 –y 2 >=2i
(54)
This matrix element is imaginary and creates an imaginary (unquenched)
admixture to the wave function, as required for magnetocrystalline anisotropy. For
degenerate |xy> and |x2 –y2 > levels, this matrix element yields the eigenfunctions
exp.(±2iφ) = cos(2φ) ± i sin(2φ), the orbital momentum <lz > = ±2, and the
orbital moment ±2μB .
In terms of Fig. 10, the spin-orbit coupling acts as a perturbation that promotes
hopping from one valley into the next and thereby creates a small net orbital motion.
As outlined above (Sect. “Spin-Orbit Coupling and Quenching”), this motion is
150
R. Skomski et al.
2 2 ˆ
Fig. 15 The three “canonical” d electron orbital-momentum
√ matrix elements: (a) <x –y |l z
|xy > = 2i, (b) <xz|lˆ z |yz>= i, and (c) <z2 | l̂x | yz>= 3 i. The dashed lines are out of
the paper plane and visualize the direction of l̂ , but the actual length of the lines is zero, because
all orbitals belong to the same atom
responsible for the small orbital contribution to the magnetic moment of itinerant
magnets, such as Fe, and for the corresponding magnetic anisotropy.
The five 3d orbitals yield a fairly large number of matrix elements such as
that in Eq. (54), but due to symmetry, many of them are zero, and only three
are nonequivalent. Figure 15 illustrates these three “canonical” matrix elements.
Figure 15(a) corresponds to Eq. (54) and is encountered only once, aside from the
conjugate complex value –2i created by interchanging xy and x2 –y2 . The matrix
element of Fig. 15(b) occurs five times, namely, in form of <xz|lˆ z |yz>, <xy|lˆ x |xz>,
<xy|lˆ y |yz>, <x2 –y2 |lˆ x |yz>, and <x2 –y2 |lˆ y |xz>, whereas that of Fig. 15(c) has two
realizations, namely, <z2 |lˆ x |yz> and <z2 |lˆ y |xz>. A physical interpretation of matrix
elements <ψ 1 |l̂ |ψ 2 > is that <ψ 1 |ψ 2 > = 0 but the angular momentum operator
rotates ψ 2 and thereby creates overlap with ψ 1 . The rotation angle is equal to π/m,
where m is the magnetic quantum number of the orbitals, so that π/4 in Fig. 15(a, c)
and π/2 in Fig. 15(b).
Crystal Fields and Band Structure
An important question is the relation between electrostatic crystal-field interaction
and the interatomic hopping that leads to band formation. In the Mott insulator limit
of negligible interatomic hopping, the energy differences Eo – Eu correspond to
the ionic CF level splittings outlined in Sect. “Crystal-Field Theory”. However,
many oxides are Bloch-Wilson insulators, whose insulating character is a bandfilling effect. This means that band effects are not negligible in many or most
oxides. Hybridization-type ligand fields, which include band formation, do not alter
the qualitative physics of crystal-field theory [17] but are often stronger than the
electrostatic crystal fields and strongly affect quantitative anisotropy predictions.
3 Anisotropy and Crystal Field
151
For example, the eg -t2g crystal-field splitting in transition-metal monoxides is of the
order of 1 eV, as compared to 3d bandwidths of about 3 eV [62].
It is important to note that properly set up band structure calculations, from firstprinciple (Sect. “First-Principle Calculations”) or based on tight-binding approximations, automatically include crystal-field effects. This is easily seen by considering a tight-binding model that is nonperturbative as regards spin-orbit coupling. The
Hamiltonian is
H=−
2 2
∇ + j Vo r − Rj + j Hso Rj
2me
(55)
where the matrix elements of Hso are those of Fig. 15. The lattice periodicity is
accounted for by the ansatz
ψkμ (r) = N
j
exp ik · Rj φμ r − Rj
(56)
where the index μ labels the orbitals, such as |xy↑>. Putting Eq. (56) into Eq. (55)
yields, in matrix notation
Eμμ (k) = Eo δμμ + Aμ δμμ + m exp (ik · Rm ) tμμ (Rm ) + Eso,μμ
(57)
Here Eo is the on-site energy, Aμ is the subband-specific crystal-field energy, and
tμμ (Rm ) is the matrix containing the interatomic hopping integrals. The crystal-field
term is easily derived by splitting the potential energy j Vo (r – Rj ) into an on-site
term Vo (r – Ri ), which enters Eo , and a crystal-field term j = i Vo (r – Rj ).
Itinerant Anisotropy
Figure 16 shows an explicit example, namely, a monatomic tight-binding spin chain
with two partially occupied ↓ subbands near the Fermi level, namely, |xy > and
|x2 –y2 >, whereas Fig. 17 illustrates the corresponding band structure and anisotropy.
In terms of the fundamental Slater-Koster hopping integrals [64], txy, xy = Vddπ and
tx2–y2 ,x2–y2 = ¾Vddσ + ¼Vddδ , whereas txy ,x2–y2 is zero by symmetry. The ratio
Vddσ :Vddπ :Vddδ is about +6:-4:+1 [65], so that the model creates two cos(ka) bands
of nearly equal widths W ≈ 2Vddπ but opposite slope. The two bands, shown as
dashed curves in Fig. 17, exhibit a crossing at k = π/a.
The solid curves in Fig. 17 differ from the dashed ones by including crystal-field
and spin-orbit interactions. First, the charge distributions of the |x2 –y2 > orbitals
(bottom row in Fig. 16) point towards each other, so that the crystal-field charges
felt by the |x2 –y2 > orbitals are more negative than those felt by the |xy> orbitals.
This yields an equal CF shift of the two bands and shifts the crossing to slightly
lower k-vectors. Second, for s || ez , which is perpendicular to the plane of the paper
in Fig. 16, Eq. (54) yields an off-diagonal spin-orbit matrix element which mixes
152
R. Skomski et al.
the bands and creates an avoided crossing near k = π/a. The gap at this degenerate
Fermi-surface crossing (DFSC), 4λ, and the derived anisotropy energy K1 (k), shown
in Fig. 17(b), are finite, in contrast to the perturbative result of Eq. (50), where the
anisotropy contribution diverges at Eo (k) = Eu (k). To appreciate this peak, is useful
to recall that typical noncubic 3d anisotropies are of the order of 0.1 meV per atom,
as compared to SOC constants λ of about 50 meV and bandwidths W in excess of
1000 meV. In other words, the avoided crossings in Fig. 17(a) may look tiny on the
scale of the bandwidth but they are huge compared to anisotropies actually realized
in solids.
The bottom panel in Fig. 17(b) shows the k-space integrated density of states as
a function of the occupancy n of the spin-down |xy> and |x2 –y2 > bands. In analogy
to Eq. (50), it is sufficient to restrict the integration to the matrix elements between
occupied (o) and unoccupied (u) states, as schematically shown in Fig. 17(a). The
anisotropy, which favors a magnetization perpendicular to the chain, also exhibits a
DFSC peak for half filling, near k = π/a, although this peak is much less pronounced
than the k-space peak.
The simple model of Fig. 16 elucidates a major aspect of itinerant anisotropy,
namely, that different pairs of 3d subbands yield positive or negative anisotropy
contributions, depending on which of the three canonical matrix elements are
realized in each magnetization direction. Including spin, this creates 10 × 10 = 100
different contributions. Each of these contributions depends on the band filling and
may further split due to the involvement of different neighbors. As a consequence,
the anisotropies exhibit a complicated oscillatory dependence on d-band filling.
Figure 18 illustrates this point for a nanoparticle with a completely filled ↑ band.
Fig. 16 Monatomic spin-chain model (top) with two orbitals per site, |xy > (center) and
|x2 –y2 > (bottom)
3 Anisotropy and Crystal Field
153
Fig. 17 Magnetic anisotropy of the spin chain of Fig. 16: (a) band structure without CF and SOC
interactions (dashed lines) and with CF and SOC interaction (solid lines) and (b) anisotropy as a
function of the electron wave vector in units of 1/a (top) and band filling of the |xy> and |x2 –
y2 > orbitals (bottom). The gray area in (a) shows the occupied states used to define the electron
count 0 ≤ n ≤ 2 in the bottom part of (b). The peaks in (b) are caused by degenerate Fermi-surface
crossing near k = π/a
Fig. 18 First-order anisotropy constant of a hexagonal nanoparticle: (a) structure and (b) tightbinding anisotropy as a function of the number of d electrons (after Ref. 66)
By comparison, the anisotropy of rare-earth atoms in a given atomic environment
yields only two minima and two maxima, given by the quarter-shell rule of
Sect. “Rare-Earth Ions”. This simplicity originates from Hund’s rules, which yield
electron clouds of well-defined shape as a function of the number of f electrons.
In the itinerant case, each k-state corresponds to a different shape of the electron
cloud. This complicated picture starts to emerge in the simplest itinerant picture,
154
R. Skomski et al.
namely, in the diatomic pair model [60, 61], where the situation is reminiscent of a
quarter-shell rule.
It is instructive to compare the contributions of the nonperturbative DFSC
anisotropy peaks with the perturbative volume anisotropy due to Eu – Eo ∼ W. The
latter corresponds to the nearly homogeneous background in the top of Fig. 17(b)
and to the constant slopes near n = 0 and n = 2 in the bottom of Fig. 17(b).
In systems where the peak contribution is strong, a very dense k-point mesh is
necessary, or else the numerical error gets very big. The relative contribution of
the peaks depends on both the order of the anisotropy and the dimensionality of
the magnet. In one-dimensional magnets, the bulk and peak contributions to K1 are
comparable, as one may guess from the bottom of Fig. 17(b). More generally, Eq.
(53) means that perturbative anisotropy contributions scale as Km = W(λ/W)2m .
The peak contributions have a strength of λ but are restricted to a small kspace volume of (l/W)d , so the corresponding anisotropy contribution scales as
λ(λ/W)d = W(λ/W)d + 1 . The peak contributions are therefore strongest in lowdimensional magnets. They are of equal importance for d = 2 m – 1, that is for K1 in
one-dimensional magnets and K2 (K1 c ) in three-dimensional magnets. The latter is
fundamentally important, because it includes the anisotropy of cubic magnets such
as Fe and Ni. The former is important from a practical viewpoint, because quasi-onedimensional reflection from lattice planes creates pronounced peaks in the density
of states [77, 78].
First-Principle Calculations
The explanation of magnetocrystalline anisotropy by Bloch and Gentile [1] led to
the first attempt by Brooks in 1940 to describe itinerant anisotropy numerically
[53]. Early attempts to compute the anisotropy of itinerant magnets [53, 79–82]
led to substantial errors, such as wrong signs of K1 in cubic magnets. The errors are
partially due to the DFSC peaks discussed above, but they also reflect the limitations
of approximations such as tight binding. The use of self-consistent first-principle
density functional theory (DFT) has improved the situation in recent decades
[83–85], although reliable anisotropy calculations have remained a challenge,
especially for cubic magnets. Second-order anisotropy calculations for noncubic
transition-metals alloys, transition-metal contributions in rare-earth intermetallics,
and ultrathin films [86–91] are better described by DFT and have typical errors of
the order of 20–50%. However, in uniaxial magnets having nearly cubic atomic
environments, such as hcp Co, the situation is comparable to cubic magnets.
The Kohn-Sham equations, which form the basis of density functional theory,
are nonrelativistic. Spin-orbit coupling needs to be added in form of Eq. (50), which
is a second-order relativistic approximation, or a fully relativistic form, starting
from the Dirac equation. The latter is implemented in many modern codes, for
example, in the Vienna Ab Initio Simulation Package (VASP). The simplest method
to compute second-order anisotropies uses the so-called magnetic force theorem
3 Anisotropy and Crystal Field
155
[92, 93]. In this approach, the energy differences between two magnetization
directions are approximated by the difference of band-energy sums along different
magnetization directions, which can be achieved by a one-step diagonalization of
the full Hamiltonian. A better approach is to use total energy calculations, where
the energy is self-consistently calculated for each spin direction.
A specific problem is Hund’s second rule, which states that intra-atomic
electron-electron exchange favors states with large orbital momentum. The effect
is parameterized by the Racah parameter B and, in itinerant magnets, is known as
orbital polarization [89, 94]. The relative importance of this intra-atomic exchange
effect is reduced by band formation, but anisotropy calculations require a very high
accuracy, so that the corresponding orbital polarization effect cannot be ignored in
general. A simple but fairly accurate approach is to add an orbital polarization term
–½BL2 to the Hamiltonian, where B is of the order of 100 meV [94]. This term
lowers the energies of |xy> and |x2 – y2 > orbitals and enhances those of |z2 > orbitals.
The example of orbital polarization shows that correlation effects are important
in the determination of the anisotropy. In a strict sense, correlation effects involve
two or more Slater determinants [17], but sometimes their definition includes
Hund’s rule correlations. The latter are of the one-electron or independent-electron
type in the sense of a single Hartree-Fock-type Slater determinant [23]. Density
functional theory is, in principle, able to describe anisotropy, because anisotropy is
a ground-state property for any given spin direction. However, very little is known
about the density functional beyond the comfort zone of the free electron-inspired
local spin density approximation [95], including gradient corrections. For example,
rare-earth anisotropy, which is largely determined by the crystal-field interaction
of 4f charge distribution, can be cast in form of a density functional [96], but
this functional looks very different from the LSDA functional and its gradient
extensions.
One approach to approximately treat correlations is LSDA+U, where a Coulomb
repulsion parameter is added to the density functional [97]. The parameter U or, in
a somewhat more accurate interpretation, U ∗ = U –J is well-defined in the sense
that it should not be used to adjust theoretical results to achieve an agreement
with the experiment. Treating U as an adjustable parameter yields substantial
errors, of the order of 1 MJ/m3 for Ni [98]. However, similar to Hund’s-rules
correlations and LSDA, the LSDA+U approximation does not go beyond a single
Stater determinant. For example, it does not specifically address many-electron
phenomena such as spin-charge separation. The merit of the approach consists in
replacing local or quasilocal LSDA-type density functionals by density functionals
that are somewhat less inadequate for highly correlated systems. In particular, U
suppresses charge fluctuations and thereby improves the accuracy of the energy
levels connected by spin-orbit matrix elements [84]. Calculations going beyond a
single Slater determinant are still in their infancy. An analytic model calculation has
yielded Kondo-like corrections to the anisotropy [96], and dynamical mean-field
theory (DMFT) is being used to investigate the effect of charge fluctuations beyond
one-electron LSDA+U [99].
156
R. Skomski et al.
Case Studies
The magnetic anisotropies of a number of cubic and hexagonal 3d compounds are
only partially understood, both quantitatively and qualitatively. In cubic crystals, the
smallness of the anisotropy constants makes numerically calculations susceptible to
errors, for example, due to electron-electron correlations. Anisotropies in hexagonal
(and trigonal) magnets are higher, but their theoretical determination is complicated
by the fact that hexagonal crystal fields (sixfold symmetry) do not quench 3d
states (two- or fourfold symmetry). This quenching behavior is one reason for the
relatively high anisotropy of hexagonal magnets like BaFe12 O19 , SrFe12 O19 , and
YCo5 , as contrasted to tetragonal 3d magnets, such as steel. Hexagonal Co also
belongs to this high-anisotropy category, given that the atomic environment of the
Co atoms is nearly cubic.
Hexagonal ferrites. The anisotropy of Ba and Sr ferrites, which are widely
used as moderate-performance permanent magnets, is poorly understood in terms
of quantitative density-functional theory, partially due to the very narrow energy
levels. Nevertheless, early research by Fuchikami [57] traces the anisotropy to Fe
atoms on sites with a trigonal environment. An intriguing aspect of the system is
that all iron atoms in MFe12 O19 = (MO)·(Fe2 O3 )6 are ferric, Fe3+ , characterized
by half-filled 3d shells and zero anisotropy in the ground state. In more detail, the
crystal-field splitting yields an S = 5/2 ground state where two ↑ electrons occupy
a low-lying |xz> and |yz> doublet (e ), two ↑ electrons occupy an excited |xy>
and |x2 –y2 > doublet (e ), and the fifth ↑ electron occupies a |z2 > singlet (a 1 ) of
intermediate energy. The first-excited spin configuration is of the low-spin type
(S = 3/2), realized by one ↑ electron from the excited e level becoming an e
↓ electron. This spin configuration supports substantial anisotropy, because it has
odd numbers of electrons in two unquenched doublets. The splitting between the
S = 3/2 and S = 5/2 levels is fairly large (about 1 eV), but the admixiture of S = 3/2
character due to spin-orbit coupling is sufficient to create an anisotropy of the order
of 0.3 MJ/m3 .
Nickel. The anisotropies of the cubic transition metals (bcc Fe, fcc Co, fcc
Ni) have remained a moderate challenge to computational physics. Calculated
anisotropy constants are often wrong by several hundred percent and may even have
the wrong sign, that is, they predict the wrong easy axis. The choice of methods,
for example, with respect to the inapplicability of the force theorem to fourth-order
anisotropies, is one question [92, 93]. For instance, when a generalized gradient
approximation is used instead of the LSDA, the results are improved for bcc Fe
but not for Ni and Co [100]. In fact, the available choice of methods and density
functionals adds a “second-principle” component to first-principle calculations,
whose only input should be the atomic positions. Another problem is numerical
accuracy, depending on the number of k-points used.
A particularly well-investigated system is nickel [80, 82–84], where problems
are exacerbated by the smallness of the magnetic anisotropy (Table 12). The
anisotropy of Ni is determined by several contributions that largely cancel each
other: DFSC effects (Sect. “Itinerant Anisotropy”) are important, and the sum of the
3 Anisotropy and Crystal Field
157
anisotropy contributions from different orbitals and k-space regions is nearly zero. It
is also known that LSDA+U-type one-electron correlations are important in Ni. An
LSDA+U or “static DMFT” calculation was performed for Fe and Ni [84]. Values
of U* = 0.4 eV and U* = 0.7 eV have been advocated for Fe and Ni, respectively,
leading to anisotropy constants of 0.02 MJ/m3 for Fe (experiment: 0.05 MJ/m3 )
and − 0.04 MJ/m3 for Ni (experiment −0.005 MJ/m3 ). The Ni anisotropy is
overestimated, but the sign is correct, and a major reason for the correct sign is
the absence of a pocket near the X point of the fcc Brillouin zone. Without U, the
Fermi level cuts the pocket and spin-orbit matrix elements between occupied and
unoccupied states, similar to Fig. 17(a), creating an unphysical positive anisotropy
contribution.
YCo5 . The intermetallic compound YCo5 , which crystallizes in the hexagonal
CaCu5 structure, has the largest anisotropy among all know iron-series transitionmetal intermetallics, about 8 MJ/m3 at low temperature and 5 MJ/m3 at room
temperature [101]. Nearly all this anisotropy arises from the Co sublattices, in spite
of Y being a relatively heavy atom. According to Table 7, the spin-orbit coupling
of Y (32 meV) is not much smaller than that of Co (68 meV), but according to Eq.
(50), the effect of atomic SOC on the anisotropy scales is λ2 s2 , and the magnitude
of the Y spin is only about 0.3 μB , as compared to about 1.4 μB for Co [89]. In
other words, the anisotropy of YCo5 is about ten times greater than that of hcp Co,
in spite of the magnetically largely inert Y.
There are two reasons for the high anisotropy of YCo5 . First, the structure of
the YCo5 consists of alternating Co and Y-Co layers, in contrast to the nearly cubic
atomic environment in hcp Co. In this framework, the Y acts as a nonmagnetic
crystal-field source with a contribution similar to a vacuum. This has been shown in
a computer experiment [101] where the Y atoms were replaced by fictitious empty
interstices without any changes to the Co positions. The replacement reduces the
anisotropy by only 13%, which confirms that the anisotropy of YCo5 is largely due
to the anisotropic distribution of the Co atoms.
A secondary reason for the high anisotropy is that the electronic structure of
YCo5 supports a fairly strong orbital moment, about 0.2 μB per Co atom [93],
as compared to about 0.1 μB per atom in hcp Co [82]. The less quenched orbital
moment in YCo5 , which translates into enhanced anisotropy, partially reflects the
presence of degenerate |xy> and |x2 – y2 > states near the Fermi level [89]. According
to Eq. (54), the mixing of these states yields an orbital moment of up to 2 μB
per atom and a disproportionally strong anisotropy contribution (Fig. 17). More
importantly, the bands are very narrow near the Fermi level, which reduces the
denominator Eo – Eu in Eq. (50).
Iron, steel, and Fe nitride. Purified iron is magnetically very soft, but steel
formation due to the addition of carbon (Fe100–x Cx , x ≈ 4) drastically enhances
the coercivity [70, 102, 103]. The underlying physics is that carbon causes a
martensitic phase transition in bcc Fe, leading to a tetragonally distorted phase [70].
Figure 19 illustrates this mechanism, which is responsible for both the mechanical
and magnetic hardnesses of steel. The carbon occupies the octahedral interstitial
sites in the middle of the faces of the bcc unit cell (a). These octahedra are strongly
158
R. Skomski et al.
Fig. 19 Martensitic distortion of bcc Fe: (a) undistorted unit cell and (b) unit cell distorted
along the c-axis (dashed line). The martensitic distortion involves spontaneous symmetry breaking
along the a-, b-, or c-axis and extends over many interatomic distances, typically over several
micrometers
√
distorted: perpendicular to the faces, the Fe-Fe distance is smaller by a factor 2
than along the face diagonals. In a hard-sphere model based on an Fe radius of
1.24 Å, the radius of the interstitial site is 0.78 Å along the face diagonals but
only 0.19 Å perpendicular to the face. The atomic radius of C is about 0.77 Å
[103], so that the interstitial occupancy requires a strong tetragonal distortion. This
distortion breaks the cubic symmetry locally and, via elastic interactions between C
atoms on different interstitial sites, macroscopically. For example, 4 at% C yields an
enhancement of the c/a ratio by 3.5% [103]. Figure 19(b) shows the C occupancy
for a tetragonal distortion along the c-axis.
The martensitic lattice strain and the chemical effect due to the presence of the
carbon atoms yield almost equal uniaxial anisotropy contributions [102], and K1
is negative for Fe1-x Cx , of the order of −0.2 MJ/m3 . Cobalt addition changes the
sign of the volume magnetoelastic constant (Sect. “Magnetoelastic Anisotropy”)
and therefore the sign of the strain effect [70]. The magnetization is as high as
2.43 T in Fe65 Co35 , and the corresponding Honda steel [104] has a coercivity of
μo Hc = 0.020 T, as compared to 0.004 T for ordinary carbon steel. Such steels
dominated permanent-magnet technology in the early twentieth century and have
recently attracted renewed attention. Substantial anisotropy, K1 = 9.5 MJ/m3 , and a
magnetization of μo Ms = 1.9 T have been predicted for tetragonally distorted FeCo with c/a = 1.23 [105], although such a strong distortion is virtually impossible
to sustain metallurgically. Experimental room-temperature anisotropies reach about
2.1 MJ/m3 [106] and require a large amount of Pt (about 75 vol.%).
The behavior of interstitial N in Fe is similar to that of C [107], but nitrogen has
the additional advantage of improving the magnetization in tetragonal superlattices
of Fe8 N, or Fe16 N2 [108]. It is well-established that α -Fe16 N2 has a very high
magnetization [109, 110], about 2.8 ± 0.4 T, but the precise value has been a
subject of debate. An LSDA+U prediction of the magnetization is 2.6 T, which
3 Anisotropy and Crystal Field
159
includes a U contribution of 0.3 T [98]. Using U as an adjustable second-principle
parameter enhances the magnetization at a rate of 0.1 T/eV [111], but very large
values of U correspond to a heavy Fermion-like behavior that is contradictory to the
band structure of Fe8 N and to explicit first-principle calculations [98]. The roomtemperature K1 of the material is about 1.6 MJ/m3 [71]. LSDA and GGA reproduce
the correct order of magnitude [112].
Other Anisotropy Mechanisms
The magnetocrystalline anisotropy of Sects. Rare-Earth Anisotropy and 5 dominates
the behavior of most magnetic materials. Less commonly considered or more exotic
anisotropy mechanisms provide the leading contributions in a few systems and
substantial corrections in others. For example, two-ion anisotropies of magnetostatic
or electronic origin are usually much smaller than single-ion anisotropies, but they
dominate if the latter are zero by symmetry or by chance. An exotic mechanism
is the anisotropy of superconducting permanent magnets [113], which is not an
anisotropy in a narrow sense but reflects the interaction of local currents with the
real-structure features after field-cooling.
Magnetostatic Anisotropy
Magnetostatic dipole-dipole interaction between atomic spins yields a magnetostatic contribution to the magnetocrystalline anisotropy (MCA). Relativistically,
both spin-obit coupling and magnetostatic interactions are of the same order in the
small parameter v/c [29], but the similarities end here, and it is customary to treat
magnetostatic anisotropy contributions separately from MCA involving spin-orbit
coupling. The magnetostatic interaction energy between two dipole moments m and
m , located at r and r , respectively, has the form
EMS =
μo 3m · R m · R − m · m R 2
4π
R5
(58)
where R = r – r . The total magnetostatic energy is obtained by summation over all
spin pairs. In continuum theory, the summation must be replaced integration, i ...
mi = ... M(r) dV, and it can be shown that EMS = ½μo H2 (r) dV or, equivalently
EMS
1
= − μo
2
M (r) · H (r) dV
(59)
where H is the self-interaction field.
In a homogeneously magnetized body, the energy EMS depends on the direction
of m = m . Figure 20 shows the “compass-needle” interpretation of this anisotropy
contribution. Neighboring spins lower their magnetostatic energy by aligning
160
R. Skomski et al.
Fig. 20 Magnetostatic contribution to the magnetocrystalline anisotropy of a layered magnet with
tetragonal symmetry. The energy of the spin configuration (a) is higher than that of (b), because
the former creates a relatively large magnetic field between the layers
parallel to the nearest-neighbor bond direction R/R, and in noncubic compounds,
this amounts to magnetic anisotropy. The corresponding contribution to K1 , which
can exceed 0.1 MJ/m3 , is especially important in some noncubic Gd-containing
magnets, because Gd combines a large atomic moment (7 μB ) with zero crystalfield anisotropy due to its half-filled 4f shell. In cubic magnets, the anisotropy arising
from Eq. (58) is exactly zero [1], because it is a second-order anisotropy.
The anisotropy of Fig. 20 is closely related to the phenomenon of shape
anisotropy. If a homogeneously magnetized magnet has the shape of an ellipsoid,
then H(r) in Eq. (59) is also homogeneous inside the magnet (demagnetizing field).
For ellipsoids of revolution magnetized along the axis of revolution, H = –N M,
where N is the demagnetizing factor, that is, N ≈ 0 for long needles, N = 1/3 for
spheres, and N ≈ 1 for plate-like ellipsoids [10, 114]. Turning the magnetization
in a direction perpendicular to the axis of revolution yields H ⊥ = – 1–2N M.
Putting H|| and H⊥ into Eq. (59) and comparing the energies EMS yields the shape
anisotropy constant:
Ksh =
μo 1 − 3 N M2
4
(60)
This constant adds to the magnetocrystalline anisotropy constant, Keff = K1 + Ksh .
However, some precautions are necessary when using this equation. Consider a
slightly elongated magnet with N = 1/4 and zero magnetocrystalline anisotropy.
Equation (60) then predicts a positive net anisotropy constant Keff = μo M2 /16,
corresponding to a preferred magnetization direction parallel to the axis of
revolution. This is contradictory to the experiment.
In fact, the “shape anisotropy” of macroscopic magnets is merely a demagnetizing field energy. The demagnetizing factor N is defined for uniform magnetization, corresponding to the Stoner-Wohlfarth model in micromagnetism, and
this nanoscale uniformity is also exploited to evaluate Eq. (59). However, in
3 Anisotropy and Crystal Field
161
Fig. 21 Micromagnetic nature of shape anisotropy in a slightly prolate but defect-free ellipsoid.
Imperfections, including nonellipsoidal edges, cause reduced nucleation fields (coercivities), which
is known as Brown’s paradox
macroscopic magnets, the magnetization state becomes nonuniform (incoherent)
due to magnetization curling [9]. The curling leads to vortex-like magnetization
states for which a shape anisotropy can no longer be meaningfully defined. Curling
reflects the strength of the magnetostatic interaction relative to the interatomic
exchange and occurs when the radius of the ellipsoid exceeds the coherence radius
Rcoh ≈ 5(A/μo Ms 2 )1/2 , or about 10 nm for a broad range of ferromagnetic materials.
Figure 21 elaborates the micromagnetic character of shape anisotropy by showing
the external nucleation field (coercivity) as a function of the particle radius.
Atomic-scale magnetism, as in Fig. 20, is realized on a sub nm length scale. On
this scale, the interatomic exchange is sufficient to ensure a parallel spin alignment,
and the magnetic anisotropy is a well-defined quantity. Elongated nanoparticles, for
example, fine-particle magnets such as Fe amalgam [13], have radii of the order
of 10 nm and are well-described by Eq. (60). Shape anisotropy is also exploited
in alnico magnets [115–118], which contain needles of high-magnetization FeCo
embedded in a nonmagnetic NiAl matrix. The radius R of the needles is smaller than
about 50 nm but substantially larger than Rcoh , which reduces the shape anisotropy
by a factor Rcoh 2 /R2 [9].
Néel’s Pair-Interaction Model
The magnetocrystalline anisotropies of Sects. “Rare-Earth Anisotropy” and 5 are
single-ion anisotropies, that is, they can be expressed in terms of atomic spin
operators such as ŝz 2 . The underlying physical phenomenon is the spin-orbit
162
R. Skomski et al.
coupling, which is separately realized in each atom and described by Eq. (21).
The single-ion mechanism does not exclude interactions between spins, such as
exchange, but the net anisotropy of a magnet is obtained by adding all single-ion
contributions. Examples of two-ion anisotropies are the magnetostatic anisotropy,
just discussed and epitomized by Eq. (58), and Néel’s phenomenological pairinteraction model [119]. The latter uses an expansion of the anisotropy energy in
direction cosines. In the lowest order, the pair energy is equal to L (cos2 ψ – 1/3),
where L is a phenomenological parameter and ψ is the angle between bond axis and
magnetization direction. Néel’s expression is reproduced by putting m = m’ in Eq.
(58), that is, by assuming a uniform magnetization direction.
Single-ion and Néel two-ion anisotropies yield anisotropy-energy expressions of
the correct symmetry, but this does not mean that they are physically equivalent. For
example, both magnetic and nonmagnetic atoms contribute to the crystal field acting
on rare-earth ions, but the latter make no contribution in the Néel model, because is
based on pairs of magnetic atoms. Nonmagnetic ligands yield big anisotropy effects
in some materials, such as Sm2 Fe17 interstitially modified by N or C [41, 120]. The
alloy crystallizes in the rhombohedral Th2 Zn17 structure, where each Sm atom is
coordinated by three 9e interstitial sites, as shown in Fig. 22(a). The anisotropy of
Sm2 Fe17 is easy plane, that is, the Sm moment lies in the x-y-plane plane, which
also contains the 9e triangle. Heating powders of Sm2 Fe17 in N2 gas (gas-phase
interstitial modification) causes the nitrogen atoms to occupy the 9e interstices,
yielding the approximate composition Sm2 Fe17 N3 . The nitrogen addition changes
the anisotropy from easy-plane (K1 = −0.8 MJ/m3 ) to easy-axis (K1 = 8.6 MJ/m3 ),
because the virtually nonmagnetic N atoms act as strongly negative crystal-field
charges and repel the tips of the 4f charge distribution, Fig. 22(b).
One- and two-ion anisotropies are difficult to distinguish experimentally, partly
because interatomic exchange keeps neighboring spins parallel. The temperature dependence of the anisotropy is sometimes used as a criterion, scaling as
K1 (T) ∼ Ms (T)2 for magnetostatic anisotropy. However, a very similar behavior
is observed in L10 magnets such as FePt and CoPt, where the anisotropy is of the
single-ion type but requires proximity spin polarization of the 5d electrons by the
3d electrons [51, 52].
Two-Ion Anisotropies of Electronic Origin
Two-ion anisotropy is sometimes equated with magnetostatic anisotropy, but there
are also quantum-mechanical two-ion mechanisms [121]. The simplest example is
the two-ion anisotropy model described by the S = 1/2 Hamiltonian:
H = –Jxx Ŝx · Ŝx –Jyy Ŝy · Ŝy –Jzz Ŝz · Ŝz
(61)
In the isotropic Heisenberg model, J xx = J yy = J zz = J , but generally J xx = J
J zz due to spin-orbit coupling. There is no single-ion anisotropy in the model,
=
because the operator equivalent O 2 0 (S) = 3 Sz 2 – S(S + 1) is zero for S = 1/2 and
yy
3 Anisotropy and Crystal Field
163
Fig. 22 Anisotropy of
Sm2 Fe17 N3 : (a) interstitial
sites surrounding the Sm3+
ion in Sm2 Fe17 (blue) and (b)
change of the easy-axis
direction due to interstitial
nitrogen (yellow). Since this
anisotropy mechanism
involves one magnetic atom
only, it cannot be cast in form
of a Néel interaction
Sz = ± 1/2, but the “combined” spin S = 1, with Sz = 0 and Sz = ±1, supports
second-order anisotropy.
In the uniaxial limit, J xx = J yy = J o + J and J zz = J o – 2 J , where J o is
the isotropic Heisenberg exchange and J is relativistically small. Diagonalization
of Eq. (61) yields a singlet (S = 0) with wave function |↓↑ – ↑↓ > and energy
3J o /4, as well as triplet (S = 1). The triplet contains the Sz = ± 1 states |↑↑ > and
|↓↓>, both of energy – J o /4 – J /2, and the Sz = 0 state |↓↑ + ↑↓>, which has the
energy – J o /4 + J . Figure 23 shows the corresponding energy levels for J o > 0
and an anisotropy splitting 3 J /2 > 0. The anisotropic part of the triplet energy can
be written as
Ea = −
J 2
3Sz 2 − S (S + 1)
(62)
Formally, this expression is a spin Hamiltonian in form of an operator equivalent,
but here the spin S is the combined spin of the two atoms.
J reflects spin-orbit coupling, very similar to singleThe parameter
ion anisotropy and Dzyaloshinski-Moriya interactions. As emphasized in the
introduction, the Heisenberg model is isotropic, even if the bond distribution
164
R. Skomski et al.
Fig. 23 Level splitting for
the two-ion model of Eq.
(61). The anisotropic triplet is
very similar to an L = 1 or
J = 1 term in ionic
crystal-field theory, except
that the two coupled spins
reside on different ions
is anisotropic, for example, in a thin film. For example, ignoring spin-orbit
coupling and trying to explain electronic two-ion anisotropy in terms of SlaterPauling-Néel distance dependences yields lattice-anisotropic exchange constants Jo (z – z ) = Jo (x – x ) but does not reproduce the spin-anisotropic exchange
constants Jzz (r – r ) = Jxx (r – r ) required in Eq. (61). Anisotropic exchange is
usually mixed with single-ion anisotropy and relatively small, as exemplified by
hexagonal Co, whose saturation magnetization decreases by about 0.5% on turning
the magnetization from the easy magnetization direction into the basal plane [122].
The small parameter involved is essentially K1 Vat /Jo , so that the effect can be
enhanced by reducing Jo . However, since Tc ∼ Jo , this strategy is limited to lowtemperature magnets [123, 124].
Starting from the isotropic Heisenberg model (J ), the addition of an anisotropy
term Ea ≈ –K1 Sz 2 and putting K1 = ∞ yields the classical single-ion Ising
model [125–129]. The model, which has greatly advanced the understanding of
thermodynamic phase transitions, is characterized by Sz = ±S, whereas Sx = Sy = 0
reflects the “squeezing” of quantum-mechanical degrees of freedom due to the high
anisotropy. The model requires S ≥ 1, because Eq. (52) yields zero anisotropy for
S = 1/2. However, the underlying quantum-fluctuations are ignored in classical
models in the first place, and it is common to interpret the Ising model as a classical
spin1/2 model. Quantum-mechanical Ising models are obtained by putting J xx =
J yy = 0 in Eq. (61) while allowing nonzero values of Sx and Sy , for example, in
a transverse magnetic field [130, 131]. Such two-ion models are important in the
context of quantum-phase transitions.
Dzyaloshinski-Moriya Interactions
An interaction phenomenon closely related to single-ion anisotropy, electronic pair
anisotropy, and anisotropic exchange is the Dzyaloshinski-Moriya (DM) interaction
HDM = − ½ ij Dij · Si × Sj [132–135], where i and j refer to neighboring
atoms. The DM vector Dij = − Dji reflects the local environment of the magnetic
atoms and is nonzero only in the absence of inversion symmetry. Like the spin-
3 Anisotropy and Crystal Field
165
orbit coupling, the DM interaction is derived from the Dirac equation and is of the
same order relativistically. Phenomenologically, the interaction favors noncollinear
spin states, because HDM = 0 if the spins Si and Sj are parallel. Micromagnetically,
the DM interactions can be expressed in terms of magnetization gradients ∇S and
then assume the form of Lifshitz invariants. The corresponding energy contributions
depend on the point group of the crystal or film and are zero even for some crystals
without inversion symmetry [136].
DM interactions occur in some crystalline materials, such as α-Fe2 O3
(haematite), in amorphous magnets, in spin glasses, and in magnetic nanostructures
[37, 135, 137]. The resulting canting is small, because D competes against the
dominant Heisenberg exchange J, but the canting is easily observed in hematite
and other canted antiferromagnets where there is no ferromagnetic background.
The micromagnetism of the DM interactions [138] and its competition with singleion anisotropy is important in the context of magnetic vortices, for example, in
MnSi [139]. The spin angles between neighboring atoms are comparable to angles
encountered in domain walls, of the order of 1◦ for material ordered at room
temperature, which reflects the common relativistic origin of both phenomena
(D in the DM interactions and K1 determining the domain-wall width). DM
noncollinearities are not be confused with noncollinearities caused by competing
Heisenberg exchange interactions.
Antiferromagnetic Anisotropy
Magnetic anisotropy is not restricted to ferromagnets, because the single-ion
mechanism is operative in each magnetic sublattice. As in ferromagnets, the
net anisotropy is obtained by adding all sublattice anisotropy contributions. The
resultant is usually nonzero; single-ion anisotropy requires a magnetic moment on
each atom, but it does not require a nonzero net magnetization. An example is CoO,
where K1 ≈ 1 MJ/m3 [78].
Antiferromagnetic anisotropy can, in principle, be extracted from the spin-flop
field. When the antiferromagnet is subjected to a sufficiently strong magnetic field
parallel to easy axis, the net magnetization jumps from zero to a finite value [129].
The corresponding spin-flop field Hsf
μo μB Hsf = 2
K1 Vat (J ∗ − K1 Vat )
(63)
reflects the competition between intersublattice exchange J * and anisotropy K1 .
Snce J * Vat K1 in most materials, Hsf Ha , and high fields are needed to
produce the spin-flop, even in fairly soft materials.
The anisotropy energy remains unchanged on reversing the magnetization direction, Ea (M) = Ea (−M). This means that there should be no odd-order anisotropy
contributions. However, exchange bias caused by the exchange coupling of a
ferromagnetic and an antiferromagnetic phase yields an apparent unidirectional
166
R. Skomski et al.
anisotropy on cooling through a blocking temperature that was first observed as
an asymmetric shift of the hysteresis loop by Meiklejohn and Bean 1956 [140], in
their study Co nanoparticles coated with an antiferromagnetic CoO layer. Exchange
bias may be best characterized as an inner-loop effect, caused by the external
field’s inability to overcome the high anisotropy field of the antiferromagnetic
phase.
Magnetoelastic Anisotropy
Straining a magnet with a cubic crystal structure yields a noncubic structure
with nonzero second-order magnetic anisotropy. This mechanism contributes, for
example, to the magnetic anisotropy of steel (Sect. “Case Studies”). The same
consideration applies to isotropic magnetic materials, such as amorphous and polycrystalline magnets, if they are rolled and extruded. However, the change in K1 is
usually very small for metallurgically sustainable strain. Magnetoelastic anisotropy
is also important in soft magnets, especially in permalloy-type materials (Fe20 Ni80 ),
where the cubic anisotropy is small and the magnetoelastic contribution, caused by
magnet processing or a substrate, often dominates the total anisotropy. Magnetoelastic anisotropy is physically equivalent to magnetocrystalline anisotropy, because
a strained lattice is merely an unstrained lattice with modified atomic positions.
For example, the atomic environment in Fig. 1 can be considered as a tetragonally
strained cubic environment.
In many cases it is sufficient to describe a uniaxially strained isotropic medium
by the magnetoelastic energy:
HME
V
=−
λs E E
3 cos2 θ − 1 ε + ε2 − ε σ
2
2
(64)
where σ is the uniaxial stress, ε = l/l denotes the elongation along the stress axis,
E is Young’s modulus, and θ is the angle between the magnetization and strain
axis. The strength of the magnetoelastic coupling is described by the saturation
magnetostriction λs .
Putting σ = 0 and θ = 0 and minimizing the magnetoelastic energy with respect
to ε yields the elongation ε = λs . A magnet which has a spherical shape in the
paramagnetic state becomes a prolate ferromagnet when λs > 0 but an oblate
ferromagnet when λs < 0. Physically, the spin alignment creates, via spin-orbit
coupling, an alignment of the atomic electron distributions and a change in lattice
parameters. Since λs is only 10–100 ppm in most ferromagnetic compounds, a
moderate stress σ = Eε can outweigh the spontaneous magnetostriction. This then
produces a magnetoelastic anisotropy energy density:
HME
V
=−
λs σ 3 cos2 θ − 1
2
(65)
3 Anisotropy and Crystal Field
167
and the magnetoelastic contribution to K1 KME = 3λs σ /2, which may be fairly
large.
For cubic crystallites, there are two independent magnetostriction coefficients in
the lowest order, and the polycrystalline average over all possible orientations is
[141]
λs =
2
3
λ100 + λ111
5
5
(66)
where the quantities λ100 and λ111 are the spontaneous magnetostriction along
the cube edge and the cube diagonal, respectively. Experimental room-temperature
values of λs , measured in parts per million (10−6 ), are −7 for Fe, −33 for Ni,
+75 for FeCo, +40 for Fe3 O4 , −1560 for SmFe2 , and + 1800 for TbFe2 , and
practically zero for Py (permalloy, Fe20 Ni80 ) [11, 70, 115]. For example, highcarbon steel (Fe94 C6 ) has E = 200 GPa and is strained by about 5% [103], so that
KME ≈ −0.1 MJ/m3 (see the discussion of steel magnets in Sect. “Case Studies”).
Low-Dimensional and Nanoscale Anisotropies
Nanostructuring opens a new dimension to anisotropy research and practical
applications. Surface and interface anisotropies become important on the nanoscale,
and it is possible to realize atomic environments not encountered in the bulk [9,
142]. Examples are thin films and multilayers, nanowires, single atoms, molecules,
and nanodots on surfaces, nanogranular thin-film, and bulk magnets [142]. Figure 24
shows some of these nanostructures, whose dimensions range from less than 1 nm,
for adatoms and monatomic nanowires, to 100 nm or more in nanostructured composites. Most structures can be produced freestanding or deposited on substrates,
and advanced techniques are available for their fabrication and characterization (see
the other chapters of this book and Refs. [15, 143, 144]).
From a theoretical viewpoint, arbitrarily small anisotropies are important in
the theory of two-dimensional phase transitions, because they can change the
universality class from Heisenberg-like to Ising-like and even create a nonzero Curie
temperature [145, 146].
Surface Anisotropy
Surface and interface anisotropies, which are closely related, play an important role
in magnetic thin films and nanostructures. Surface anisotropies easily dominate the
bulk anisotropy in nanostructures with cubic or amorphous crystal structures, but
surface and interface contributions are also of interest in noncubic systems. For
example, L10 -ordered magnets such as FePt and CoPt can be considered as naturally
occurring multilayers. In line with other 3d anisotropies, the sign and magnitude of
surface anisotropies are difficult to predict, but some crude rules of thumb exist for
168
R. Skomski et al.
Fig. 24 Anisotropic
nanostructures: (a) thin films
(L10 -FePt/MgO), (b)
free-standing Pd zigzag
nanowire, (c) monatomic Fe
nanowire on Pt(001), and (d)
Co adatom on an insulating
substrate. First-principle
calculations often use
periodic arrays of supercells
with sufficiently big
airgaps (a)
the anisotropy as a function of band filling [60, 61]. For example, anisotropy often
changes sign between Fe and Co, the Fe preferring an easy axis perpendicular to the
Fe-Fe bonds (perpendicular to the plane).
Surface anisotropy tends to dominate when the thin-film thickness is in the range
of a few atomic layers. Phenomenologically [88, 147]
K1 = KS /t + KV
(67)
where t is the film thickness, KS is the surface anisotropy, and KV includes the bulk
magnetocrystalline and shape anisotropies. Typical iron-series surface anisotropies
are of the order of 0.1–1 mJ/m2 [147], or 0.03–0.3 meV per surface transitionmetal atom, which corresponds to bulk equivalents of 0.5–5 MJ/m3 . When KV and
KS favor in-plane and perpendicular anisotropy, respectively, then there is a spinreorientation transition from perpendicular to in-plane as the thickness exceeds
KS /|KV |. Note that Eq. (67) does not mean that the anisotropy is limited to the
surface: the equation is asymptotic, with small contributions from subsurface atoms
and from atoms deeper in the bulk.
Thin-film, multilayer, surface, and interface anisotropies have the same physical
origin as bulk anisotropies, mostly single-ion anisotropy with magnetostatic correc-
3 Anisotropy and Crystal Field
169
Fig. 25 Effect of surface
index on the surface of bcc
Fe: (a) fourth-order in-plane
anisotropy for a (001) surface
and (b) second-order in-plane
anisotropy for a (011)
surface. Gray atoms are
subsurface atoms [148]
tions. The anisotropic distribution of exchange bonds at the surface does not create
magnetic anisotropy. The Heisenberg Hamiltonian is isotropic, even if the exchange
bonds Jij = J(ri – rj ) are anisotropic. Only relative angles between neighboring spins
matter, and the Heisenberg model is silent about the easy magnetization directions.
Both the easy axes and the strength of the anisotropy depend on the index of the
surface, and there is no reason to expect that the anisotropy axis should necessarily
be normal to the surface. For example, the (100) surface of bcc Fe, Fig. 25(a), has
fourfold in-plane symmetry and yields a fourth-order anisotropy contribution. By
comparison, the (011) surface, Fig. 25(b), has a twofold in-plane symmetry, which
yields two nonzero lowest-order anisotropy constants , K1 and K1 [148]. Surface
defects often yield substantial anisotropy contributions [88, 144]. Stepped surfaces
are an example, which can also be considered as high-index surfaces [144, 149].
Random Anisotropy in Nanoparticles, Amorphous, and Granular
Magnets
Many magnetic materials are characterized by random easy axes n(r), so that the
uniaxial anisotropy-energy expression K1 sin2 θ must be replaced by
Ha = –
K1 (n · s)2 dV
(68)
where s = M(r)/Ms . Atomically, K1 in Eq. (68) is the same as the K1 in Sect. “Lowest-Order Anisotropies”, the only difference being the randomness of the local
c-axis. Random anisotropy is important in a variety of materials, including hard
and soft-magnetic polycrystalline solids [150–155], amorphous magnets [124, 137,
156], spin glasses [135], and nanoparticles [143, 157]. One example is the approach
to saturation in polycrystalline materials (Sect. “Anisotropy Measurements”).
Nanoparticles and nanoclusters are defined very similarly, but in a strict sense,
the former are random objects, whereas the latter are characterized by well-defined
atomic positions. Typical nanoparticles contain surface patches with many different
indexes, and the corresponding anisotropy contributions add.
The net anisotropy of nanoparticles is generally biaxial, involving both K1 and
K1 , and there is generally no physical justification for considering nanoparticles as
170
R. Skomski et al.
uniaxial magnets. This can be seen from Eq. (3): aside from accidental degeneracies,
there is always one axis of lowest energy. However, uniaxiality goes beyond
having an axis of lowest energy (easy axis), because it also requires the absence
of “secondary” anisotropy axes perpendicular to the easy axis. The secondary
anisotropy is important, because it causes hysteresis loops to deviate from the
uniaxial predictions.
Consider a nanoparticle with a highly disordered surface, so that each of the
NS surface atoms yields an anisotropy contribution ±Ko , where Ko is 0.03–0.3 meV
(Sect. “Surface Anisotropy”) and ± refers to orthogonal easy axes. For NS = ∞, the
surface anisotropy would average to zero, but in patches of finite NS , the averaging
is incomplete. The addition of NS random contributions
√ ±Ko creates a Gaussian
distribution√of net anisotropies of the order of ±Ko / NS per surface atom [9], or
Keff = Ko NS /N averaged over all N atoms in the particle. Here the negative sign
means that the easiest axis switches into a direction perpendicular to the reference
axis (z-axis). Since Ns ∼ R2 and N ∼ R3 , Keff scales as 1/R.
Atomic-scale random-anisotropy effects in bulk solids were first discussed in
the context of amorphous magnets, which exhibit random-field [158], randomanisotropy [159], and random-exchange spin glasses [135, 137]. In less than four
dimensions, the ground state of random-anisotropy magnets does not exhibit longrange ferromagnetic order [135]. However, this does not preclude the use of
random-anisotropy magnets as nanostructured magnetic materials, where hysteretic
properties are important [155, 160] and true equilibrium is rarely reached. The
coercivity and remanence of atomic-scale random-anisotropy magnets were first
investigated in the late 1970s [151, 156], but a very similar situation is encountered
in nanocrystalline magnets [161, 162].
The random anisotropy in Eq. (68) creates magnetic hysteresis. In the case
of noninteracting random-anisotropy grains, which also includes noninteracting
nanoparticles, the M(H) loops are obtained by adding the Zeeman interaction –μo
Ms H·s dV to Eq. (68), finding the M(H) loop for each direction n, and then
averaging over all n. In terms of Ha = 2 K1 /μo Ms , the behavior near remanence is
M(H) = Ms (1/2 + 2H/3Ha ). In particular, the remanence ratio Mr /Ms = M(0)/Ms
is equal to 1/2. Performing the same analysis for cubic magnets with iron-type
(K1 > 0) and nickel-type (K1 < 0) anisotropy yields the remanence ratios 0.832 and
0.866, respectively. Replacing the easy-axis anisotropy by easy-plane anisotropy
yields a very similar curve for H > 0, namely, M(H) = Ms (π/4 + H/3Ha ), and
the same asymptotic behavior (Sect. “Anisotropy Measurements”). However, the
coercive behavior is very different: random easy-axis anisotropy yields Hc = 0.479
Ha , whereas easy-plane anisotropy leads to Hc = 0.
Intergranular exchange modifies the hysteresis loops, creating some coercivity
in the easy-plane ensembles but reducing the coercivity in the case of easyaxis anisotropy. The exchange energy density, A(∇σ m )2 , is largest for rapidly
varying magnetization directions σ m , so that exchange effects are most pronounced
grain with small radius R. In the weak-coupling limit, A/R2 K1 , there are
quantitative corrections to the hysteresis loop [9], but the strong-coupling behavior
is qualitatively different.
3 Anisotropy and Crystal Field
171
In the limit of infinite exchange, all grain magnetizations would be parallel,
σ m (r) = σ mo , and the average anisotropy of Eq. (68) would be zero by symmetry
for isotropic magnets. Large but finite exchange means that N grains are coupled
ferromagnetically, where N increases with A. Each grain yields an anisotropy
contribution ±K1 , but as in the above nanoparticle
√ case, the anisotropy does not
average to zero but exhibits a distribution ±K1 / N. This yields the total energy
density:
η=
1
A
− K1 √
2
L
N
(69)
where L is the magnetic correlation length, that is, the radius of the correlated
regions. In d dimensions, it is given by N = (L/R)d . Putting this expression into
Eq. (69) and minimization with respect to L yields the scaling relation
L ∼ R (δo /R)4/(4–d)
(70)
where δ o = (A/K)1/2 is the domain-wall-width parameter. Equation (70) shows that
d = 4 is a marginal dimension below which small grains (R < δ o ) yield intergranular
correlations (L√> R). In three dimensions, L ∼ 1/R3 .
Since K1 / N can be considered as an effective anisotropy, the formation of
correlated regions reduces the coercivity:
Hc ∼ Ha (R/δo )2d/(4–d)
(71)
In three dimensions, this means that the coercivity of random-anisotropy magnets
scales as R6 [156]. This dependence helps to reduce the coercivity of soft-magnetic
materials [155]. For example, K1 is virtually zero for amorphous alloys Fe40 Ni40 B20
and Gd25 Co75 [37]. Random anisotropy magnets having large grain sizes are in a
weak-coupling regime and exhibit high coercivities of the order of 2K1 /μo Ms , and
there is a fairly sharp transition between the strong-coupling (small R) and weakcoupling (large R) regimes.
Giant Anisotropy in Low-Dimensional Magnets
Very high anisotropies per atom are possible in small-scale nanostructures such
as adatoms on surfaces or monatomic wires. These high anisotropies indicate
unquenched orbital moments , due to either high spin-orbit coupling or high crystalfield symmetry. The former is realized for Co atoms on Pt(111) [163], where
a giant magnetic anisotropy of about 9 meV per Co atom has been measured.
Platinum is predisposed toward strong anisotropy, because it is close to the onset
of ferromagnetism and possesses a spin-orbit coupling of about 550 meV. A single
atom of Fe or Co easily spin-polarizes several Pt atoms, which then make large
contributions to the anisotropy. Atomically thin nanowires, such as the zigzag wire
172
R. Skomski et al.
in Fig. 24(b), may support very high anisotropy, partly due to pronounced van-Hove
peaks in the density of states. In terms of Eq. (53), van-Hove singularities near
the Fermi level correspond to small energy differences Eu – Eo . For example, an
anisotropy of 5.36 meV per atom has been predicted for free-standing ladders of Pd
atoms [164].
An upper limit to the anisotropy per atom is given by the spin-orbit coupling
constant, λ ≈ 50 meV for the late iron-series transition metals. This huge value
corresponds to 140 MJ/m3 for dense-packed atoms. It is unlikely that this anisotropy
could be exploited in nanotechnology, because anisotropy is defined as anisotropy
energy per unit volume and the requirement of isolated or freestanding wires leads to
a dilution of the anisotropy. Densification is incompatible with such huge anisotropy,
because crystal formation involves interactions of the order 1000 meV, which tend
to quench the orbital moment.
Quenching is ineffective in free-standing monatomic nanowire however, and
anisotropy energies of 20–60 meV have been predicted or experimentally inferred
for these structures. In 3d systems, anisotropies as high 6–20 meV/atom have been
calculated for free-standing linear monatomic Co wires [165]. Some monatomic
4d and 5d wires exhibit larger anisotropies, up to 60 meV per atom in stretched
Rh and Pd, respectively [166]. The high anisotropy of freestanding monatomic
nanowires indicates that some levels undergo little or no quenching. The wires
have C∞ symmetry, which leaves the states with nonzero Lz , namely, |xz> and
|yz> (Lz = ± 1) and |xy> and |x– y2 > (Lz = ± 2), completely unquenched so long
as the spin is parallel to the symmetry axis of the wire (z-axis). Figure 26 compares
the corresponding level splitting with the tetragonal one in Fig. 6. Physically, the
electrons freely orbit around the wires, because there are no in-plane crystal-field
charges that could perturb this motion. The corresponding wave functions, |xz > ± i
|yz > and |xy > ± i |x– y2 >, yield anisotropy energies of up to λ and 2λ, respectively,
depending on the number of electrons in the system.
Configurations similar to Fig. 6 also exist in a few crystalline environments.
Recent experiments have indicated that a Co ad-atom deposited on MgO shows
the giant magnetic anisotropy of 58 meV [167]. This huge anisotropy requires a
degeneracy between two levels of equal |Lz |. Co adatoms on MgO(001) have C4
symmetry. Due to Hund’s rules, the Co2+ ion (3d7 ) has one electron in the xy-xz
doublet, and this degeneracy yields a large orbital moment, <Lz > ≈ 1, and a huge
anisotropy.
The example of Co on MgO shows high anisotropy energies can also be obtained
in some crystalline environments. The C4 argument can be extended to vertically
embedded but laterally isolated wires. Such configurations might conceivably be
used for magnetic recording. In terms of thermal stability, 50 meV corresponds to
580 K, or about 2kB T per atom. For magnetic recording, one would need about 50
kB T, or chain lengths of 25 strongly exchange-coupled 3d atoms. Heavier elements
have stronger spin-orbit couplings but cannot be used for this purpose, because
3 Anisotropy and Crystal Field
173
Fig. 26 Crystal-field
splitting in an insulating
free-standing nanowire. Since
states with the same quantum
number |Lz | > 0 form
degenerate states (doublets),
there is no quenching, and
large magnetocrystalline
anisotropies K1 Vo ∼ λ are
possible
their interatomic exchange is too small to ensure ferromagnetic alignment at room
temperature. It is uncertain whether any of the approaches outlined in this section
could be used to improve areal recording densities.
Multiferroic aspects of magnetic anisotropy are an important aspect of current
research in solid-state physics and nanoscience. Electric-field control of magnetic
anisotropy in magnetic nanostructures could enable entirely new device concepts,
such as energy-efficient electric field-assisted magnetic data storage. Due to screening by conduction electrons in metals, there is no electric-field dependent bulk
anisotropy, but the surface anisotropy changes via the filling of the 3d orbitals, which
is modified by the electric field. This was demonstrated L10 -FePd and FePt thin
films immersed in a liquid electrolyte [168], where the coercivity can be modified by
an applied electric field. A common scenario is that an electric field yields a modest
change in K1 , which modifies the coercivity of the films and could be exploited for
magnetization switching [169, 170]. Similar mechanisms are realized in nanowires
on substrates, Fig. 24 [171]. For example, the application of an electric field has
been predicted to change the sign of K1 of organometallic vanadium-benzene wires
[172]. Mechanical strain and adsorbate atoms on thin films may have a similar
effect [16].
Acknowledgments This chapter has benefited from discussions with B. Balamurugan, C. Binek,
R. Choudhary, J. Cui, P. A. Dowben, A. Enders, O. Gutfleisch, G. C. Hadjipanayis, H. Herper,
X. Hong, S. S. Jaswal, P. Kharel, M. J. Kramer, P. Kumar, A. Laraoui, L. H. Lewis, S.-H. Liou,
J.-P. Liu, R. W. McCallum, O. N. Mryasov, D. Paudyal, R. Sabirianov, S. S. Sankar, T. Schrefl,
D. J. Sellmyer, J. E. Shield, A. K. Solanki, and A. Ullah. The underlying work was or has been
supported by ARO (W911NF-10-2-0099), DOE (DE-FG02-04ER46152), NSF EQUATE (OIA2044049), partially NSF-DMREF (1729288), HCC, and NCMN.
174
R. Skomski et al.
Appendices
Appendix A: Spherical Harmonics
Separating radial (r) and angular (θ , φ) degrees of freedom, any function f (θ , φ)
can be expanded into spherical harmonics Yl m (θ , φ). The present chapter uses this
expansion to describe (i) atomic wave functions ψ(r), as in Figs. 4 and 11, (ii) atomic
charge densities n(r), (iii) crystal-field potentials V(r) and operator equivalents O m l ,
and (iv) magnetic anisotropy energies Ea (θ , φ). These quantities differ by radial part
and physical meaning, but their angular dependences are all described by
m
Yl m (θ, φ) = Nl exp (imφ) Pl m (cos θ )
(72)
where the Pl m are the the associated Legendre polynomials. Concerning sign and
magnitude of the normalization factor N l m , we use the convention
(2l + 1) (l − m)!
4π (l + m)!
Nl m =
(73)
It is sometimes useful to express Eq. (1) in terms of Cartesian coordinates or
“direction cosines” x, y, and z. Last but not least, the complex functions exp.(imφ)
may be replaced by real functions, using exp.(±imφ) = cos (mφ) ± i sin(mφ). These
real spherical harmonics, also known as tesseral harmonics, are often convenient,
because charge densities, crystal-field potentials, and anisotropy energies are real
by definition. However, the distinction remains important in quantum mechanics,
because complex and real spherical harmonics correspond to unquenched and
quenched wave functions, respectively.
A very frequently occurring function is
Y2 0 =
1
2
5 3 cos2 θ –1
4π
(74a)
or
Y2
0
1
=
2
5 3z2 − r 2
4π
r2
(74b)
Note that the Cartesian coordinates require a factor 1/rl , which ensures that the
Yl are dimensionless and that the expansion is in terms of direction cosines x/r,
y/r, and z/r. Up to the sixth order, there are Table 16 lists real and complex spherical
harmonics up to the sixth order.
m
3 Anisotropy and Crystal Field
175
Table A.16 Spherical harmonics
√ in several representations. For m = 0, the real representation
requires an additional factor 1/ 2, because the normalization behavior of cos(mφ) ± sin(mφ)
differs from that of exp.(imφ). Furthermore, the minus sign in N l m is not used for m = 0.
The following formulae can be used to extract the full spherical harmonics from the table:
Yl m = π-1/2 fN fP exp.(imφ), Yl m = π-1/2 fN fR /rl (m = 0), and Yl m = (2π)-1/2 fN fR /rl (m = 0).
Anisotropy energies involve even-order spherical harmonics only (gray rows)
Function
0
Y0
Y1 1
Y1 0
Y1 – 1
Y2 2
Y2 1
Y2 0
Y2 – 1
Y2 – 2
Y3 3
Y3 2
Y3 1
Y3 0
Y3 – 1
Y3 – 2
Y3 – 3
Y4 4
Y4 3
Y4 2
Y4 1
Y4 0
Y4 – 1
Y4 – 2
Y4 – 3
Y4 – 4
Y5 5
Y5 4
Y5 3
Y5 2
Y5 1
Y5 0
Y5 – 1
Y5 – 2
Y5 – 3
Y5 – 4
Y5 – 5
Y6 6
Y6 5
Y6 4
Y6 3
Y6 2
Y6 1
Y6 0
Y6 – 1
Y6 – 2
Y6 – 3
Y6 – 4
Y6 – 5
Y6 – 6
fN =
S ࣨm
fP = Plm
fR = rlYlm/ࣨ m
1/2
1
1
x
sinT
– 1/2 · 3/2
z
cosT
1/2 · 3
y
sinT
1/2 · 3/2
x2 – y2
sin2T
1/4 · 15/2
xz
sinT cosT
– 1/2 · 15/2
3 z2 – r2
3 cos2T – 1
1/4 · 5
yz
sinT cosT
1/2 · 15/2
2 xy
sin2T
1/4 · 15/2
x3 – 3 xy2
sin3T
– 1/8 · 35
2
z
(x2 – y2)
sin T cosT
1/4 · 105/2
x (5 z2 – r2)
sinT cos2T – cosT
– 1/8 · 21
z (5 z2 – 3r2)
cos3T – 3
1/4 · 7
2
y (5 z2 – r2)
sin
cos
–
cos
T
T
T
1/8 · 21
2
2 xyz
sin T cosT
1/4 · 105/2
3 x2y – y3
sin3T
1/8 · 35
x4 – 6 x2y2 + y4
sin4T
3/16 · 35/2
3
xz (x2 – 3 y3)
sin T cosT
– 3/8 · 35
2
2
2
(x
– y2) (7 z2 – r2)
sin
cos
–
1
T
T
3/8 · 5/2
3
xz (7 z2 – 3 r2)
sinT cos T – 3 cosT
– 3/8 · 5
4
2
3/16
35
z4 – 30 z2r2 + 3 r4
35 cos T – 30 cos T + 3
yz (7 z2 – 3r2)
sinT cos3T – 3 cosT
3/8 · 5
2
2
2
xy (7 z2 – r2)
sin T cos T – 1
3/8 · 5/2
yz (3 x2 – y2)
sin3T cosT
3/8 · 35
4 xy (x2 – y2)
sin4T
3/16 · 35/2
x (x4 – 10 x2y2 + 5 y4)
sin5T
– 3/32 · 77
z (x4 – 6 x2y2 + y4)
sin4T cosT
3/16 · 385/2
2
3
2
– 3 y2)·(9 z2 – r2)
x
(x
cos
–
1
sin
T
T
– 1/32 · 385
z (x2 – y2) (3 z2 – r2)
sin2T cos3T – cosT
1/8 · 1155/2
z (21 z4 – 12 z2r2 + r4)
sinT 21 cos4T – 14 cos2T + 1)
– 1/16 · 165/2
z (63 z4 – 70 z2r2 + 15 r4)
63 cos5T – 70 cos3T + 15 cosT
1/16 · 11
y (21 z4 – 12 z2r2 + r4)
sinT 21 cos4T – 14 cos2T + 1)
1/16 · 165/2
2
3
2 xyz (3z2 – r2)
sin
cos
–
cos
T
T
T
1/8 · 1155/2
y (3x2 – y2) (9z2 – r2)
sin3T cos2T – 1
1/32 · 385
4 xyz (x2 – y2)
sin4T cosT
3/16 · 385/2
y (5 x4 – 10 x2y2 + y4)
sin5T
3/32 · 77
x6 – 15 x4y2 + 15 x2y4 – y6
sin6T
1/64 · 3003
5
x (x4 – 10 x2y2 + 5 y4)
sin
cos
T
T
– 3/32 · 1001
4
4
2
(x – 6 x2y2 + y4)(11 z2 – r2)
sin T cos T – 1
3/32 · 91/2
3
3
xz (x2 – 3 y2)(11 z2 – 3 r2)
sin T cos T – 3 cosT
– 1/32 · 1365
(x2 – y2)(33 z4 – 18 z2r2 + r4)
sin2T 33 cos4T – 18 cos2T + 1)
1/64 · 1365
5
3
xz (33 z4 – 30 z2r2 + 5 r4)
– 1/16 · 273/2 sinT (33 cos T – 70 cos T + 5 cosT )
6
4
2
231
z6 – 315 z4r2 + 105 z2r2 – 5 r6
231
cos
–
315
cos
+
105
cos
–
5
T
T
T
1/32 · 13
5
3
yz (33z4 – 30z2r2 + 5r4)
sinT (33 cos T – 70 cos T + 5 cosT )
1/16 · 273/2
2
4
2
2
xy (33 z4 – 18 z2r2 + r4)
sin T 33 cos T – 18 cos T + 1)
1/64 · 1365
3 yz (3 x2 – y2)(11 z2 – 3r2)
sin3T cos3T – 3 cosT
1/32 · 1365
4 xy (x2 – y2)(11 z2 – r2)
sin4T cos2T – 1
3/32 · 91/2
zy (5 x4 – 10 x2y2 + y4)
sin5T cosT
3/32 · 1001
6
xy (6 x4 – 20 x2y2 – 6 y4)
sin T
1/64 · 3003
176
R. Skomski et al.
Appendix B: Point Groups
Table A.17 Less common space and point groups. The space groups in bold characters are
frequently encountered in magnetism and separately considered in the main text of the chapter
(Table 1)
Crystal system
Triclinic
Triclinic
Monoclinic
Monoclinic
Monoclinic
Orthorhombic
Point group
C1 (1)
Ci (1)
C2 (2)
Cs (m)
C2h (2/m)
D2 (222)
Orthorhombic
C2v (mm2)
Orthorhombic
D2h (mmm)
Tetragonal
Tetragonal
Tetragonal
Tetragonal
C4 (4)
S4 (4)
C4h (4/m)
D4 (422)
Tetragonal
C4v (4 mm)
Tetragonal
D2d (42m)
Tetragonal
D4h (4/mmm)
Trigonal
Trigonal
Trigonal
Trigonal
Trigonal
Hexagonal
Hexagonal
Hexagonal
Hexagonal
Hexagonal
Hexagonal
Hexagonal
Cubic
C3 (3)
S6 (3)
D3 (32)
C3v (3 m)
D3d (3m)
C6 (6)
C3h (6)
C6h (6/m)
D6 (622)
C6v (6 mm)
D3h (6m2)
D6h (6/mmm)
T (23)
Space group
P1
P1
P2, P21 , C2
Pm, Pc, Cm, Cc
C2/m, C2/c, P2/m, P21 /m, P2/c, P21 /c
P222, P2221 , P21 21 2, P21 21 21 , C2221 , C222, F222, I222,
I21 21 21
Pmm2, Pmc21 , Pcc2, Pma2, Pca21 , Pnc2, Pmn21 , Pba2,
Pna21 , Pnn2, Cmm2, Cmc21 , Ccc2, Amm2, Aem2, Ama2,
Aea2, Fmm2, Fdd2, Imm2, Iba2, Ima2
Pnma, Pmmm, Pnnn, Pccm, Pban, Pmma, Pnna, Pmna,
Pcca, Pbam, Pccn, Pbcm, Pnnm, Pmmn, Pbcn, Pbca, Cmcm,
Cmce, Cmmm, Cccm, Cmme, Ccce, Fmmm, Fddd, Immm,
Ibam, Ibca, Imma
P4, P41 , P42 , P43 , I4, I41
P4, I4
P4/m, P42 /m, P4/n, P42 /n, I4/m, I41 /a
P422, P421 2, P41 22, P41 21 2, P42 22, P42 21 2, P43 22,
P43 21 2, I422, I41 22
P4mm, P4bm, P42 cm, P42 nm, P4cc, P4nc, P42 mc, P42 bc,
I4mm, I4cm, I41 md, I41 cd
P42m, P42c, P421 m, P421 c, P4m2, P4c2, P4b2, P4n2, I4m2,
I4c2, I42m, I42d
P4/mmm, P42 /mnm, I4/mmm, P4/mcc, P4/nbm, P4/nnc,
P4/mbm, P4/mnc, P4/nmm, P4/ncc, P42 /mmc, P42 /mcm,
P42 /nbc, P42 /nnm, P42 /mbc, P42 /nmc, P42 /ncm, I4/mcm,
I41 /amd, I41 /acd
P3, P31 , P32 , R3
P3, R3
P32 12, P312, P321, P31 12, P31 21, P32 21, R32
P3m1, P31m, P3c1, P31c, R3m, R3c
R3m, R3c, P31m, P31c, P3m1, P3c1
P6, P61 , P65 , P62 , P64 , P63
P6
P63 /mmc, P6/m, P63 /m
P622, P61 22, P65 22, P62 22, P64 22, P63 22
P63 mc, P6mm, P6cc, P63 cm
P6m2, P6c2, P62m, P62c
P6/mcc, P63 /mcm
P21 3, P23, F23, I23, I21 3
(continued)
3 Anisotropy and Crystal Field
177
Table A.17 (Continued)
Crystal system
Cubic
Cubic
Cubic
Cubic
Point group
Td (43m)
Th (m3)
O (432)
Oh (m3m)
Space group
F43m, I43m, P43m, P43n, F43c, I43d
Pa3, Pm3, Pn3, Fm3, Fd3, Im3, Ia3
P432, P42 32, F432, F41 32, I432, P43 32, P41 32, I41 32
Fm3m, Im3m, Pm3m, Pn3m, Fd3m, Ia3d, Ia3d, Pn3n,
Pm3n, Fm3c, Fd3c
Appendix C: Hydrogen-Like Atomic 3d Wave Functions
Hydrogen-like 3d wave functions are obtained by solving the Schrödinger equation
for n = 3 (third shell) and l = 2 (d electrons). There are 2 l + 1 = 5 different orbitals,
and each can be occupied by up to two electrons. Explicitly,
where N =
√
|xy> = R3d (r)sin2 θ sin 2φ
(75)
|x 2 − y 2 > = N R3d (r) sin2 θ cos 2φ
(76)
|xz> = 2N R3d (r) sin θ cos θ cos φ
(77)
|z2 > = R3d (r) 3 sin2 θ − 1
(78)
|yz > = 2N R3d (r) sin θ cos θ sin φ
(79)
15/16π, ao = 0.529 Å, and
R3d (r) =
4Z 5/2 r 2
Zr
exp −
ao
81ao2 30ao3
(80)
Aside from the real set of wave functions, there exist complex wave functions of
the type exp.(±imφ). The two sets of wave functions are linear combinations of
each other, and both are solutions of the Schrödinger equation. However, they are
nonequivalent with respect to orbital moment and magnetic anisotropy.
More generally, Ψ (r, φ, θ ) = Rn l (r) Yl m (φ, θ ), where it is convenient to express
the radial wave functions in terms of the parameter ro = ao /Z:
2
r
R1s = exp −
ro
ro 3
178
R. Skomski et al.
R2s
1
= 2 2ro 3
R2p
R3s
r
1
r
= exp −
2ro
2 6ro 3 ro
2
= 81 3ro 3
R3p
r
r
2−
exp −
ro
2ro
r
r
r2
27 − 18 + 2 2 exp −
ro
3ro
ro
r
r
r2
6 − 2 exp −
ro
3ro
ro
4
= 81 6ro 3
R3d =
r2
4
r
exp
−
3ro
81 30ro 3 ro 2
R4f =
r3
1
r
exp
−
4ro
768 35ro 3 ro 3
From the radial wave functions, the following averages are obtained:
<r 2 > =
n2 ro 2 2
5n + 1 − 3l (l + 1)
2
<r> =
ro 2
3 n − l (l + 1)
2
<1/r> =
<1/r 2 > =
<1/r 3 > =
1
n2 r
o
2
n3 ro 2 (2l + 1)
2
n3 ro 3 l (l + 1) (l + 2)
3 Anisotropy and Crystal Field
179
These formulae have numerous applications. For example, <r> and the square
root of <r2 > are used to estimate shell radii, <1/r> gives the electronic energy,
and <1/r3 > determines the strength of the spin-orbit coupling on which magnetocrystalline anisotropy relies.
References
1. Bloch, F., Gentile, G.: Zur Anisotropie der Magnetisierung ferromagnetischer Einkristalle. Z.
Phys. 70, 395–408 (1931)
2. Jäger, E., Perthel, R.: Magnetische Eigenschaften von Festkörpern. Akademie-Verlag, Berlin
(1983)
3. Buschow, K.H.J., van Diepen, A.M., de Wijn, H.W.: Crystal-field anisotropy of Sm3+ in
SmCo5 . Solid State Commun. 15, 903–906 (1974)
4. Sankar, S.G., Rao, V.U.S., Segal, E., Wallace, W.E., Frederick, W.G.D., Garrett, H.J.:
Magnetocrystalline anisotropy of SmCo5 and its interpretation on a crystal-field model. Phys.
Rev. B. 11, 435–439 (1975)
5. Cadogan, J.M., Li, H.-S., Margarian, A., Dunlop, J.B., Ryan, D.H., Collocott, S.J., Davis,
R.L.: New rare-earth intermetallic phases R3 (Fe,M)29 Xn : (R = Ce, Pr, Nd, Sm, Gd; M = Ti,
V, Cr, Mn; and X = H, N, C) (invited). J. Appl. Phys. 76, 6138–6143 (1994)
6. Wirth, S., Wolf, M., Margarian, A., Müller, K.-H.: Determination of anisotropy constants
for monoclinic ferromagnetic compounds. In: Missell, F.P., et al. (eds.) Proc. 9th Int.
Symp. Magn. Anisotropy and Coercivity in RE-TM Allyos, pp. 399–408. World Scientific,
Singapore (1996)
7. Andreev, A.V., Deryagin, A.V., Kudrevatykh, N.V., Mushnikov, N.V., Re’imer, V.A., Terent’ev, S.V.: Magnetic properties of Y2 Fe14 B and Nd2 Fe14 B and their hydrides. Zh. Eksp.
Teor. Fiz. 90, 1042–1050 (1986)
8. Kronmüller, H.: Theory of nucleation fields in inhomogeneous ferromagnets. Phys. Status
Solidi B. 144, 385–396 (1987)
9. Skomski, R.: Nanomagnetics. J. Phys. Condens. Matter. 15, R841–R896 (2003)
10. Skomski, R., Coey, J.M.D.: Permanent Magnetism. Institute of Physics, Bristol (1999)
11. Evetts, J.E. (ed.): Concise Encyclopedia of Magnetic and Superconducting Materials. Pergamon, Oxford (1992)
12. Sucksmith, W., Thompson, J.E.: The magnetic anisotropy of cobalt. Proc. Roy. Soc. London.
A 225, 362–375 (1954)
13. Chikazumi, S.: Physics of Magnetism. Wiley, New York (1964)
14. Thole, B.T., van der Laan, G., Sawatzky, G.A.: Strong magnetic dichroism predicted in the
M4,5 X-ray absorption spectra of magnetic rare-earth materials. Phys. Rev. Lett. 55, 2086–
2089 (1985)
15. Stöhr, J.: Exploring the microscopic origin of magnetic anisotropies with X-ray magnetic
circular dichroism (XMCD) spectroscopy. J. Magn. Magn. Mater. 200, 470–497 (1999)
16. Sander, D.: The magnetic anisotropy and spin reorientation of nanostructures and nanoscale
films. J. Phys. Condens. Matter. 16, R603–R636 (2004)
17. Ballhausen, C.J.: Ligand Field Theory. McGraw-Hill, New York (1962)
18. Bethe, H.: Termaufspaltung in Kristallen. Ann. Phys. 3, 133–208 (1929)
19. Griffith, J.S., Orgel, L.E.: Ligand-field theory. Q. Rev. Chem. Soc. 11, 381–393 (1957)
20. Newman, D.J., Ng, B.: The superposition model of crystal fields. Rep. Prog. Phys. 52, 699–
763 (1989)
21. Skomski, R.: The screened-charge model of crystal-field interaction. Philos. Mag. B. 70, 175–
189 (1994)
22. Lawrence, J.M., Riseborough, P.S., Parks, R.D.: Valence fluctuation phenomena. Rep. Prog.
Phys. 44, 1–84 (1981)
180
R. Skomski et al.
23. Fulde, P.: Electron Correlations in Molecules and Solids. Springer, Berlin (1991)
24. Reinhold, J.: Quantentheorie der Moleküle. Teubner, Stuttgart (1994)
25. Goodenough, J.B.: Magnetism and the Chemical Bond. Wiley, New York (1963)
26. Herzberg, G.: Atomic Spectra and Atomic Structure. Dover, New York (1944)
27. Pauling, L., Wilson, E.B.: Introduction to Quantum Mechanics. McGraw-Hill, New York
(1935)
28. Racah, G.: Theory of complex spectra. II. Phys. Rev. 62, 438–462 (1942)
29. Jones, W., March, N.H.: Theoretical Solid State Physics I. Wiley & Sons, London (1973)
30. Bychkov, Y.A., Rashba, E.I.: Properties of a 2D electron gas with lifted spectral degeneracy.
JETP Lett. 39, 78–81 (1984)
31. Skomski, R.: The itinerant limit of metallic anisotropy. IEEE Trans. Magn. 32(5), 4794–4796
(1996)
32. Montalti, M., Credi, A., Prodi, L., Teresa Gandolfi, M.: Handbook of Photochemistry. CRC
Press, Boca Raton (2006)
33. Taylor, K.N.R., Darby, M.I.: Physics of Rare Earth Solids. Chapman and Hall, London (1972)
34. Cole, G.M., Garrett, B.B.: Atomic and molecular spin-orbit coupling constants for 3d
transition metal ions. Inorg. Chem. 9, 1898–1902 (1970)
35. van Vleck, J.H.: On the anisotropy of cubic ferromagnetic crystal. Phys. Rev. 52, 1178–1198
(1937)
36. Skomski, R., Istomin, A.Y., Starace, A.F., Sellmyer, D.J.: Quantum entanglement of
anisotropic magnetic nanodots. Phys. Rev. A. 70, 062307 (2004)
37. Coey, J.M.: Magnetism and Magnetic Materials. University Press, Cambridge (2010)
38. Hutchings, M.T.: Point-charge calculations of energy levels of magnetic ions in crystalline
electric fields. Solid State Phys. 16, 227–273 (1964)
39. Yamada, M., Kato, H., Yamamoto, H., Nakagawa, Y.: Crystal-field analysis of the magnetization process in a series of Nd2 Fe14 B-type compounds. Phys. Rev. B. 38, 620–633 (1988)
40. Herbst, J.F.: R2 Fe14 B materials: intrinsic properties and technological aspects. Rev. Mod.
Phys. 63, 819–898 (1991)
41. Coey, J.M.D., Sun, H.: Improved magnetic properties by treatment iron-based rare-earth
intermetallic compounds in ammonia. J. Magn. Magn. Mater. 87, L251–L254 (1990)
42. Duc, N.H., Hien, T.D., Givord, D., Franse, J.J.M., de Boer, F.R.: Exchange interactions in
rare-earth transition-metal compounds. J. Magn. Magn. Mater. 124, 305–311 (1993)
43. Skomski, R.: Finite-temperature behavior of anisotropic two-sublattice magnets. J. Appl.
Phys. 83, 6724–6726 (1998)
44. Akulov, N.: Zur Quantentheorie der Temperaturabhängigkeit der Magnetisierungskurve. Z.
Phys. 100, 197–202 (1936)
45. Callen, E.R.: Temperature dependence of ferromagnetic uniaxial anisotropy constants. J.
Appl. Phys. 33, 832–835 (1962)
46. Callen, H.R., Callen, E.: The present status of the temperature dependence of the magnetocrystalline anisotropy and the l(l+1) power law. J. Phys. Chem. Solids. 27, 1271–1285
(1966)
47. O’Handley, R.C.: Modern Magnetic Materials. Wiley, New York (2000)
48. Kronmüller, H., Fähnle, M.: Micromagnetism and the Microstructure of Ferromagnetic
Solids. University Press, Cambridge (2003)
49. Kumar, K.: RETM5 and RE2 TM17 permanent magnets development. J. Appl. Phys. 63, R13–
R57 (1988)
50. Skomski, R.: Exchange-controlled magnetic anisotropy. J. Appl. Phys. 91, 8489–8491 (2002)
51. Skomski, R., Kashyap, A., Sellmyer, D.J.: Finite-temperature anisotropy of PtCo magnets.
IEEE Trans. Magn. 39(5), 2917–2919 (2003)
52. Skomski, R., Mryasov, O.N., Zhou, J., Sellmyer, D.J.: Finite-temperature anisotropy of
magnetic alloys. J. Appl. Phys. 99, 08E916-1-4 (2006)
53. Brooks, H.: Ferromagnetic anisotropy and the itinerant electron model. Phys. Rev. 58, 909–
918 (1940)
3 Anisotropy and Crystal Field
181
54. Bruno, P.: Tight-binding approach to the orbital magnetic moment and magnetocrystalline
anisotropy of transition-metal monolayers. Phys. Rev. B. 39, 865–868 (1989)
55. Skomski, R., Kashyap, A., Enders, A.: Is the magnetic anisotropy proportional to the orbital
moment? J. Appl. Phys. 109, 07E143-1-3 (2011)
56. White, R.M.: Quantum Theory of Magnetism. McGraw-Hill, New York (1970)
57. Fuchikami, N.: Magnetic anisotropy of Magnetoplumbite BaFe12 O19 . J. Phys. Soc. Jpn. 20,
760–769 (1965)
58. Weller, D., Moser, A., Folks, L., Best, M.E., Lee, W., Toney, M.F., Schwickert, M., Thiele,
J.-U., Doerner, M.F.: High Ku materials approach to 100 Gbits/inz. TEEE Trans. Magn. 36,
10–15 (2000)
59. Brooks, M.S.S., Johansson, B.: Density functional theory of the ground-state magnetic
properties of rare earths and actinides. In: Buschow, K.H.J. (ed.) Handbook of Magnetic
Materials, vol. 7, pp. 139–230. Elsevier, Amsterdam
60. Wang, D.-S., Wu, R.-Q., Freeman, A.J.: First-principles theory of surface magnetocrystalline
anisotropy and the diatomic-pair model. Phys. Rev. B. 47, 14932–14947 (1993)
61. Skomski, R.: Magnetoelectric Néel Anisotropies. IEEE Trans. Magn. 34(4), 1207–1209
(1998)
62. Mattheiss, L.F.: Electronic structure of the 3d transition-metal monoxides: I. energy band
results. Phys. Rev. B. 5, 290–306 (1972).; “II. Interpretation”, ibidem 306–315
63. Newnham, R.E.: Properties of Materials: Anisotropy, Symmetry, Structure. University Press,
Oxford (2004)
64. Slater, J.C., Koster, G.F.: Simplified LCAO method for the periodic potential problem. Phys.
Rev. 94, 1498–1524 (1954)
65. Sutton, A.P.: Electronic Structure of Materials. University Press, Oxford (1993)
66. Skomski, R., Kashyap, A., Solanki, A., Enders, A., Sellmyer, D.J.: Magnetic anisotropy in
itinerant magnets. J. Appl. Phys. 107, 09A735-1-3 (2010)
67. Yamada, Y., Suzuki, T., Kanazawa, H., Österman, J.C.: The origin of the large perpendicular
magnetic anisotropy in Co3 Pt alloy thin films. J. Appl. Phys. 85, 5094 (1999)
68. Lewis, L.H., Pinkerton, F.E., Bordeaux, N., Mubarok, A., Poirier, E., Goldstein, J.I., Skomski,
R., Barmak, K.: De Magnete et meteorite: cosmically motivated materials. IEEE Magn. Lett.
5, 5500104-1-4 (2014)
69. Fisher, J.E., Goddard, J.: Magnetocrystalline and uniaxial anisotropy in electrodeposited
single crystal films of cobalt and nickel. J. Phys. Soc. Jpn. 25, 413–418 (1968)
70. Bozorth, R.M.: Ferromagnetism. van Nostrand, Princeton (1951)
71. Takahashi, H., Igarashi, M., Kaneko, A., Miyajima, H., Sugita, Y.: Perpendicular uniaxial
magnetic anisotropy of Fe16 N2 (001) single crystal films grown by molecular beam Epitaxy.
IEEE Trans. Magn. 35(5), 2982–2984 (1999)
72. Coene, W., Hakkens, F., Coehoorn, R., de Mooij, B.D., De Waard, C., Fidler, J., Grössinger,
R.: J. Magn. Magn. Mater. 96, 189–196 (1991)
73. Kneller, E.F., Hawig, R.: The exchange-spring magnet: a new material principle for permanent
magnets. IEEE Trans. Magn. 27, 3588–3600 (1991)
74. Klemmer, T., Hoydick, D., Okumura, H., Zhang, B., Soffa, W.A.: Magnetic hardening and
coercivity mechanisms in L10 ordered FePd Ferromagnets. Scr. Met. Mater. 33, 1793–1805
(1995)
75. Willoughby, S.D., Stern, R.A., Duplessis, R., MacLaren, J.M., McHenry, M.E., Laughlin,
D.E.: Electronic structure calculations of hexagonal and cubic phases of Co3 Pt. J. Appl. Phys.
93, 7145–7147 (2003)
76. Coey, J.M.D.: New permanent magnets: manganese compounds. J. Phys. Condens. Matter.
26, 064211-1-6 (2014)
77. Lessard, A., Moos, T.H., Hübner, W.: Magnetocrystalline anisotropy energy of transitionmetal thin films: a nonperturbative theory. Phys. Rev. B. 56, 2594–2604 (1997)
78. Skomski, R., Wei, X., Sellmyer, D.J.: Band-structure and correlation effects in the Co(111)
planes of CoO. J. Appl. Phys. 103, 07C908-1-3 (2008)
182
R. Skomski et al.
79. Fletcher, G.C.: Calculations of the first ferromagnetic anisotropy coefficient, gyromagnetic
ratio and spectroscopic splitting factor for nickel. Proc. Phys. Soc. A. 67, 505–519 (1954).
An erratum, yielding reductions of SOC constant and K1 by factors of 2 and 4, respectively,
was subsequently published as a letter to the editor: G. C. Fletcher, Proc. Phys. Soc. 78, 145
(1961)
80. Kondorski, E.M., Straube, E.: Magnetic anisotropy of nickel. Zh. Eksp. Teor. Fiz. 63, 356–365
(1972) [Sov. Phys. JETP 36, 188 (1973)]
81. Takayama, H., Bohnen, K.P., Fulde, P.: Magnetic surface anisotropy of transition metals.
Phys. Rev. B. 14, 2287–2295 (1976)
82. Trygg, J., Johansson, B., Eriksson, O., Wills, J.M.: Total energy calculation of the Magnetocrystalline anisotropy energy in the ferromagnetic 3d metals. Phys. Rev. Lett. 75,
2871–2875 (1995)
83. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: First-principle calculation of the
Magnetocrystalline anisotropy energy of Iron, cobalt, and nickel. Phys. Rev. B. 41, 11919–
11937 (1990)
84. Yang, I., Savrasov, S.Y., Kotliar, G.: Importance of correlation effects on magnetic anisotropy
in Fe and Ni. Phys. Rev. Lett. 87, 216405 (2001)
85. Zhao, X., Nguyen, M.C., Zhang, W.Y., Wang, C.Z., Kramer, M.J., Sellmyer, D.J., Li, X.Z.,
Zhang, F., Ke, L.Q., Antropov, V.P., Ho, K.M.: Exploring the structural complexity of
intermetallic compounds by an adaptive genetic algorithm. Phys. Rev. Lett. 112, 045502-1-5
(2014)
86. Gay, J.G., Richter, R.: Spin anisotropy of ferromagnetic films. Phys. Rev. Lett. 56, 2728–2731
(1986)
87. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: First-principle calculation of the
magnetic anisotropy energy of (Co)n /(X)m multilayers. Phys. Rev. B. 42, 7270–7273
(1990)
88. Johnson, M.T., Bloemen, P.J.H., den Broeder, F.J.A., de Vries, J.J.: Magnetic anisotropy in
metallic multilayers. Rep. Prog. Phys. 59, 1409–1458 (1996)
89. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: Magnetocrystalline anisotropy of YCo5
and related RECo5 compounds. Phys. Rev. B. 53, 14415–14433 (1996)
90. Manchanda, P., Kumar, P., Kashyap, A., Lucis, M.J., Shield, J.E., Mubarok, A., Goldstein, J.,
Constantinides, S., Barmak, K., Lewis, L.-H., Sellmyer, D.J., Skomski, R.: Intrinsic properties
of Fe-substituted L10 magnets. IEEE Trans. Magn. 49, 5194–5198 (2013)
91. Kumar, P., Kashyap, A., Balamurugan, B., Shield, J.E., Sellmyer, D.J., Skomski, R.:
Permanent magnetism of intermetallic compounds between light and heavy transition-metal
elements. J. Phys. Condens. Matter. 26, 064209-1-8 (2014)
92. Wang, X., Wang, D., Wu, R., Freeman, A.J.: Validity of the force theorem for magnetocrystalline anisotropy. J. Magn. Magn. Mater. 159, 337–341 (1996)
93. Richter, M.: Band structure theory of magnetism in 3d-4f compounds. J. Phys. D. Appl. Phys.
31, 1017–1048 (1998)
94. Eriksson, O., Johansson, B., Albers, R.C., Boring, A.M., Brooks, M.S.S.: Orbital magnetism
in Fe, Co, and Ni. Phys. Rev. B. 42, 2707–2710 (1990)
95. Lieb, E.H.: Density functionals for coulomb systems. Int. J. Quantum Chem. 24, 243–277
(1983)
96. Skomski, R., Manchanda, P., Kashyap, A.: Correlations in rare-earth transition-metal permanent magnets. J. Appl. Phys. 117, 17C740-1-4 (2015)
97. Anisimov, V.I., Zaanen, J., Andersen, O.K.: Band theory and Mott insulators: Hubbard U
instead of stoner I. Phys. Rev. B. 44, 943–954 (1991)
98. Lai, W.Y., Zheng, Q.Q., Hu, W.Y.: The giant magnetic moment and electronic correlation
effect in ferromagnetic nitride Fe16 N2 . J. Phys. Condens. Matter. 6, L259–L264 (1994)
99. Zhu, J.-X., Janoschek, M., Rosenberg, R., Ronning, F., Thompson, J.D., Torrez, M.A.,
Bauer, E.D., Batista, C.D.: LDA+DMFT approach to Magnetocrystalline anisotropy of strong
magnets. Phys. Rev. X. 4, 021027-1-7 (2014)
3 Anisotropy and Crystal Field
183
100. Freeman, A.J., Wu, R.-Q., Kima, M., Gavrilenko, V.I.: Magnetism, magnetocrystalline
anisotropy, magnetostriction and MOKE at surfaces and interfaces. J. Magn. Magn. Mater.
203, 1–5 (1999)
101. Larson, P., Mazin, I.I.: Calculation of magnetic anisotropy energy in YCo5 . J. Magn. Magn.
Mater. 264, 7–13 (2003)
102. Skomski, R., Sharma, V., Balamurugan, B., Shield, J.E., Kashyap, A., Sellmyer, D.J.:
Anisotropy of doped transition-metal magnets. In: Kobe, S., McGuinness, P. (eds.) Proc.
REPM’10, pp. 55–60. Jozef Stefan Institute, Ljubljana (2010)
103. Fast, J.D.: Gases in Metals. Macmillan, London (1976)
104. Yensen, T.D.: Development of magnetic material. Electr J. 18, 93–95 (1921)
105. Burkert, T., Nordström, L., Eriksson, O., Heinonen, O.: Giant magnetic anisotropy in
tetragonal FeCo alloys. Phys. Rev. Lett. 93, 027203-1-4 (2004)
106. Andersson, G., Burkert, T., Warnicke, P., Björck, M., Sanyal, B., Chacon, C., Zlotea, C.,
Nordström, L., Nordblad, P., Eriksson, O.: Perpendicular Magnetocrystalline anisotropy in
Tetragonally distorted Fe-Co alloys. Phys. Rev. Lett. 96, 037205-1-4 (2006)
107. Jack, K.W.: The iron—nitrogen system: the crystal structures of ε-phase iron nitrides. Act
Crystallogr. 5, 404–411 (1952)
108. The α”-Fe16 N2 unit cell contains two Fe8 N formula units. However, quoting unit cells is an
awkward notation to describe stoichiometries and rarely used. For example, one generally
writes Nd2 Fe14 B and Sm2 Co17 rather than Nd8 Fe56 B4 and Sm6 Co51
109. Coey, J.M.D., O’Donnell, K., Qinian, Q., Touchais, E., Jack, K.H.: The magnetization of
α”Fe16 N2 . J. Phys. Condens. Matter. 6, L23–L28 (1994)
110. Kim, T.K., Takahashi, M.: New magnetic material having ultrahigh magnetic moment. Appl.
Phys. Lett. 20, 492–494 (1972)
111. Ji, N., Liu, X., Wang, J.-P.: Theory of giant saturation magnetization in α”-Fe16 N2 : role of
partial localization in ferromagnetism of 3d transition metals. New J. Phys. 12, 063032-1-8
(2010)
112. Ke, L.-Q., Belashchenko, K.D., van Schilfgaarde, M., Kotani, T., Antropov, V.P.: Effects of
alloying and strain on the magnetic properties of Fe16 N2 . Phys. Rev. B. 88, 024404-1-9 (2013)
113. Trommler, S., Hänisch, J., Matias, V., Hühne, R., Reich, E., Iida, K., Haindl, S., Schultz, L.,
Holzapfel, B.: Architecture, microstructure and Jc anisotropy of highly oriented biaxially textured co-doped BaFe2 As2 on Fe/IBAD-MgO-buffered metal tapes. Supercond. Sci. Technol.
25, 84019-1-7 (2012)
114. Osborn, J.A.: Demagnetizing factors of the general ellipsoid. Phys. Rev. 67, 351–357 (1945)
115. McCurrie, R.A.: Ferromagnetic Materials—Structure and Properties. Academic Press, London (1994)
116. Skomski, R., Liu, Y., Shield, J.E., Hadjipanayis, G.C., Sellmyer, D.J.: Permanent magnetism
of dense-packed nanostructures. J. Appl. Phys. 107, 09A739-1-3 (2010)
117. Skomski, R., Manchanda, P., Kumar, P., Balamurugan, B., Kashyap, A., Sellmyer, D.J.:
Predicting the future of permanent-magnet materials (invited). IEEE Trans. Magn. 49, 3215–
3220 (2013)
118. Zhou, L., Miller, M.K., Lu, P., Ke, L., Skomski, R., Dillon, H., Xing, Q., Palasyuk,
A., McCartney, M.R., Smith, D.J., Constantinides, S., McCallum, R.W., Anderson, I.E.,
Antropov, V., Kramer, M.J.: Architecture and magnetism of alnico. Acta Mater. 74, 224–233
(2014)
119. Néel, L.: Anisotropie Magnétique Superficielle et Surstructures d’Orientation. J. Phys.
Radium. 15, 225–239 (1954)
120. Skomski, R.: Interstitial modification. In: Coey, J.M.D. (ed.) Rare-Earth—Iron Permanent
Magnets, pp. 178–217. University Press, Oxford (1996)
121. Jensen, J., Mackintosh, A.R.: Rare Earth Magnetism: Structures and Excitations. Clarendon,
Oxford (1991)
122. Wijn, H.P.J. (ed.): Magnetic Properties of Metals: d-Elements, Alloys, and Compounds.
Springer, Berlin (1991)
184
R. Skomski et al.
123. de Jongh, L.J., Miedema, A.R.: Experiments on simple magnetic model systems. Adv. Phys.
23, 1–260 (1974)
124. Sellmyer, D.J., Nafis, S.: Phase transition behavior in a random-anisotropy system. Phys. Rev.
Lett. 57, 1173–1176 (1986)
125. Ising, E.: Beitrag zur Theorie des Ferromagnetismus. Z. Phys. 31, 253–258 (1925)
126. Brush, S.G.: History of the Lenz-Ising model. Rev. Mod. Phys. 39, 883–893 (1967)
127. Yeomans, J.M.: Statistical Mechanics of Phase Transitions. University Press, Oxford (1992)
128. Binek, C.: Ising-Type Antiferromagnets: Model Systems in Statistical Physics and in the
Magnetism of Exchange Bias Springer Tracts in Modern Physics 196. Springer, Berlin (2003)
129. Skomski, R.: Simple Models of Magnetism. University Press, Oxford (2008)
130. Sachdev, S.: Quantum Phase Transitions. University Press, Cambridge (1999)
131. Skomski, R.: On the Ising character of the quantum-phase transition in LiHoF4 . AIP Adv. 6,
055704-1-5 (2016)
132. Stevens, K.W.H.: A note on exchange interactions. Rev. Mod. Phys. 25, 166 (1953)
133. Dzyaloshinsky, I.: A thermodynamic theory of ‘weak’ ferromagnetism of antiferromagnetics.
J. Phys. Chem. Solids. 4, 241–255 (1958)
134. Moriya, T.: Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120,
91–98 (1960)
135. Fischer, K.-H., Hertz, A.J.: Spin Glasses. University Press, Cambridge (1991)
136. Ullah, A., Balamurugan, B., Zhang, W., Valloppilly, S., Li, X.-Z., Pahari, R., Yue, L.P., Sokolov, A., Sellmyer, D.J., Skomski, R.: Crystal structure and Dzyaloshinski–Moriya
micromagnetics. IEEE Trans. Magn. 55, 7100305-1-5 (2019)
137. Moorjani, K., Coey, J.M.D.: Magnetic Glasses. Elsevier, Amsterdam (1984)
138. Skomski, R., Honolka, J., Bornemann, S., Ebert, H., Enders, A.: Dzyaloshinski–Moriya
micromagnetics of magnetic surface alloys. J. Appl. Phys. 105, 07D533-1-3 (2009)
139. Bak, P., Jensen, H.H.: Theory of helical magnetic structures and phase transitions in MnSi
and FeGe. J. Phys. C. 13, L881–L885 (1980)
140. Meiklejohn, W.H., Bean, C.P.: New magnetic anisotropy. Phys. Rev. 102, 1413–1414 (1956)
141. Kneller, E.: Ferromagnetismus. Springer, Berlin (1962)
142. Balamurugan, B., Mukherjee, P., Skomski, R., Manchanda, P., Das, B., Sellmyer, D.J.:
Magnetic nanostructuring and overcoming Brown’s paradox to realize extraordinary hightemperature energy products. Sci. Rep. 4, 6265-1-6 (2014)
143. Balamurugan, B., Das, B., Shah, V.R., Skomski, R., Li, X.Z., Sellmyer, D.J.: Assembly of
uniaxially aligned rare-earth-free nanomagnets. Appl. Phys. Lett. 101, 122407-1-5 (2012)
144. Enders, A., Skomski, R., Honolka, J.: Magnetic surface nanostructures. J. Phys. Condens.
Matter. 22, 433001-1-32 (2010)
145. Bander, M., Mills, D.L.: Ferromagnetism of ultrathin films. Phys. Rev. B. 38, 12015–12018
(1988)
146. Shen, J., Skomski, R., Klaua, M., Jenniches, H., Manoharan, S.S., Kirschner, J.: Magnetism
in one dimension: Fe on Cu(111). Phys. Rev. B. 56, 2340–2343 (1997)
147. Gradmann, U.: Magnetism in ultrathin transition metal films. In: Buschow, K.H.J. (ed.)
Handbook of Magnetic Materials, vol. 7, pp. 1–95. Elsevier, Amsterdam (1993)
148. Sander, D., Skomski, R., Schmidthals, C., Enders, A., Kirschner, J.: Film stress and domain
wall pinning in Sesquilayer iron films on W(110). Phys. Rev. Lett. 77, 2566–2569 (1996)
149. Chuang, D.S., Ballentine, C.A., O’Handley, R.C.: Surface and step magnetic anisotropy. Phys.
Rev. B. 49, 15084–15095 (1994)
150. Coehoorn, R., de Mooij, D.B., Duchateau, J.P.W.B., Buschow, K.H.J.: Novel permanent
magnetic materials made by rapid quenching. J. Phys. 49(C-8), 669–670 (1988)
151. Callen, E., Liu, Y.J., Cullen, J.R.: Initial magnetization, remanence, and coercivity of the
random anisotropy amorphous ferromagnet. Phys. Rev. B. 16, 263–270 (1977)
152. Schneider, J., Eckert, D., Müller, K.-H., Handstein, A., Mühlbach, H., Sassik, H., Kirchmayr,
H.R.: Magnetization processes in Nd4 Fe77 B19 permanent magnetic materials. Materials Lett.
9, 201–203 (1990)
3 Anisotropy and Crystal Field
185
153. Manaf, A., Buckley, R.A., Davies, H.A.: New nanocrystalline high-remanence Nd-Fe-B
alloys by rapid solidification. J. Magn. Magn. Mater. 128, 302–306 (1993)
154. Müller, K.-H., Eckert, D., Wendhausen, P.A.P., Handstein, A., Wolf, M.: Description of
texture for permanent magnets. IEEE Trans. Magn. 30, 586–588 (1994)
155. Herzer, G.: Nanocrystalline soft magnetic materials. J. Magn. Magn. Mater. 112, 258–262
(1992)
156. Alben, R., Becker, J.J., Chi, M.C.: Random anisotropy in amorphous ferromagnets. J. Appl.
Phys. 49, 1653–1658 (1978)
157. Balasubramanian, B., Skomski, R., Li, X.-Z., Valloppilly, S.R., Shield, J.E., Hadjipanayis,
G.C., Sellmyer, D.J.: Cluster synthesis and direct ordering of rare-earth transition-metal
Nanomagnets. Nano Lett. 11, 1747–1752 (2011)
158. Imry, Y., Ma, S.-K.: Random-field instability of of the ordered state of continuos symmetry.
Phys. Rev. Lett. 35, 1399–1401 (1975)
159. Harris, R., Plischke, M., Zuckermann, M.J.: New model for amorphous magnetism. Phys.
Rev. Lett. 31, 160–162 (1973)
160. Skomski, R.: Spin-glass permanent magnets. J. Magn. Magn. Mater. 157-158, 713–714
(1996)
161. Chudnovsky, E.M., Saslow, W.M., Serota, R.A.: Ordering in ferromagnets with random
anisotropy. Phys. Rev. B. 33, 251–261 (1986)
162. Richter, J., Skomski, R.: Antiferromagnets with random anisotropy. Phys. Status Solidi B.
153, 711–719 (1989)
163. Gambaradella, P., Rusponi, S., Veronese, M., Dhesi, S.S., Grazioli, C., Dallmeyer, A., Cabria,
I., Zeller, R., Dederichs, P.H., Kern, K., Carbone, C., Brune, H.: Giant magnetic anisotropy of
single co atoms and nanoparticles. Science. 300, 1130–1133 (2003)
164. Kumar, P., Skomski, R., Manchanda, P., Kashyap, A.: Ab-initio study of anisotropy and
nonuniaxial anisotropy coefficients in Pd nanochains. Chem. Phys. Lett. 583, 109–113 (2013)
165. Dorantes-Dávila, J., Pastor, G.M.: Magnetic anisotropy of one-dimensional nanostructures of
transition metals. Phys. Rev. Lett. 81, 208–211 (1998)
166. Mokrousov, Y., Bihlmayer, G., Heinze, S., Blügel, S.: Giant Magnetocrystalline anisotropies
of 4d transition-metal Monowires. Phys. Rev. Lett. 96, 147201-1-4 (2006)
167. Rau, I.G., Baumann, S., Rusponi, S., Donati, F., Stepanow, S., Gragnaniello, L., Dreiser, J.,
Piamonteze, C., Nolting, F., Gangopadhyay, S., Albertini, O.R., Macfarlane, R.M., Lutz, C.P.,
Jones, B.A., Gamberdella, P., Heinrich, A.J., Brune, H.: Reaching the magnetic anisotropy
limit of a 3d metal atom. Science. 344, 988–992 (2014)
168. Weisheit, M., Fähler, S., Marty, A., Souche, Y., Poinsignon, C., Givord, D.: Electric fieldinduced modification of magnetism in thin-film Ferromagnets. Science. 315, 349–351 (2007)
169. Shiota, Y., Maruyama, T., Nozaki, T., Shinjo, T., Shiraishi, M., Suzuki, Y.: Voltage-assisted
magnetization switching in ultrathin Fe80 Co20 alloy layers. Appl. Phys. Express. 2, 0630011-3 (2009)
170. Manchanda, P., Kumar, P., Fangohr, H., Sellmyer, D.J., Kashyap, A., Skomski, R.: Magnetoelectric control of surface anisotropy and nucleation modes in L10 -CoPt thin films. IEEE
Magn. Lett. 5, 2500104-1-4 (2014)
171. Manchanda, P., Skomski, R., Prabhakar, A., Kashyap, A.: Magnetoelectric effect in Fe linear
chains on Pt (001). J. Appl. Phys. 115, 17C733-1-3 (2014)
172. Manchanda, P., Kumar, P., Skomski, R., Kashyap, A.: Magnetoelectric effect in organometallic vanadium-benzene wires. Chem. Phys. Lett. 568, 121–124 (2013)
4
Electronic Structure: Metals and Insulators
Hubert Ebert, Sergiy Mankovsky , and Sebastian Wimmer
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Electronic Structure Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin Density Functional Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Band Structure Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Relativistic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Adiabatic Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Itinerant Magnetism of Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Stoner Model of Itinerant Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Slater-Pauling Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Heusler Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Total Electronic Energy and Magnetic Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Total Electronic Energy and Magnetic Ground State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Exchange Coupling Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magneto-Crystalline Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Excitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnon Dispersion Relations Based on the Rigid Spin Approximation . . . . . . . . . . . . . . .
Spin Spiral Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Excitation Spectra Based on the Dynamical Susceptibility . . . . . . . . . . . . . . . . . . . . . . . . .
Finite-Temperature Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Methods Relying on the Rigid Spin Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Methods Accounting for Longitudinal Spin Fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . .
Coherent Treatment of Electronic Structure and Spin Statistics . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
188
190
190
196
201
204
208
208
211
213
217
218
220
224
230
231
233
237
240
240
242
244
249
H. Ebert () · S. Mankovsky · S. Wimmer
München, Department Chemie, Ludwig-Maximilians-Universität, München, Germany
e-mail: [email protected]; [email protected];
[email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_4
187
188
H. Ebert et al.
Abstract
This chapter gives an overview on the various methods used to deal with
the electronic properties of magnetic solids. This covers the treatment of
noncollinear magnetism, structural and spin disorder, as well as relativistic and
many-body effects. An introduction to the Stoner theory for itinerant or band
magnetism is followed by a number of examples with an emphasis on transition
metal-based systems. The direct connection of the total electronic energy in the
ground state and its magnetic configuration is considered next. This includes
mapping the dependence of the energy on the spin configuration on a simplified
spin Hamiltonian as provided, for example, by the Heisenberg model. Another
important issue in this context is magnetic anisotropy. As it is shown, considering
excitations from a suitable reference state provides a powerful tool to search
for stable phases, while calculating the wave vector- and frequency-dependent
susceptibility gives a sound basis to understand the dynamical properties of
magnetic solids. Finally, magnetism at finite temperature is dealt with starting
from a pure classical treatment of the problem and ending with schemes that deal
with quantum mechanics and statistics in a coherent way.
Introduction
Theory and modeling always played an important role for the understanding and
development of magnetism [1, 2, 3, 4]. An early example for this is the presence of
ring currents suggested by Ampère to explain the properties of permanent magnetic
materials. Another example is the introduction of the molecular field by Weiss when
discussing magnetism at finite temperature. Another important milestone in the theory of magnetism is the Bohr-van Leeuwen theorem [3] that unambiguously made
clear that magnetism is a quantum mechanical phenomenon and for that reason
requires a corresponding description. In line with this, Heisenberg’s investigation on
the relation between the energy and the spin configuration clearly demonstrated that
spontaneous spin-magnetic ordering is connected with the exchange interaction and
is not due to the much weaker dipole-dipole interaction. Interestingly, the existence
of the electronic spin follows directly from the Dirac equation that is the proper
relativistic counterpart of Schrödinger’s equation [5]. Another important direct
consequence of Dirac’s equation is the presence of spin-orbit coupling [5] that
gives rise to many technologically important phenomena as the magnetocrystalline
anisotropy, magnetostriction [6, 7], anomalous Hall [8], and magneto-optical Kerreffect [9]. Although these phenomena were discovered already in the nineteenth
century, a proper explanation could be given only much later on the basis of quantum
mechanics. Spin-orbit coupling is also the origin of many other important new
effects that are exploited in the field of spintronics as, for example, the spin Hall
effect [10] and spin-orbit torque [11]. In addition, one has to mention the spin-orbitinduced anisotropy in the magnetic exchange interaction. Apart from adding to the
magnetocrystalline anisotropy, this includes the so-called Dzyaloshinskii-Moriya
4 Electronic Structure: Metals and Insulators
189
interaction that is responsible for chiral spin configurations leading in particular to
skyrmionic spin structures [12]. Starting from a description of electron-electron
interaction in the framework of quantum electrodynamics, one is led to the Breit
interaction that can be seen as a current-current interaction [13]. This correction to
the isotropic electron-electron Coulomb interaction is anisotropic and leads to the
magnetic shape anisotropy [14].
Apart from explaining magnetic phenomena on a fundamental level in detail,
theory nowadays provides in most cases a quantitative description of these. Corresponding numerical studies are in general based on a treatment of the underlying
electronic structure within the framework of spin-density functional theory [15] to
deal with electronic exchange and correlation. In spite of the many successes of
this ab initio approach to magnetism, there are several limitations. This concerns
first of all the impact of strong correlations or many-body effects that are often
discussed on the basis of simplified models or hybrid schemes as the LDA+DMFT
[16] (see section “Spin Density Functional Theory”). In addition, there are still open
questions. For example, a coherent definition of orbital magnetism [17] and with this
a prescription for its calculation were suggested only recently. Another important
field to mention in this context is magnetism at finite temperature. Although the
necessary quantum-statistical formalism [4] is available, one is in practice most
often forced to use approximate schemes. These include in particular statistical or
dynamical simulations on the basis of quasi-classical spin Hamiltonians. Ab initio
theory is still extremely helpful in this case as it provides realistic parameters. This
holds true, for example, for the exchange coupling tensor [18] and the anisotropy
constant entering the extended Heisenberg Hamiltonian or for the Gilbert damping
parameter occurring within the Landau-Lifshitz-Gilbert equation [19].
Starting from a basic knowledge in solid state theory [20], this chapter gives
an overview on the methods used to deal with the electronic properties of magnetic solids (section “Electronic Structure Theory”). This covers in particular
the treatment of noncollinear magnetism, structural and spin disorder, as well as
relativistic and many-body effects. An introduction to the Stoner theory for itinerant
or band magnetism is followed by a number of examples with an emphasis on
transition metal-based systems (section “Itinerant Magnetism of Solids”). The direct
connection of the total electronic energy of the ground state and its magnetic
configuration is considered next (section “Total Electronic Energy and Magnetic
Configuration”). This includes the mapping of the complex energy landscape representing the dependence on the spin configuration on a simplified spin Hamiltonian
as provided, for example, by the Heisenberg model. Another important issue of this
section will be magnetic anisotropy. As it will be shown, considering excitations
from a suitable reference state provides a very powerful tool to search for stable
phases. Calculation of the wave vector- and frequency-dependent susceptibility
provides a sound basis for understanding the dynamical properties of magnetic
solids (section “Excitations”). Finally, magnetism at finite temperature is dealt with
in the last section that presents a series of methods starting from a pure classical
treatment of the problem and leading to schemes that deal with quantum mechanics
and statistics in a coherent way (section “Finite-Temperature Magnetism”).
190
H. Ebert et al.
Electronic Structure Theory
Dealing with the electronic structure of magnetic solids requires to account for
exchange and correlation and to solve the resulting electronic structure problem.
Spin density functional theory provides a powerful and flexible platform to tackle
the first issue while still allowing for extensions aiming at an improved treatment
of correlation effects. Concerning the second issue, there are many band structure
schemes now available that provide the necessary accuracy. In particular, they allow
dealing with two important aspects that are often of crucial importance for magnetic
properties, namely, disorder and relativistic effects.
Spin Density Functional Theory
Most computational investigations on the electronic structure of magnetic materials
are nowadays based on density functional theory (DFT) [15] or extensions to it.
The major goal of this approach is to reduce the complicated many-body problem
connected with the electron system of an atom, molecule, or solid effectively to a
single-particle problem. The formal basis for this tremendous simplification is laid
by the theorems of Kohn and Hohenberg [21] that introduce the electron density
n(r) as the basic system variable:
1. The total ground state energy E of any many-body system is a functional of the
density n(r)
E[n] = F [n] +
d 3 r n(r) Vext (r),
(1)
where Vext (r) is an arbitrary external potential, in general the Coulomb potential
of the nuclei, and F [n] itself is a functional of the density n(r) but does not
depend on Vext (r).
2. For any many-electron system, the functional E[n] for the total energy has a
minimum equal to the ground-state energy E0 = E[n0 ] at the ground-state
density n0 (r).
Applying the variational principle to the minimal property of the energy functional
Kohn and Sham [22] derived Schrödinger-like single-particle equations whose
solution allows calculating any property of the system. For this purpose, the
functional F [n] is split into three parts:
F [n] = T [n] +
3
d r
d 3r n(r) n(r )
+ Exc [n],
|r − r |
(2)
with the two first terms representing the kinetic and Coulomb or Hartree energy
of the electrons. The last term is a universal functional Exc [n] that represents all
4 Electronic Structure: Metals and Insulators
191
exchange and correlation effects. Introducing an auxiliary system of noninteracting
particles with the same density n(r) as the real one, the corresponding kinetic
energy can easily be expressed leading to the formal definition of the corresponding
exchange and correlation energy functional Exc [n] as containing all remaining
many-body effects. As it is common in electronic structure theory of solids in Eq. (2)
and the following atomic Rydberg units, (h̄ = 1, me = 1/2, e2 = 2, and c = 2/α
with α the fine-structure constant) are used.
As the functional Exc [n] is universal, the resulting scheme can in principle
be applied without modification to spin-magnetic, i.e., spin-polarized, systems.
However, as the available formulations for the functional are far too complicated
to be applied to real systems, Exc [n] has to be represented in practice by a suitable
approximation. For this, it is advantageous to replace DFT by the corresponding
spin-density functional theory (SDFT) that was introduced by von Barth and Hedin
[23] and Rajagopal and Callaway [24]. Restricting to the situation of collinear
magnetism with the spin-quantization axis along the global magnetization, this leads
to the following Schrödinger-like single-particle equations:
−∇ 2 + Vσeff (r) φiσ (r) = iσ φiσ (r).
(3)
Here φiσ (r) and iσ are the wave function and energy of the single-particle state i
with spin character σ (up or down). While these quantities have a priori no physical
meaning, they can nevertheless be used to determine the central properties of the
system. This applies in particular for the spin densities
nσ (r) =
Nσ
|φiσ (r)|2 ,
(4)
i=1
as the basic variables of the system. In Eq. (4), the summation runs over all Nσ states
with their energy iσ below the Fermi energy EF that in turn is determined by the
requirement
N=
d 3 r n↑ (r) + n↓ (r) ,
(5)
where N = N↑ + N↓ is the total number of electrons. Obviously, the corresponding
particle density n(r) and spin magnetization m(r), given by the relations:
n(r) = n↑ (r) + n↓ (r)
(6)
m(r) = n↑ (r) − n↓ (r),
(7)
may also be chosen as alternative basic variables of a spin-polarized system.
The spin-dependent effective potential Vσeff (r) entering Eq. (3) is determined
by the requirement that the total energy E[n↑ , n↓ ] that now has to be seen as
192
H. Ebert et al.
a functional of the spin densities takes a minimum. This leads finally to the
expression:
Vσeff (r)
=
Vσext (r) + 2
d 3r n(r )
+ Vσxc (r),
|r − r |
(8)
with
Vσxc (r) =
δExc [n↑ , n↓ ]
.
δnσ (r)
(9)
Equations (3), (8), and (9) constitute the coupled Kohn-Sham equations that
obviously have to be solved self-consistently. With this accomplished, the total
energy of the system can be obtained from:
E[n↑ , n↓ ] =
Nσ
iσ −
d 3r
d 3r σ =↑,↓ i=1
−
n(r) n(r )
|r − r |
d 3 r Vσxc (r) nσ (r) + Exc [n↑ , n↓ ].
(10)
σ =↑,↓
When the impact of spin-orbit coupling is included into the formalism (see
section “Relativistic Effects”), the corresponding expression for E[n↑ , n↓ ] is very
convenient for investigations on the magnetic anisotropy. In this case, one can in
general neglect the change of nσ (r) with the orientation of the magnetization, and
one has to consider only the first term in Eq. (10) representing the single-particle
energies of the system.
The collinear formulation given above is adequate for most situations. In case of
noncollinear magnetic configurations with the orientation of the spin magnetization
changing with position, one has in principle to use the spin-matrix formulation
of SDFT [23]. This implies that the wave functions φiσ (r) in Eq. (3) have to be
replaced by spinors φi (r), i.e., two component wave functions, that in general will
have no pure spin character. However, very often one can still assume a uniform
orientation of the magnetization within an atomic cell. In this case the potential is
also spin-diagonal in a local frame of reference with the spin-quantization axis along
the orientation m̂ of the local spin moment m. Accordingly, one can represent the
spin-dependent part of the effective 2 × 2 potential matrix function by:
V spin (r) = σ · m̂ B(r),
(11)
where σ is the vector of 2 × 2 Pauli spin matrices [5] and the effective field
B(r) = (V ↑ (r) − V ↓ (r))/2 is given by the difference of the spin-up and spin-down
potentials in the local frame. Alternatively, the spin-diagonal potential in the local
frame may be related to the 2 × 2 potential matrix function in the global frame by
4 Electronic Structure: Metals and Insulators
193
means of a transformation matrix U (θ, φ) such that: σ · m̂ = U † (θ, φ) σz U (θ, φ).
This matrix can be obtained from Eq. (80) (see below) by setting θ and φ according
to the orientation m̂ and q = 0. In either case, the form of the spin-dependent
potential again implies spinor or two-component wave functions.
As indicated, the major problem of SDFT is that the functional Exc [n↑ , n↓ ]
for the exchange-correlation is not known. A useful expression for this, that can
be justified for slowly varying densities, is supplied by the local spin-density
approximation (LSDA):
LSDA
Exc
[n↑ , n↓ ]
=
d 3 r n(r) (n↑ (r), n↓ (r)).
(12)
2
2
0
0
-2
-2
E (eV)
Ejkσ (eV)
Here (n↑ (r), n↓ (r)) is the exchange-correlation energy per electron for the
homogeneous free electron gas with uniform spin densities n↑ (r) and n↓ (r) that can
be determined with high accuracy [25]. A fit to numerical results for (n↑ (r), n↓ (r))
allows giving explicit expressions for the corresponding spin-dependent exchangecorrelation potential Vσxc (r) = Vσxc [n↑ (r), n↓ (r)].
As an example for a spin-polarized solid, Fig. 1 gives the band structure or
dispersion relation Ej kσ and the density of states (DOS) nσ (E) of bcc-Fe as calculated within LSDA (see section “Band Structure Methods”). These curves clearly
show the exchange splitting due to the spin dependency of the exchange-correlation
potential when compared with results for the corresponding paramagnetic state.
Integrating nσ (E) up to the Fermi energy EF obviously gives the number of
electrons Nσ with spin character σ . The corresponding spin-magnetic moment
M = N↑ − N↓ is given in Table 1 for bcc-Fe, fcc-Ni, and hcp-Co [27]. Obviously,
-4
-4
-6
-6
-8
-8
Γ
Δ
wave vector k
X
3
2
1
n↓(E) (sts./eV)
0
1
2
n↑(E) (sts./eV)
3
Fig. 1 Dispersion relation Ej kσ (left) and density of states nσ (E) (right) of ferromagnetic bcc-Fe
as calculated within LSDA. Results for the majority (↑) and minority (↓) spin states are given in
red and blue, respectively. In addition, results for the corresponding paramagnetic state are given
in black [26]
194
H. Ebert et al.
Table 1 Spin-magnetic moment of bcc-Fe, fcc-Ni, and hcp-Co calculated on the basis of the
LSDA and the GGA, compared with experimental values. The calculated magnetic moments given
in the last two columns have been obtained for the theoretical equilibrium and experimental lattice
parameter, respectively. (All data taken from [27])
Magnetic moment (μB )
Fe
Co
Ni
Wigner-Seitzradius (a.u.)
LSDA 2.59
GGA 2.68
Expt. 2.67
LSDA 2.54
GGA 2.63
Expt. 2.62
LSDA 2.53
GGA 2.63
Expt. 2.60
Bulk modulus
(Mbar)
2.64
1.74
1.68
2.68
2.14
1.91
2.50
2.08
2.86
Cohesive energy
(eV)
7.32
6.31
4.28
5.98
4.52
4.39
5.45
4.18
4.44
(atheo )
2.08
2.20
–
1.50
1.63
–
0.59
0.65
–
(aexpt )
2.14
2.17
2.22
1.62
1.63
1.72
0.61
0.63
0.61
the results depend to some extent on the chosen functional for Vσxc (r) (LSDA or
GGA) and the lattice parameter. Nevertheless, the calculated moments are in fairly
good agreement with experiment.
Although LSDA turned out to be astonishingly successful for many situations,
it nevertheless shows severe limitations. For example, LSDA leads in general to an
over-binding as can be seen from the Wigner-Seitz radius given in Table 1 that is too
small when compared to experiment. Furthermore, calculations on ferromagnetic Fe
led to a lower total energy for the fcc instead of the bcc structure (see section “Total
Electronic Energy and Magnetic Ground State”). These deficits could be removed
when the generalized gradient approximation (GGA) was introduced that expresses
GGA [n , n ] not only in terms of the spin densities n (r) but also of their
Exc
↑ ↓
σ
gradients ∇nσ (r). A more systematic route to derive accurate exchange-correlation
energies and corresponding potentials is supplied by the optimized potential method
that leads to a functional expressed in terms of the Kohn-Sham orbitals [15].
Unfortunately, this approach is numerically much more demanding than the very
efficient LSDA or GGA schemes in particular when a reliable representation of
correlation is incorporated. Accordingly, the development of parametrizations for
the exchange-correlation potential that are at the same time efficient and sufficiently
accurate is a field of ongoing research.
A major problem of LSDA and comparable SDFT schemes is the accurate
treatment of correlation effects in case of moderate or strong correlations as they
occur in systems with narrow energy bands. A way to cure this problem is the
use of the GW method [28] that is applicable to moderately correlated systems.
Application to Ni, for example, showed in particular a narrowing of the d-band
when compared to LSDA-based results [29] as expected from photoemission [30].
A scheme that is applicable also to strongly correlated materials at a much lower
4 Electronic Structure: Metals and Insulators
195
numerical cost is the LDA+U [31] that accounts for static correlations by adding
corrections to the LDA or LSDA Hamiltonian that depend on the Hubbard Coulomb
parameter U . In case that U is much larger than the band width W , a situation
typically met in oxides, the correction term leads in particular to a splitting into
an upper and lower Hubbard band. Dynamical correlations, on the other hand, are
accounted for by the dynamical mean field theory (DMFT) that when merged with
the LSDA leads to the combined Hamiltonian [32, 33]:
H = HLSDA +
1 σσ
Umm n̂ilmσ n̂ilm σ 2
il
−
mσ m σ
1 †
†
Jmm ĉilmσ
ĉilm σ̄ ĉilm
σ̄ ĉilmσ
2
il
−
mσ m
il
mσ
Δl n̂ilmσ
.
(13)
i=id ,l=d
Here, ĉ, ĉ† , and n̂ are creation, annihilation, and particle density operators that refer
to atomic orbitals labeled by the quantum numbers l, m, and σ and site index i, with
σ σ and J
Umm
mm corresponding to Coulomb and exchange integrals, respectively.
The quantity Δl represents the so-called double counting term that takes care that
static correlations are not accounted for twice – by the LSDA Hamiltonian HLSDA
and by its complementary DMFT counterpart HDMFT . As Eq. (13) indicates, the
DMFT correction is usually restricted to the correlated subsystem of the system as,
for example, d-states in case of transition metals (i = id , l = d). Furthermore,
the Coulomb and exchange integrals are assumed to be site diagonal (singlesite approximation) leading for the many-body problem to the same situation as
for the Anderson impurity model (AIM). Accordingly, all the various many-body
techniques available to deal with the AIM can also be used when dealing with
the combined LSDA+DMFT Hamiltonian. In most cases, Eq. (13) is dealt with
by calculating in a first step the one-electron Green function (see section “Band
Structure Methods”) associated with HLSDA . In a next step, the single-site problem
is solved by a so-called impurity solver that allows representing the impact of
HDMFT in terms of a corresponding complex and energy-dependent self-energy
ΣDMFT (E). Finally, making use of the Dyson equation (see Eq. (21) below), the
one-electron Green function of the system is updated.
Figure 2 shows typical results for the spin-dependent self-energy ΣDMFT (E) of
ferromagnetic Ni. The characteristics of these curves lead to the various correlationinduced features expected from photoemission [30]: the real part gives rise to a
renormalization of the energy bands leading to band narrowing, reduction of the
exchange splitting, and the occurrence of a satellite structure at 6 eV binding energy.
The imaginary part, on the other hand, implies a finite lifetime of the electronic state
that increases for the d-states with distance from the Fermi energy. This is reflected
in the DOS curves by a smearing-out of its structure when compared to the LSDA
result.
196
H. Ebert et al.
Fig. 2 Left: real (red) and imaginary (blue) parts of the spin-resolved self-energy ΣDMFT (E)
of ferromagnetic Ni. Right: corresponding DOS obtained on the basis of plain LSDA and
LSDA+DMFT [26]
Band Structure Methods
Most methods for calculating the electronic structure of solids on the basis of the
Kohn-Sham equation (3) assume three-dimensional periodicity for the potential
Vσeff (r). This implies that the corresponding solutions ψj k (r) are Bloch states that
transform under a lattice translation as:
ψj k (r + R n ) = eik·R n ψj k (r)
(14)
and accordingly can be labeled by the wave vector k and an additional band index
j . A direct consequence of this is that Bloch states for different wave vectors k
are orthogonal. This property simplifies the solution of the band structure problem
tremendously when using the variational principle to solve Eq. (3) and transforming
that way the problem to solving an algebraic eigenvalue problem. Constructing in
this case the basis functions such that they obey Eq. (14), one is led to a secular
equation with finite dimension for each k-vector
H k − Ej k S k α j k = 0
(15)
with the Hamilton and overlap matrices, H k and S k , referring to the basis functions
and Ej k and α j k the associated eigenvalues and eigenvectors, respectively.
Obviously one still has great freedom to construct a suitable basis set, and
accordingly there is a large number of methods and corresponding computer
codes available [34]. One route to set up an electronic structure method is to
take the tightly bound core states as frozen. This allows introducing effective
pseudo-potentials that do not show the singularity of the Coulomb potential when
4 Electronic Structure: Metals and Insulators
197
approaching the atomic nucleus. As a consequence, one may use even plane
waves as basis functions. Higher accuracy and flexibility, however, can be achieved
by using a technique called projector-augmented wave method (PAW) [35] that
mediates between pseudo-potential and so-called all-electron methods, with the
latter aiming to solve the Kohn-Sham equations directly. Again one may distinguish
between methods using analytical or numerical basis functions. Within the LMTO
method [36], for example, the Kohn-Sham equation is solved numerically at a fixed
energy Eν and angular momentum l for a spherical potential inside an atomic cell
that is approximated by a sphere. A linear combination of these solutions φlν (r)
and their energy derivatives φ̇lν (r) are augmented outside the atomic cell in a way
that leads to a decaying muffin-tin orbital that solves the Kohn-Sham equation also
within all neighboring atomic cells. The Bloch sum of such muffin-tin orbitals
is obviously a suitable basis function. By construction, it is energy independent,
but with an appropriate choice of Eν it will account, within the energy regime of
interest, for the energy dependence of the exact solution up to first order. This is a
common feature of all so-called linear methods [36] like the LAPW [36, 37]. The
special choice of the basis function of the LMTO method has the additional feature
that it is minimal: this means that one can restrict the angular momentum expansion
of the basis function in line with chemical intuition, i.e., for transition metals, one
should go at least up to d-states with l = 2. Solving the resulting secular equation
or algebraic eigenvalue problem, one gets finally the energy eigenvalue Ej k of the
Bloch states together with a corresponding representation of their wave functions
ψj k (r) in terms of the radial functions φlν (r) and their energy derivatives φ̇lν (r):
ψj k (r) =
jk
jk
Alm φlν (r) Ylm (r̂) + Blm φ̇lν (r) Ylm (r̂),
(16)
lm
jk
jk
where the expansion coefficients Alm and Blm are given by the eigenvectors α j k
and Ylm (r̂) are spherical harmonics. A similar representation of the Bloch wave
functions is obtained for most other all-electron band structure methods.
Although most band structure methods are formulated as k-space methods
assuming three-dimensional periodicity, they can nevertheless be applied to situations with lower dimensionality or symmetry by means of the super-cell technique.
This is illustrated for the case of a random substitutionally disordered binary alloy
Ax B1−x in Fig. 3. Instead of dealing with an, in principle, infinitely large unit cell
that represents the random distribution of the A- and B-atoms on the geometric
lattice, periodic boundary conditions are imposed implying the use of a finite size
super-cell that is enlarged when compared to the unit cell of the underlying lattice.
To achieve a reliable representation of the configurational average of a disordered
alloy, obviously the unit cell has to be large enough and a representative average has
to be taken concerning the atomic configuration within the super-cell [38] using,
for example, the concept of a special quasi-random structure [39]. As indicated by
the lower panel of Fig. 3, the same type of reasoning can be applied when dealing
with the problem of a disordered spin configuration of a solid that may occur due to
thermal spin fluctuations [40]. The super-cell technique is also frequently applied
198
H. Ebert et al.
Fig. 3 Top row: representation of a random substitutionally disordered binary alloy Ax B1−x (left)
by means of the super-cell technique (middle) and by means of an effective medium theory (right).
Bottom row: corresponding application of these schemes to the problem of a disordered spin
configuration
in case of reduced dimensionality of the system as, for example, when dealing with
surfaces or impurities in a host system. In the latter case obviously the size of the
super-cell has to be chosen large enough to avoid an interaction of impurities in
neighboring super-cells.
Instead of representing the electronic structure of a solid in terms of Bloch
states with associated wave functions ψj k (r) and energy eigenvalues Ej k , this can
be done by means of the corresponding retarded single-particle Green function
G+ (r, r , E). With the solutions to the band structure problem available, the
Green function G+ (r, r , E) can be given via the so-called Lehmann spectral
representation [41]
+
ψj k (r) ψj†k (r )
G (r, r , E) = lim
→0
jk
E − Ej k + i
,
(17)
that allows straightforwardly to derive convenient expressions, for example, for the
density of states n(E) and electron density n(r), respectively:
1
n(E) = − π
d 3 r G+ (r, r, E)
(18)
4 Electronic Structure: Metals and Insulators
1
n(r) = − π
199
EF
dE G+ (r, r, E).
(19)
Although Eq. (17) is frequently used, it is not very convenient as one needs in
principle the whole spectrum connected with the underlying electronic Hamiltonian
to get the Green function for a given energy E. An alternative to this is offered
by the multiple scattering theory-based KKR (Korringa-Kohn-Rostoker) formalism
that leads to the following expression for G+ (r, r , E) [42]:
G+ (r, r , E) =
L,L
i× ii
ZLi (r, E) τLL
(E) ZL (r , E)
−δii ZLi (r < , E) JLi× (r > , E)
(20)
L
that in particular does not require Bloch translational symmetry for the system
considered. The functions ZLi (r, E) and JLi (r, E) in Eq. (20) are regular and
irregular solutions, respectively, to the Kohn-Sham equation with angular character
L = (l, m) for r in the atomic cell at site i and a specific normalization [42]. The
ii (E) is the so-called scattering path operator that transfers a wave with
quantity τLL
character L coming in at site i into a wave going out from site i with character L
with all intermediate scattering events accounted for. From Eq. (18), it is obvious
ii (E) determines in particular the variation of the
that the site-diagonal quantity τLL
density of states ni (E) at site i with energy E.
The use of the Green function offers many advantages when dealing with
embedded subsystems, response functions, spectroscopy, disorder, or the manybody problem. To a large extent, this is due to the Dyson equation that allows to
express the Green function G+ (r, r , E) of a complex system on the basis of that of
a simpler reference system (G+
0 (r, r , E)) and the arbitrary perturbing Hamiltonian
Hpert (r) that connects the two systems:
G+ (r, r , E) = G+
0 (r, r , E) +
Ω
d3 r G+
0 (r, r , E)
Hpert (r ) G+ (r , r , E),
(21)
with Ω the region for which Hpert (r) has to be accounted for. For a substitutional
impurity, this would include the atomic cell of the impurity and the region of the
neighboring host atoms that are distorted by the impurity.
The Green function formalism is particularly useful when dealing with the
electronic structure of disordered systems. By using the concept of the molecular
field, Soven [43] introduced the Coherent Potential Approximation (CPA) approach
when dealing with disordered substitutional alloys. The corresponding hypothetical
effective CPA medium plays the role of the molecular field and is constructed such
that it represents the configurational average for the alloy as accurate as possible
200
H. Ebert et al.
(see Fig. 3). The standard CPA is a so-called single-site theory implying that the
occupation of neighboring lattice sites is uncorrelated, i.e., short-range order is
excluded. Within the KKR approach, the CPA medium is therefore determined by
requiring that for an Ax B1−x alloy, the embedding of an A- or B-atom into the CPA
medium should on the average lead to no additional scattering:
ii
ii
x τ ii
A + (1 − x) τ B = τ CPA .
(22)
2
2
0
0
0
-2
-2
-2
-4
-4
-6
-6
-6
-8
-8
Δ
Pd
-8
2
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2
n↑(E) (sts./eV)
n↑(E) (sts./eV)
2
2
0
0
0
-2
-2
-2
-4
-4
-6
-6
-6
-8
-8
wave vector k
X
E (eV)
E (eV)
Γ
Ni
-4
Γ
Δ
wave vector k
X
Ni
Pd
E (eV)
-4
E (eV)
2
E (eV)
E (eV)
Here the component-projected scattering path operators τ ii
α represent the single-site
embedding of the component α into the CPA medium according to Eq. (21). When
using these quantities together with the corresponding component-related wave
functions in Eq. (20), one gets obviously access to component-specific properties as
the partial DOS of an alloy. Corresponding results for the disordered ferromagnetic
alloy fcc-Ni0.8 Pd0.2 are shown in Fig. 4. The left column gives the spin-resolved
band structure in terms of the Bloch spectral function AB
σ (k, E) that can be seen
as the Fourier transform of the real space Green function G+
σ (r, r , E), while the
-8
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2
n↓(E) (sts./eV)
n↓(E) (sts./eV)
Fig. 4 Left: spin-resolved Bloch spectral function AB
σ (k, E) of the disordered ferromagnetic alloy
fcc-Ni0.8 Pd0.2 calculated on the basis of the CPA. Middle and right column: corresponding spinresolved partial density of states nασ (E) for α = Ni or Pd, respectively. The top and bottom row
give results for spin up and down, respectively. As a reference, the dispersion relation Ej kσ of pure
Ni
Ni is superimposed as a black line to AB
σ (k, E) (left). In addition, nσ (E) for pure ferromagnetic
Pd
Ni (middle) and of nσ (E) for pure paramagnetic Pd (right) are included in the figures as dashed
lines [26]
4 Electronic Structure: Metals and Insulators
201
middle and right columns give the spin-resolved partial density of states nασ (E) for
α = Ni or Pd, respectively. Comparison of AB
σ (k, E) with the dispersion relation
Ej kσ of fcc-Ni and of nNi
σ (E) for the alloy with that for pure Ni clearly shows the
smearing-out of these curves for the alloy in particular in the regime of the d-states.
This reflects the fact that for the alloy the wave vector k is not a good quantum
number. In fact the width of the AB
σ (k, E) functions can be interpreted as a measure
for the electronic lifetime to be used in the calculation of the residual resistivity on
the basis of the Boltzmann formalism [44]. The right column of Fig. 4 gives results
for the partial DOS nPd
σ (E) of Pd in fcc-Ni0.8 Pd0.2 and for pure paramagnetic Pd
that clearly shows that Pd gets spin-polarized due to hybridization leading to an
induced spin-magnetic moment of about 0.21 μB .
It is important to note that the concept of the CPA is not restricted to alloys but
can be applied to any type of disorder. Making use of the alloy analogy, the CPA is
used, for example, within the disordered local moment (DLM) model [45] to perform an average over spin configurations connected with thermal spin fluctuations
(see section “Coherent Treatment of Electronic Structure and Spin Statistics”).
Relativistic Effects
A most coherent way to account for the influence of relativistic effects is to work on
the basis of the Dirac Hamiltonian [5]:
ĤD = −icα · ∇ +
1 2
c (β − 1) + V̄ (r) + β σ · B(r) + eα · A(r).
2
(23)
Here c is the speed of light, αi and β are the standard 4 × 4 Dirac matrices, cα
is the electronic velocity operator, σi are the 4 × 4 spin matrices, and the local
potential may involve a spin-independent part V̄ (r), an effective magnetic field
(B(r)) coupling only to the spin [46] and the vector potential (A(r)) coupling
to the electronic current, where the effective fields B(r) and A(r) combine
exchange-correlation and possible external contributions. This approach implies a
four-component electronic wave function ψ(r, E) or bi-spinor, respectively, with a
large and small component [5]. Corresponding versions based on this framework
have been worked out for several band structure methods [47, 48] assuming in
general a spin-dependent potential only (A(r) = 0). The appealing feature of these
schemes is that they treat all relativistic effects and magnetic ordering on the same
footing.
Alternatively, one may apply a Foldy-Wouthuysen transformation [5] to the Dirac
equation. Considering only a spherical scalar potential V (r) in Eq. (23) and keeping
only terms up to the order of 1/c2 , this leads to three relativistic corrections when
compared to the Schrödinger Hamiltonian [49]:
Hmass = −β
1 4
p
c2
(24)
202
H. Ebert et al.
1 2
∇ V (r)
2c2
1 1 ∂V (r)
= 2
σ · l.
c r ∂r
HDarwin =
(25)
HSOC
(26)
The first two corrections, mass enhancement (Hmass ) and Darwin (HDarwin ) terms,
do not involve the spin operator σ , and for that reason, they are called scalar
relativistic. In general, these corrections give rise to a downward shift in energy
for s- and p-states and, in response to this, to an upward shift in energy for d-states
[50]. For a paramagnetic metal, this influences the density of states at the Fermi
level and this way the tendency toward spontaneous formation of spin magnetism
via the Stoner mechanism (see section “Stoner Model of Itinerant Magnetism”).
The third correction term is the spin-orbit coupling (HSOC ), which is given in its
commonly simplified form, that holds in case of a spherically symmetric scalar
potential V (r). As HSOC couples the electronic orbital and spin degrees of freedom,
this term will lead to the removal of degeneracies for any material. For a spinpolarized material, it leads furthermore to a reduction in symmetry as reflected by
many important properties as, for example, the magnetocrystalline anisotropy [51,6]
or galvanomagnetic [8] and magneto-optical [9] properties.
In practice, most electronic structure calculations are based on relativistic
correction schemes that first of all aim to eliminate the small component from
the Dirac equation leading to a two-component formalism [52, 53]. Within socalled scalar relativistic calculations, spin-orbit coupling is ignored leading to no
technical changes when compared to a nonrelativistic calculation on the basis of
the LSDA that treat spin-up and spin-down states separately. However, including
HSOC does not allow this simplification any more requiring higher computational
effort. While HSOC can still be accounted for when setting up the basis functions of
a band structure scheme [52, 54], it is in most cases seen as a correction to a scalar
relativistic Hamiltonian and therefore accounted for only in the variational step of a
standard band structure scheme [36, 37].
As an illustration, Fig. 5 gives the dispersion relation Ej k of ferromagnetic Ni
calculated in a non-, scalar, and fully relativistic way. The left panel of the figure
shows the impact of the scalar relativistic corrections Hmass and HDarwin . The
middle panel shows results that account in addition for the spin-diagonal part of
HSOC proportional to lz σz [55]. Although this term leaves spin as a good quantum
number, it obviously removes many degeneracies and band crossings giving rise, for
example, to spin-orbit-induced orbital magnetic moments [55]. Accounting finally
for the full spin-orbit coupling HSOC (right panel), one finds further removals of
band crossings and a coupling of the two spin systems. This spin mixing gives rise
to so-called hot spots in the band structure [56] that are important, for example, for
spin-flip relaxation processes.
The relativistic corrections mentioned so far concern only the kinetic part of
the electronic Hamiltonian, but not the effective potential. Starting from a fully
relativistic framework in fact corresponding corrections have to be expected when
4 Electronic Structure: Metals and Insulators
203
-0.5
-0.5
-1
Ejkσ (eV)
-1
-1.5
-1.5
-2
-2
NREL
-2.5
Γ
Δ
FREL
SOC zz
X
Γ
Δ
wave vector k
X
Γ
Δ
-2.5
X
Fig. 5 Dispersion relation Ej k of ferromagnetic Ni. Left: spin-resolved curves Ej kσ for non(black lines) and scalar (up and down triangles) relativistic mode. Middle: relativistic mode that
accounts only of the spin-diagonal part of HSOC proportional to lz σz (black lines) and scalar
relativistic mode (up and down triangles). Right: fully (black lines) and scalar (up and down
triangles) relativistic mode [26]
considering the free electron gas as a reference system [15]. The impact of such
corrections to the exchange-correlation potential has been monitored in particular
for paramagnetic materials [57]. For magnetic materials, a coherent derivation
of nonrelativistic spin-density functional theory could be given by starting from
a relativistic formalism [24]. Later on, a relativistic formulation of spin-density
functional theory assuming a spin-dependent exchange-correlation potential that
couples only to the spin degree of freedom was worked out [58, 46] leading
to Eq. (23) with A(r) = 0. This simplification seems to be acceptable for
many magnetically ordered transition metal systems. On the other hand, spin-orbit
coupling gives for spin-polarized materials automatically rise to orbital magnetism
[59] that can be associated with a corresponding orbital current. Accordingly,
relativistic spin-density functional theory should in principle be replaced by a
current density formalism with the four current as a basic system variable [15],
i.e., one has B(r) = 0 and A(r) = 0 in Eq. (23). While various developments have
been made in this direction [60], there are no functionals for the corresponding
exchange-correlation available at the moment. As a consequence, the feedback of
orbital magnetism or polarization on the electronic Hamiltonian has been accounted
for primarily by means of hybrid schemes like the OP- [61], the LSDA+U- [62], or
the LSDA+DMFT-formalism [63].
Another consequence of a fully relativistic formalism is a modification of the
Coulomb potential and the occurrence of the Breit interaction [13]. The latter one
can be seen as a current-current interaction and accordingly can be represented
by a corresponding vector potential A(r) in Eq. (23). For magnetic materials, the
Breit interaction contributes not only to the total energy but also to its magnetic
anisotropy. In fact it has been pointed out that the Breit interaction is the quantum
204
H. Ebert et al.
mechanical source for the classical dipole-dipole interaction giving rise to the
magnetic shape anisotropy [14].
Adiabatic Dynamics
The features of the electronic band structure determine not only the ground-state
properties of a material but also the dynamical properties of electrons in the presence
of external time-dependent perturbations. When a perturbation varies slowly in time,
the dynamics of the electronic subsystem can be efficiently described using the
concept of the Berry phase [64] arising during the adiabatic evolution of electronic
quantum states. According to its definition, the Berry phase is connected to a
parameter-dependent Hamiltonian [65,66,67]. For systems with a periodic effective
potential Veff (r + R n ) = Veff (r) with R n a lattice translation vector, leading to a
Bloch-like solution of the eigenvalue problem (see Eq. (14)):
ψj k (r) = eik·r uj k (r),
(27)
2
p̂
the unitary transformation of the Hamiltonian H = 2m
+ Veff (r) leads to a
momentum-dependent Hamiltonian (for details see [67])
H(k) = e−ik·r Heik·r =
(p̂ + h̄k)2
+ Veff (r).
2m
(28)
This allows treating the Brillouin zone as the parameter space of the transformed
Hamiltonian H(k) with cell-periodic eigenfunctions uj k (r). The Berry phase is
expressed as an integral over the path C in the parameter space of the electronic
momentum k,
γj (k) =
dk · Aj (k).
(29)
C
The integral is gauge-invariant for a closed path, while the Berry vector potential,
or the Berry connection
Aj (k) = i uj k
∂
uj k ,
∂k
(30)
is a gauge-dependent quantity. This value, treated in analogy to electrodynamics as
a vector potential, gives access, using Stokes’ theorem, to a gauge-invariant quantity
called Berry curvature playing the role of a magnetic field in the parameter space:
Ω j (k) = ∇ k × Aj (k),
or, alternatively,
(31)
4 Electronic Structure: Metals and Insulators
Ωj,μν (k) = ∂μ Aj,ν − ∂ν Aj,μ = −2 ∂μ uj k | ∂ν uj k ,
γj =
dS · Ω j (k),
S
205
(32)
(33)
with the integration over an arbitrary surface enclosed by the path C.
In case of a translation-invariant crystal, a closed path in Eq. (29) occurs due
to a torus topology of the Brillouin zone as any two points k and k + G are
fully equivalent for any reciprocal lattice vector G, leading to an integration over
a Brillouin zone in Eq. (29).
Concerning applications of the Berry phase concept, we focus here on the
electron dynamics in the presence of an electric field E entering the Hamiltonian
trough a time-dependent uniform vector potential A(t), preserving the translation
symmetry of the system. The group velocity of a state (j, k) is given to first
order by an expression [65, 68] consisting of two terms, the usual band dispersion
contribution as well as a so-called anomalous velocity proportional to the Berry
curvature of the bands,
v j (k) =
∂j (k)
e
− E × Ω j (k),
h̄∂k
h̄
(34)
with the Berry curvature Eq. (32) given in the form [69, 67]:
Ω j (k) = i ∇ k uj k | × |∇ k uj k .
(35)
The Berry curvature is nonzero either in non-centrosymmetric systems or in systems
with broken time-reversal symmetry and vanishes in the case when both symmetries
are present, leading to elimination of the anomalous velocity in Eq. (34) [67]. Thus,
due to these symmetry properties of the Berry curvature, the anomalous Hall effect
can be observed in ferromagnetic systems. One has to stress the crucial role of
the spin-orbit interaction required for a nonzero Berry curvature in FM systems,
leading to avoided crossings of the energy bands which give in turn the most
pronounced contributions in the vicinity of the Fermi surface that can be seen in
Fig. 6a [70].
As one can see in Eq. (34), the anomalous velocity is always transverse to the
electric field giving rise to a Hall current. The corresponding contribution to the
Hall conductivity associated with the anomalous velocity term was demonstrated
first by Karplus and Luttinger [72] and is called Karplus-Luttinger mechanism. In
terms of the Berry curvature, it is given by the expression [65, 68, 69, 73]:
KL
σαβ
=
e2
h̄
BZ
dDk γ
f (j (k))αβγ Ωj (k),
(2π )D n
(36)
with D the dimensionality of the system and αβγ the Levi-Civita tensor. As one
can see, this contribution is determined by all occupied states, and it is the only term
206
H. Ebert et al.
Fig. 6 (a) Fermi surface in the (010) plane (solid lines) and integrated Berry curvature −Ω z (k)
in atomic units (color map) of fcc Fe. (From Yao et al. [70]); (b) Berry curvature projected onto
the (k3 , k1 ) plane for Mn3 Ge, where k3 and k1 are aligned with the kz and kx axes, respectively.
It is calculated by integrating along the k2 direction. (From Ref. [71]); (c) Energy dispersion of
Mn3 Ge along k2 with (k3 , k1 ) fixed at the point with the largest Berry curvature, indicated by a
black dashed circle in (b). (From Ref. [71]). (Figure are printed with permission from [70] by
the American Physical Society; Figure (b) and (c) reprinted with permission from [71] by the
American Physical Society.)
contributing to the AHC for insulating systems. It is independent on the nature of the
impurities and their concentration and therefore is called intrinsic contribution to the
AHE. Note, however, that this contribution is not the only intrinsic contribution in
metals. The so-called side-jump mechanism suggested by Berger [74] is contributed
by the electrons at the Fermi surface and also yields a Hall conductivity that is
independent of the impurity concentration [69].
Note that the anomalous thermoelectric transport driven by statistical forces due
to a temperature gradient ∇T (the same concerns also the transport driven by
chemical potential gradients ∇μ) cannot rely on the anomalous velocity term as
it vanishes in the absence of an electric field. A corresponding theory giving the
intrinsic contribution to transverse thermoelectric transport has been reported by
Xiao et al. [75]. It is based on the generalization to finite temperatures of a theory
giving a Berry-phase correction to the orbital magnetization [76, 77, 17]:
4 Electronic Structure: Metals and Insulators
M(r) =
BZ
j
+
207
dDk
f (j (k))mn (k)
(2π )D
1
dDk e
Ω j (k)log 1 + e−β(j (k)−μ) ,
D
β
BZ (2π ) h̄
(37)
j
where mn (k) = (e/2h̄)i ∇ k uj,k |[j (k) − H(k)] × |∇ k uj,k is the orbital moment
of state n. By doing some transformations, this expression can be written also in the
form [67]:
Mz (r) =
BZ
j
dDk
1
f (j (k))mn,z (k) +
D
(2π )
e
df ()σxy ().
(38)
It has two different contributions associated with the self-rotation of the wave packet
representing an electron and with the center-of-mass motion, respectively. The first
term, obviously, occurs by treating the carrier as a wave packet having finite spread
in the phase space. The second term comes from the Berry-phase correction to the
electron density of states [67].
A crucial point for thermoelectric transport is that the conventional expression
used for the current density is incomplete as it is derived for the carrier treated as a
point particle. Having a corresponding expression for the local current J obtained by
treating the carrier as a wave packet, introducing the concept of a transport current
j = J − ∇ × M(r),
(39)
and using the expression Eq. (37) for the orbital magnetization density M(r), one
obtains the transport current as given by:
dDk
g(r, k)ṙ
D
BZ (2π )
dDk e
1
Ω(k)log(1 + e−β((k)−μ) ).
−∇ ×
β BZ (2π )D h̄
j = −e
(40)
(41)
With this expression, it is straightforward to calculate various thermoelectric
responses to statistical forces [67]. Thus, the expression for an anomalous Nernst
conductivity αxy is given in terms of the intrinsic anomalous Hall conductivity σxy :
αxy
1
=−
e
d
−μ
∂f
σxy ()
.
∂μ
T
(42)
Talking about the AHE in antiferromagnets (AFM), one has to distinguish the
systems according to their symmetry. The AHE does not occur in collinear AFMs
symmetric with respect to time-reversal symmetry T combined with a half-magnetic
208
H. Ebert et al.
unit-cell translation Ta/2 or spatial inversion I, i.e., either Ta/2 T symmetry or
IT (see [78]). However, in AFM materials having a symmetry violating these
conditions, the AHE can be observed. Accordingly, a large intrinsic AHE was
predicted in Mn3 Ir [79] and Mn3 Ge [80] by performing first-principles electronic
structure calculations. As already indicated, the avoided band crossings near the
Fermi surface give the dominant contribution to the AHE. An example of a
calculated k-resolved Berry curvature for Mn3 Ge [71] is plotted in Fig. 6b. It is
dominating in the area highlighted in red giving rise to the large AHC. In order
to investigate the origin of the hot spot at (0.127, 0.428), the band structure is
plotted in Fig. 6c along k2 varying from 0 to 1. As one can see, the Fermi level
crosses two small gaps around k2 = 0 and 0.5. This implies that the entanglement
between occupied and unoccupied states must be very strong around these two
points, giving a large contribution to the Berry curvature and in turn to the AHE.
Finally, it should be noted that also the real-space topology of a material as, for
example, the presence of a noncoplanar, chiral spin texture can give rise to an
intrinsic AHC. This phenomenon, that requires neither a finite external magnetic
field, nor a finite net magnetic moment, nor even spin-orbit coupling, is commonly
termed topological or chirality-induced Hall effect. The newly emerging field of
topological antiferromagnetic spintronics [81] deals with this and related response
phenomena at the intersection of antiferromagnetic spintronics with topology.
Itinerant Magnetism of Solids
Historically there was for a long time a heated discussion whether the model of
itinerant or band magnetism is a suitable platform to discuss the magnetic properties
of a specific solid or whether the assumption of local magnetic moments is more
appropriate. With the advances in electronic structure theory to provide schemes
that allow treating the electronic structure of solids from wide-band metallic solids
up to narrow-band oxides, including localized systems showing the formation of
electronic multiplets, the competition between these extreme models gets more
or less obsolete. Accordingly, this section gives only a brief introduction to the
theory of itinerant or band magnetism on the basis of the Stoner model followed
by discussing the main electronic features of two prototype class of materials:
disordered alloys of transition metals and the Heusler alloys.
Stoner Model of Itinerant Magnetism
A rather simple criterion for the spontaneous formation of ferromagnetic order is
provided by the Stoner model for itinerant magnetism. Starting point is the spinprojected DOS for a paramagnetic solid as sketched in Fig. 7. Application of an
external magnetic field Bext gives rise to the Zeeman splitting ΔEZ = 2μB Bext for
spin-up and spin-down states with μB the Bohr magneton. The flip of the spin for
some states in the vicinity of the Fermi energy EF reestablishes a common Fermi
4 Electronic Structure: Metals and Insulators
209
E
E
E
EF
EF
EF
2 μBBext
n↓(E)
n↑(E)
n↓(E)
n↑(E)
n↓(E)
n↑(E)
Fig. 7 Left: spin-dependent DOS n(EF ) at the Fermi energy EF for a paramagnetic metal. Middle:
Zeeman splitting ΔEZ = 2 μB Bext due to an external magnetic field Bext . Right: spin flip leads to
a common Fermi energy EF and a finite spin-magnetic moment M = (N↑ − N↓ )
energy for both spin systems leading to a net spin-magnetic moment M = (N↑ −
N↓ ). For small magnetic fields Bext , the resulting Pauli spin susceptibility χ =
M μB /Bext is determined by the density of states n(EF ) at the Fermi energy EF :
χ0 = 2μB n(EF ).
(43)
An imbalance of N↑ and N↓ also changes the total energy due to the exchange
interaction. The corresponding spin-dependent correction for the electron energies
Ej kσ for spin σ (with σ = ±1/2) may be written as [82]:
Ej kσ = Ej k + sign (σ ) M I
(44)
with the Stoner exchange integral I originally seen as a parameter. Accounting for
this correction, in addition one is led to the enhanced spin susceptibility
χ = S χ0 ,
(45)
with S the Stoner enhancement factor:
S=
1
.
1 − I n(EF )
(46)
The paramagnetic state remains stable as long as I n(EF ) < 1 holds. However, when
I n(EF ) approaches the value 1, the enhancement factor S and with this the induced
magnetic moment diverge indicating an instability. In fact, I n(EF ) > 1 for the
paramagnetic reference state implies that the increase of kinetic energy associated
with the flip of the spin for electrons at the Fermi energy is more than compensated
by the resulting change in the exchange-correlation energy even without an external
field leading to a stabilization of the ferromagnetic state [83]. Accordingly, the
Stoner criterion I n(EF ) > 1 indicates the spontaneous formation of ferromagnetic
spin order for a solid.
210
H. Ebert et al.
Linear response theory allows deriving explicit expressions for the static [84,
85, 86] and dynamic [87, 88] spin susceptibility of arbitrary systems that confirm
Eqs. (43) and (45) even for inhomogeneous, i.e., non-bulk, systems. Working within
the framework of SDFT provides in particular a clear prescription for the calculation
of the Stoner exchange integral [89, 84]:
I=
d 3 r γ (r)2 K(r),
(47)
with the induced spin polarization γ (r) and the exchange-correlation kernel K(r):
γ (r) =
|ψj k (r)|2 δ(EF − Ej k )/n(EF )
(48)
jk
K(r) = −
1
2
δ 2 Exc
δm2
(49)
.
m(r)=0
Corresponding numerical results for the Stoner exchange-correlation integral I
(left) and density of states n(EF ) (right) are given in Fig. 8 for the 3d, 4d, and 5d
transition metal rows. As one notes, I varies smoothly within a transition metal
row. Accordingly, the product I n(EF ) primarily reflects the variation of n(EF )
with atomic number. In line with experiment, the Stoner criterion for ferromagnetic
ordering is met only for the late 3d transition metals Fe, Co, and Ni. Because of
the increase of the d-band width when going from a 3d metal to the corresponding
isoelectronic 4d or 5d metal, the Stoner product decreases and with this the tendency
toward ferromagnetic ordering. This trend is best seen for the sequence of fcc-metals
Ni-Pd-Pt that leads from a ferromagnet to strongly enhanced Pauli paramagnets with
a Stoner enhancement factor of 5.96 and 2.16, respectively.
1
3d
4d
5d
2
n(EF) (1/eV)
I (eV)
0.9
0.8
0.7
0.6
3d
4d
5d
1.5
1
0.5
0.5
0.4
0
Ti
Zr
V Cr Mn Fe
Nb Mo Tc Ru
Co
Rh
Ni
Pd
Cu
Ag
Ti V Cr Mn Fe
Zr Nb Mo Tc Ru
Co Ni
Rh Pd
Cu
Ag
Hf
Ta
Ir
Pt
Au
Hf Ta W
Ir
Au
W
Re
Os
Re Os
Pt
Fig. 8 Stoner exchange-correlation integral I (left) and density of states n(EF ) (right) at the Fermi
energy EF for the 3d, 4d, and 5d transition metal rows [26]
4 Electronic Structure: Metals and Insulators
211
As the Stoner integral I in general does not change much with the atomic
environment, the Stoner factor is primarily determined by the DOS n(EF ) at the
Fermi energy EF . As the d-band width W of the transition metals decreases with
coordination number, this leads usually to an increase of n(EF ). This implies
a corresponding increase of the tendency toward ferromagnetic ordering with
reduced dimensionality. Accordingly, magnetically ordered surface layers have been
predicted by theory for the paramagnetic metals V and Pd while the experimental
situation is unclear. For free and deposited transition metal clusters, many SDFTbased calculations led to finite spin-magnetic moments as, for example, for free
Ru13 , Rh13 , and Pd13 clusters [90]. Additional calculations for the paramagnetic
state gave a large peak for the DOS near the top of the valence band with the Fermi
energy located at its maximum, i.e., the magnetic ordering could be explained on
the basis of the Stoner criterion. These results are fully in line with experimental
data for RuN and PdN clusters; for example, mRh ≈ 0.8 μB /atom in Rh9 , mPd <
0.4 μB /atom in Pd13 and mRu < 0.32 μB /atom in Ru10 clusters [91]. In accordance
with the Stoner criterion, an increase of cluster size led to a decrease of magnetic
moments as, for example, mRh ≈ 0.16 μB /atom in Rh34 and mRu < 0.09 μB /atom
in Ru1115 clusters.
The Stoner criterion was also applied successfully to deposited atoms. For example, in line with experiment, self-consistent LSDA-based calculations predicted a
finite and vanishing spin-magnetic moment for Fe and Ni atoms, respectively, on a
chalcogenide topological insulator surface [92].
Slater-Pauling Curve
The substitutional magnetic alloys of 3d transition metals are often seen as prototype
materials for itinerant metallic magnetism. The experimental data on the average
magnetic moment M per atom of these systems is summarized by the well-known
Slater-Pauling curve shown in Fig. 9 (top). There are two main branches to be seen:
one leading from Fe to Cu with a slope of −45◦ and another one from Fe to Cr
with a slope of +45◦ . M is obviously given by M = N↑ − N↓ , where Nσ is the
average number of valence electrons with spin character σ . With the total number
of valence electrons Z = N↑ + N↓ , one gets the simple relation M = 2 N↑ − Z. On
the basis of the outdated rigid-band model that postulates a common electronic band
structure for both components of a binary alloy, all Ni and Co alloys are considered
to be strong ferromagnets with their spin-up band filled. Accordingly, N↑ is constant
and M decreases with Z explaining the right main branch of the Slater-Pauling
curve. For Z = 8.25, one may assume that the Fermi energy is at the top of the
spin-up band. Decreasing Z, one can now expect that N↓ stays constant leading to
M = Z − 2 N↓ . This obviously gives a simple explanation for the second branch
including in particular the Fe-Cr and Fe-V alloys.
Using the tight-binding version of the CPA combined with the Hartree-Fock
approximation, Hasegawa and Kanamori [94] could already give an alternative qualitative explanation for the Slater-Pauling curve avoiding the unrealistic assumptions
212
H. Ebert et al.
2.5
FeCo
magnetic moment (µ B)
Experiment
FeTi
FeNi
2
CoFe
1.5
FeMn
NiCo
NiFe
CoCr
1
FeCr
0.5
CoMn
NiMn
NiCu
NiCr
0
NiV
24
25
2.5
2
FeSc
1.5
27
26
electron number / atom
28
FeCo
Theory
magnetic moment (µ B)
NiTi
FeV
FeNi
FeCu
CoFe
NiFe
CoMn(2)
CoCr
NiCo
1
CoMn (1) NiMn
FeTi
0.5
FeV
NiCu
NiTi
FeCr
NiCr
0
NiV
24
25
27
26
electron number / atom
28
Fig. 9 Top: experimental Slater-Pauling curve, i.e., the average magnetic moment corresponding
to the saturation magnetization of Fe-, Ni-, and Co-based alloys vs. average number of electrons
per atom. Bottom: corresponding theoretical results obtained by means of the KKR-CPA. The fcc,
instead of hcp, structure is assumed for Co-based alloys. For Co-Mn, two solutions, CoMn(1) with
a Mn local moment parallel to the bulk magnetization and CoMn(2) with an antiparallel moment,
are obtained. (All data taken from [93])
of the rigid-band model. This approach could be improved a lot by Akai using
the KKR-CPA within the framework of SDFT leading for most alloy systems to
a quantitative agreement between theory and experiment [93] (see bottom panel of
Fig. 9). Most importantly, this implies that the average moments are well described
by the effective CPA medium.
As found by experiment, the curves on the left-hand side, connected with Febased bcc alloys, have a slope of about +45◦ although there is some spread. The
common feature of the alloys belonging to these subbranches is that the solute
atoms (Sc, Ti, V, and Cr) have negative local magnetic moments; i.e., they are
aligned antiparallel to the moments of the host (Fe, Co, and Ni). These negative
4 Electronic Structure: Metals and Insulators
213
moments can be associated with the appearance of hole states above or at the top
of the majority-spin d-bands. These hole states have a large amplitude at the solute
atoms causing a negative local moment there, as well as the rapid decrease of the
average moment with the solute concentration. The existence of the hole states in
the majority-spin band also affects the concentration dependence of the average
moment. In the case that no holes exist in the majority-spin band, the main origin
of the concentration dependence of the average moment is the reduced number of
available d-electrons as determined by the average number of valence electrons. In
fact, the straight line with a slope of −45◦ in the right half of the Slater-Pauling
curve, that is mainly associated with Ni-based fcc alloys, is explained this way.
Contrarily, in the case that holes exist in the majority-spin band, it is mainly the
missing number of majority d-states that causes the concentration dependence.
Thus, the slope of the average moment against the total number of electrons varies
depending on the solute atoms; the smaller the difference in the number of the
valence electrons, the steeper the slope. The above discussion, however, is useful
only for simple cases where the magnetic state is rather stable, typically in the
region of the strong ferromagnetism of Ni. More delicate situations as, for example,
the Ni-Fe, Ni-Mn, and Fe-Mn alloys in the vicinity where ferromagnetism becomes
instable, however, need a more detailed and specific discussion [93].
An important feature of the CPA calculations is that the alloy components
essentially keep their intrinsic properties. Fe, Co, and Ni, for example, have in
general in the various alloys a spin moment close to that of the pure elements (2.2,
1.7, and 0.6 μB , respectively; see Table 1). This is fully in line with results of neutron
scattering experiments or XMCD (X-ray magnetic circular dichroism) experiments
that give access to the spin and orbital moment in an element-specific way via the
XMCD sum rules [9]. Figure 10 shows corresponding results for the average spinand orbital magnetic moment per atom in fcc-Fex Ni1−x together with componentspecific data as calculated via the relativistic KKR-CPA on the basis of the LSDA
and LSDA+DMFT, respectively, in comparison with experiment [95]. As one notes,
the individual spin-magnetic moments of Fe and Ni in fcc-Fex Ni1−x show only a
rather weak concentration dependence. This also applies for the spin-orbit-induced
orbital moments that are defined by the expectation value of the angular momentum
operator lz (see, e.g., the discussion in [17]). These findings are in full agreement
with the individual moments determined via XMCD at the L2,3 -edges of Fe and Ni
[95]. Here, it is interesting to note that the theoretical results for the spin moment
hardly depend on the computational mode, i.e., whether the calculations are based
on the LSDA or the LSDA+DMFT. The orbital moment, on the other hand, depends
strongly on the computational mode. In particular it is found that inclusion of
correlation effects via the DMFT improves agreement with experiment.
Heusler Alloys
There are plenty of experimental and theoretical investigations on Heusler alloys in
the literature because of their very rich variety of magnetic properties including
214
H. Ebert et al.
1.5
0.08
1
LSDA
LSDA+DMFT
Expt
0.5
0
0
0.2
0.4
xFe
0.6
0.04
0
0
0.8
3
mspin (μB)
morb (μB)
LSDA
LSDA+DMFT
Expt
1
0
0.8
0.2
0.4
xFe
0.1
0.05
0.6
0
0
0.8
0.2
0.4
xFe
0.6
0.8
0.25
1
Ni
0.2
morb (μB)
0.8
LSDA
LSDA+DMFT
Expt
Ni
0.15
0.6
0.4
0
0
0.6
0.15
2
0.2
0.4
xFe
LSDA
LSDA+DMFT
Expt
Fe
2.5
1.5
0.2
0.2
Fe
mspin (μB)
LSDA
LSDA+DMFT
Expt
0.12
morb (μB)
mspin (μB)
2
LSDA
LSDA+DMFT
Expt
0.2
0.4
xFe
0.1
0.05
0.6
0.8
0
0
0.2
0.4
xFe
0.6
0.8
Fig. 10 Top row: average spin- (left) and orbital (right) magnetic moment per atom in fccFex Ni1−x as calculated on the basis of the LSDA and LSDA+DMFT, respectively, in comparison
with experiment. In addition, the individual moments calculated for Fe (middle row) and Ni (bottom row) are given. The experimental component-specific data stem from XMCD measurements.
(All data taken from [95])
ferromagnetism, antiferromagnetism, helimagnetism, and Pauli paramagnetism.
Depending on their crystal structure, one can distinguish two families of Heusler
alloys: semi-Heuslers of the type XYZ with C1b structure and full-Heuslers of the
type X2 YZ with L21 structure. Most of the Heusler alloys are metals; however, for
some systems also, half-metallic and semiconducting behavior has been observed.
The half-metallicity found in certain Heusler magnets [62, 87, 96, 97, 98]
attracted especially strong interest over the last 30 years because of its possible
4 Electronic Structure: Metals and Insulators
215
Fig. 11 Left: spin-resolved DOS for Co2 MnSi showing a bandgap for the minority states. Right:
comparison of the spin polarization obtained by in situ SRUPS on a Co2 MnSi thin film with
the calculated DOS-derived spin polarization, the calculated UPS spin polarization including
broadening effects and considering only bulk states, and the calculated total UPS spin polarization
including broadening effects with additional surface state contributions [99]
use in spintronics and magneto-electronics. Half-metallic materials exhibit metallic
behavior only for one spin direction, while for the other spin direction the Fermi
level is located in a bandgap. Figure 11 shows the spin-resolved DOS for Co2 MnSi
as a representative example. Defining the spin polarization p of a material in terms
of the spin-dependent density of states nσ (E) according to
p=
n↑ (E) − n↓ (E)
n↑ (E) + n↓ (E)
(50)
one has for half-metallic materials 100% spin polarization at the Fermi energy that
should lead to a fully spin-polarized electric current.
For real materials, however, the spin polarization may be influenced in various
ways. Mavropoulos et al. [100], for example, demonstrated the impact of spin
mixing caused by spin-orbit coupling for the Heusler alloys of the type XMnSb.
As expected, this influence increased with increasing atomic number of the element
X, as reflected by the ratio n↓ (EF )/n↑ (EF ) = 0.25, 0.30, 0.35, 0.75, and 2.70 found
for the series X = Co, Fe, Ni, Pd, and Pt. As a consequence, one can expect a spin
polarization well below 100% for compounds with heavier elements.
The influence of the surface on the spin polarization has been studied by
Galanakis [101] investigating the (001) surfaces of the semi-Heusler alloys
NiMnSb, CoMnSb, and PtMnSb and for the full-Heusler alloys Co2 MnGe,
Co2 MnSi, and Co2 CrAl. In general, a rather strong modification of electronic
and magnetic properties has been found for the surface region. In the case of semiHeuslers, the Ni-, Co-, or Pt-terminated surface has a rather large DOS at the Fermi
level for minority-spin states, while for the MnSb-terminated surfaces the calculated
properties are close to those of the bulk. Nevertheless, half-metallicity disappears
also in this case due to surface states, resulting in a spin polarization at the Fermi
216
H. Ebert et al.
level of 38%, 46%, and 46% for NiMnSb, CoMnSb, and PtMnSb, respectively. A
similar behavior was found for the Co-terminated surfaces of the full-Heusler alloys
Co2 MnGe, Co2 MnSi, and Co2 CrAl. In the case of the MnGe-terminated (001)surface of Co2 MnGe, the spin polarization vanishes also due to the surface states,
although the CrAl termination of Co2 CrAl leads to a very high spin polarization of
around 84%.
In experiment, the spin polarization of a material can be determined among
others by tunneling experiments or by spin-resolved photoemission [102, 103, 104].
Figure 11 shows corresponding experimental and theoretical spin-polarization
curves of Co2 MnSi [99] for the UPS-regime (hν = 21.2 eV) that showed for the
surface regime a value of 93% at the Fermi level, the largest observed so far.
NiMnSb was among the first semi-Heusler compounds predicted to be halfmetallic [105] and was intensively investigated since then by experiment as well
as theory [106, 107, 108]. The Sb atoms with the atomic configuration 5s2 5p3 lead
to the formation of a deep lying narrow s- and a p-band at around 12 and 3 − 5 eV,
respectively, below the Fermi energy EF . Accordingly, these states are not involved
in the formation of the bandgap near EF . The Ni and Mn d-states hybridize with
each other as well as with the sp-states of Sb leading to the formation of bonding
and antibonding bands. For the paramagnetic state of NiMnSb, the Fermi level
lies in the middle of an antibonding band mainly associated with the d-states of
Mn. Accordingly, exchange splitting of the antibonding band leads to a gain in
energy accompanied by formation of a strong magnetic moment for the Mn atom.
As a result, the Fermi energy moves to the energy gap separating bonding and
antibonding minority-spin states. Due to this, there are nine minority-spin states per
unit cell below EF , one due to the Sb-like s-band, three due to the Sb-like p-bands,
and five due to the Ni-like d-bands, that are all occupied. As atoms forming the alloy
contribute 22 electrons per unit cell, the majority-spin band contains 22 − 9 = 13
electrons, resulting in a moment of 4 μB per unit cell.
Plotting the total spin-magnetic moment per unit cell, M, of NiMnSb together
with that of other semi-Heusler half-metallic compounds as a function of the total
number of valence electrons Z, one can see that M – in analogy to the Slater-Pauling
curve for the binary transition metal alloys – follows the relation: M = Z −18 [107]
(see Fig. 12 (left)). This relation is a consequence of the complete occupation of the
nine minority-spin bands and follows directly from the definitions Z = N↑ + N↓
and M = N↑ − N↓ that lead to mt = Z − N↓ [98].
The occurrence of half-metallicity in the case of the prototype full-Heusler
compounds Co2 MnSi and Co2 MnGe was also predicted by electronic structure
calculations [109, 110]. Similar to the case of semi-Heusler alloys, the sp-bands
of Si and Ge are located well below EF . For that reason, they do not participate in
the formation of the energy gap that is caused by the hybridization of Mn and Co dstates. As a result, for the spin-polarized state of the material, the Fermi level is again
located within the minority-spin energy gap, so that the minority band contains 1 sband and 3 p-bands derived from the sp-element and 8 Co-related d-bands, which
are fully occupied by 12 electrons [98]. As shown by Galanakis et al. [111], the
spin-magnetic moment of the full-Heusler alloys accordingly follows the relation
M = Z − 24. Corresponding theoretical and experimental data are summarized in
4 Electronic Structure: Metals and Insulators
217
6
NiMnSe
NiMnTe
0
18
-2
20
22
Z
24
20
M
CoTiSb
Co2MnSb
Co2FeSi
Co2FeAl
Ni2MnAl
Rh2MnGe
Rh2MnSn
Rh2MnPb
Co2VAl
Fe2MnAl
2
=
Z18
=
M
16
CoVSb
FeMnSb
CoCrSb
NiVSb
Co2MnAs
Rh2MnIn
Rh2MnTl
Co2TiSn
Fe2CrAl
Co2TiAl
Fe2VAl
Mn2VGe
24
2
CoMnSb
IrMnSb
NiCrSb
M
M
RhMnSb
4
CoFeSb
NiFeSb
Co2CrAl
Fe2MnSi
Ru2MnSi
Ru2MnGe
Ru2MnSn
Z-
NiMnSb
PdMnSb
PtMnSb
4
0
Co2MnSi
Co2MnGe
Co2MnSn
6
Co2MnAl
Co2MnGa
Rh2MnAl
Rh2MnGa
Ru2MnSb
Mn2VAl
22
24
26
Z
28
30
32
Fig. 12 Left: calculated total spin moments per unit cell for several semi-Heusler alloys.
Experimental values are given for NiMnSb (3.85 μB ), PdMnSb (3.95 μB ), PtMnSb (4.14 μB ), and
CoTiSb (nonmagnetic) [107]. Right: calculated total spin moments for several full-Heusler alloys.
Experimental values are given for Co2 MnAl (4.01 μB ), Co2 MnSi (5.07 μB ), Co2 MnGa (4.05 μB ),
Co2 MnGe (5.11 μB ), Co2 MnSn (5.08 μB ), Co2 FeSi (5.9 μB ), Mn2 VAl (1.82 μB ), and Fe2 VAl
(nonmagnetic) [111]. In both figures the dashed line represents the corresponding Slater-Pauling
curve. The open circles represent the compounds deviating from this curve
the Slater-Pauling plot in Fig. 12 (right). The difference to the Slater-Pauling curve
of the binary transition metal alloys in Fig. 9 is due to the fixed number of minorityspin electrons in the half-metallic Heusler compounds. In this case, increasing Z
leads to a filling of the majority band, while for the ferromagnetic transition metal
alloys, the relation M = 10 − Z for the right branch is a result of the full occupation
of five majority-spin d-states and charge neutrality achieved by filling the minorityspin d-states.
Within the family of Heusler compounds, there are furthermore the so-called
inverse full-Heusler compounds that have a similar chemical formula X2 YZ but
crystallize in the so-called Xα structure [98]. The prototype of the inverse fullHeusler compounds is Hg2 TiCu [112]. As Fig. 13 demonstrates, the magnetic
moments per unit cell as a function of valence electrons also follow in this case
corresponding Slater-Pauling rules.
Total Electronic Energy and Magnetic Configuration
The calculation of the electronic total energy allows to seek for the magnetic ground
state of a solid. This applies to the crystal structure as well as to the specific magnetic
ordering or spin configuration, respectively. For many purposes, it is helpful to
represent the neighborhood of the ground state or a suitable magnetic reference
state by mapping the complex configuration dependence of the total energy of the
system on an approximate spin Hamiltonian. An issue in this context is spin-orbit
coupling that removes energetic degeneracies of competing spin configurations, may
218
H. Ebert et al.
M
2
0
-2
-4
M = Z-18
Ti2NiAl
Ti2CuAl
Ti2CoAl Ti2CoSi
Ti2FeAs
Ti2FeSi
Sc2NiAl
V2MnAl
Sc2CoSi
V2CrSi
Ti2FeAl
Sc2NiSi
V2CrAl
Ti2MnAs
Ti2MnAl
Ti2CrAl
Sc2MnAl
Ti2VAs
Sc2FeAl
Sc2CrSi Ti TiSi
V2VAl
2
Sc2VAs
Sc2MnSi
Sc
CrAs
Ti2VAl
2
Ti2VSi
Sc2VSi
Sc2VAl
Sc2CrAl
13
15
17
19
21
23
M = Z-24
4
2
M
4
0
-2
V2CoSi
V2FeAs
Cr2CrAs
Cr2FeAl
Mn2CrAl
Cr2CoAs
Mn2FeSi
Mn2MnAs
Mn2CoAl
Cr2CrSi
V2MnSi
V2FeAl
-4
Cr2CrAl
20
22
Mn2NiAs
Mn2CoAs
Mn2NiSi
Mn2FeAs
Mn2CoSi
Cr2NiSi
V2NiAs
Cr2CoAl
Cr2MnAs
V2NiAl
Mn
MnAl
2
Cr2MnSi
Mn2CrSi
V2CoAl
V2FeSi
V2MnAs
Cr2MnAl
24
Z
26
28
Cr2NiAl
Cr2CoSi
Cr2FeAs
Mn2FeAl
Mn2MnSi
Mn2CrAs
30
Z
Fig. 13 Total spin-magnetic moments per unit cell (in μB ) as a function of the total number
of valence electrons Z in the unit cell for several compounds. The lines represent the two of the
various forms of the Slater-Pauling rule [98]. The compounds within the frames follow one of these
rules and are perfect half-metals, while the rest of the alloys slightly deviate. For this reason, their
total spin-magnetic moment is represented by an open red circle. The sign of the spin-magnetic
moments has been chosen so that the half-metallic gap is in the spin-down band
lead to the anisotropic Dzyaloshinsky-Moriya exchange interaction, and gives rise
to magnetocrystalline anisotropy.
Total Electronic Energy and Magnetic Ground State
Access to the electronic total energy Etot (see Eq. (10)) in principle allows to
determine the magnetic ground state of any solid. As a corresponding example
for this, results of LSDA-based calculations for Fe in the para- (PM), ferro- (FM),
as well as antiferromagnetic (AFM) state with bcc and fcc structure are shown in
Fig. 14 [27]. Obviously, the use of the LSDA led to an over-binding, i.e., to a lattice
parameter that is too small and a bulk modulus that is too large when compared
with experiment (see Table 1). Most importantly, however, the paramagnetic fccphase was found as the ground state instead of the ferromagnetic bcc-phase. Use of
the GGA on the other hand improved the situation very much giving in particular
the ferromagnetic bcc-phase as the ground state.
Another example for a search for the magnetic ground state configuration is given
by Fig. 15. In this case, the so-called fixed spin moment method was used to explore
the dependency of the total energy Etot on the lattice parameter and average atomic
moment of disordered fcc-Fex Ni1−x alloys. As one notes, a double minimum occurs
in the vicinity of the concentration x = 0.65 indicating the competition between a
low-volume, low-spin moment phase and a high-volume, high-spin moment phase.
In fact, this has been seen as a possible explanation for the occurrence of the invar
effect for that composition. Another study on the invar effect, however, stressed the
importance of a noncollinear spin configuration [114].
4 Electronic Structure: Metals and Insulators
219
400
600
200
E (meV)
E (meV)
PM bcc
FM bcc
0
PM bcc
400
200
PM fcc
AFM fcc
PM fcc
FM bcc
0
-200
60
65
70
75
Volume (a.u.)
70
80
80
75
85
Volume (a.u.)
90
Fig. 14 Total energy Etot of paramagnetic (PM) bcc and fcc, ferromagnetic (FM) bcc, and
antiferromagnetic (AFM) fcc-Fe as a function of the volume as obtained within the LSDA (left)
and GGA (right), respectively. The curves are shifted in energy so that the minima of the FM-bcc
curves, corresponding to Etot = 0, coincide. (All data taken from [27])
2.0
2.0
M avg ( μ Β/atom )
2.5
M avg ( μ Β/atom )
2.5
1.5
1.0
1.0
0.5
0.0
6.3
1.5
Fe 60 Ni 40
6.4
6.5
6.6
a (a.u.)
6.7
6.8
6.9
0.5
0.0
6.3
Fe 65 Ni 35
6.4
6.5
6.6
6.7
6.8
6.9
a (a.u.)
Fig. 15 Total energy or binding surfaces for the disordered ferromagnetic alloys Fe60 Ni40 and
Fe65 Ni35 . (All data taken from [113])
The occurrence of a noncollinear spin configuration is quite common for actinide
compounds but also for a number of transition metal-based systems [115]. A
prominent example for this is Mn3 Sn with its hexagonal unit cell given in Fig. 16.
Scalar relativistic calculations gave for all shown noncollinear spin configurations
a total energy well below that of competing collinear spin configurations. Ignoring
spin-orbit coupling, only the angle between the spin moments is relevant leading
to a degeneracy for all four spin configurations. However, if spin-orbit coupling is
taken into account, the degeneracy is removed and the resulting moment may deviate
to some extent from the orientation found for the scalar relativistic calculations
(see thin arrows for configuration (c) and (d) in Fig. 16). The presence of spinorbit coupling implies that the spin-magnetic moment is accompanied by an orbital
one. It is interesting to note that the orientation of the spin-orbit-induced orbital
magnetic moment may and will deviate from that of the spin moment if symmetry of
the system allows. Another interesting property of spin-compensated noncollinear
220
H. Ebert et al.
Mn
Sn
z=1/4
z=3/4
(a)
(b)
1
2
1
3
2
3
2
(c)
(d)
1
1
3
2
3
Fig. 16 Crystal and magnetic structure of Mn3 Sn. Rotations of the magnetic moments leading
to weak ferromagnetism in the structure in (c) and (d) are shown only for atoms in the z = 0.25
plane (thin arrows). Moments of the atoms in the z = 0.75 plane are parallel to the moments of the
corresponding atoms of the z = 0.25 plane. (All data taken from [115])
antiferromagnets having certain magnetic space groups is the occurrence of the
anomalous Hall effect that is usually not expected for an antiferromagnet [79].
However, group theoretical considerations unambiguously show that the Hall effect
may even show up for spin-compensated solids [116]. The first observations of the
AHE [117] as well as of its thermoelectric analog, the anomalous Nernst effect
[118], in a non-collinear antiferromagnet could in fact both be made in Mn3 Sn. The
occurrence of spin-polarized currents in this and related materials currently attracts a
lot of attention as well [119,120]. Finally, a criterion for the instability of a collinear
spin structure with respect to a transition to a noncollinear one was formulated by
Sandratskii and Kübler: If the collinear magnetic structure under consideration is
not distinguished by symmetry compared with the noncollinear structures obtained
with infinitesimal deviations of the magnetic moments from collinear directions, this
structure is unstable [121].
Exchange Coupling Parameters
When dealing with competing magnetic configurations, it is not always possible or
not necessary to perform full ab-initio calculations. In these cases, one may adopt a
multi-scale approach that uses the classical Heisenberg Hamiltonian:
4 Electronic Structure: Metals and Insulators
H=−
221
Jij m̂i · m̂j ,
(51)
i=j
with m̂i(j ) the orientation of the magnetic moment on the lattice site i(j ), for
corresponding simulations. The isotropic exchange coupling parameters Jij , on
the other hand, are calculated in an ab initio way. This can be done by applying
a corresponding version of the so-called Connolly-Williams method [122]. This
implies to calculate the total energy for many different magnetic configurations
within a super-cell and to determine the exchange coupling parameters by fitting
the energy on the basis of Eq. (51). This way one obviously achieves a mapping
of the complicated energy landscape E({m̂i }) on the rather simple expression in
Eq. (51) involving only pair interactions that is easy to handle. Accordingly, a more
accurate representation of the energy as a function of the magnetic configuration
can therefore be expected by a cluster expansion as suggested by various authors
[123, 124].
Another approach to determine the exchange coupling parameters Jij in Eq. (51)
is to consider the change of the single-particle energy ΔEij if two magnetic
moments on sites i and j change their relative orientation. The necessary formal
developments started with the work of Oguchi et al. who expressed the difference
in energy between the ferro- and antiferromagnetic state of a solid making use of
multiple scattering theory and Lloyd’s formula [125]. Lichtenstein et al. [126]
extended this approach dealing with the coupling energy ΔEij associated with an
individual pair (i, j ) of atoms. If ΔEij is expressed to lowest order with respect to
the orientation angle of the moments m̂i and m̂j , one gets a one-to-one mapping of
the exchange coupling energy ΔEij to the Heisenberg Hamiltonian in Eq. (51), with
the exchange coupling constants Jij given by [126]:
Jij =
1
4π
EF
ij
ji
−1
−1
−1
τ↑ tj−1
dE Trace ti↑
− ti↓
↑ − tj ↓ τ↓ ,
(52)
ij
where ti↑(↓) is the spin-dependent single site scattering matrix for site i and τ↑(↓) is
the spin-dependent scattering path operator matrix connecting sites i and j . Results
for the isotropic exchange coupling parameter Jij of Fe and Co as a function of
the distance Rij between sites i and j that have been obtained using an analogous
expression derived within the LMTO-GF formalism [127] are shown in Fig. 17. The
big advantage of this approach is that it can be applied with comparable effort to
more complex systems like disordered substitutional alloys [128], Heusler alloys
[129, 130, 131], diluted magnetic semiconductors [132, 133, 134, 135, 136, 137],
magnetic surface films [127,138], or finite deposited clusters [139,140]. In addition,
it should be mentioned that an approach similar to that leading to Eq. (52) was
worked out by Katsnelson and Lichtenstein [141] that allows for an improved
treatment of correlated systems.
If spin-orbit coupling is accounted for, the exchange coupling parameter in the
Heisenberg Hamiltonian has to be replaced by a corresponding tensor:
222
H. Ebert et al.
20
16
20
16
Fe
12
Jij (meV)
Jij (meV)
12
Co
8
8
4
4
0
0
-4
-4
1
2
1.5
2.5
Rij (units of lattice parameter)
1
2
1.5
2.5
Rij (units of lattice parameter)
Fig. 17 Isotropic exchange coupling constants Jij of Fe and Co as a function of the distance Rij
between sites i and j . (All data taken from [127])
H=−
i=j
=−
i=j
−
m̂i J m̂j +
ij
K(m̂i )
(53)
i
Jij m̂i · m̂j −
m̂i J S m̂j
ij
i=j
D ij · m̂i × m̂j +
Ki (m̂i ),
i=j
(54)
i
with the accompanying single-site magnetic anisotropy represented by the term
K(m̂i ). In Eq. (53), the coupling tensor J has been decomposed into its isotropic
ij
part Jij , its traceless symmetric part J S , and its antisymmetric part. The latter
ij
one, that may occur in case of systems without inversion symmetry, is often
represented in terms of the so-called Dzyaloshinsky-Moriya (DM) vector D ij , with
βγ
γβ
Dijα = 12 (Jij − Jij ) and a cyclic sequence of the Cartesian indices α, β, and
γ . A corresponding generalization of the nonrelativistic expression for Jij given in
Eq. (52) to its relativistic tensor form was worked out by various authors [18, 142]
and applied in particular to cluster systems [124, 143] with the interest focusing on
the impact of the DM interaction.
It should be emphasized once more that Eq. (51) and extensions to it supply an
approximate mapping of the complicated energy landscape E({m̂i }) of a system
calculated in an ab initio way onto a simplified analytical expression. This implies
corresponding limitations [124] in particular due to the use of the rigid spin
approximation (RSA) [144]. It is interesting to note that a coupling tensor of the
same shape as in Eq. (53) occurs for the indirect coupling of nuclear spins mediated
by conduction electrons. In this case the mentioned restrictions do not apply. As a
consequence, the linear response formalism on the basis of the Dyson equation can
4 Electronic Structure: Metals and Insulators
223
be used without restrictions to determine the corresponding nuclear spin-nuclear
spin coupling tensor [145].
Another approach using the same idea as the Connolly-Williams method is based
on the total energy ΔE(q, θ ) = E(q, θ ) − E(0, θ ) calculated for noncollinear spin
spirals (see section “Spin Spiral Calculations”) that are characterized by the wave
vector q and tilt angles (θ, φ(R i ) = q · R i ) for the moment mi on the atomic
position R i . In the case of a small tilt angle θ , ΔE(q, θ ) can be represented in terms
of the Fourier transform J (q) of the real space exchange coupling parameters [146]:
E(q, θ ) = E0 (θ ) −
θ2
J (q).
2
(55)
Accordingly, performing an inverse Fourier transformation, one can determine the
real-space interatomic exchange coupling parameters J0j . In the case of a simple
Bravais lattice, this is given by the expression
J0j =
1 −iqR 0j
e
J (q).
N q
(56)
Uhl et al. [147] applied this scheme among others for a study of the invar system
Fe3 Pt. From their numerical results for ΔE(q, θ ), they evaluated the spin-wave
stiffness constant A and the exchange parameter J0 that allows to give an estimate
for the Curie temperature on the basis of the mean field approximation (MFA) (see
section “Finite-Temperature Magnetism”). Similar work was also done for twodimensional systems as for example magnetic surface films. Within their study on
an Fe film on W(110), Heide et al. also accounted for the spin-orbit coupling [148].
This way they could determine not only the spin-wave stiffness constant A but also
the Dzyaloshinsky-Moriya interaction vectors D.
As discussed in section “Spin Density Functional Theory”, dealing with systems
with narrow electronic energy bands, in order to go beyond the local spin-density
approximation, the LDA+U or +DMFT methods can be used to properly account
for strong electronic correlation in these materials. In order to adapt the method for
the treatment of exchange interactions formulated within the LSDA [126, 18, 142],
a corresponding theory was developed by Katsnelson and Lichtenstein [141] that
employs an analog of the local force theorem to derive expressions for effective
exchange parameters, Dzyaloshinsky-Moriya interaction, and magnetic anisotropy
in highly correlated systems. The authors demonstrated for the particular case of
ferromagnetic Fe that treating correlation effects beyond the LSDA (within the
LDA+Σ approach) in the exchange interactions results in a spin-wave spectrum
and spin-wave stiffness which are in better agreement with experiment than those
obtained within plain LSDA. The important role of additional contributions to the
exchange coupling of the correlation interactions has been demonstrated, e.g., for
half-metallic ferromagnetism in CrO2 [149], magnetic properties of CaMnO3 [150],
and magnetic properties of transition metal oxides [151]. Another important
feature of the exchange interactions observed in various materials are their strong
224
H. Ebert et al.
orientation dependence that would require to go beyond the Heisenberg model when
considering the finite-temperature or spin-wave properties of these materials. This
problem was recently discussed by different groups [151, 152, 153, 154].
Magneto-Crystalline Anisotropy
Magnetocrystalline anisotropy denotes the dependence of the total energy of
a system on the orientation m̂ of its magnetization with the anisotropic part
EA (m̂) of the energy taking a minimum for m̂ along a so-called easy direction
of the magnetization. Usually, EA (m̂) is split into the intrinsic material-specific
magnetocrystalline anisotropy (MCA) energy EMCA (m̂) and the extrinsic shape
anisotropy energy Eshape (m̂) determined by the shape of the sample:
EA (m̂) = EMCA (m̂) + Eshape (m̂).
(57)
Considering the difference in energy ΔEX (m̂, m̂ ), with X=A, MCA or shape,
respectively, for the magnetization oriented along directions m̂ and m̂ , respectively,
one has accordingly ΔEX (m̂, m̂ ) = EX (m̂) − EX (m̂ ).
A convenient phenomenological representation of the magnetocrystalline
anisotropy energy can be given by an expansion in terms of spherical harmonics
Ylm (m̂)
EMCA (m̂) =
m=l
κlm Ylm (m̂)
(58)
l even m=−l
or alternatively by an expansion in powers of the direction cosines (α1 , α2 , α3 ) =
(m̂ · x̂, m̂ · ŷ, m̂ · ẑ):
EMCA (m̂) = b0 +
i,j
bij αi αj +
bij kl αi αj αk αl + . . .
(59)
i,j,k,l
Assuming degeneracy of the energy upon time-reversal, i.e., flip of the magnetization, only terms that are even with respect to the orientation m̂ of the magnetization
can occur in these equations. Further restrictions on the expansions are imposed by
the crystal symmetry of the investigated material [7]. Considering, for example,
a hexagonal system with the expansion up to sixth order, the corresponding
hex (m̂) are given by:
expressions for EMCA
hex
EMCA
(m̂) = K̃0 + K̃1 Y20 (θ, φ) + K̃2 Y40 (θ, φ)
+K̃3 Y60 (θ, φ) + K̃4 Y64 (θ, φ)
(60)
= K0 + K1 (α12 + α22 ) + K2 (α12 + α22 )2 + K3 (α12 + α22 )3
+K4 (α12 − α22 ) (α14 − 14α12 α22 + α24 ),
(61)
4 Electronic Structure: Metals and Insulators
225
2
where the coefficients are interconnected by the relations K̃1 = 21
(7K1 + 8K2 +
8
16
1
8K3 ), K̃2 = 385 (11K2 + 18K3 ), K̃3 = 231 K3 and K̃4 = 10,395 K4 .
The shape anisotropy energy Eshape (m̂) is usually associated with the classical
dipole-dipole interaction of the individual magnetic moments mν on the lattice sites
[6, 155]. Accordingly, it can be determined straightforwardly by a corresponding
lattice summation:
Edip (m̂) =
mν mν c2
νν Rn
[R nνν · m̂]2
1
−
3
,
|R nνν |3
|R nνν |2
1
(62)
with R nνν = R n + ρ ν − ρ ν , where a periodic system has been considered with
lattice vectors R n and ρ ν the basis vectors within the unit cell.
For transition metal systems, the intrinsic part of the magnetic anisotropy
energy EMCA (m̂) has to be ascribed to the spin-orbit coupling. Accordingly, the
corresponding energy difference ΔESOC (m̂, m̂ ) can be determined by total energy
calculations with the magnetization oriented along directions m̂ and m̂ , respectively, and taking the difference. Obviously, this implies a full SCF calculation for
both orientations and taking the difference of large numbers to get ΔESOC (m̂, m̂ ).
The problems connected with this approach can be avoided by making use of the
so-called magnetic force theorem that allows to approximate ΔESOC (m̂, m̂ ) by
the difference of the single particle or band energies (see Eq. (10)) for the two
orientations obtained using a frozen spin-dependent potential [6, 156]:
ΔESOC (m̂, m̂ ) = −
EFm̂
dE N m̂ (E) − N m̂ (E)
1 − nm̂ (EFm̂ ) (EFm̂ − EFm̂ )2 + O(EFm̂ − EFm̂ )3 .
2
(63)
Here EFm̂ is the Fermi energy for the magnetization along m̂ while nm̂ (E) and
E
N m̂ (E) =
dE nm̂ (E ) are the corresponding DOS and integrated DOS,
respectively.
This approach is used extensively for compounds and layered systems and leads
typically to anisotropy energies that deviate less than 10% from results obtained
from full SCF calculations. By using, in the case of layered systems, layer-resolved
data for the DOS in Eq. (63), a corresponding layer decomposition of the anisotropy
energy ΔESOC (m̂, m̂ ) could be achieved [157]. Application of this scheme for the
spatial decomposition of ΔESOC (m̂, m̂ ) shows that the dominating contributions
originate in general from the interface or surface layers, respectively.
Equation (63) implies that spin-orbit coupling is accounted for within the
underlying electronic structure calculations. Instead one can start from a scalar
relativistic calculation and treat HSOC as a perturbation. Solovyev et al. [158] used
this approach on the basis of the Green function method in combination with the
Dyson equation Eq. (21). This allowed to write the spin-orbit-induced correction
226
H. Ebert et al.
ESOC (m̂) to the single-particle energies as a sum of two-site interactions:
ESOC (m̂) =
EF
dE δN(E) =
Eij (m̂)
(64)
ij
with
Eij (m̂) = −
1
2π
EF
ij
j
ji
i
dE Trace G0 (m̂) HSOC G0 (m̂) HSOC
,
(65)
ij
where G0 (m̂) are real space structural Green function matrices corresponding to the
scattering path operator in Eq. (20) [42]. In contrast to Eq. (63), Eq. (65) provides
a unique spatial or component-wise decomposition of the magnetocrystalline
energy. Solovyev et al. [158] used this approach for a detailed study of the
ordered compounds TX with T = Fe, Co and X = Pd, Pt having CuAu structure.
This way they could in particular show that the hybridization between the T
and X sublattices essentially determines their magnetocrystalline anisotropy. In
addition, an expression analogous to Eq. (65) allowed to demonstrate and discuss
the interconnection between the energy correction E(m̂) and the spin-orbit-induced
orbital magnetic moment μorb represented by the expectation value of the angular
momentum operator l. Using a similar approach as sketched here, this relation
was already investigated before by Bruno [159] and also by van der Laan [160].
Assuming a strong ferromagnet with the majority band filled, the relation:
1
ESOC (m̂) = − C ζ σ · l ,
4
(66)
was derived, where C is a constant and ζ represents the strength of the spin-orbit
coupling. This equation was used in numerous experimental studies that exploited
the XMCD (X-ray magnetic circular dichroism) and the associated sum rules [9]
to determine in an element specific way the change of the angular momentum Δl
when changing the orientation of the magnetization from m̂ to m̂ to get a component
resolved estimate for the corresponding anisotropy energy ΔESOC (m̂, m̂ ).
A further approach to calculate the spin-orbit-induced anisotropy energy is to
consider the torque T (θ ) exerted on a magnetic moment m when the magnetization
is tilted by the angle θ away from its equilibrium orientation (easy axis). The
corresponding expression for T (θ ),
T (θ ) =
j k occ
ψj k
∂HSOC
ψj k ,
∂θ
(67)
was given first by Wang et al. [161] for the case that the electronic structure is
represented in terms of Bloch states. A more general expression was obtained on
the basis of multiple scattering theory [162]:
4 Electronic Structure: Metals and Insulators
Tαm̂û = −
1
π
EF
dE
∂
−1
0
ln
det
t(
,
m̂)
−
G
∂α û
227
(68)
where the torque component with respect to a rotation of the magnetization around
an axis û is considered and where t(m̂) is the single-site t-matrix for an orientation
of the moments along m̂ and G0 is the corresponding free electron Green function
matrix.
On the basis of Eq. (67) or (68), respectively, the anisotropy energy
ΔESOC (m̂, m̂ ) is obtained from the torque by integrating along a path connecting
m̂ and m̂ . This approach is especially suited when dealing with systems with
uniaxial anisotropy. Neglecting in this case the dependence on φ, the anisotropy
energy can be represented as ESOC (θ ) = K0 + K1 sin2 (θ ) + K2 sin4 (θ ) with the
torque given by:
T (θ ) =
dESOC (θ )
= K1 sin(2θ ) + 2K2 sin(2θ ) sin2 θ.
dθ
(69)
For the special setting θ = π/4 and φ = 0, one has therefore:
ESOC (π/2) − ESOC (0) = K1 + K2 = T (π/4).
(70)
This implies that if the contribution K1 sin2 (θ ) to ESOC (θ ) dominates, a situation
often met, K1 and with this ESOC (θ ) can be obtained from a single calculation for
the special settings. Otherwise, K1 and K2 can be obtained by a fit to a sequence of
calculations for varying angles θ .
The contribution ΔEdip (m̂, m̂ ) to the total anisotropy energy that is associated
with the dipole-dipole interaction of the individual magnetic moments is usually
treated classically by evaluating a corresponding Madelung sum (see Eq. (62)) [155,
163, 164]. While for most cases ΔESOC (m̂, m̂ ) is much larger than ΔEdip (m̂, m̂ ),
both contributions are often found for layered systems to be in the same order
of magnitude. As ΔEdip (m̂, m̂ ) always favors an in-plane orientation of the
magnetization while ΔESOC (m̂, m̂ ) in general favors an out-of-plane orientation,
one may have a flip of the easy axis from out-of-plane to in-plane with increasing
thickness of the magnetic layers. Such a behavior has been found, for example, for
Con Pdm multilayers as shown in Fig. 18 [163]. Similar results were obtained for the
magnetic surface layer system Fen /Au(001) that shows a change from out-of-plane
to in-plane anisotropy if the number n of Fe layers is larger than 3 [155].
In particular in cases for which ΔESOC (m̂, m̂ ) and ΔEdip (m̂, m̂ ) are of the
same order of magnitude, it seems questionable to treat the first contribution
quantum mechanically and the second one in a classical way. Although it was
pointed out already nearly 30 years ago that ΔEdip (m̂, m̂ ) is caused by the Breit
interaction [14], there is only little numerical work done in this direction [165,166].
Including a vector potential in the Dirac equation Eq. (23) that represents the
corresponding current-current interaction such numerical work has been done on
magnetic surface films and multilayer systems. It turned out in all investigated
228
H. Ebert et al.
Co1Pd2
0.34
Co4Pd2
0
0
0.4
ΔE (meV)
Co2Pd4
1
ΔE (meV/unit cell)
0.68
Co1Pd5
2
Kt (mJ/m )
2
0
-0.4
-0.8
Co3Pd3
-1
2
4
6
8 10 12 14
t (Å)
-0.34
-1.2
1
2 3 4
5 6
Number of Fe layers
Fig. 18 Left: calculated total anisotropy energy ΔE of Con Pdm multilayers with (111)-oriented
fcc structure as a function of the thickness t of the magnetic Co layers. Corresponding experimental
data for the product of the anisotropy energy density K and Co thickness t are shown for
polycrystalline films deposited at two different temperatures (triangles up and down). (All data
taken from [163]) Right: SOC-induced (ΔESOC ; circles) and dipole-dipole (ΔEdip ; triangles)
contributions to the total anisotropy energy (ΔE; squares) for the magnetic surface layer system
Fen /Au(001) as a function of the number n of Fe layers. (All data taken from [155])
cases that the classical treatment on the basis of the dipole-dipole interaction leads
to results for ΔEdip (m̂, m̂ ) that are very close to those of a coherent quantummechanical calculation that accounts also for the Breit interaction.
Starting from the 1950s, compounds of rare-earth (RE) with 3d transition metal
(TM) elements, as, for example SmCo5 , or Nd2 Fe14 B, attracted much attention
because of their strong magnetic anisotropy. In these materials, the MCA is
primarily determined by the RE sublattice, while the TM sublattice is responsible
for the magnetic ordering [167]. For that reason, the simplified two-sublattice
Hamiltonian
f
d-f
H = Hd + HCEF
+ Hex
(71)
f
is often used to discuss their properties, where Hd and HCEF
characterize the TM
d-f
and RE, respectively, sublattices while Hex describes the exchange interactions
f
between the two. Within the single-ion model, HCEF
accounts for the interaction
of the aspherical 4f-charge with the crystalline electric field (CEF). Due to strong
spin-orbit interaction for the 4f-electrons, rotation of the magnetization leads to a
rotation of their aspherical charge cloud. This in turn results in a dependency of the
electrostatic energy on the orientation of the 4f-magnetic moment as described by
the Hamiltonian [168, 169, 170]
f
HCEF
=
n,m
n
Am
n θJ n r
4f
Onm
4f .
(72)
4 Electronic Structure: Metals and Insulators
229
Here Am
n are crystal field parameters for the angular momentum quantum numbers
n and m determined by the charge contribution in the system excluding the 4felectrons, θJ n are Stevens’ factors depending on the total angular momentum
quantum number J , Onm 4f are the expectation values of the Stevens’ operators, and
r n 4f are the expectation values of r n calculated for the 4f-states of the RE atom.
As the quantities θJ n and Onm 4f are all tabulated, calculation of the crystal field
f
n
parameters Am
n together with r 4f allows to fix HCEF and with this to determine
the corresponding phenomenological anisotropy constants Ki [168, 169].
While the first calculations in the field have been done adopting a spherical
approximation for the potential [171,172], later work clearly demonstrated the need
to use a nonspherical potential. Such calculations have been performed, for example,
by Richter et al. [173] on SmCo5 representing the itinerant s-, p-, and d-electrons via
band states, while the localized Sm 4f-states are treated within the atomic like socalled open shell scheme. Hummler and Fähnle report on corresponding calculations
on the CEF parameters for the whole RECo5 series with RE=Ce . . . Yb [170].
Their results for A02 r 2 4f and A04 r 4 4f are plotted in Fig. 19 in comparison with
experiment. This type of calculations on bulk materials led in general to satisfying
agreement with experiment and clearly showed that the naive point charge model
is completely inadequate for an estimate of the CEF parameters: point charges
chosen according to the chemical valency of the elements are much too high when
compared to ionic charges obtained from self-consistent calculations. In addition,
it turned out that the parameters Am
n are determined by about 80% by the charge
distribution on the RE site while the point charge model assumes a lattice of ionic
point charges surrounding the RE site.
Corresponding work has also been performed in order to investigate the magnetic
anisotropy at the surface or interface of RE-based compounds. Calculations of
0
0
-100
A4 < r >4f (K)
A2 < r >4f (K)
-10
-20
4
-300
0
0
2
-200
-400
-30
-40
-500
Ce
Pr
Nd
Er
Sm Gd Dy
Yb
Pm Eu
Tb Ho Tm
Ce
Pr
Nd
Er
Sm Gd Dy
Yb
Pm Eu
Tb Ho Tm
Fig. 19 Comparison of theoretical and experimental values for A02 r 2 4f (left) and A04 r 4 4f
(right) parameters for the series of RECo5 compounds with RE = Ce . . . Yb. The full circles (full
squares) are theoretical results for the experimental lattice parameters (for the lattice parameters
fixed to those of GdCo5 ). Experimental values are shown as open squares and crosses. (All data
taken from [174])
230
H. Ebert et al.
A02 r 2 have been done, for example, for the Nd sites of the (001)-surface of
Nd2 Fe14 B [175]. It turned out that the sign of A02 r 2 depends on the positions of
the Nd atoms in the unit cell supplying this way an explanation for the different
coercivity of crystalline and sintered Nd2 Fe14 B. Similar calculations have been
performed also to investigate the impact of Dy impurities on the coercivity of
Nd2 Fe14 B [176]. From these, it was found that the parameter A02 r 2 for Dy
atoms in the surface region of Nd2 Fe14 B also may have a positive or negative sign
depending on its position, leading finally to a decrease of the coercivity of sintered
samples.
P. Novák et al. introduced a scheme for the calculation of the crystal field
parameters that avoids the assumption of an inert 4f-charge cloud and allows for
the hybridization of the 4f-states with the surrounding electronic states [177].
The approach is based on a local Hamiltonian represented in the basis of Wannier
functions and expanded in a series of spherical tensor operators. Applications to RE
impurities in yttrium aluminate showed that the calculated crystal field decreases
continuously as the number of 4f-electrons increases and that the hybridization of
4f-states with the states of the oxygen ligands is important. This method has been
successfully applied also to calculate crystal field parameters for RE impurities in
LaF3 [178].
Dealing with ferrimagnetic materials composed of several, inequivalent magnetic sublattices, calculating the magnetic anisotropy may become more complicated as an additional canting between the sublattices introduced by an external
field may play a significant role and should be taken into account to get a
reasonable agreement with experiment. This was demonstrated for the RE-TM
ferrimagnet GdCo5 [179], where the authors report a first-principles magnetizationversus-field (FPMVB) approach giving temperature-dependent magnetization as
a function of an externally applied magnetic field in excellent agreement with
experiment.
Excitations
Many dynamical as well as finite-temperature properties of the magnetization of a
solid can be understood and described on the basis of magnetic excitations. In the
low-energy, small-wave vector regime, one has to deal with the collective magnon
excitations that can be investigated by various techniques. An approximate approach
builds on the use of calculated exchange coupling parameters in combination with
the so-called rigid spin approximation. More accurate results can be expected from
self-consistent spin-spiral or frozen-magnon calculations that also allow exploring
the magnetic phase space in an efficient way. Both approaches, however, do not give
access to single-particle or Stoner excitations. On the other hand, using the concept
of the dynamical susceptibility depending on frequency and wave vector, a coherent
description of magnon and Stoner excitations is achieved.
4 Electronic Structure: Metals and Insulators
231
Magnon Dispersion Relations Based on the Rigid Spin
Approximation
When considering the magnetization dynamics of solids, one usually assumes the
magnetization to be collinear inside an atomic cell i oriented along the common
direction m̂i implying a coherent rotation of the magnetization within the cell during
progress of time (rigid spin approximation (RSA)) [144]. As a consequence, the
equation of motion for the magnetization can be replaced by the equation of motion
for the local magnetic moments mi = mi m̂i that can be written as [144, 180, 181]:
2μB 1 ∂E
d
m̂i = −
× m̂i ,
h̄ mi ∂ m̂i
dt
(73)
where the right-hand side represents the torque acting on the magnetic moment mi .
Making use of the harmonic approximation for the energy, Eq. (73) yields for the
spin waves uλν (q) = uλν eiqR n with wave vector q the following eigenvalue problem
[181]:
h̄ ωλ (q) uλν =
2μB νν J (q) uλν ,
mν (74)
ν
for solids with translational symmetry. Here the eigenvectors uλν numbered by the
index λ represent small deviations of magnetic moments from the direction of the
ground state and the J νν (q) are the Fourier transforms of the interatomic exchange
coupling parameters with ν labeling the basis atoms within a unit cell.
Solution of the eigenvalue problem in Eq. (74) obviously yields the frequencies
ωλ (q) of the various collective spin-wave eigenmodes that can be compared with
magnon excitation energies as deduced, for example, from neutron scattering.
Corresponding results obtained for Fe and Ni are given in Fig. 20 in comparison
with experiment. Although good agreement between theory and experiment is
achieved, one has to stress that the theoretical results depend on the method used to
calculate the J νν (q) parameters. The data shown by a solid line were obtained using
the exchange coupling parameters Jij calculated on the basis of the Lichtenstein
formula Eq. (52) [182]. In the case of a lattice with one atom per unit cell, the
magnon energy spectra E(q) possess only a single branch. Therefore, the bccFe and fcc-Ni magnon spectra in Fig. 20 could be obtained by a simple Fourier
transformation [182]
E(q) =
4μB J0j (1 − eiq·R j ).
m
(75)
j
The minima of E(q) for bcc-Fe along the Γ − H and H − N directions to be
seen in Fig. 20 are so-called Kohn anomalies which occur due to long-range RKKY
interactions. It turned out that these minima appear only if the summation in Eq. (75)
232
H. Ebert et al.
600
Expt. 1
Expt. 2
Halilov et al.
Pajda et al.
Fe
400
300
200
300
200
Expt
Halilov et al.
Pajda et al.
100
100
0
Γ
Ni
400
E(q) (meV)
E(q) (meV)
500
500
N
Γ
P
H
0
L
N
Γ
X
W
K
Γ
Fig. 20 Magnon dispersion relations for bcc-Fe (left) and fcc-Ni (right) along high-symmetry
directions in the Brillouin zone, in comparison with experiment (open symbols [182]). Solid lines
represent the results by Pajda et al. [182], while full circles show the results by Halilov et al. [180]
700
Co
E(q) (meV)
E(q) (meV)
Expt
Theory
25
500
400
300
20
15
200
10
100
5
0
Γ
Gd
30
600
M
K
Γ
A
L
0
Γ
M
K
Γ
A
Fig. 21 Left: magnon dispersion relation for hcp-Co along high-symmetry lines in the Brillouin
zone [180]. Right: magnon dispersion relation for hcp-Gd (full lines) in comparison with
experimental data. (All data taken from [183])
is performed over a sufficiently large number of atom shells around the central
atomic site with index 0.
As an example for a lattice with a multiatom basis, Fig. 21 (left) displays the
magnon spectrum calculated for hcp-Co [180]. As there are two atomic sites in the
unit cell, solving the eigenvalue problem Eq. (74) leads to two magnon branches.
A similar approach was applied to hcp-Gd [183]. The corresponding experimental
data shown in Fig. 21 (right) were obtained at T = 78 K. This was accounted for
in the calculations within the RPA (see section “Methods Relying on the Rigid Spin
Approximation”) leading to a simple rescaling of the magnon energies proportional
to the temperature-dependent average magnetization.
4 Electronic Structure: Metals and Insulators
233
Spin Spiral Calculations
Usually calculations of the electronic structure for magnetic systems are performed
assuming a collinear spin-magnetic structure and using the smallest unit cell
corresponding to the space group of the system. However, this configuration does
not have to correspond to the ground state of the system. A possible way to
search for the proper magnetic ground state is to consider incommensurate spinspiral configurations. Furthermore, within the adiabatic approximation, spin spirals
can be seen as a representation of transverse spin fluctuations. Therefore, selfconsistent calculations on static spin spirals or so-called frozen magnons give
access to the energies of spin-wave excitations that can be used in particular to
investigate the finite-temperature magnetism. Considering a corresponding spin
spiral characterized by the wave vector q and the tilt angles θν and φν , the variation
of the spin-magnetic moment mnν from site to site may be expressed via:
mnν = mν [cos(q · R n + φν ) sin θν , sin(q · R n + φν ) sin θν , cos θν ],
(76)
where ν labels the atomic site in the unit cell located at lattice vector R n .
Because of broken translational and rotational symmetry, the presence of a spin
spiral in principle implies an increased unit cell compared to a collinear spin
configuration. However, as shown by Brinkman and Elliot [184, 185] as well as
Herring [186], one can make use of the fact that a spin-spiral structure characterized
by the wave vector q is invariant with respect to a so-called generalized translation:
Tn = {α(φ)|αR |t n },
(77)
if spin-orbit coupling is neglected. Here, the vector t n specifies a spatial translation
combined with a spatial rotation αR and a spin rotation about the ẑ axis by the angle
α(φ) = α(q · t n ). This property allows to formulate the generalized Bloch theorem
[187]:
Tn ψj k (r) = e−ik·t n ψj k (r),
(78)
that specifies the behavior under a generalized translation for the two-component
eigenfunctions ψj k of a Hamiltonian with a noncollinear spin-dependent potential
of the form (see also Eq. (11)):
V (r) =
nν
q†
Unν (θν , φν )
↑
Vnν (r)
0
↓
0 Vnν (r)
q
Unν (θν , φν ).
(79)
Here n specifies the Bravais lattice vector R n , ν gives the position ρ ν of an atom
q
in the unit cell, and Unν is a spin-transformation matrix that connects the global
frame of reference of the crystal to the local frame of the atom site at R n + ρ ν
that has its magnetic moment mnν tilted away from the global z-direction. The
234
H. Ebert et al.
x
ϕ=qR
m
θ
z
y
q
Fig. 22 Geometry of a spin spiral with the wave vector q along the z-direction
q
transformation Unν is characterized by the Euler angles θnν and φnν as it is shown
in Fig. 22. Assuming a collinear alignment of the spin density within the atomic cell
at (n, ν), it is natural to use a local frame of reference with its z-axis oriented along
q
mnν . The corresponding transformation matrices Unν occurring in Eq. (79) can be
q
written as a product of two independent rotation matrices Unν = Un (θν , φν , q) =
Uν (θν , φν ) UqR n , where the matrix UqR n depends only on the translation vector R n
[187]:
q
Unν
=
cos θ2ν sin θ2ν
− sin θ2ν cos θ2ν
i
0
e 2 φν
− 2i φν
0e
i
e 2 q·Rn
0
− 2i q·Rn
0e
.
(80)
Results for the wave vector-dependent energy and spin-magnetic moments per
atom obtained by Uhl et al. [188] from corresponding spin-spiral calculations
for γ -Fe are shown in Fig. 23. This work was based on the LSDA and used the
augmented spherical wave (ASW) band structure method in combination with the
atomic sphere approximation (ASA). In addition, noncollinearity within an atomic
cell was neglected. The investigations on γ -Fe by Kurz et al. [189], on the other
hand, avoided these simplifications by the use of the LAPW band structure method.
The noncollinear magnetic structure was imposed by a constraining magnetic field
applied to the magnetic moments of the atoms. Furthermore, the GGA was used
for the exchange-correlation potential. In spite of the various technical differences
between the two studies, the results shown in Fig. 23 agree fairly well and justify
the approach used by Uhl et al. as well as many others.
The implementation of the spin-spiral method within the KKR band structure
method allows dealing not only with ordered materials but also with random alloys
[190]. Figure 24 gives as an example the energy (left) and individual spin moments
(right) for a spin-spiral magnetic structure in Fe0.5 Mn0.5 as a function of the wave
vector q directed along the [001] direction. One can see a transition from the
antiparallel alignment of the Fe and Mn magnetic moments at small q to a parallel
alignment when q approaches the boundary of the Brillouin zone at |q| = 2π/a.
As it is seen from the left part of Fig. 24, the latter magnetic configuration is
energetically more stable.
Apart from exploring the magnetic phase space by performing self-consistent
spin-spiral calculations, the technique can also be used to get access to magnon
4 Electronic Structure: Metals and Insulators
235
0
2
Kurz et al.
Uhl et al.
1.5
-20
mspin (μB)
E(q) (meV)
-10
-30
1
-40
Kurz et al.
Uhl et al.
0.5
-50
-60
Γ
X
0
Γ
W
X
W
Fig. 23 The energies of the spin-spiral structure with respect to the energy of the FM state (left)
and spin-magnetic moments per atom (right) calculated for γ -Fe for the wave vector q varying
along Γ − X − W in the Brillouin zone. Open diamonds represent the results obtained by Uhl et al.
[188] while full circles represent the results obtained by Kurz et al. [189]. (All data taken from
[188] and [189])
3
0.2
MFe
MMn
Mtotal
2.5
Mn and Fe parallel
Mspin (μB)
Espin spiral (eV)
2
0.1
Mn and Fe
antiparallel
1.5
1
0.5
0
0
-0.5
0
0.2
0.4
qz
0.6
0.8
1
-1
0
0.2
0.4
qz
0.6
0.8
1
Fig. 24 Left: energy of a spin spiral in a Fe0.5 Mn0.5 alloy calculated for the wave vectors q =
2π
a (0, 0, qz ) along the [001] direction. Right: local magnetic moments on Fe and Mn atoms as a
function of the wave vector q. (All data taken from [190])
excitation energies h̄ω(q). In the case of simple lattices, the energy ΔE(q, θ ) of a
spin spiral with wave vector q and tilt angle θ can be used directly to get h̄ω(q)
from the expression [191]:
4 ΔE(q, θ )
.
θ→0 m sin2 θ
h̄ω(q) = lim
(81)
This approach was used, for example, by Halilov et al. [180] to calculate the magnon
energy spectra for Fe and Ni represented in Fig. 20. Obviously, the results are in a
reasonably good agreement with those of Pajda et al. [182] that are based on a
236
H. Ebert et al.
calculation of the real space exchange coupling parameters Jij via the Lichtenstein
formula Eq. (52). As one notes, the minima for Fe along the Γ − H and H − N
directions are given by both approaches. However, the magnons obtained by the
spin-spiral calculations are softer, because of the self-consistent relaxation within
the electronic structure calculations.
As another example, Fig. 25 represents spin-wave dispersion curves obtained for
the full-Heusler alloys Cu2 MnAl, Pd2 MnSn, Ni2 MnSn, and the L12 -type ferromagnet MnPt3 [192]. The simplified approach used in this work did not fully account for
the magnetic sublattices of the investigated systems providing for that reason only
the first magnon branch. Nevertheless, this already led to values for the Curie temperature calculated within the RPA approach (see section “Methods Relying on the
Rigid Spin Approximation”) in reasonable agreement with experiment. As a general
trend, one can see in Fig. 25 that the calculated magnon energies h̄ω(q) are too high
when compared with experiment. The authors attribute this to the treatment of electronic correlations on the basis of the GGA. In fact, previous work on the series of
Heusler alloys Co2 Mn1−x Fex Si [62] clearly demonstrated the impact of correlation
effects by comparison of results based on the LSDA, LSDA+U, and LSDA+DMFT.
200
Expt (4.2 K)
Theory
Cu2MnAl
60
Pd2MnSn
Expt (50 K)
Theory
E(q) (meV)
E(q) (meV)
150
100
20
50
0
Γ
80
[100]
[110]
Γ
X
Ni2MnSn
[111]
L
0
Γ
200
Theory
Expt (50 K)
[100]
[110]
[111]
Γ
X
MnPt3
L
Expt (80 K)
Theory
150
E(q) (meV)
E(q) (meV)
60
40
20
0
Γ
40
100
50
[100]
[111]
[110]
X
Γ
L
0
Γ
[100]
[110]
X M
[111]
Γ
R
Fig. 25 Calculated (solid lines) spin-wave dispersion curves h̄ω(q) in the first Brillouin zone
along high-symmetry directions for the L21 -type full-Heusler and L12 -type ferromagnets. As indicated, the experimental data stem from neutron diffraction measurements at various temperatures.
(All data taken from [192])
4 Electronic Structure: Metals and Insulators
237
The use of the spin-spiral technique for the calculation of the full magnon energy
spectrum in case of complex compounds was demonstrated by Şaşıoğlu et al. [193].
In this case, a set of spin-spiral calculations is required to obtain all exchange
coupling parameters J νν (q) that enter the eigenvalue problem Eq. (74).
Excitation Spectra Based on the Dynamical Susceptibility
Despite the many successful applications of the adiabatic approach for the investigation of spin-wave excitations, one has to stress that it has severe limitations. In
particular it can be applied only to systems for which single-particle or the so-called
Stoner excitations can be neglected [194, 195]. Figure 26 gives a simplified picture
of the exchange-split band structure of an itinerant ferromagnet in the vicinity of its
Fermi level. Excitation of an electron from an occupied majority-spin state below
the Fermi level to an empty minority state may not only be associated with a spin
flip but also with a change of the electronic wave vector. As it is visualized in the
right panel of Fig. 26, these Stoner excitations lead to a broad continuum that in
general overlaps and hybridizes with the discrete magnon dispersion spectrum. For
this and other reasons, a more sophisticated description of spin-wave excitations was
worked out making use of the linear response formalism within the framework of
time-dependent density functional theory [196,197,87,198]. This allows expressing
the magnetization Δmi (r, q, ω) induced by a magnetic field B(r, q, ω), with its
time and spatial dependency expressed by the wave vector q and frequency ω,
respectively, in terms of a corresponding susceptibility tensor [2]:
Δmi (r, q, ω) =
j
(82)
Ω
Minority-spin
δ
ΔE = ε
U
Majority-spin
Electron wave vector, k
EF
Excitation energy, ε
Electron energy, E
Δk = q
d 3 r χ ij (r, r , q, ω) B j (r , q, ω).
Stoner continuum
U
Magnons
0
δ
Excitation wave vector, q
Fig. 26 Left: schematic representation of Stoner excitations in an itinerant ferromagnet. A
majority electron is excited from an occupied state below the Fermi level to an unoccupied minority
state above the Fermi level. Right: continuum of Stoner excitations for a metallic ferromagnet
238
H. Ebert et al.
Within spin-density functional theory and making use of circular coordinates, one
may write in particular for the transverse susceptibility tensor element χ ± a Dysonlike equation [87]:
χ ± (r, r , q, ω) = χ0± (r, r , q, ω)
d 3 r χ0± (r, r , q, ω) Kxc (r ) χ ± (r , r , q, ω), (83)
+
Ω
xc (r)
where χ0± is the unenhanced susceptibility, Kxc (r) = Bm(r)
is the exchangecorrelation kernel function, and Bxc (r) and m(r) are the local exchange-correlation
field and magnetization, respectively. As discussed, for example, by Bruno [199] as
well as by Katsnelson and Lichtenstein [200], it is the second term in Eq. (83) that
gives rise to the enhancement of the transverse susceptibility.
The dynamical susceptibility gives not only access to the energetics of magnetic
excitations but also to their lifetime characterizing this way the dissipation of the
energy. Outside the Stoner continuum, the loss tensor associated with χ + shows
peaks at frequencies corresponding to the excitation of spin waves. Inside the Stoner
continuum, these show a finite width due to the hybridization with the Stoner
excitations. In this case, one has approximately:
χλ+ (q, ω) ≈
Aλ (q)
,
(ω − ω0λ (q))2 + βλ (q)2
(84)
with the amplitude Aλ (q), the spin-wave energy ω0λ (q), and inverse lifetime βλ (q).
As an example, Fig. 27 shows the magnon dispersion curves ω0 (q) together with
the corresponding broadening for the magnon states as deduced from the dynamical
susceptibility as calculated for bcc-Fe and fcc-Co [203]. As one can clearly see
from the given width, Stoner excitations have only a very small influence on the
long-period magnons.
Results for ω0λ (q) and βλ (q) for the Heusler alloy Co2 MnSi that has three
magnetic sublattices are shown in Fig. 28. The corresponding eigenvectors (EV)
Fig. 27 Spin waves of bcc-Fe (left) and fcc-Co (right). Solid circles correspond to ω0 (q), while
the error bars denote full width at half maximum of the peak. Solid line denote spin-wave energies
obtained using the magnetic force theorem [203]
4 Electronic Structure: Metals and Insulators
600
EV 1
EV 2
EV 3
400
200
0
Γ
EV 2
EV 3
150
β(q) (meV)
ω(q) (meV)
800
239
100
50
[ξ00] X
K
[ξξ0]
Γ [ξξξ] L
Γ
[ξ00] X
K
[ξξ0]
Γ [ξξξ] L
Fig. 28 Energies ω0λ (left) and inverse lifetimes βλ (middle) of three spin-wave modes in
Co2 MnSi together with the corresponding eigenvectors (right); arrows indicate the orientations
of the magnetic moments. The basis atoms are Co at (1/4, 1/4, 1/4)a and (3/4, 3/4, 3/4)a and Mn
at (1/2, 1/2, 1/2)a. The parameter β1 of EV 1 does not exceed 5 meV and is not shown. (All data
taken from [87])
of the resulting spin-wave modes are given on the right-hand side of the figure. The acoustic mode that is lowest in energy has a vanishing energy for
q = 0, and its value for βλ is very small (therefore not shown in Fig. 28).
The optical modes, on the other hand, appear at higher energies where the
continuum density is appreciable. Accordingly, their inverse lifetime βλ is quite
large and depends strongly and non-monotonously on the wave vector q. For
the Heusler alloy Cu2 MnAl, only one magnetic sublattice has to be considered, and accordingly, there is only one acoustic spin-wave mode. In contrast
to Co2 MnSi, a more pronounced damping is found in this case. The influence
of the Stoner excitations can also be seen for the spin-wave energies. Calculating these by use of the so-called adiabatic approach [202] that neglects the
hybridization, the spin-wave energies are higher and in less good agreement with
experiment [202].
Comparable studies based on the dynamical susceptibility were done (i) to
investigate the Landau damping in Fe(100) and Fe(110) films and the effect of
the substrate on this [201], (ii) to study the Landau damping of spin waves
and large Rh moments induced by the AFM magnons in FeRh [203], and
(iii) on acoustic magnons in the long-wavelength limit in order to analyze the
Goldstone violation in many-body perturbation theory [204]. The concept of
the dynamic spin susceptibility has been applied also to paramagnetic systems
at finite temperatures by Staunton et al. [205, 206]. Due to the use of the
multiple scattering formalism, investigations on alloys could be made, for example, on paramagnetic Cr0.95 V0.05 and antiferromagnetic Cr0.95 Re0.05 above the
Néel temperature TN . While the work sketched here was primarily based on
the linear response formalism applied within the framework of time-dependent
density functional theory, similar work on quasiparticle and collective electronic
excitations in solids was done using techniques from many-body perturbation
theory [207].
240
H. Ebert et al.
Finite-Temperature Magnetism
Dealing with the impact of finite temperatures in a quantitative way is a big
challenge for theory. Accordingly, many different techniques on various levels of
sophistication are in use for that purpose. Most of these employ the adiabatic
approximation that decouples the electronic and magnetic degrees of freedom.
One type of such approaches, that proved to be astonishingly successful for
many situations, starts from the properties of low-energy magnetic excitations by
calculating real-space exchange coupling parameters or the energies of spin spirals.
In a second step, this information is used in combination with classical statistical
methods including in particular the Monte Carlo method to deduce temperaturedependent magnetic properties. More advanced schemes, however, are based on a
coherent description of the electronic structure and statistics. While the disordered
local moment (DLM) method still relies on the adiabatic approximation, this does
not apply, for example, to the functional-integral method or various many-body
approaches used within the dynamical mean field theory (DMFT) that account for
finite temperature in a coherent way.
Methods Relying on the Rigid Spin Approximation
Within standard Stoner theory, a spin-dependent but collinear electronic structure is
assumed, and finite temperatures are accounted for only via the Fermi distribution
function. Accordingly, the resulting critical temperatures are much too high. More
successful approaches to deal with magnetism at finite temperatures, on the other
hand, allow for transverse spin excitations. A simple model accounting for this was
suggested already by P. Weiss who considered a magnet as a system of localized
magnetic moments that order spontaneously due to an effective molecular or Weiss
field hW (T ) = w m(T ) n̂ that depends on the average magnetic moment m(T ) on
an atomic site. The factor w is the molecular or Weiss field constant:
w=
3kB TC
,
m20
(85)
which is determined by the magnetic moment m0 = m(0) at T = 0 K and the critical
temperature TC . Quite general, the temperature-dependent magnetic moment m(T )
along n̂ is determined by the statistical average over all possible orientations ê with
the probability distribution for the local magnetic moments given by:
P n̂ (ê) = e−β hW n̂·ê
d ê e−β hW n̂·ê
,
(86)
where hW = w m(T ) and β = 1/(kB T ). The various techniques discussed in
section “Exchange Coupling Parameters” allow to deduce the Weiss or mean field
4 Electronic Structure: Metals and Insulators
241
constant from the calculated exchange coupling parameters Jij . Within this mean
field approach (MFA), application of classical spin statistics leads to:
TCMFA =
2 2
J0j =
J0 ,
3kB
3kB
(87)
j =0
MFA in case of an elemental ferromagnet, with J =
for
0
the Curie temperature TC
j =0 J0j [126].
A more accurate approach to deal with finite-temperature magnetism is provided
by the random-phase approximation Green function (RPA-GF) method which also
can be based on a combination of the Heisenberg model and SDFT calculations
(see section “Exchange Coupling Parameters”). The decoupling scheme suggested
by Tyablikov leads to an approximate expression for the one-particle Green magnon
function [208, 209]:
Gm (z) =
1
1 ,
N q z − E(q)
(88)
with E(q) the magnon energies that allows expressing the critical, i.e., Curie or
Néel, respectively, temperature TcRPA as [182]:
6
1
= − lim
RPA
z→0 m
k B Tc
Gm (z),
(89)
in case of a ferro- or antiferromagnet, respectively. The MFA approach accounts
for all spin-wave excitations with the same weight leading in general to an
overestimation of the Curie temperature. The RPA-GF approach, on the other hand,
accounts in particular for the low-energy excitations in a much more adequate
manner. As a consequence, more accurate results for the critical temperature are
normally obtained this way when compared with experiment. This behavior has
been found within numerical work on the elemental ferromagnets Fe, Co, and Ni
[182] but also for other systems, as, for example, the Heusler alloys (Ni,Cu)2 MnSn,
(Ni,Pd)2 MnSn [210], NiMnSb, CoMnSb, Co2 MnSi, and Co2 CrAl [193], for Gdbased intermetallic compounds GdX (X = Mg, Rh, Ni, Pd) [211] as well as for
zincblende half-metallic ferromagnets GaX X = N, P, As, Sb [212]. The exchange
coupling parameters in these works have been obtained either from spin-spiral
calculations [193,212] or in real space, on the basis of the force theorem [182,211].
In all cases, reasonable agreement with experiment could be achieved, except for
fcc-Ni, a system for which the application of a Heisenberg Hamiltonian with fixed
spin moments seems questionable.
Corresponding investigations on the spin-wave spectra and Curie temperatures
have also been performed for L21 -type full-Heusler FM alloys and L12 -type XPt3
alloys [192]. It was found that the Curie temperatures are in good agreement with
experiment when the Stoner gap was large enough so that the magnon regime is
well separated from the Stoner excitations. In this case, single-particle spin-flip
242
H. Ebert et al.
Stoner excitations make only a small contribution to the excitation spectra at
low energies, so that magnon excitations make the dominating contribution to
thermodynamics. It is interesting to note that the RPA-GF formalism fulfills the
Mermin-Wagner theorem [213], i.e., in the absence of magnetic anisotropy it leads
for two-dimensional systems to TC = 0 K [214]. Corresponding studies by Pajda
et al. on the Curie temperature in Fe and Co films that included a finite magnetic
anisotropy led to an oscillating behavior as a function of film thickness [127]. These
oscillations could be ascribed to the oscillating behavior of the exchange coupling
parameters due to thickness-dependent quantum well states. An extension of the
RPA-GF method (RPA-CPA) to deal within the CPA also with disordered alloys
was worked out by Bouzerar and Bruno [215].
Starting from the Stratonovich-Hubbard functional integral method [83], Kübler
[191] derived an expression for the critical, i.e., Curie or Néel, respectively,
temperature:
⎡
⎤−1
2 2⎣1 1 ⎦
k B Tc =
L
(90)
3 ν ν N q,n jn (q)
that is applicable to systems with several atoms per unit cell. Here Lν is the socalled local moment of atom ν that should be found in a self-consistent way [191].
However, in the case of well-localized magnetic moments, values of mν at T = 0
K can be used as a reasonable approximation [216]. Finally, jn (q) in Eq. (90)
is the exchange function after diagonalization of jνν (q) [191]. This expression
accounts in particular for the fact that different atomic types have in general different
moments, while the expressions in Eqs. (88) and (89) that are often applied to
multicomponent systems are based on the assumption that Lν = m0 with m0 , the
average saturation moment. As is demonstrated by the results in Table 2, Eq. (90)
gives in general results for the critical temperature that are in good agreement with
experiment.
An important alternative to the RPA-GF scheme for dealing with magnetic
properties at finite temperatures is provided by the Monte Carlo method that is
used to deal with the statistical aspect of the problem. Corresponding work again
is in general based on the Heisenberg Hamiltonian (see Eq. (51)) with the exchange
parameters calculated in an ab initio way. Numerous successful applications have
been done, both for ordered compounds [218] and for disordered alloys such as
diluted magnetic semiconductors like Ga1−x Mnx As [135, 219, 220] or Heusler
alloys [129, 221]. In most cases, good agreement with RPA-based results as well as
with corresponding experimental data for the critical temperature could be achieved.
Methods Accounting for Longitudinal Spin Fluctuations
Despite the good results often obtained via the RPA and MC methods, one has
to stress that they are based on a classical spin Hamiltonian and therefore are
4 Electronic Structure: Metals and Insulators
243
Table 2 Structure, magnetic moments on M1 and M2 sublattices, as well as critical Curie or Néel
temperature calculated via the MC (TcMC ), RPA (TcRPA ), or MFA (TcMFA ) approaches in comparison
with experiment. (All data taken from [191] and [217])
System
Fe
Co
Ni
FeNi
CoNi
FeNi3
CoNi3
NiMnSb
Mn2 VAl
Co2 FeSi
Mn3 Al
Mn3 Ga
Mn3 Ga
RhMn3
Structure
bcc
fcc
fcc
CuAu
CuAu
AuCu3
AuCu3
C1b
L21
L21
L21
L21
DO22
AuCu3
M1 (μB )
2.330
1.410
0.630
2.551
1.643
2.822
1.640
3.697
−0.769
2.698
−2.258
−2.744
−2.829
3.066
M2 (μB )
–
–
–
0.600
0.673
0.588
0.629
0.303
1.374
1.149
1.128
1.363
2.273
–
TcMC/RPA (K)
1060MC
1080MC
510MC
972RPA
1149RPA
986RPA
733RPA
968RPA
580RPA
1058RPA
196RPA
314RPA
762RPA
1059RPA
TcMFA (K)
1460
1770
660
1130
1538
1290
925
1281
663
1267
342
482
1176
–
Exp
Tc (K)
1043
1388
633
790
1140
870
920
730
760
1100
–
–
–
855
suitable only for systems with their magnetic moments depending only weakly on
the temperature. As the latter assumption is not always fulfilled, an extension of
the Heisenberg Hamiltonian was suggested that is meant to account for longitudinal
fluctuations [222, 217] that express the total energy by an expansion in even powers
of the magnetic moments per atom m:
E(M, q, θ ) =
n
An m2n +
Jn (q, θ ) m2n .
(91)
n
Here the functions Jn (q, θ ) are proportional to the energy difference between the
ferromagnetic and the spin-spiral states specified by wave vector q and tilt angle θ .
Monte Carlo simulations for bcc-Fe, fcc-Co and fcc-Ni performed on this basis
[217] led to a rather good agreement with experiment as one can see from Table 2.
Another model to account for temperature-induced longitudinal spin fluctuations
was suggested by Ruban et al. [223] that also led for Fe and Ni to a rather realistic
description of the magnetic properties at finite temperatures.
An important class of materials, for which longitudinal spin fluctuations are
of great importance, are alloys and compounds composed of magnetic and otherwise nonmagnetic elements. Such systems exhibit so-called covalent magnetism
[224, 225], i.e., the magnetization of the nonmagnetic component is caused by
the spontaneously magnetized atoms via a spin-dependent hybridization of the
electronic states. Ležaić et al. [129, 221], for example, emphasized the need
to account for longitudinal fluctuations of the magnetic moment induced on the
Ni atoms for a proper description of the temperature dependence of the spin
polarization at the Fermi energy EF when performing Monte Carlo investigations
244
H. Ebert et al.
on the half-metallic Heusler alloy NiMnSb. For that purpose, they used an extended
Heisenberg Hamiltonian:
Hext =
1
Jij mi · mi +
(a m2i + b m4i ),
2
ij
(92)
i∈N i
that allows accounting for transverse and longitudinal magnetic fluctuations connected with the temperature-dependent induced magnetic moment on the Ni atom.
The Heusler alloy NiMnSb was also investigated by Sandratskii et al. [212] using
the spin-spiral approach and treating the magnetic moment of Ni atom as being
induced.
Mryasov et al. [226] found that the induced magnetic moment of Pt plays a
crucial role for the magnetic anisotropy of FePt at finite temperatures. To account
for this, a renomalization of the Fe-Fe exchange interactions according to:
J˜ij = Jij + I
Pt
χPt
m0Pt
2
Jiν Jνj
(93)
ν∈Pt
was suggested. Here I Pt characterizes the local exchange interaction of the Pt
atoms, m0Pt is the Pt magnetic moment in the ordered ferromagnetic state, and
χPt is the partial spin susceptibility of Pt. The impact of a renormalization of
the Fe-Fe interactions due to the induced Rh magnetic moment has also been
demonstrated for the stabilization of the ferromagnetic state and for the control of
the antiferromagnet-ferromagnet phase transition of FeRh [227].
Another scheme to account within Monte Carlo simulations for the impact of
the induced magnetic moments on nonmagnetic alloy components leading to a
renomalization of the exchange interactions between the magnetic components
was worked out by Polesya et al. [128]. Figure 29 (left) shows corresponding
results for the Curie temperatures of the ferromagnetic alloy Fex Pd1−x that are in
very good agreement even for the Pd-rich side of the alloy system. Corresponding
calculations have been done for FeRh to investigate its antiferromagnet-ferromagnet
phase transition [229]. The right panel of Fig. 29 shows results of Kudrnovský
et al. [228] for fcc-Ni1−x Pdx . Investigating the finite-temperature magnetism of
various Ni-based transition metal alloys, these authors concluded that Bruno’s
formulation [199] for the renormalized RPA gives the most satisfying agreement
with experiment.
Coherent Treatment of Electronic Structure and Spin Statistics
The methods sketched above that are based on the Heisenberg model obviously
have problems to account for longitudinal spin fluctuations. In addition, they consist
in an incoherent combination of electronic structure calculations and classical
4 Electronic Structure: Metals and Insulators
T (K)
400
Theory
Expt. 1
Expt. 2
Expt. 3
Expt. 4
RPA
MFA
rRPA
Expt.
600
T (K)
500
245
300
400
200
200
100
0
0
0.05
0.1
xFe
0.15
0.2
0
0
0.2
0.4
0.6
0.8
1
xNi
Fig. 29 Calculated Curie temperature TC of disordered alloys for fcc-Fex Pd1−x (left) using the
Monte Carlo method and fcc-Nix Pd1−x (right) using the MFA, RPA, and renormalized RPA (rRPA)
approaches, in comparison with experiment. (All data taken from [128] and [228])
statistics. These problems can be avoided by the use of the disordered local moment
(DLM) method [230] that deals with the temperature-dependent magnetization
within self-consistent electronic structure calculations. The slow dynamics of the
magnetic moments spatially localized on the atoms when compared to the fast
electron propagation and relaxation time scale allows to make use of the adiabatic
approximation. Assuming ergodicity for the system of local magnetic moments, the
time average required to calculate the average magnetization at a given temperature
can be replaced by the average over the ensemble of all orientational configurations
characterized by a set of unit vectors {êi }. Within the DLM theory, this is determined
by the corresponding single-site probabilities P n̂ (êi ) obtained from Eq. (86) for
the Weiss field hW = hn̂ . Following Györffy et al. [230], hn̂ is determined by
approximating the free energy F corresponding to the microscopic Hamiltonian
H of the
considered system by the free energy F0 based on the trial Hamiltonian
H0n̂ = j hn̂ · êj with hn̂ = hn̂ n̂ and using the Feynman-Peierls inequality [231]:
F ≤ F0 + H − H 0 ,
(94)
where the canonical distribution Eq. (86) is used to calculate the average. Using the
Weiss field hn̂ as a variational parameter to minimize the right-hand side of Eq. (94)
one is led to [162]:
3
h =
4π
n̂
d êi êi Hn̂
êi ,
(95)
where Hn̂ êi denotes the average of Hn̂ with the restriction that the orientation
of the moment on site i is fixed to êi [230]. Self-consistent electronic structure
calculations for a given temperature result in a temperature-dependent magnetic
moment that automatically accounts for longitudinal fluctuations. The resulting
246
H. Ebert et al.
magnetization M(T ) = M(T )n̂ can be obtained from
M(T ) = mloc (T )
d êi P n̂ (êi ) n̂ · êi ,
(96)
with the local moment mloc to be determined self-consistently.
Based on a relativistic formulation of the DLM method (RDLM), Staunton et al.
[162] worked out a scheme to investigate the magnetocrystalline anisotropy at finite
temperature. This implies in particular a corresponding extension of Eq. (68) that
allows calculating the magnetic torque together with the Weiss field. Results for
the temperature-dependent magnetization (M(T )) and uniaxial magnetocrystalline
anisotropy (K(T )) that have been obtained by an application of this approach
to L10 -ordered FePt are shown in Fig. 30 [232]. In line with experiment, the
orientation of the easy axis was found to be perpendicular to the Fe- and Pt-layers for
all temperatures. Also the temperature dependence of the anisotropy constant K(T )
is in good agreement with experiment. The single-ion anisotropy model, on the other
hand, fails to give a correct qualitative description for K(T ). Similar RDLM-based
investigations have been performed also for the L10 -ordered FePd [162]. As in
the case of FePt, the easy axis in FePd is perpendicular to the Fe- and Pd-layers,
with the uniaxial magnetocrystalline anisotropy also showing the scaling behavior
K(T ) ∼ [M(T )/M(0)]2 .
Buruzs et al. [233] used the RDLM method to investigate the temperaturedependent magnetic properties of Co films deposited on Cu(100) (Con /Cu(100))
with the thickness n of the Co film varying from 1 to 6. The resulting Curie
temperatures are given in the Table 3. In agreement with experiment, it was
found that the magnetization is oriented parallel to the surface for almost all
temperatures below the Curie temperature, except for the system with n = 2.
Based on the calculation of the anisotropy constant, a temperature-induced reversal
ΔESOC (meV)
M(T)
0.8
-0.5
0.6
0.4
-1.5
0.2
0
-1
-2
200
400
600
800
Temperature T (K)
0.2
0.4
0.6
0.8
2
(M(T))
Fig. 30 RDLM calculations on FePt. Left: the magnetization M(T ) versus T for the magnetization along the easy [001] axis (filled squares). The full line shows the mean field approximation
to a classical Heisenberg model for comparison. Right: the magnetic anisotropy energy ΔESOC
as a function of the square of the magnetization M(T ). The filled circles show the RDLM-based
results, the full line give K(T ) ∼ [M(T )/M(0)]2 , and the dashed line is based on the single-ion
model function. (All data taken from [232])
4 Electronic Structure: Metals and Insulators
Table 3 Calculated Curie
temperatures (K) for
Con /Cu(100) [233]
247
n
TC
1
1330
2
933
3
897
4
960
5
945
6
960
of the anisotropy direction from in-plane to out-of-plane was predicted. A more
detailed investigation led to the conclusion that the spin reorientation is driven by a
competition of exchange and single-site anisotropies [18].
Zhuravlev et al. [234] used the RDLM method to investigate the origin of
the anomalous temperature dependence of the magnetocrystalline anisotropy in
(Fe1−x Cox )2 B alloys. In contrast to a conventional monotonous variation of the
MCA energy with increasing temperature, the anisotropy in (Fe1−x Cox )2 B shows a
non-monotonous temperature dependence due to increasing magnetic disorder. This
behavior of the experimental data was quantitatively reproduced by the RDLMbased calculations. It turned out that the observed temperature dependence of the
anisotropy is caused by a modification of the electronic structure induced by spin
fluctuations which can result in a selective suppression of spin-orbit-induced hot
spots (see section “Relativistic Effects”). In contrast to the expectations based on
other models, this peculiar electronic mechanism may lead to an increase, rather
than decrease, of the anisotropy with decreasing magnetization.
Another approach to account for temperature-induced charge and spin fluctuations when dealing with the magnetic properties of itinerant-electron magnets
at finite temperature is based on the functional integral method [4]. Within
this approach, the corresponding auxiliary exchange field is introduced using
the Hubbard-Stratonovich transformation [235]. This allows to describe the spin
fluctuation contribution to the free energy in a rather simple way. Based on the
functional-integral method, Kakehashi has proposed to use the dynamical coherentpotential approximation (CPA) theory to go beyond the adiabatic theories in
metallic magnetism [236], which was first formulated for a model Hamiltonian.
In order to apply this approach to realistic systems, the dynamical CPA theory
has been extended by a combination with the LMTO band structure method
(see section “Band Structure Methods”) [237]. Using this approach, the hightemperature susceptibility is expected to follow the Curie-Weiss law:
χCW (T ) =
m2eff
.
3(T − TC )
(97)
Using the atomic exchange coupling parameter J , the values J = 0.9, 0.94, and
0.9 eV for Fe, Co, and Ni and the Curie temperatures TC = 1930, 2250, and 620 K,
respectively, are found. The corresponding effective magnetic moments meff = 3.0,
expt
3.0, and 1.93 μB are in reasonable agreement with experiment (meff = 3.2 , 3.15,
and 1.6 μB , respectively)
The importance of correlation effects for finite-temperature magnetism has been
investigated within the framework of the dynamical mean field theory (DMFT) (see
section “Spin Density Functional Theory”) [16,32]. As demonstrated by Kakehashi
248
H. Ebert et al.
1
2
χ meff 3TC
M(T)/M(0)
Fe
0
0
0.5
-1
Ni
1
T/TC
1.5
2
0
1000
TCW (K)
1
500
0.2
CPA+DMFT
Peschard 1925
Chechernikov 1962
0.4
0.6
xNi
0.8
1
Fig. 31 Left: temperature dependence of the magnetization M(T )/M(0) and the inverse ferromagnetic susceptibility for Fe (open squares) and Ni (open circles) compared with experimental
results for Fe (squares) and Ni (circles). Right: CPA+DMFT-based and experimental values for
the Curie-Weiss temperature of Fe1−x Nix alloys as a function of Ni concentration. (All data taken
from [16, 32] (left) and [239] (right))
[238], concerning the treatment of finite temperatures within electronic structure
calculations, this approach is essentially equivalent to the dynamical CPA theory
mentioned above. Figure 31 (left) shows results for the calculated temperature
dependence of the magnetic moment and the inverse ferromagnetic susceptibility of
Fe and Ni. The effective magnetic moments are found to be meff = 3.09 μB for Fe
and meff = 1.5 μB for Ni. The corresponding estimated Curie temperatures are 1900
and 700 K for Fe and Ni, respectively. A combination of the DMFT with the CPA
alloy theory that treats substitutional disorder and electronic correlations on equal
footing has been used by Poteryaev et al. [239] to investigate the magnetic properties
of Fe1−x Nix alloys. The calculated Curie temperatures shown in Fig. 31 (right) are
obviously in good agreement with experiment. Also in line with experiment, a linear
variation with temperature has been found for the calculated inverse magnetic susceptibilities at high temperatures. Recently, Patrick and Staunton have put forward a
computational scheme for the description of finite-temperature magnetic properties
of RE-TM compounds [240]. The correlation of the 4f -electrons of the RE atoms
is treated by applying the self-interaction correction (SIC) method, and the RDLM
approach is used to describe the magnetic disorder in the system. This theory was
successfully applied to the calculation of magnetic moments as a function of temperature as well as the Curie temperatures of the rare-earth cobalt (RECo5 ) family for
RE = Y-Lu, demonstrating rather good agreement with experiment. Based on these
calculations, the authors proposed a mechanism responsible for the strengthening of
Re-TM as well as TM-TM interactions in the light ReCo5 compounds, where the
RE variation exhibits a strong impact on the Co-Co interactions.
Acknowledgments Financial support by the DFG through the SFB 689 and 1277 is gratefully
acknowledged.
4 Electronic Structure: Metals and Insulators
249
References
1. Mattis, D.C.: The Theory of Magnetism I, Statics and Dynamics. Springer, Berlin (1981)
2. White, R.M.: Quantum Theory of Magnetism. Springer, Berlin (2007)
3. Kübler, J.: Theory of Itinerant Electron Magnetism. International Series of Monographs on
Physics. OUP, Oxford (2009)
4. Kakehashi, Y.: Modern Theory of Magnetism in Metals and Alloys. Springer, Berlin (2012)
5. Rose, M.E.: Relativistic Electron Theory. Wiley, New York (1961)
6. Blügel, S.: Magnetische Anisotropie und Magnetostriktion (Theorie). In: 30. Ferienkurs des
Instituts für Festkörperforschung 1999 “Magnetische Schichtsysteme”, editor: Institut für
Festkörperforschung, C1.1, Forschungszentrum Jülich GmbH, Jülich (1999)
7. Chikazumi, S.: Physics of Ferromagnetism. Oxford University Press, Oxford (2009)
8. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A.H., Ong, N.P.: Anomalous Hall effect. Rev.
Mod. Phys. 82, 1539 (2010)
9. Ebert, H.: Magneto-optical effects in transition metal systems. Rep. Prog. Phys. 59, 1665
(1996)
10. Sinova, J., Valenzuela, S.O., Wunderlich, J., Back, C.H., Jungwirth, T.: Spin Hall effects. Rev.
Mod. Phys. 87, 1213 (2015)
11. Garello, K., Miron, I., Avci, C., Freimuth, F., Mokrousov, Y., Blügel, S., Auffret, S.,
Boulle, O., Gaudin, G., Gambardella, P.: Symmetry and magnitude of spin-orbit torques in
ferromagnetic heterostructures. Nat. Nanotechnol. 8, 587 (2013)
12. Heinze, S., von Bergmann, K., Menzel, M., Brede, J., Kubetzka, A., Wiesendanger, R.,
Bihlmayer, G., Blügel, S.: Spontaneous atomic-scale magnetic skyrmion lattice in two
dimensions. Nat. Phys. 7, 713 (2011)
13. Bethe, H., Salpeter, E.: Quantum Mechanics of One- and Two-Electron Atoms. Springer, New
York (1957)
14. Jansen, H.J.F.: Magnetic anisotropy in density-functional theory. Phys. Rev. B 38, 8022
(1988)
15. Engel, E., Dreizler, R.M.: Density Functional Theory – An Advanced Course. Springer, Berlin
(2011)
16. Lichtenstein, A.I., Katsnelson, M.I., Kotliar, G.: Finite-temperature magnetism of transition metals: an ab initio dynamical mean-field theory. Phys. Rev. Lett. 87, 067205
(2001)
17. Shi, J., Vignale, G., Xiao, D., Niu, Q.: Quantum Theory of Orbital Magnetization and Its
Generalization to Interacting Systems. Phys. Rev. Lett. 99, 197202 (2007)
18. Udvardi, L., Szunyogh, L., Palotás, K., Weinberger, P.: First-principles relativistic study of
spin waves in thin magnetic films. Phys. Rev. B 68, 104436 (2003)
19. Brataas, A., Tserkovnyak, Y., Bauer, G.E.W.: Scattering theory of Gilbert damping. Phys.
Rev. Lett. 101, 037207 (2008)
20. Ashcroft, N., Mermin, N.: Solid State Physics. Saunders College Publishers, New York (1976)
21. Hohenberg, P., Kohn, W.: Inhomogenous electron gas. Phys. Rev. 136, B 864 (1964)
22. Sham, L.J., Kohn, W.: One-particle properties of an inhomogeneous interacting electron gas.
Phys. Rev. 145, 561 (1966)
23. von Barth, U., Hedin, L.: A local exchange-correlation potential for the spin polarized case.
I. J. Phys. C: Solid State Phys. 5, 1629 (1972)
24. Rajagopal, A.K., Callaway, J.: Inhomogeneous electron gas. Phys. Rev. B 7, 1912 (1973)
25. Ceperley, D.M., Alder, B.J.: Ground state of the electron gas by a stochastic method. Phys.
Rev. Lett. 45, 566 (1980)
26. Ebert, H., et al.: The Munich SPR-KKR package, version 7.7, https://www.ebert.cup.unimuenchen.de/en/software-en/13-sprkkr (2017)
27. Leung, T.C., Chan, C.T., Harmon, B.N.: Ground-state properties of Fe, Co, Ni, and their
monoxides: results of the generalized gradient approximation. Phys. Rev. B 44, 2923 (1991)
28. Aryasetiawan, F., Gunnarsson, O.: The GW method. Rep. Prog. Phys. 61, 237
(1998)
250
H. Ebert et al.
29. Aryasetiawan, F.: Self-energy of ferromagnetic nickel in the GW approximation. Phys. Rev.
B 46, 13051 (1992)
30. Liebsch, A.: Effect of self-energy corrections on the valence-band photoemission spectra of
Ni. Phys. Rev. Lett. 43, 1431 (1979)
31. Anisimov, V.I., Zaanen, J., Andersen, O.K.: Band theory and Mott insulators: Hubbard U
instead of Stoner I. Phys. Rev. B 44, 943 (1991)
32. Kotliar, G., Savrasov, S.Y., Haule, K., Oudovenko, V.S., Parcollet, O., Marianetti, C.A.:
Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 78, 865
(2006)
33. Held, K., Nekrasov, I.A., Keller, G., Eyert, V., Blümer, N., McMahan, A.K., Scalettar, R.T.,
Pruschke, T., Anisimov, V.I., Vollhardt, D.: Realistic investigations of correlated electron
systems with LDA + DMFT. Phys. Stat. Sol. (B) 243, 2599 (2006)
34. http://psi-k.net/software/
35. Blöchl, P.E.: Projector augmented-wave method. Phys. Rev. B 50, 17953 (1994)
36. Andersen, O.K.: Linear methods in band theory. Phys. Rev. B 12, 3060 (1975)
37. Singh, D.: Plane Waves, Pseudopotentials and the LAPW Method. Kluwer Academic,
Amsterdam (1994)
38. Ku, W., Berlijn, T.,Lee, C.-C.: Unfolding first-principles band structures. Phys. Rev. Lett. 104,
216401 (2010)
39. Zunger, A., Wei, S.-H., Ferreira, L.G., Bernard, J.E.: Special quasirandom structures. Phys.
Rev. Lett. 65, 353 (1990)
40. Kováčik, R., Mavropoulos, P., Wortmann, D., Blügel, S.: Spin-caloric transport properties of
cobalt nanostructures: spin disorder effects from first principles. Phys. Rev. B 89, 134417
(2014)
41. Economou, E.N.: Green’s Functions in Quantum Physics. Springer Series in Solid-State
Sciences, vol 7. Springer, Berlin (2006)
42. Ebert, H., Ködderitzsch, D., Minár, J.: Calculating condensed matter properties using the
KKR-Green’s function method – recent developments and applications. Rep. Prog. Phys. 74,
096501 (2011)
43. Soven, P.: Coherent-potential model of substitutional disordered alloys. Phys. Rev. 156, 809
(1967)
44. Butler, W.H., Stocks, G.M.: Calculated electrical conductivity and thermopower of silverpalladium alloys. Phys. Rev. B 29, 4217 (1984)
45. Staunton, J., Gyorffy, B.L., Pindor, A.J., Stocks, G.M., Winter, H.: The ‘disordered local
moment’ picture of itinerant magnetism at finite temperatures. J. Magn. Magn. Mater. 45, 15
(1984)
46. MacDonald, A.H., Vosko, S.H.: A relativistic density functional formalism. J. Phys. C: Solid
State Phys. 12, 2977 (1979)
47. Feder, R., Rosicky, F., Ackermann, B.: Relativistic multiple scattering theory of electrons by
ferromagnets. Z. Physik B 52, 31 (1983)
48. Ebert, H.: Two ways to perform spin-polarized relativistic linear muffin-tin-orbital calculations. Phys. Rev. B 38, 9390 (1988)
49. Reiher, M., Wolf, A.: Relativistic Quantum Chemistry: The Fundamental Theory of Molecular Science. Wiley-VCH, New York (2009)
50. Pyykkö, P.: Relativistic quantum chemistry. Adv. Quantum. Chem. 11, 353 (1978)
51. Bruno, P.: Physical origins and theoretical models of magnetic anisotropy. In: Magnetismus
von Festkörpern und Grenzflächen, editor: Forschungszentrum Jülich GmbH, Institut für
Festkörperforschung, 24.1, Forschungszentrum Jülich GmbH, Jülich (1993)
52. Koelling, D.D., Harmon, B.N.: A technique for relativistic spin-polarised calculations J. Phys.
C: Solid State Phys. 10, 3107 (1977)
53. Ebert, H., Freyer, H., Vernes, A., Guo, G.-Y.: Manipulation of the spin-orbit coupling using
the Dirac equation for spin-dependent potentials. Phys. Rev. B 53, 7721 (1996)
54. Victora, R.H., MacLaren, J.M.: Predicted spin and orbital contributions to the magnetic
structure of Co/2X superlattices. J. Appl. Phys. 70, 5880 (1991)
4 Electronic Structure: Metals and Insulators
251
55. Ebert, H., Freyer, H., Deng, M.: Manipulation of the spin-orbit coupling using the Dirac
equation for spin-dependent potentials. Phys. Rev. B 56, 9454 (1997)
56. Pickel, M., Schmidt, A.B., Giesen, F., Braun, J., Minár, J., Ebert, H., Donath, M., Weinelt,
M.: Spin-orbit hybridization points in the face-centered-cubic cobalt band structure. Phys.
Rev. Lett. 101, 066402 (2008)
57. MacDonald, A.H., Daams, J.M., Vosko, S.H., Koelling, D.D.: Influence of relativistic
contributions to the effective potential on the electronic structure of Pd and Pt. Phys. Rev.
B 23, 6377 (1981)
58. Ramana, M.V., Rajagopal, A.K.: Relativistic spin-polarised electron gas. J. Phys. C: Solid
State Phys. 12, L845 (1979)
59. Ebert, H., Battocletti, M., Gross, E.K.U.: Current density functional theory of spontaneously
magnetised solids. Europhys. Lett. 40, 545 (1997)
60. Diener, G.: Current-density-functional theory for a nonrelativistic electron gas in a strong
magnetic field. J. Phys.: Cond. Mat. 3, 9417 (1991)
61. Ebert, H., Battocletti, M.: Spin and orbital polarized relativistic multiple scattering theory –
with applications to Fe, Co, Ni and Fex Co1−x . Solid State Commun. 98, 785 (1996)
62. Chadov, S., Fecher, G.H., Felser, C., Minár, J., Braun, J., Ebert, H.: Electron correlations in Co2 Mn1−x Fex Si Heusler compounds. J. Phys. D: Appl. Phys. 42, 084002
(2009)
63. Chadov, S., Minár, J., Katsnelson, M.I., Ebert, H., Ködderitzsch, D., Lichtenstein, A.I.: Orbital
magnetism in transition metal systems: the role of local correlation effects. Europhys. Lett.
82, 37001 (2008)
64. Berry, M.V.: Quantal phase factors accompanying adiabatic changes. Proc. R. Soc. Lond. A:
Math. Phys. Eng. Sci. 392, 45 (1984)
65. Chang, M.-C., Niu, Q.: Berry phase, hyperorbits, and the Hofstadter spectrum: semiclassical
dynamics in magnetic Bloch bands. Phys. Rev. B 53, 7010 (1996)
66. Wu, B., Liu, J., Niu, Q.: Geometric phase for adiabatic evolutions of general quantum states.
Phys. Rev. Lett. 94, 140402 (2005)
67. Xiao, D., Chang, M.-C., Niu, Q.: Berry phase effects on electronic properties. Rev. Mod.
Phys. 82, 1959 (2010)
68. Sundaram, G., Niu, Q.: Wave-packet dynamics in slowly perturbed crystals: gradient corrections and Berry-phase effects. Phys. Rev. B 59, 14915 (1999)
69. Bruno, P.: The Berry phase in magnetism and the anomalous Hall effect. In: Kronmüller, H.,
Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials, vol. 1, pp. 540–
558. Wiley, Chichester (2007)
70. Yao, Y., Kleinman, L., MacDonald, A.H., Sinova, J., Jungwirth, T., Wang, D.-S., Wang, E.,
Niu, Q.: First principles calculation of anomalous Hall conductivity in ferromagnetic bcc Fe.
Phys. Rev. Lett. 92, 037204 (2004)
71. Zhang, Y., Sun, Y., Yang, H., Železný, J., Parkin, S.P.P., Felser, C., Yan, B.: Strong anisotropic
anomalous Hall effect and spin Hall effect in the chiral antiferromagnetic compounds Mn3 X
(X = Ge, Sn, Ga, Ir, Rh, and Pt). Phys. Rev. B 95, 075128 (2017)
72. Karplus, R., Luttinger, J.M.: Hall effect in ferromagnetics. Phys. Rev. 95, 1154 (1954)
73. Jungwirth, T., Niu, Q., MacDonald, A.H.: Anomalous Hall effect in ferromagnetic semiconductors. Phys. Rev. Lett. 88, 207208 (2002)
74. Berger, L.: Side-jump mechanism for the Hall effect of ferromagnets. Phys. Rev. B 2, 4559
(1970)
75. Xiao, D., Yao, Y., Fang, Z., Niu, Q.: Berry-phase effect in anomalous thermoelectric transport.
Phys. Rev. Lett. 97, 026603 (2006)
76. Xiao, D., Shi, J., Niu, Q.: Berry phase correction to electron density of states in solids. Phys.
Rev. Lett. 95, 137204 (2005)
77. Thonhauser, T., Ceresoli, D., Vanderbilt, D., Resta, R.: Orbital magnetization in periodic
insulators. Phys. Rev. Lett. 95, 137205 (2005)
78. Šmejkal, L., Jungwirth, T., Sinova, J.: Route towards Dirac and Weyl antiferromagnetic
spintronics. Phys. Status Solidi (RRL): Rapid Res. Lett. 11, 1700044 (2017)
252
H. Ebert et al.
79. Chen, H., Niu, Q., MacDonald, A.H.: Anomalous Hall effect arising from noncollinear
antiferromagnetism. Phys. Rev. Lett. 112, 017205 (2014)
80. Kübler, J., Felser, C.: Non-collinear antiferromagnets and the anomalous Hall effect. Europhys. Lett. 108, 67001 (2014)
81. Šmejkal, L., Mokrousov, Y., Yan, B., MacDonald, A.H.: Topological antiferromagnetic
spintronics. Nat. Phys. 14, 242 (2018)
82. Stoner, E.C.: Collective electron specific heat and spin paramagnetism in metals. Proc. Roy.
Soc. (Lond.) A 154, 656 (1936)
83. Moriya, T.: Spin Fluctuations in Itinerant Electron Magnetism. Springer Series in Surface
Sciences, vol. 56. Springer, Berlin (1985)
84. Janak, J.F.: Uniform susceptibilities of metallic elements. Phys. Rev. B 16, 255 (1977)
85. Matsumoto, M., Staunton, J.B., Strange, P.: A new formalism for the paramagnetic
spin susceptibility of metals using relativistic spin-polarized multiple-scattering theory: a
temperature-dependent anisotropy effect. J. Phys.: Cond. Mat. 2, 8365 (1990)
86. Mankovsky, S., Ebert, H.: Theoretical description of the high-field susceptibility of magnetically ordered transition metal systems with applications to Fe, Co, Ni, and Fe1−x Cox . Phys.
Rev. B 74, 54414 (2006)
87. Buczek, P., Ernst, A., Bruno, P., Sandratskii, L.M.: Energies and lifetimes of magnons in
complex ferromagnets: a first-principle study of heusler alloys. Phys. Rev. Lett. 102, 247206
(2009)
88. Şaşıoğlu, E., Schindlmayr, A., Friedrich, C., Freimuth, F., Blügel, S.: Wannier-function
approach to spin excitations in solids. Phys. Rev. B 81, 054434 (2010)
89. Gunnarsson, O.: Band model for magnetism of transition metals in the spin-density-functional
formalism. J. Phys. F: Met. Phys. 6, 587 (1976)
90. Reddy, B.V., Khanna, S.N., Dunlap, B.I.: Giant magnetic moments in 4d clusters. Phys. Rev.
Lett. 70, 3323 (1993)
91. Cox, A.J., Louderback, J.G., Apsel, S.E., Bloomfield, L.A.: Magnetism in 4d-transition metal
clusters. Phys. Rev. B 49, 12295 (1994)
92. Vondráček, M., Cornils, L., Minár, J., Warmuth, J., Michiardi, M., Piamonteze, C., Barreto,
L., Miwa, J.A., Bianchi, M., Hofmann, P., Zhou, L., Kamlapure, A., Khajetoorians, A.A.,
Wiesendanger, R., Mi, J.-L., Iversen, B.-B., Mankovsky, S., Borek, S., Ebert, H., Schüler, M.,
Wehling, T., Wiebe, J., Honolka, J.: Nickel: the time-reversal symmetry conserving partner of
iron on a chalcogenide topological insulator. Phys. Rev. B 94, 161114 (2016)
93. Dederichs, P.H., Zeller, R., Akai, H., Ebert, H.: Ab-initio calculations of the electronic
structure of impurities and alloys of ferromagnetic transition metals. J. Magn. Magn. Mater.
100, 241 (1991)
94. Hasegawa, H., Kanamori, J.: An application of the coherent potential approximation to
ferromagnetic alloys. J. Phys. Soc. Jpn. 31, 382 (1971)
95. Minár, J., Mankovsky, S., Šipr, O., Benea, D., Ebert, H.: Correlation effects in fcc-Fex Ni1x
alloys investigated by means of the KKR-CPA. J. Phys.: Cond. Mat. 26, 274206 (2014)
96. Miura, Y., Nagao, K., Shirai, M.: Atomic disorder effects on half-metallicity of the
full-Heusler alloys Co2 (Cr1−x Fex )Al: a first-principles study. Phys. Rev. B 69, 144413
(2004)
97. Galanakis, I., Mavropoulos, P., Dederichs, P.H.: Electronic structure and Slater-Pauling
behaviour in half-metallic Heusler alloys calculated from first principles. J. Phys. D: Appl.
Phys. 39, 765 (2006)
98. Galanakis, I.: Heusler Alloys. Properties, Growth, Applications. Springer Series in Material
Science, vol. 222. Springer International Publishing, Cham (2016)
99. Jourdan, M., Minár, J., Braun, J., Kronenberg, A., Chadov, S., Balke, B., Gloskovskii, A.,
Kolbe, M., Elmers, H., Schönhense, G., Ebert, H., Felser, C., Kläui, M.: Direct observation of
half-metallicity in the Heusler compound Co2 MnSi. Nat. Commun. 5, 3974 (2014)
100. Mavropoulos, P., Galanakis, I., Popescu, V., Dederichs, P.H.: The influence of spin-orbit
coupling on the band gap of Heusler alloys. J. Phys.: Cond. Mat. 16, S5759 (2004)
101. Galanakis, I.: Surface properties of the half-and full-Heusler alloys. J. Phys.: Cond. Mat. 14,
6329 (2002)
4 Electronic Structure: Metals and Insulators
253
102. Meservey, R., Tedrow, P.: Spin-polarized electron tunneling. Phys. Rep. 238, 173 (1994)
103. Mazin, I.I.: How to define and calculate the degree of spin polarization in ferromagnets. Phys.
Rev. Lett. 83, 1427 (1999)
104. Nadgorny, B.E.: Handbook of Spin Transport and Magnetism. Taylor and Francis Group,
Boca Raton (2012)
105. de Groot, R.A., Mueller, F.M., Engen, P.G.V., Buschow, K.H.J.: New class of materials: halfmetallic ferromagnets. Phys. Rev. Lett. 50, 2024 (1983)
106. Otto, M.J., van Woerden, R.A.M., van der Valk, P.J., Wijngaard, J., van Bruggen, C.F., Haas,
C., Buschow, K.H.J.: Half-metallic ferromagnets. I. Structure and magnetic properties of
NiMnSb and related inter-metallic compounds. J. Phys.: Cond. Mat. 1, 2341 (1989)
107. Galanakis, I., Dederichs, P.H., Papanikolaou, N.: Origin and properties of the gap in the halfferromagnetic Heusler alloys. Phys. Rev. B 66, 134428 (2002)
108. Braun, J., Ebert, H., Minár, J.: Correlation and chemical disorder in Heusler compounds: a
spectroscopical study. In: Spintronics. Fundamentals and Theory, vol. 1, Springer (2013)
109. Ishida, S., Akazawa, S., Kubo, Y., Ishida, J.: Band theory of Co2 MnSn, Co2 TiSn and Co2 TiAl.
J. Phys. F: Met. Phys. 12, 1111 (1982)
110. Fujii, S., Sugimura, S., Ishida, Asano, S.: Hyperfine fields and electronic structures of the
Heusler alloys Co2 MnX (X=Al, Ga, Si, Ge, Sn). J. Phys.: Cond. Mat. 2, 8583 (1990)
111. Galanakis, I., Dederichs, P.H., Papanikolaou, N.: Slater-Pauling behavior and origin of the
half-metallicity of the full-Heusler alloys. Phys. Rev. B 66, 174429 (2002)
112. Ozdogan, K., Galanakis, I.: First-principles electronic and magnetic properties of the halfmetallic antiferromagnet. J. Magn. Magn. Mater. 321, L34 (2009)
113. Schröter, M., Ebert, H., Akai, H., Entel, P., Hoffmann, E., Reddy, G.G.: First-principles
investigations of atomic disorder effects on magnetic and structural instabilities in transitionmetal alloys. Phys. Rev. B 52, 188 (1995)
114. van Schilfgaarde, M., Abrikosov, I.A., Johansson, B.: Origin of the Invar effect in iron-nickel
alloys. Nature 400, 46 (1999)
115. Sandratskii, L.M.: Noncollinear magnetism in itinerant-electron systems: theory and applications. Adv. Phys. 47, 91 (1998)
116. Seemann, M., Ködderitzsch, D., Wimmer, S., Ebert, H.: Symmetry-imposed shape of linear
response tensors. Phys. Rev. B 92, 155138 (2015)
117. Nakatsuji, S., Kiyohara, N., Higo, T.: Large anomalous Hall effect in a non-collinear
antiferromagnet at room temperature. Nature 527, 212 (2015)
118. Ikhlas, M., Tomita, T., Koretsune, T., Suzuki, M.-T., Nishio-Hamane, D., Arita, R., Otani, Y.,
Nakatsuji, S.: Large anomalous Nernst effect at room temperature in a chiral antiferromagnet.
Nat. Phys. 13, 1085 (2017)
119. Zhang, W., Han, W., Yang, S.-H., Sun, Y., Zhang, Y., Yan, B., Parkin, S.S.P.: Giant facetdependent spin-orbit torque and spin Hall conductivity in the triangular antiferromagnet
IrMn3 . Sci. Adv. 2, e1600759 (2016)
120. Železný, J., Zhang, Y., Felser, C., Yan, B.: Spin-polarized current in noncollinear antiferromagnets. Phys. Rev. Lett. 119, 187204 (2017)
121. Sandratskii, L.M., Kübler, J.: Magnetic structures of uranium compounds: effects of relativity
and symmetry. Phys. Rev. Lett. 75, 946 (1995)
122. Connolly, J.W.D., Williams, A.R.: Density-functional theory applied to phase transformations
in transition-metal alloys. Phys. Rev. B 27, 5169 (1983)
123. Drautz, R., Fähnle, M.: Spin-cluster expansion: parametrization of the general adiabatic
magnetic energy surface with ab initio accuracy. Phys. Rev. B 69, 104404 (2004)
124. Antal, A., Lazarovits, B., Udvardi, L., Szunyogh, L., Újfalussy, B., Weinberger, P.: Firstprinciples calculations of spin interactions and the magnetic ground states of Cr trimers on
Au(111). Phys. Rev. B 77, 174429 (2008)
125. Oguchi, T., Terakura, K., Hamada, N.: Magnetism of iron above the Curie temperature. J.
Phys. F: Met. Phys. 13, 145 (1983)
126. Liechtenstein, A.I., Katsnelson, M.I., Antropov, V.P., Gubanov, V.A.: Local spin density
functional approach to the theory of exchange interactions in ferromagnetic metals and alloys.
J. Magn. Magn. Mater. 67, 65 (1987)
254
H. Ebert et al.
127. Pajda, M., Kudrnovský, J., Turek, I., Drchal, V., Bruno, P.: Oscillatory Curie temperature of
two-dimensional ferromagnets. Phys. Rev. Lett. 85, 5424 (2000)
128. Polesya, S., Mankovsky, S., Šipr, O., Meindl, W., Strunk, C., Ebert, H.: Finite-temperature
magnetism of Fex Pd1−x and Cox Pt1−x alloys. Phys. Rev. B 82, 214409 (2010)
129. Ležaić, M., Mavropoulos, P., Enkovaara, J., Bihlmayer, G., Blügel, S.: Thermal collapse of
spin polarization in half-metallic ferromagnets. Phys. Rev. Lett. 97, 026404 (2006)
130. Buchelnikov, V.D., Entel, P., Taskaev, S.V., Sokolovskiy, V.V., Hucht, A., Ogura, M., Akai, H.,
Gruner, M.E., Nayak, S.K.: Monte Carlo study of the influence of antiferromagnetic exchange
interactions on the phase transitions of ferromagnetic Ni-Mn-X alloys (X=In,Sn,Sb). Phys.
Rev. B 78, 184427 (2008)
131. Buchelnikov, V.D., Sokolovskiy, V.V., Herper, H.C., Ebert, H., Gruner, M.E., Taskaev, S.V.,
Khovaylo, V.V., Hucht, A., Dannenberg, A., Ogura, M., Akai, H., Acet, M., Entel, P.:
First-principles and Monte Carlo study of magnetostructural transition and magnetocaloric
properties of Ni2+x Mn1−x Ga. Phys. Rev. B 81, 094411 (2010)
132. Sato, K., Dederichs, P.H., Katayama-Yoshida, H.: Curie temperatures of dilute magnetic
semiconductors from LDA+U electronic structure calculations. Physica B 376–377, 639
(2006)
133. Toyoda, M., Akai, H., Sato, K., Katayama-Yoshida, H.: Curie temperature of GaMnN and
GaMnAs from LDA-SIC electronic structure calculations. Phys. Stat. Sol. (C) 3, 4155 (2006)
134. Nayak, S.K., Ogura, M., Hucht, A., Akai, H., Entel, P.: Monte Carlo simulations of diluted
magnetic semiconductors using ab initio exchange parameters. J. Phys.: Cond. Mat. 21,
064238 (2009)
135. Bouzerar, G., Kudrnovský, J., Bergqvist, L., Bruno, P.: Ferromagnetism in diluted magnetic
semiconductors: a comparison between ab initio mean-field, RPA, and Monte Carlo treatments. Phys. Rev. B 68, 081203 (2003)
136. Eriksson, O., Bergqvist, L., Sanyal, B., Kudrnovský, J., Drchal, V., Korzhavyi, P., Turek, I.:
Electronic structure and magnetism of diluted magnetic semiconductors. J. Phys.: Cond. Mat.
16, S5481 (2004)
137. Sato, K., Bergqvist, L., Kudmovsky, J., Dederichs, P.H., Eriksson, O., Turek, I., Sanyal,
B., Bouzerar, G., Katayama-Yoshida, H., Dinh, V.A., Fukushima, T., Kizaki, H., Zeller,
R.: First-principles theory of dilute magnetic semiconductors. Rev. Mod. Phys. 82, 1633
(2010)
138. Maccherozzi, F., Sperl, M., Panaccione, G., Minár, J., Polesya, S., Ebert, H., Wurstbauer, U.,
Hochstrasser, M., Rossi, G., Woltersdorf, G., Wegscheider, W., Back, C.H.: Evidence for a
magnetic proximity effect up to room temperature at Fe/(Ga, Mn)As interfaces. Phys. Rev.
Lett. 101, 267201 (2008)
139. Polesya, S., Šipr, O., Bornemann, S., Minár, J., Ebert, H.: Magnetic properties of free Fe
clusters at finite temperatures from first principles. Europhys. Lett. 74, 1074 (2006)
140. Šipr, O., Polesya, S., Minár, J., Ebert, H.: Influence of temperature on the systematics of
magnetic moments of free Fe clusters. J. Phys.: Cond. Mat. 19, 446205 (2007)
141. Katsnelson, M.I., Lichtenstein, A.I.: First-principles calculations of magnetic interactions in
correlated systems. Phys. Rev. B 61, 8906 (2000)
142. Ebert, H., Mankovsky, S.: Anisotropic exchange coupling in diluted magnetic semiconductors: ab initio spin-density functional theory. Phys. Rev. B 79, 045209 (2009)
143. Mankovsky, S., Bornemann, S., Minár, J., Polesya, S., Ebert, H., Staunton, J.B., Lichtenstein,
A.I.: Effects of spin-orbit coupling on the spin structure of deposited transition-metal clusters.
Phys. Rev. B 80, 014422 (2009)
144. Antropov, V.P., Katsnelson, M.I., Harmon, B.N., van Schilfgaarde, M., Kusnezov, D.: Spin
dynamics in magnets: equation of motion and finite temperature effects. Phys. Rev. B 54,
1019 (1996)
145. Ebert, H.: Relativistic theory of indirect nuclear spin-spin coupling. Phil. Mag. 88, 2673
(2008)
146. Sandratskii, L.M., Bruno, P.: Exchange interactions and Curie temperature in (Ga,Mn)As.
Phys. Rev. B 66, 134435 (2002)
4 Electronic Structure: Metals and Insulators
255
147. Uhl, M., Sandratskii, L.M., Kübler, J.: Spin fluctuations in γ -Fe and in Fe3 Pt Invar from
local-density-functional calculations. Phys. Rev. B 50, 291 (1994)
148. Heide, M., Bihlmayer, G., Blügel, S.: Dzyaloshinskii-Moriya interaction accounting for the
orientation of magnetic domains in ultrathin films: Fe/W(110). Phys. Rev. B 78, 140403
(2008)
149. Solovyev, I.V., Kashin, I.V., Mazurenko, V.V.: Mechanisms and origins of half-metallic
ferromagnetism in CrO2 . Phys. Rev. B 92, 144407 (2015)
150. Keshavarz, S., Kvashnin, Y.O., Rodrigues, D.C.M., Pereiro, M., Di Marco, I., Autieri, C.,
Nordström, L., Solovyev, I.V., Sanyal, B., Eriksson, O.: Exchange interactions of CaMnO3 in
the bulk and at the surface. Phys. Rev. B 95, 115120 (2017)
151. Logemann, R., Rudenko, A.N., Katsnelson, M.I., Kirilyuk, A.: Exchange interactions in
transition metal oxides: the role of oxygen spin polarization. J. Phys.: Condens. Matter 29,
335801 (2017)
152. Katanin, A.A., Poteryaev, A.I., Efremov, A.V., Shorikov, A.O., Skornyakov, S.L., Korotin,
M.A., Anisimov, V.I.: Orbital-selective formation of local moments in α-iron: first-principles
route to an effective model. Phys. Rev. B 81, 045117 (2010)
153. Kvashnin, Y.O., Cardias, R., Szilva, A., Di Marco, I., Katsnelson, M.I., Lichtenstein, A.I.,
Nordström, L., Klautau, A.B., Eriksson, O.: Microscopic origin of Heisenberg and NonHeisenberg exchange interactions in ferromagnetic bcc Fe. Phys. Rev. Lett. 116, 217202
(2016)
154. Szilva, A., Thonig, D., Bessarab, P.F., Kvashnin, Y.O., Rodrigues, D.C.M., Cardias, R.,
Pereiro, M., Nordström, L., Bergman, A., Klautau, A.B., Eriksson, O.: Theory of noncollinear
interactions beyond Heisenberg exchange: applications to bcc Fe. Phys. Rev. B 96, 144413
(2017)
155. Szunyogh, L., Újfalussy, B., Weinberger, P.: Magnetic anisotropy of iron multilayers on
Au(001): first-principles calculations in terms of the fully relativistic spin-polarized screened
KKR method. Phys. Rev. B 51, 9552 (1995)
156. Razee, S.S.A., Staunton, J.B., Pinski, F.J.: First-principles theory of magnetocrystalline
anisotropy of disordered alloys: application to cobalt platinum. Phys. Rev. B 56, 8082 (1997)
157. Újfalussy, B., Szunyogh, L., Weinberger, P.: Magnetic anisotropy in Fe/Cu(001) overlayers
and interlayers: the high-moment ferromagnetic phase. Phys. Rev. B 54, 9883 (1996)
158. Solovyev, I.V., Dederichs, P.H., Mertig, I.: Origin of orbital magnetization and magnetocrystalline anisotropy in TX ordered alloys (where T =Fe,Co and X =Pd,Pt). Phys. Rev. B 52,
13419 (1995)
159. Bruno, P.: Tight-binding approach to the orbital magnetic moment and magnetocrystalline
anisotropy of transition-metal monolayers. Phys. Rev. B 39, 865 (1989)
160. van der Laan, G.: Determination of the element-specific magnetic anisotropy in thin films and
surfaces. J. Phys.: Cond. Mat. 13, 11149 (2001)
161. Wang, X., Wu, R., Wang, D.-S., Freeman, A.J.: Torque method for the theoretical determination of magnetocrystalline anisotropy. Phys. Rev. B 54, 61 (1996)
162. Staunton, J.B., Szunyogh, L., Buruzs, A., Gyorffy, B.L., Ostanin, S., Udvardi, L.: Temperature
dependence of magnetic anisotropy: an ab initio approach. Phys. Rev. B 74, 144411 (2006)
163. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: First-principles calculation of the
magnetic anisotropy energy of (Co)n /(X)m multilayers. Phys. Rev. B 42, 7270 (1990)
164. Weinberger, P.: Magnetic Anisotropies in Nanostructured Matter. Condensed Matter Physics.
Chapman and Hall/CRC Press, Boca Raton (2008)
165. Stiles, M.D., Halilov, S.V., Hyman, R.A., Zangwill, A.: Spin-other-orbit interaction and
magnetocrystalline anisotropy. Phys. Rev. B 64, 104430 (2001)
166. Bornemann, S., Minár, J., Braun, J., Ködderitzsch, D., Ebert, H.: Ab-initio description of the
magnetic shape anisotropy due to the Breit interaction. Solid State Commun. 152, 85 (2012)
167. Buschow, K., van Diepen, A., de Wijn, H.: Crystal-field anisotropy of Sm3+ in SmCo5 . Solid
State Commun. 15, 903 (1974)
168. Yamada, M., Kato, H., Yamamoto, H., Nakagawa, Y.: Crystal-field analysis of the magnetization process in a series of Nd2 Fe14 B-type compounds. Phys. Rev. B 38, 620 (1988)
256
H. Ebert et al.
169. Herbst, J.F.: R2 Fe14 B materials: intrinsic properties and technological aspects. Rev. Mod.
Phys. 63, 819 (1991)
170. Hummler, K., Fähnle, M.: Full-potential linear-muffin-tin-orbital calculations of the magnetic
properties of rare-earth transition-metal intermetallics. I. Description of the formalism and
application to the series R Co5 (R =rare-earth atom). Phys. Rev. B 53, 3272 (1996)
171. Hummler, K., Fähnle, M.: Ab initio calculation of local magnetic moments and the crystal
field in R2 Fe14 B (R =Gd, Tb, Dy, Ho, and Er). Phys. Rev. B 45, 3161 (1992)
172. Coehoorn, R.: Supermagnets, Hard Magnetic Materials. Nato ASI Series, Series C, chapter 8,
vol. 331, p. 133. Kluwer Academic Publishers, Dardrecht (1991)
173. Richter, M., Oppeneer, P.M., Eschrig, H., Johansson, B.: Calculated crystal-field parameters
of SmCo5 . Phys. Rev. B 46, 13919 (1992)
174. Hummler, K., Fähnle, M.: Full-potential linear-muffin-tin-orbital calculations of the magnetic
properties of rare-earth transition-metal intermetallics. II. Nd2 Fe14 B. Phys. Rev. B 53, 3290
(1996)
175. Moriya, H., Tsuchiura, H., Sakuma, A.: First principles calculation of crystal field parameter
near surfaces of Nd2 Fe14 B. J. Appl. Phys. 105, 07A740 (2009)
176. Tanaka, S., Moriya, H., Tsuchiura, H., Sakuma, A., Diviš, M., Novák, P.: First principles study
on the local magnetic anisotropy near surfaces of Dy2 Fe14 B and Nd2 Fe14 B magnets. J. Appl.
Phys. 109, 07A702 (2011)
177. Novák, P., Knížek, K., Kuneš, J.: Crystal field parameters with Wannier functions: application
to rare-earth aluminates. Phys. Rev. B 87, 205139 (2013)
178. Novák, P., Kuneš, J., Knížek, K.: Crystal field of rare earth impurities in LaF3 . Opt. Mater.
37, 414 (2014)
179. Patrick, C.E., Kumar, S., Balakrishnan, G., Edwards, R.S., Lees, M.R., Petit, L., Staunton,
J.B.: Calculating the magnetic anisotropy of rare-earth–transition-metal ferrimagnets. Phys.
Rev. Lett. 120, 097202 (2018)
180. Halilov, S.V., Eschrig, H., Perlov, A.Y., Oppeneer, P.M.: Adiabatic spin dynamics from spindensity-functional theory: application to Fe, Co, and Ni. Phys. Rev. B 58, 293 (1998)
181. Grotheer, O., Ederer, C., Fähnle, M.: Fast ab initio methods for the calculation of adiabatic
spin wave spectra in complex systems. Phys. Rev. B 63, 100401 (2001)
182. Pajda, M., Kudrnovský, J., Turek, I., Drchal, V., Bruno, P.: Ab initio calculations of exchange
interactions, spin-wave stiffness constants, and Curie temperatures of Fe, Co, and Ni. Phys.
Rev. B 64, 174402 (2001)
183. Turek, I., Kudrnovský, J., Drchal, V., Bruno, P.: Exchange interactions, spin waves, and
transition temperatures in itinerant magnets. Phil. Mag. 86, 1713 (2006)
184. Brinkman, W.F., Elliot, R.J.: Theory of spin-space groups. Proc. R. Soc. (Lond.) A 294, 343
(1966)
185. Brinkman, W.F., Elliot, R.J.: Space group theory for spin waves. J. Appl. Phys. 37, 1457
(1966)
186. Herring, C.: Magnetism: exchange interactions among itinerant electrons In: Rado, G., Suhl,
H. (eds.) Magnetism, vol. IV, p. 191. Academic Press, New York (1966)
187. Sandratskii, L.M.: Symmetry analysis of electronic states for crystals with spiral magnetic
order. I. General properties. J. Phys.: Cond. Mat. 3, 8565 (1991)
188. Uhl, M., Sandratskii, L., Kübler, J.: Electronic and magnetic states of γ -Fe. J. Magn. Magn.
Mater. 103, 314 (1992)
189. Kurz, P., Förster, F., Nordström, L., Bihlmayer, G., Blügel, S.: Ab initio treatment of
noncollinear magnets with the full-potential linearized augmented plane wave method. Phys.
Rev. B 69, 024415 (2004)
190. Mankovsky, S., Fecher, G.H., Ebert, H.: Electronic structure calculations in ordered and
disordered solids with spiral magnetic order. Phys. Rev. B 83, 144401 (2011)
191. Kübler,J.: Ab initio estimates of the Curie temperature for magnetic compounds. J. Phys.:
Condens. Matter 18, 9795 (2006)
192. Galanakis, I., Sasioglu, E.: Ab-initio calculation of effective exchange interactions, spin
waves, and Curie temperature in L21 - and L12 -type local moment ferromagnets. J. Mater.
Sci. 47, 7678 (2012)
4 Electronic Structure: Metals and Insulators
257
193. Şaşıoğlu, E., Sandratskii, L.M., Bruno, P., Galanakis, I.: Exchange interactions and temperature dependence of magnetization in half-metallic Heusler alloys. Phys. Rev. B 72, 184415
(2005)
194. Edwards, D.M., Katsnelson, M.I.: High-temperature ferromagnetism of sp electrons in narrow
impurity bands: application to CaB6 . J. Phys.: Condens. Matter 18, 7209 (2006)
195. Buczek, P., Ernst, A., Sandratskii, L.M.: Spin dynamics of half-metallic Co2 MnSi. J. Phys.:
Conf. Ser. 200, 042006 (2010)
196. Savrasov, S.Y.: Linear response calculations of spin fluctuations. Phys. Rev. Lett. 81, 2570
(1998)
197. Qian, Z., Vignale, G.: Spin dynamics from time-dependent spin-density-functional theory.
Phys. Rev. Lett. 88, 056404 (2002)
198. Lounis, S., dos Santos Dias, M., Schweflinghaus, B.: Transverse dynamical magnetic
susceptibilities from regular static density functional theory: evaluation of damping and g
shifts of spin excitations. Phys. Rev. B 91, 104420 (2015)
199. Bruno, P.: Exchange interaction parameters and adiabatic spin-wave spectra of ferromagnets:
a “renormalized magnetic force theorem”. Phys. Rev. Lett. 90, 087205 (2003)
200. Katsnelson, M.I., Lichtenstein, A.I.: Magnetic susceptibility, exchange interactions and spinwave spectra in the local spin density approximation. J. Phys.: Condens. Matter 16, 7439
(2004)
201. Buczek, P., Ernst, A., Sandratskii, L.M.: Interface electronic complexes and landau damping
of magnons in ultrathin magnets. Phys. Rev. Lett. 106, 157204 (2011)
202. Tajima, K., Ishikawa, Y., Webster, P.J., Stringfellow, M.W., Tocchetti, D., Zeabeck, K.R.A.:
Spin waves in a heusler alloy Cu2 MnAl. J. Phys. Soc. Jpn. 43, 483 (1977)
203. Buczek, P., Ernst, A., Sandratskii, L.M.: Different dimensionality trends in the Landau
damping of magnons in iron, cobalt, and nickel: time-dependent density functional study.
Phys. Rev. B 84, 174418 (2011)
204. Müller, M.C.T.D., Friedrich, C., Blügel, S.: Acoustic magnons in the long-wavelength limit:
investigating the Goldstone violation in many-body perturbation theory. Phys. Rev. B 94,
064433 (2016)
205. Staunton, J.B., Poulter, J., Ginatempo, B., Bruno, E., Johnson, D.D.: Incommensurate
and commensurate antiferromagnetic spin fluctuations in Cr and Cr alloys from ab initio
dynamical spin susceptibility calculations. Phys. Rev. Lett. 82, 3340 (1999)
206. Staunton, J.B., Poulter, J., Ginatempo, B., Bruno, E., Johnson, D.D.: Spin fluctuations in
nearly magnetic metals from ab initio dynamical spin susceptibility calculations: application
to Pd and Cr95 V5 . Phys. Rev. B 62, 1075 (2000)
207. Schindlmayr, A., Friedrich, C., Sasioglu, E., Blügel, S.: First-principles calculation of
electronic excitations in solids with SPEX. Z. Phys. Chem. 224, 357 (2010)
208. Tyablokov, S.V.: Methods of Quantum Theory of Magnetism. Plenum Press, New York
(1967)
209. Callen, H.B.: Green function theory of ferromagnetism. Phys. Rev. 130, 890 (1963)
210. Bose, S.K., Kudrnovský, J.,Drchal, V., Turek, I.: Magnetism of mixed quaternary Heusler
alloys: (N i, T )2 MnSn (T = Cu, P d) as a case study. Phys. Rev. B 82, 174402 (2010)
211. Rusz, J., Turek, I., Diviš, M.: Random-phase approximation for critical temperatures of
collinear magnets with multiple sublattices: GdX compounds (X = Mg, Rh, Ni, Pd). Phys.
Rev. B 71, 174408 (2005)
212. Sandratskii, L.M., Singer, R., Şaşıoğlu, E.: Heisenberg Hamiltonian description of multiplesublattice itinerant-electron systems: general considerations and applications to NiMnSb and
MnAs. Phys. Rev. B 76, 184406 (2007)
213. Mermin, N.D., Wagner, H.: Absence of ferromagnetism or antiferromagnetism in one- or
two-dimensional isotropic Heisenberg models. Phys. Rev. Lett. 17, 1133 (1966)
214. Bruno, P.: Magnetization and Curie temperature of ferromagnetic ultrathin films: the influence
of magnetic anisotropy and dipolar interactions (invited). Mater. Res. Soc. Symp. Proc. 231,
299 (1991)
215. Bouzerar, G., Bruno, P.: RPA-CPA theory for magnetism in disordered Heisenberg binary
systems with long-range exchange integrals. Phys. Rev. B 66, 014410 (2002)
258
H. Ebert et al.
216. Kübler, J., Fecher, G.H., Felser, C.: Understanding the trend in the Curie temperatures of
Co2 -based Heusler compounds: ab initio calculations. Phys. Rev. B 76, 024414 (2007)
217. Rosengaard, N.M., Johansson, B.: Finite-temperature study of itinerant ferromagnetism in Fe,
Co, and Ni. Phys. Rev. B 55, 14975 (1997)
218. Jakobsson, A., Şaşıoğlu, E., Mavropoulos, P., Ležaić, M., Sanyal, B., Bihlmayer, G., Blügel,
S.: Tuning the Curie temperature of FeCo compounds by tetragonal distortion. Appl. Phys.
Lett. 103, 102404 (2013)
219. Bergqvist, L., Korzhavyi, P.A., Sanyal, B., Mirbt, S., Abrikosov, I.A., Nordström, L.,
Smirnova, E.A., Mohn, P., Svedlindh, P., Eriksson, O.: Magnetic and electronic structure of
(Ga1−x Mnx )As. Phys. Rev. B 67, 205201 (2003)
220. Bergqvist, L., Eriksson, O., Kudrnovský, J., Drchal, V., Korzhavyi, P., Turek, I.: Magnetic
percolation in diluted magnetic semiconductors. Phys. Rev. Lett. 93, 137202 (2004)
221. Ležaić, M., Mavropoulos, P., Bihlmayer, G., Blügel, S.: Exchange interactions and localmoment fluctuation corrections in ferromagnets at finite temperatures based on noncollinear
density-functional calculations. Phys. Rev. B 88, 134403 (2013)
222. Uhl, M., Kübler, J.: Exchange-coupled spin-fluctuation theory: application to Fe, Co, and Ni.
Phys. Rev. Lett. 77, 334 (1996)
223. Ruban, A.V., Khmelevskyi, S., Mohn, P., Johansson, B.: Temperature-induced longitudinal
spin fluctuations in Fe and Ni. Phys. Rev. B 75, 054402 (2007)
224. Williams, A.R., Zeller, R., Moruzzi, V.L., Gelatt, C.D., Kubler, J.: Covalent magnetism: an
alternative to the Stoner model. J. Appl. Phys. 52, 2067 (1981)
225. Mohn, P., Schwarz, K.: Supercell calculations for transition metal impurities in palladium. J.
Phys.: Cond. Mat. 5, 5099 (1993)
226. Mryasov, O.N., Nowak, U., Guslienko, K.Y., Chantrell, R.W.: Temperature-dependent magnetic properties of FePt: effective spin Hamiltonian model. Europhys. Lett. 69, 805 (2005)
227. Mryasov, O.N.: Magnetic interactions and phase transformations in FeM, M = (Pt,Rh) ordered
alloys. Phase Transit. 78, 197 (2005)
228. Kudrnovský, J., Drchal, V., Bruno, P.: Magnetic properties of fcc Ni-based transition metal
alloys. Phys. Rev. B 77, 224422 (2008)
229. Polesya, S., Mankovsky, S., Ködderitzsch, D., Minár, J., Ebert, H.: Finite-temperature
magnetism of FeRh compounds. Phys. Rev. B 93, 024423 (2016)
230. Gyorffy, B.L., Pindor, A.J., Staunton, J., Stocks, G.M., Winter, H.: A first-principles theory
of ferromagnetic phase transitions in metals. J. Phys. F: Met. Phys. 15, 1337 (1985)
231. Feynman, R.P.: Slow electrons in a polar crystal. Phys. Rev. 97, 660 (1955)
232. Staunton, J.B., Ostanin, S., Razee, S.S.A., Gyorffy, B.L., Szunyogh, L., Ginatempo, B.,
Bruno, E.: Temperature dependent magnetic anisotropy in metallic magnets from an ab initio
electronic structure theory: L10 -ordered FePt. Phys. Rev. Lett. 93, 257204 (2004)
233. Buruzs, A., Weinberger, P., Szunyogh, L., Udvardi, L., Chleboun, P.I., Fischer, A.M.,
Staunton, J.B.: Ab initio theory of temperature dependence of magnetic anisotropy in layered
systems: applications to thin Co films on Cu(100). Phys. Rev. B 76, 064417 (2007)
234. Zhuravlev, I.A., Antropov, V.P., Belashchenko, K.D.: Spin-fluctuation mechanism of anomalous temperature dependence of magnetocrystalline anisotropy in itinerant magnets. Phys.
Rev. Lett. 115, 217201 (2015)
235. Hubbard, J.: Calculation of partition functions. Phys. Rev. Lett. 3, 77 (1959)
236. Kakehashi, Y.: Monte Carlo approach to the dynamical coherent-potential approximation in
metallic magnetism. Phys. Rev. B 45, 7196 (1992)
237. Kakehashi, Y., Shimabukuro, T., Tamashiro, T., Nakamura, T.: Dynamical coherent-potential
approximation and tight-binding linear muffintin orbital approach to correlated electron
system. J. Phys. Soc. Jpn. 77, 094706 (2008)
238. Kakehashi, Y.: Many-body coherent potential approximation, dynamical coherent potential
approximation, and dynamical mean-field theory. Phys. Rev. B 66, 104428 (2002)
239. Poteryaev, A.I., Skorikov, N.A., Anisimov, V.I., Korotin, M.A.: Magnetic properties of
Fe1−x Nix alloy from CPA+DMFT perspectives. Phys. Rev. B 93, 205135 (2016)
4 Electronic Structure: Metals and Insulators
259
240. Patrick, C.E., Staunton, J.B.: Rare-earth/transition-metal magnets at finite temperature: selfinteraction-corrected relativistic density functional theory in the disordered local moment
picture Phys. Rev. B 97, 224415 (2018)
Hubert Ebert studied physics and received his Ph.D. from the
Ludwig-Maximilians-University Munich in 1986. After a postdoc stay at the University of Bristol (UK), he worked for several
years at the central laboratory for research and development of
Siemens Company in Erlangen. Since 1993 he is professor for
theoretical physical chemistry at the university of Munich.
Sergiy Mankovsky studied physics in Moscow institute of
Physics and Technology. In 1992 he received the degree of candidate of physico-mathematical sciences (an analogue of the Ph.D.)
from the Kurdyumov Institute for Metal Physics of the N.A.S. of
Ukraine, where he worked as a research scientist until 2001. Since
2001 he works at the Ludwig-Maximilians-University Munich.
Sebastian Wimmer received his Ph.D. from the LudwigMaximilians-Universität München in 2018. Until 2019 he worked
in the group of Prof. Dr. Hubert Ebert, focusing on the firstprinciples description of spintronic and spincaloritronic linear
response properties of metals and alloys.
5
Quantum Magnetism
Gabriel Aeppli and Philip Stamp
Contents
Spin Paths and Spin Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Importance of Decoherence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Quantum Relaxation in Dipolar Nets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Large-Scale Coherence and Entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Future Directions and Open Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
262
265
268
271
276
277
Abstract
Macroscopic quantum effects have been familiar since the discovery of superfluids and superconductors over 100 years ago. In the last few decades, it
has been understood how large-scale quantum effects can also show up in
“spin space.” The collective tunneling of many spins was observed in magnetic
nanomolecules and in insulating dipolar-coupled spin arrays, and the tunneling
of ferromagnetic domain walls has also been cleanly identified. To see largescale coherence or entanglement effects, the decoherence caused by interactions
with the environment (particularly with nuclear spins) must be controlled.
Theory indicates ways of doing this, and systems ranging from classic magnetic
compounds to deterministically doped silicon will make the job easier. Coherent
G. Aeppli ()
Physics Department (ETHZ), Institut de Physique (EPFL) and Photon Science Division (PSI),
ETHZ, EPFL and PSI, Zürich, Lausanne and Villigen, Switzerland
e-mail: [email protected]
P. Stamp ()
Pacific Institute of Theoretical Physics, University of British Columbia, Vancouver, BC, Canada
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_5
261
262
G. Aeppli and P. Stamp
control of quantum spin arrays, and large-scale quantum spin superpositions, is
a likely prospect for the future.
Magnetism has its microscopic origin in quantum mechanics which determines
basic parameters such as single-site anisotropy and intersite couplings. However,
once the parameters have been fixed, classical reasoning is usually all that is
required to model and engineer real materials and devices. In this sense, magnetism
is not different than any other branch of condensed matter physics: quantum
mechanics is essential to describe very small objects such as individual nuclei,
atoms, and molecules, but the emergent behavior of macroscopic ensembles of
such objects can almost always be explained in classical terms. There are a few
striking exceptions, most notably for fractional quantum Hall systems and Bose
condensates. These depend for their existence on “macroscopic wave functions”
[1], which correlate simultaneously the dynamics of all the particles into a singlewave function Ψ ∼ |Ψ (r, t)|eiϕ(r,t) , with a phase ϕ(r, t) varying in space and time.
So, can any other kinds of large-scale quantum behavior be expected in physics, not
entailing a Laughlin state, Bose condensation, or supercurrents? We report here on
some of the remarkable ways this can happen, not in real space but in spin space.
These developments are relatively recent and constitute a new class of quantum
phenomena – and in the last three decades or so, experiments have begun to confront
theory in a rigorous way.
Spin Paths and Spin Phase
There is a remarkable way of formulating quantum mechanics, found by Feynman
during his doctoral studies [2], which allows a neat appreciation of these developments. One writes the usual amplitude Gba , for the transition between two quantum
states |ψa and |ψb , as a sum of “amplitudes” over all possible paths between
them – the system simultaneously explores all these paths. The probability that
the transition will then occur is the usual “amplitude squared,” i.e., |Gba |2 . The
amplitude or weighting factor for the μ-th path is ei Aμ /h̄ , exponentiating a phase
ϕ = Aμ /h̄, where Aμ is the “action” of this path. Crucially, the action is just that
for the corresponding
classical system (e.g., a particle) moving along the same path,
i.e., Aμ = dtL, where L is the classical Lagrangian.
This sum over paths, or “path integral,” brings out vividly the essential role of
phase interference between different paths. It has been enormously fruitful, both
as a pedagogical guide [3] and in advancing our understanding of basic quantum
physics and quantum field theory [4]. An obvious question, which notoriously
worried Feynman, is how to then deal with quantum spins, which have no classical
limit (if the spin quantum number S is finite, then when h̄ → 0, the spin moment
h̄S disappears!). The answer turned out to be interesting [5, 6, 7]. Imagine a unit
sphere in spin space, with a “particle” of unit charge moving on its surface – its
coordinate n̂(t) = S(t)/S defines the spin direction (see Fig. 1). The fake charge
5 Quantum Magnetism
263
accumulates a conventional “potential phase” by coupling to a potential Ho (S) (the
spin Hamiltonian) on the sphere. Crucially, however, the charge also couples to a
fake monopole of strength q = h̄S, situated at the center of the sphere. This adds a
“kinetic” or Berry phase [6], having the same form as the Aharonov-Bohm phase [8]
for a charge moving in real space (see Fig. 1a). One then sums over paths on the spin
sphere, to reproduce exactly the dynamics of a quantum spin.
Tunneling spins: As an example, suppose Ho (S) has local potential minima – at
low energy, the spin must then quantum tunnel between these minima to move at
all. For simplicity, we imagine two paths, connecting the lowest states | ↑, | ↓ in
these energy minima (Fig. 1b). This is just like the famous “two-slit” problem – the
tunneling amplitude o will be controlled by the interference due to the different
phases ϕ1 , ϕ2 accumulated along the two paths [3]. Two points are interesting here.
First, as discussed by Bogachek and Krive [10], and later others [11, 12], one can
manipulate the two paths by applying a transverse magnetic field H⊥
o , thereby giving
oscillations in the tunneling amplitude (Fig. 1c). Second, by considering only the
lowest state in each potential well, we have in effect truncated the Hamiltonian
Ho (S) to a two-level system (justified if the temperature is low enough that higher
spin states are inactive). We can write Ho (S) → Ho (τ̂ ) = o τ̂x , where τ̂ is the
Pauli vector. The eigenstates are just the symmetric and antisymmetric combinations
of | ↑ and | ↓. If the paths are oppositely directed but otherwise symmetric, then
ϕ1 = −ϕ2 = π S, and the transition amplitude (and hence the tunneling splitting)
between the two eigenstates is then ∼ (eiϕ +e−iϕ ) ∝ cos(π S), giving a gap between
the lowest and first excited states if S is integer-valued, but no gap if S is half-integer.
This last result was of course already noted by Kramers [13] in 1930; spin paths are
not required to understand it! Interest in spin phases was really set off by Haldane’s
remarkable prediction [9] that a large gap existed in the spectrum of an integer spin
chain (but not a half-integer one).
If we add an extra “longitudinal” field Hoz to lift the degeneracy of the minima
(i.e., of the two states | ↑ and | ↓), by an “energy bias,” o = gμB Sz Hoz . Then our
two-level system (or qubit, in the language of quantum information) has
Hamiltonian:
Ho (τ̂ ) = o τ̂ x + o τ̂ z
(1)
We see the tunneling term o tries to drive quantum transitions between | ↑ and
| ↓, but is hindered by the longitudinal field bias o , which pushes the states | ↑,
| ↓ out of resonance. Notice that we can manipulate both o and o with external
fields.
Quantum Ising Spin Networks: Now imagine an interacting “spin net” of twolevel systems τ̂i , with i = 1, 2, ..N, and each with its own tunneling amplitude i
and bias i . Adding interspin interactions between them generalizes (1) to:
HoQI =
i τ̂ix + i τ̂iz +
Vij τ̂iz τ̂jz
i
ij
(2)
264
G. Aeppli and P. Stamp
a
b
flux
ω
n(t)
q
c 1E+1
1E+0
1E-1
1E-2
90
1E-3
o
1E-4
50
Δ(K)
1E-5
o
o
20
1E-6
o
1E-7
7
1E-8
0
o
1E-9
1E-10
0
Fig. 1 (Continued)
1
2
H (T)
3
4
5
5 Quantum Magnetism
265
We only include longitudinal couplings (coupling the z-components of the spins) in
the interaction Vij ; in many interesting magnetic systems, the other interaction terms
are strongly suppressed relative to these. Notice that the total longitudinal
field bias
blocking quantum transitions on the i-th site is now ξi = i + j Vij τjz , instead of
just i .
This “Quantum Ising” model describes a large variety of physical systems which,
at low temperature, truncate to a set of interacting two-level systems. The best
known examples in quantum magnetism are dipolar-coupled magnetic molecules
and ions and quantum spin glasses. However, H0QI also describes a set of N
interacting qubits – and if we can manipulate the couplings in it, we have a toy
model for a quantum computer. In quantum computation (reviewed in [14, 15, 16]),
one creates and manipulates N-qubit states which are “entangled,” i.e., which cannot
in general be written as products over separate spins. Such states are fundamentally
nonclassical, as first discussed by Schrödinger and Einstein in 1935. The “quantum
information” in them is encoded in the 2N −1 relative “spin phases” between the
different qubits. To make and use such states is one of the great goals in this field –
but it will be hard. To see why, we must first understand the main obstacle in the
way.
The Importance of Decoherence
Decoherence arises when a quantum system interacts with its environment, and their
Feynman paths and quantum phases become entangled. Even if they later decouple,
averaging over the unknown environmental states then “smears” over states of
the system, destroying phase coherence over some timescale τφ (the “decoherence
time”). Decoherence in many-particle systems is usually much larger than expected,
Fig. 1 Path integrals on the spin sphere. In (a), we show a spin S as a unit charge, moving on the
unit sphere around a magnetic monopole of strength q = h̄S, along a path which
has coordinate
n̂(t) = S(t)/S, and encloses a solid angle . The kinetic phase φB = q/h̄ A · dn, where A
is the monopole vector potential (compare the “Aharonov-Bohm” phase [8] accumulated by a
charged particle moving through an ordinary magnetic vector potential A(r) in real space). For
a closed path, Gauss’s theorem then shows that the enclosed flux from the fake monopole is just
φB = S. In (b), a biaxial (easy ẑ-axis, hard x̂-axis) potential field Ho (S) is added (dark areas
are regions of higher potential). The spin moves preferentially between the two minimum energy
states at the poles by tunneling along the pair of minimum action paths (shown as dashed lines),
with amplitudes 12 μ eiϕμ respectively, where μ = 1, 2 labels the paths, and the μ are real.
An external field H⊥
o , applied along the hard x̂-axis, pulls the two states, and the paths between
them, toward x̂, thereby reducing the enclosed area
on the unit sphere. In this “symmetric”
˜
case, ϕ1 = −ϕ2 = ϕ(H⊥
o ) and 1 = 2 = o . The total tunneling amplitude is then just the
⊥ ) cos ϕ(H⊥ ). (c) shows the resulting oscillations in
˜ o (H⊥
sum 12 o (eiϕ + e−iϕ ), i.e., )
=
(H
o
o
o
o
˜ o (H⊥
o ), for a typical biaxial potential (easy axis ẑ, hard axis x̂), as a function of transverse field
⊥
H⊥
o . If Ho is rotated away from x̂ by an angle φ (shown here in degrees), then |ϕ1 | = |ϕ2 |, and
1 = 2 , i.e., one path is favored over the other, and the oscillations are lost
266
G. Aeppli and P. Stamp
a
b
+ E(t)
ε(t)
t=0
t
γk
- E(t)
c
Vo Dipolar-Dominated
Regime
Li Ho x Y1-x F4
0.2 K
Mn-12
0.1 K
Quantum
ξ o Relaxation
Regime
Fe-8
1K
Fig. 2 (Continued)
Δo
Quantum
Coherence
Regime
γk
5 Quantum Magnetism
267
and still somewhat mysterious (witness the debate over the saturation of dephasing
times in mesoscopic conductors [17, 18]). Moreover, superpositions and entangled
states are extraordinarily sensitive to even very small environmental interactions.
To get a feeling for decoherence, let us go back to our toy tunneling spin, and now
couple it to a bath of “satellite” spins. In the real world, these satellite spins describe
localized modes (defects, nuclear and paramagnetic spins, etc.) which couple to the
central spin [19]. There are also delocalized environmental modes like electrons,
photons, phonons, etc., which also cause decoherence [21, 22], and which can be
described as a bath of oscillators [20]).
How do the satellite spins and oscillators dephase the “central” tunneling spin?
We explain this again using path integral language (Fig. 2). There are three main
decoherence mechanisms:
(i) Typically the environmental modes have their own dynamics, creating an
extra fluctuating “noise” field on the central system. This adds random phases
to each path, eventually destroying phase coherence between them (“noise
decoherence”). The noise can even push the central spin in and out of resonance
(Fig. 2a).
(ii) When the central spin tunnels, it causes a sudden “kick” perturbation on
the satellite spins, giving them an extra phase which is entangled with the
central spin phase – thence dephasing the central spin dynamics (“topological
decoherence” [23]).
(iii) The field on the k-th satellite, from the central spin, flips with the central
spin between two (in general noncollinear) orientations (Fig. 2b). Between
flips, the satellite precesses in these fields. Summing over all central spin
paths, each involving a different accumulated satellite precessional phase,
Fig. 2 The role of decoherence: In (a), we see the effect of a randomly fluctuating environmental
noise bias ε(t) (black curve) on a tunneling two-level qubit with tunneling matrix element o .
The two levels having adiabatic energies ±E(t), with E 2 (t) = 2o + ε2 (t), are shown as red
and blue curves. The system can only make transitions when near “resonance” (i.e., when |ε(t)|
is ∼ o or less, the regions shown in green). In (b), we show schematically the motion of a
satellite spin, in the presence of a qubit which is flipping between two different states | ↑ and
| ↓. When the qubit flips, the qubit field acting on the k-th satellite spin rapidly changes, from
↑
↓
γk to γk (or vice versa). Between flips, the spin precesses around the qubit field, accumulating
an extra “precessional” phase. Averaging over this phase gives precessional decoherence. The
sudden change of qubit field also perturbs the satellite spin phase, giving further decoherence
(the “topological decoherence” mechanism [23]). (c) shows the important parameters for a spin
network – the typical tunneling splitting o , the characteristic energy Vo of interspin interactions,
and the energy scale ξo governing interactions with the environment (which in insulating magnetic
systems at very low T comes from the coupling to nuclear spins). For definiteness, we show the
parameter range covered in this space by experiments in crystals of Mn-12 and F e-8 molecular
magnets, and in LiH ox Y1−x F4 ; energy scales are in temperature units. In these systems, o is
controlled by varying the transverse field, and Vo is varied by changing x (in LiH ox Y1−x F4 ), or
by diluting the molecules in solution (in F e-8 and Mn-12)
268
G. Aeppli and P. Stamp
gives “precessional” decoherence. In an oscillator bath, the central spin flip
slightly shifts the oscillator wave functions – for metallic environments, this
gives Anderson’s “orthogonality catastrophe” [24], a very strong decoherence
mechanism.
Notice that decoherence may involve very little energy transfer – it is not
necessarily a dissipative process. In magnetic systems, the worst low-T decoherence
will come from very low-energy localized modes, particularly nuclear spins,
which cause very little dissipation, but lots of precessional decoherence [19, 25].
Delocalized modes like phonons, photons, and electrons cause strong decoherence
(and strong dissipation) at higher energy, where they have a high density of states.
Thus, at intermediate energies (typically around 0.01 − 0.5 K), decoherence is at a
minimum. This “window” of low decoherence is of great practical importance – it
also exists for many other solid-state systems [26].
The basic problem with our toy model (2) for a quantum computer is now
clear – it ignores decoherence. If we couple each of the spins in the spin net to
an environment, there are now three main energy scales (Fig. 2c). A “quantum”
parameter o (the typical value of i ) drives the dynamics, along with interspin
interactions of typical strength Vo ; but a coupling of each spin to the environment,
having some effective energy scale ξo , destroys phase coherence. If we could switch
off ξo (i.e., stay in the Vo − o plane in Fig. 2c), we would have perfect quantum
behavior, with Vij correlating the entangled dynamics of vast numbers of qubits –
this would be true macroscopic quantum spin entanglement. But how close are we
to this goal?
Quantum Relaxation in Dipolar Nets
In fact most work has been done on systems with dipolar interspin couplings, having
non-negligible environmental interactions – i.e., near the Vo − ξo plane in Fig. 2c.
These systems are very complex – but a simple theoretical picture can be given.
Consider first a single spin qubit τi . The net bias i on τiz now includes a dynamic
contribution from the nuclear spin environment. This typically fluctuates over an
energy range ξo ∼ Eo , where Eo defines the energy width of the multiplet of nuclear
spin states coupled to S; this width is easy to calculate if the hyperfine couplings are
known. Then if i is within a “tunneling window” of width ξo around zero bias, the
fluctuating field can actually bring the qubit to resonance (recall Fig. 2a), where it
can make inelastic (i.e., incoherent) transitions [25].
Now consider an interacting network – assuming here for definiteness that go =
1 (the “dipolar-dominated” regime in Fig. 2c). As the resonant spins
o /Vo
tunnel, a “hole” of width ξo should appear in the distribution of longitudinal
fields
in the system, around zero (see Fig. 3a). The interaction contribution j Vij τj to i
then plays a key role – it slowly changes as the τj relax, bringing more spins into
the tunneling window (hole “refilling”). The total spin distribution is then predicted
5 Quantum Magnetism
a
269
[P (ξ,t) - P ( ξ,t)]
εo
Vo
2ξo
0
Fig. 3 (Continued)
ξ
270
G. Aeppli and P. Stamp
Fig. 3 Collective tunneling dynamics of dipolar nets, when o
ξo , Vo (incoherent tunneling
relaxation regime). In (a), we show the short-time evolution of the distribution function M(ξ, t) =
P↑ (ξ, t) − −P↓ (ξ, t), where Pσ (ξ, t) is the normalized probability that a spin in a state |σ = | ↑
or | ↓ finds itself in a bias field ξ at time t. Different colors show the distribution at different times.
At short times, a “tunneling hole” of width ξo appears, driven by inelastic tunneling transitions
involving nuclear spins. The dipolar interactions gradually modify the shape and width of the hole
at later times. This figure was produced by Monte Carlo simulations for a sample starting in a
strongly annealed state. (b) shows measurements of the function M(ξ = o , t = 0) on a strongly
annealed F e-8 crystal (from Ref. [30]), obtained by extracting the square root relaxation rate
−1
Γsqrt ≡ τQ
∼ (2o /Vo )ξo N (o ), where N (ξ ) is the “density of states” of spins in a bias energy ξ
(see text). If one lets M(ξ, t) relax for a time tW before examining it, the tunneling hole is revealed.
Closer examination (lower graph) shows that for small initial magnetization Min (i.e., strong
annealing), the hole has an intrinsic linewidth, revealed at short waiting times tW . This linewidth
ξo is caused by the nuclear spins (see text). (c) shows the tunneling matrix element extracted from
measurements of relaxation in a transverse field H⊥
o (from ref. [31]). These experiments found the
−1
oscillatory dependence of τQ
, and thence |o |, on H⊥
o (compare text, and Fig. 1c)
to relax, with a characteristic “square root” relaxation [27] ∼ (t/τQ )1/2 . One gets
−1
(o ) ∼ (2o /Vo )ξo N(o ), where N(ξ ) is the “density of states” of spins in a bias
τQ
energy ξ , i.e., ξo N (H ) is the number of spins in the tunneling window, centered at
the external field bias energy o = gμB SHoz .
Many experiments, using ensembles of magnetic nanomolecules such as F e-8
and Mn-12 (which truncate to two-level systems at low T ), have now tested this
theory. Square root relaxation was found at short times [28, 29, 30]. Wernsdorfer et
−1
al. [30] found the time-evolving hole, of width ξo , by measuring τQ
(ξ = o ) for
many different values of o (Fig. 3b). In strongly annealed samples (where M(ξ )
is a known Gaussian independent of sample shape), they also extracted 2o from
measurements of τQ , and showed how it oscillated in a transverse field (Fig. 3c),
and then confirmed this in independent “Landau-Zener” relaxation measurements.
We emphasize these oscillations are not evidence for coherent tunneling, quite the
5 Quantum Magnetism
271
contrary – the experiments observe incoherent relaxation rates! In a very striking
result [32], the nuclear isotopes were varied (substituting 2 H for 1 H , or 57 F e for
56 F e). This changed the hole width and the relaxation rate, giving independent
measurements of ξo which were consistent with the calculated value. This is fairly
direct evidence for the nuclear spin-mediated tunneling mechanism. The Leiden
group [33] has also done low-T NMR on Mn-12, seeing not only how nuclear
spins control the tunneling dynamics but also how the nuclear dynamics in turn is
controlled by the molecular tunneling dynamics.
Notice these are all results for short-time quantum relaxation. At longer times,
multi-spin correlations intervene, causing a breakdown of the square root – the
system moves into the quantum spin glass regime, of fundamental interest [35, 36].
Only quantum tunneling, simultaneously involving many spins, allows the system
to escape local potential minima. Experiments on the insulator LiH o0.44 Y0.56 F4 , in
which the lowest magnetic doublets (i.e., two-level systems) of J = 8 H o3+ ions
interact primarily via dipolar interactions, have looked at this tunneling relaxation
(Fig. 4a). Remarkably, much of the relaxation (here to a ferromagnetic ground state)
goes via collective dissipative tunneling of domain walls (see Fig. 4a); this purely
quantum effect has been definitively confirmed by observing its dependence on an
applied transverse field [37]. For these dense H o arrays, we can also reinterpret
the long-time relaxation as a quantum optimization process. One relaxes the system
toward the ground state not by reducing the temperature (as in thermal annealing
optimization protocols [38]), but by “quantum annealing” – exposing the system at
very low T to a large transverse field H⊥
o , allowing it to quantum relax, and then
reducing H⊥
o to zero, thence freezing the dynamics [39]. One then reads the final
state – which is the “solution” to the problem of energy optimization.
Finally, one can also explore the regime where ξo
Vo , i.e., where dipolar
interspin interactions are unimportant, and the nuclear environment dominates
completely. Here, a “toothcomb” structure was expected in the quantum relaxation
rate, reflecting the level structure of nuclear spins [25]. This was recently found
(Fig. 4b) in experiments [40] on dilute concentrations of H o ions in LiH ox Y1−x F4
(there was also an interesting catch – residual inter-H o interactions can cause pairs
of spins to co-flip, giving a doubling of the teeth).
We see that the study of the quantum relaxation of a spin net reveals the essential
physics governing the incoherent spin dynamics. Now we can turn to the coherent
spin dynamics.
Large-Scale Coherence and Entanglement
The acid test of our understanding comes with large-scale quantum effects – where
decoherence must be rigorously suppressed. Of course the traditional view has been
that too many microstates are involved in any macrostate for this to be possible [41].
However, the modern picture is different.
Macroscopic tunneling: In pioneering work, predictions of macroscopic tunneling between different flux states in superconducting SQUID rings [20] were
272
Fig. 4 (Continued)
G. Aeppli and P. Stamp
5 Quantum Magnetism
c
273
70
0.11 K
0.11 K
with pump
0.15 K
χ'' (emu/mole Ho)
60
50
40
30
20
10
0
1
10
100
1000
f (Hz)
Fig. 4 Tunneling dynamics of the LiH ox Y1−x F4 system. In (a), we show typical behavior for
the rate of microscopic domain wall depinning in LiH o0.44 Y0.56 F4 , as a function of inverse
temperature (from Ref. [37]); the crossover from thermal activation to T -independent tunneling
−1
relaxation occurs when T ∼ 50 mK. (b) shows the relaxation rate τQ
(H ) of the H o spins in a
−3
very dilute system (x = 2 × 10 ), from Ref. [40]. The main peaks in the “toothcomb” pattern,
each separated by the H o hyperfine energy, come from nuclear spin-mediated tunneling of single
H o ions between the lowest doublet states. The n = 8, 9 peaks shown in the inset come from coflip tunneling of pairs of H o ions, mediated by residual inter-H o interactions. (c) shows spectral
hole burning for x=0.045, from Ref. [57]. The absorptive part of the magnetic susceptibility is
measured as a function of frequency in the linear response regime using a probe amplitude of 0.04
Oe, giving a broad maximum centered at a frequency which depends strongly on temperature. The
same spectroscopy in the presence of a 0.2 Oe pump at 5 Hz shifts the spectrum, cuts off its tails,
and most important, inserts a sharp hole at 5 Hz.
quantitatively verified in the 1980s [42, 43]. In magnets, similar tunneling was
predicted for large domain walls pinned to defects [44], and also found in
experiments [45,46]. In experiments on large domain walls in mesoscopic Ni wires
(of thickness ∼20 − 80 nm), one sees a crossover to a T -independent escape rate
of the walls from a pinning center, in an applied field; the dependence of the rate
on field can be compared with theory [46]. This tunneling involved roughly 107
spins – not far short of the number of Cooper pairs involved in SQUID tunneling.
A revealing set of experiments [46] also looked at microwave absorption between
different levels – these represented the quantized dynamics of the wall center of
mass, trapped in the pinning potential well. A fairly detailed picture can be given of
these experiments [47, 48].
Such tunneling phenomena involve a collective degree of freedom (SQUID flux,
magnetic domain wall position) which does indeed involve a huge number of
274
G. Aeppli and P. Stamp
microstates. So how can it happen? One reason is that all electronic microstates
(Bogoliubov quasiparticles, magnons) are strongly gapped, by energies EG ∼
10 K. To such high-energy excitations, the collective coordinate tunneling, over a
timescale τo , seems very slow. The amplitude to excite them is then exponentially
small, ∼ O(e−EG τo ), by elementary time-dependent perturbation theory. Of course
there are also many very low-energy excitations (defects, paramagnetic impurities,
nuclear spins, etc. – the “satellite spins” discussed in section “The Importance
of Decoherence”), which can entangle with the collective tunneling coordinate.
However, they cause little dissipation, because their energy is so low – their direct
effect on “single-shot” tunneling is then rather weak.
Large-scale coherence: Coherent superpositions, on the other hand, require phase
coherence between many successive tunneling events [21,22]. Now, the low-energy
environmental microstates are indeed very dangerous [19]. So is macroscopic state
superposition feasible?
In superconductors, the answer to this question came a considerable time ago,
including early experimental evidence for macroscopic flux state superpositions
[49, 50, 51]. The decoherence times τφ for superconducting qubits have undergone
a spectacular rise since the year 2000 so that today, they are viewed as leading
candidates for the fundamental building blocks of quantum computers.
Analogous macroscopic superpositions in magnets - e.g., of “giant qubit”
superpositions of two different magnetization states in a large magnetic particle –
have not yet been confirmed. Some years ago, absorption experiments in very large
ferritin molecules (with Neél vector ∼ 23, 000 μB ) showed sharp resonances,
attributed to coherent tunneling of individual ferritin molecules [53]. However, these
results were hard to understand theoretically, and no other group has confirmed
them; and experiments on much smaller molecules like Mn-12 or F e-8 have
never seen coherence. The basic difficulty is that low-energy decoherence from the
hyperfine coupling to nuclear spins is expected to be large (in ferritin, the hyperfine
coupling to a single 57 F e nucleus is much larger than the tunnel splitting!), and at
higher energies, phonon contributions are not negligible [25, 19].
However, experiments may have simply been looking in the wrong place. The
“window of opportunity” between nuclear spin and phonon energy scales is actually
very wide; by applying strong transverse fields, one can increase o to values
much higher than hyperfine couplings (compare Fig. 1c), but still much lower than
most phonon energies, and optimize the decoherence rate. By combining isotopic
purification with choice of material, one can also remove almost all nuclear spins.
Experimentalists like to define a decoherence Q-factor Qφ = o τφ , which tells us
how many coherent oscillations the system can show before decoherence sets in.
Elementary theory [25, 52] then indicates that for an spin S in an insulator, we have
2 /SK E , where K is the anisotropy energy per electronic spin
an optimal Qφ ∼ θD
o o
o
of the magnet, θD the Debye energy, and Eo is again the spreading of the nuclear
multiplet. For example, if θD = 300 K, Ko = 1 K, and S = 106 , then by reducing
0.1 K, we should get “mesoscopic” coherent dynamics (i.e., Qφ > 1).
Eo to
Large-scale entanglement: We next turn to multi-qubit entangled states, but now
involving microscopic spins. It is estimated that quantum information processing
5 Quantum Magnetism
275
Fig. 5 Design for a nuclear
spin-based quantum
computer, from Kane [54].
Two cells in a
one-dimensional array,
containing 31P donors and
electrons in a Si host, are
separated by a barrier from
metal gates on the surface.
The “A gates” control the
resonance frequency of the
nuclear spin qubits, while the
“J gates” control the
electron-mediated coupling
between adjacent nuclear
spins
(QUIP) will be possible, using error correction [14,16], if the single qubit coherence
Q-factor Qφ > 104 − 105 . This should easily be possible with microscopic spins –
note from above that the optimal Qφ ∼ O(1/S). Thus, theory clearly indicates
that QUIP is feasible with microscopic magnetic qubits, provided electronic decoherence is absent (e.g., magnetic ions in insulators or semiconductors, or perhaps
insulating molecular crystals). Many proposals have appeared along these lines – as
an example, consider that due to Kane [54] in Fig. 5, using networks of nuclear spins
in semiconductors to do the computation. Reasonable estimates of decoherence rates
[54] give very small numbers here – problems should only arise, as before, from
very low-energy excitations (e.g., 1/f noise from charge defects). Again, applying
strong fields should help [55], and measurements of the decoherence rates will be
crucial, in this and other designs [56].
Experiments over the last years have given very long spin relaxation times for the
spins associated with isolated impurities and quantum dots in silicon. While this is
very interesting for quantum information science, the finding of sharp resonances in
a strongly interacting many-body system using spectral hole-burning [57, 64, 69]
in LiH o0.045 Y0.995 F4 (see Fig. 4c) is important for the science of disordered
magnetism. The decoherence was remarkably small, in spite of the long-range interH o dipole interactions. The data were explained as a collective effect, involving
tunneling of large clusters of H o spins. These results are both surprising and
exciting – they indicate that we may be close to manipulating entangled mesoscopic
spin states.
To do fully fledged quantum computations will require controlling individual
spins or groups of spins, i.e., control of the parameters i , i , and Vij . Control of i
and i can obviously be done by varying transverse and longitudinal external fields –
in fact, all quantum logic operations can be implemented by varying only one of
these three parameters, and one can also use timed pulses in creative ways (which
also help with decoherence [58,59]), so control of Vij is not crucial. One possibility
276
G. Aeppli and P. Stamp
is to use engineered heterostructures to control the local fields [54]; another would
be to use magnetic STM tips, although the practicality of this for anything other than
demonstration experiments involving a very small number of qubits is questionable.
A more difficult problem will be to measure the quantum state of the spin qubits,
without affecting their operation. One can imagine many possibilities – for example,
bringing in superconducting or magnetic devices, whose tunneling into the qubit
depends on its polarization [60], or using optical methods. Over the last two decades,
great progress has been made on both adiabatic and gated quantum computation.
For solid-state implementations, the leading contenders have been superconducting
qubits, whose decoherence times have dramatically improved, and for which very
complex circuits can be constructed. It is beyond the scope of this chapter on
magnetism to review these developments on magnetism, except to mention that
the quantum annealer manufactured by D-wave systems [62] is built to simulate
the transverse field Ising model, and can therefore be viewed as a programmable
version of LiH o1−x Yx F4 . Equally relevant here and for the future of magnetism are
experiments showing control of the magnetic interactions between the very simple
S=1/2 spins associated with either donors [34] or quantum dots [67] in silicon.
Future Directions and Open Problems
When many of the authors of this volume began graduate work in the 1980s, the
idea of large-scale quantum phenomena in magnetic systems was hardly a topic for
discussion – for ∼70 years, quantum mechanics was only used in magnetism to
discuss atomic and nuclear spins and the microscopic interactions (exchange, spinorbit, etc.) operating on them. Now we are discussing and even seeing quantum
phenomena at much larger scales, where magnetic variables were previously treated
as classical – it is in this sense that the “quantum” is being put back into magnetism.
We are on the threshold of a very different era – in which coherent spin states, having
no classical analogue, may come to play a role as important as the macroscopic wave
function in superconductors.
As always, it is difficult to make predictions about a fast-moving field. The
preparation and readout of coherent multi-spin states may well require techniques
from spintronics [61], implemented on submicron scales. The key challenges
here will be (i) to marry spintronics with the science of collective quantum spin
states, under progressively less extreme experimental conditions of magnetic field
and temperature, and (ii) to understand how to suppress electronic decoherence
in conducting magnets. This latter is a hard problem because spin current is not
a conserved quantity, which has made it difficult to find a rigorous theory of
spin dynamics in conducting magnets, although moving to antiferromagnets where
mesoscopic quantum tunneling of domain walls has been identified (in the common
metal chromium) [68] may be a promising route. The development of new materials
having the correct mix of optical, electronic, and magnetic properties will be crucial,
and theory will need to be developed to model candidate materials and devices.
It is sobering that even for an insulator as simple as LiH ox Y1−x F4 , many key
5 Quantum Magnetism
277
discoveries – e.g., the coherent hole burning for x=0.045 – were unexpected, and
dictated by such factors as the availability of samples with particular compositions
at sale prices.
We also note that there are many interesting spin systems apart from electronic
magnets. Magnetic superfluids like 3 H e (see, e.g., [63]), or spin-1 Bose-Einstein
condensates (BECs) of alkali atoms [65, 66], and quantum Hall spin condensates,
offer examples where a spin coherent wave function coupled to another order
parameter can display very rich dynamics, and where large-scale quantum effects
are worth exploring.
Finally, we emphasize a crucial change of perspective in the field. Condensed
matter physicists are accustomed to dealing with only one- and two-spin correlation
functions, whereas quantum information theory requires manipulation and measurement of multi-qubit correlations. We still have a great deal to learn about these, and
the next decades promise to be a very exciting one for quantum magnetism.
References
1. London. F.: The λ-phenomenon of liquid H e and the Bose-Einstein degeneracy. Nature 141,
643 (1938)
2. Feynman, R.P.: Spacetime approach to non-relativistic quantum mechanics. Rev. Mod. Phys.
20, 367 (1948)
3. Feynman, R.P.: The Feynman Lectures on Physics, vol. 3. Addison-Wesley (1965)
4. Shapere. A., Wilczek. F.: Geometric Phases in Physics. World Scientific (1989)
5. Klauder, J.R.: Path integrals and stationary phase approximations. Phys. Rev. D19, 2349 (1979)
6. Berry, M.V.: Quantal Phase factors accompanying adiabatic changes. Proc. Roy. Soc. A392, 45
(1984)
7. Auerbach. A.: Interacting Electrons and Quantum Magnetism. Springer (1994)
8. Aharonov. Y, Bohm. D.: Significance of electromagnetic potentials. Phys. Rev. 115, 485 (1959)
9. Haldane, F.D.M.: Non-linear field thery of large-spin Heisenberg antiferromagnets: semiclassically quantized solitons of the 1-d easy-axis Néel state. Phys. Rev. Lett. 50, 1153 (1983)
10. Bogachek, E.N., Krive, I.V.: Quantum oscillations in small magnetic particles. Phys. Rev. B46,
14559 (1992)
11. Garg. A.: Topologically Quenched Tunnel Splitting in Spin Systems without Kramers’
Degeneracy. Europhys. Lett. 22, 205 (1993)
12. Tupitsyn, I.S., Barbara, B.: Quantum tunneling of Magnetisation in molecular complexes with
large spins: effect of the Environment. In: Miller, J.S., Drillon, M. (eds.) Magnetoscience- from
Molecules to Materials, vol. 3, pp. 109–168. Wiley (2001)
13. Kramers. H.A.: Théorie générale de la rotation paramagnétique dans les cristaux. Proc. Acad.
Sci. Amst. 33, 959 (1930)
14. Bennett, C.H., DiVincenzo, D.: Quantum information and computation. Nature 404, 247
(2000)
15. Lo, H.-K., Popescu, S., Spiller, T.: Introduction to Quantum Computation and Information.
World Scientific (1998)
16. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge
University Press (2000)
17. Mohanty, P., Jariwala, E.M.Q., Webb, R.A.: Intrinsic decoherence in mesocopic systems. Phys.
Rev. Lett. 78, 3366 (1997)
18. Aleiner, I.L., Altshuler, B.L., Gershenson, M.E.: Interaction effects and phase relaxation in
disordered systems. Waves Random Media 9, 201 (1999)
278
G. Aeppli and P. Stamp
19. Prokof’ev, N.V., Stamp, P.C.E.: Theory of the spin bath. Rep. Prog. 63, 669 (2000)
20. Caldeira, A.O., Leggett, A.J.: Quantum tunneling in a dissipative system. Ann. Phys. 149, 374
(1984)
21. Leggett, A.J., Chakravarty. S, Dorsey, A.T., Fisher, M.P.A., Garg, A., Zwerger, W.: Dynamics
of the dissipative 2-state system. Rev. Mod. Phys. 59, 1 (1987)
22. Weiss, U.: Quantum Dissipative Systems. World Scientific (1999)
23. Prokof’ev, N.V., Stamp, P.C.E.: Giant spins and topological decoherence: a Hamiltonian
approach. J. Phys. CM 5, L663 (1993)
24. Anderson, P.W.: Infrared catastrophe in fermi gases with local scattering potentials. Phys. Rev.
Lett. 18, 1049 (1967).
25. Prokof’ev, N.V., Stamp, P.C.E.: Quantum relaxation of magnetisation in magnetic Particles. J.
Low Temp. Phys. 104, 143 (1996)
26. Dubé, M., Stamp, P.C.E.: Mechanisms of decoherence at low Temperatures. Chem. Phys. 268,
257 (2001)
27. Prokof’ev, N.V., Stamp, P.C.E.: Low-T relaxation in a system of Magnetic molecules. Phys.
Rev. Lett. 80, 5794 (1998)
28. Ohm, T., et al.: Local field dynamics in a resonant quantum tunneling system of magnetic
molecules. Europhys. J. B6, 595 (1998)
29. Thomas, L., et al.: Non-exponential scaling of the magnetisation relaxation in Mn12 acetate.
Phys. Rev. Lett. 83, 2398 (1999)
30. Wernsdorfer, W., Ohm, T., Sangregorio, C., Sessoli, R., Mailly, D., Paulsen, C.: Observation of
the distribution of molecular spin states by resonant quantum tunneling of the magnetisation.
Phys. Rev. Lett. 82, 3903 (1999)
31. Wernsdorfer, W., Sessoli, R.: (1999) Quantum phase interference and parity effects in Magnetic
molecular clusters. Science 284, 133
32. Wernsdorfer, W., Caneschi, A., Sessoli, R., Gatteschi, D., Cornia, A., Paulsen, C.: Effect of
nuclear spins on the quantum relaxation of the magnetisation for the molecular magnet F e-8.
Phys. Rev. Lett. 84, 2965 (2000)
33. Morello, A., et al.: Quantum tunneling of magnetisation in Mn12-ac studied by Mn-55 NMR.
Polyhedron 22, 1745–1749 (2003)
34. Veldhorst, M., Yang, C.H., Hwang, J.C.C., Huang, W., Dehollain, J.P., Muhonen, J.T.,
Simmons, S., Laucht, A., Hudson, F.E., Itoh, K.M., Morello, A., Dzurak, A.S.: A two-qubit
logic gate in silicon. Nature 526, 410(2015)
35. Sachdev, S.: Quantum Phase Transitions. Ch. 16. Cambridge University Press (1999)
36. Thill, M.J., Huse, D.A.: Equilibrium behaviour of quantum Ising spin glass. Physica 214A, 321
(1995)
37. Brooke, J., Rosenbaum, T.F., Aeppli, G.: Tunable quantum tunneling of magnetic domain
walls. Nature 413, 610 (2001)
38. Kirkpatrick, S., Gelatt, C.D., Vecchi, M.P.: Optimisation by simulated annealing. Science 220,
671 (1983)
39. Brooke, J., Bitko, D., Rosenbaum, T.F., Aeppli, G.: Quantum annealing of a disordered magnet.
Science 284, 779 (1999)
40. Giraud, R., Wernsdorfer, W., Tkachuk, A.M., Mailly, D., Barbara, B.: Nuclear spin driven
quantum relaxation in LiY0.998 H o0.002 F4 . Phys. Rev. Lett. 87, 057203 (2001)
41. van Kampen, N.G.: Ten Theorems about Quantum Mechanical Measurements. Physica A153,
97 (1988)
42. Washburn, S., Webb, R.A., Voss, R.F., Faris, S.M.: Effects of dissipation and temperature on
macroscopic quantum tunneling. Phys. Rev. Lett. 54, 2712 (1985)
43. Clarke, J., Cleland, A.N., Devoret, M., Esteve, D., Martinis, J.M.: Quantum mechanics of
a macroscopic variable: the phase difference of a Josephson junction. Science 239, 992
(1998)
44. Stamp, P.C.E.: Quantum dynamics and tunneling of domain walls in ferromagnetic insulators.
Phys. Rev. Lett. 66, 2802 (1991)
45. Paulsen, C., et al.: Macroscopic Quantum Tunnelling Effects of Bloch Walls in Small
Ferromagnetic Particles. Europhys. Lett. 19, 643 (1992)
5 Quantum Magnetism
279
46. Hong, K., Giordano, N.: Effect of microwaves on domain wall motion in this Ni wires.
Europhys. Lett. 36, 147 (1996)
47. Tatara, G., Fukuyama, H.: Macroscopic quantum tunneling of a domain wall in a ferromagnetic
metal. Phys. Rev. Lett. 72, 772 (1994)
48. Dubé, M., Stamp, P.C.E.: Effect of Phonons and Nuclear Spins on the tunneling of a Domain
wall. J. Low Temp. Phys. 110, 779 (1998)
49. Van der Wal, C.H., ter Haar, A.J.C., Wilhelm, F.K., Shouten, R.N.: Harmans, C.J.P.M.,
Orlando, T.P., Lloyd, S., Mooij, J.E.: Quantum Superposition of Macroscopic PersistentCurrent states. Science 290, 773 (2000)
50. Vion, D., Aasime, A., Cottet, A., Joyez, P., Pothier, H., Urbina, C., Esteve, D., Devoret, M.:
Manipulating the Quantum State of an Electrical Circuit. Science 296, 886 (2002)
51. Chiorescu, I., Nakamura, Y., Harmans, C.J.P.M., Mooij, J.E.: Coherent quantum dynamics of
a superconducting flux qubit. Science 299, 1869 (2003)
52. Stamp, P.C.E., Tupitsyn, I.S.: Coherence window in the dynamics of quantum nanomagnets.
Phys. Rev. B69, 014401 (2004)
53. Gider, S., et al.: Classical and quantum magnetic phenomena in natural and artificial ferritin
proteins. Science 268, 77 (1995)
54. Kane, B.L.: A Si-based nuclear spin quantum computer. Nature 393, 133 (1998)
55. Privman, V., Vagner, I.D., Kventsel, G.: Quantum computation in quantum-Hall systems Phys.
Lett. A236, 141 (1998)
56. Barrett, S.D., Milburn, G.J.: Measuring the decoherence rate in a semiconductor charge qubit.
Phys. Rev. B68, 155307 (2003)
57. Ghosh, S., Parthasarathy, R., Rosenbaum, T.F., Aeppli, G.: Coherent spin oscillations in a
disordered magnet. Science 296, 2195 (2002)
58. Viola, L., Lloyd, S.: Dynamical supression of decoherence in 2-state quantum systems. Phys.
Rev. A58, 2733 (1998)
59. Wu, L.A., Lidar, D.A.: Creating decoherence free subspaces using strong and fast pulses. Phys.
Rev. Lett. 88, 207902 (2002)
60. Manoharan, H.C., Lutz, C.P., Eigler, D.M.: Quantum Mirages formed by coherent projection
of electronic structure. Nature 403, 512 (2000)
61. Wolf, S.A., Awschalom, D.D., Buhrman, R.A., Daughton, J.M., von Molnar, S., Roukes, M.L.,
Chtchelkanova, A.Y., Treger, D.M.: Spintronics: a spin-based electronics vision for the future.
Science 294, 1488 (2001)
62. Johnson, M.W., Amin, M.H.S, Gildert, S., Lanting, T., Hamze, F., Dickson, N., Harris, R.,
Berkley, A.J., Johansson, J., Bunyk, P., Chapple, E.M., Enderud, C., Hilton, J.P., Karimi, K.,
Ladizinsky, E., Ladizinsky, N., Oh, T., Perminov, I., Rich, C., Thom, M.C., Tolkacheva, E.,
Truncik, C.J.S., Uchaikin, S., Wang, J., Wilson, B., Rose, G.: Quantum annealing with
manufactured spins. Nature 473 194 (2011)
63. Vollhardt, D., Wölfle, P.: The superfluid phases of Helium 3. Taylor & Francis (1990)
64. Schmidt, M.A., Silevitch, D.M., Aeppli, G., Rosenbaum, T.F.: Using thermal boundary
conditions to engineer the quantum state of a bulk ferromagnet. PNAS 111, 3689 (2014)
65. Stamper-Kurn, D.M., Miesner, H.J., Chikkatur, A.P., Inouye, S., Stenger, J., Ketterle, W.:
Quantum tunneling across spin domains in a BEC. Phys.Rev. Lett. 83, 661 (1999)
66. Ketterle, W.: Spinor condenstates and light scattering from Bose-Einstein condensates. Proc.
Les Houches summer school LXXII (1999)
67. Watson, T.F., Philips, S.G.J., Kawakami, E., Ward, D.R., Scarlino, P., Veldhorst, M., Savage,
D.E., Lagally, M.G., Mark Friesen, Coppersmith, S.N., Eriksson, M.A., Vandersypen, L.M.K.:
A programmable two-qubit quantum processor in silicon. Nature 555, 263 (2018)
68. Shpyrko, O., Isaacs, E., Logan, J., Yejun Feng, Aeppli. G, Jaramillo, R., Kim, H.C.,
Rosenbaum, T.F., Zschack, P., Sprung, M., Narayanan, S., Sandy, A.R.: Direct measurement
of antiferromagnetic domain fluctuations. Nature 447, 68–71 (2007). https://doi.org/10.1038/
nature05776
69. Silevitch, D.M., Tang, C., Aeppli, G., Rosenbaum, T.F.: Tuning high-Q nonlinear dynamics in a
disordered quantum magnet. Nat. Commun. 10, 4001 (2019). https://doi.org/10.1038/s41467019-11985-1
280
G. Aeppli and P. Stamp
Gabriel Aeppli is professor at ETHZ and EPFL, and head of
Photon Science at PSI. After working at NEC, AT&T, IBM and
MIT on problems from liquid crystals to magnetic data storage,
he co-founded the London Centre for Nanotechnology and BioNano Consulting. His focus is on implications and development
of photon science and nanotechnology for information processing
and health care.
Philip Stamp received his PhD from the Univ of Sussex. After
postdoctoral work in Massachusetts, Grenoble, and Santa Barbara, he held positions in the Univ of British Columbia, and as
a Spinoza Professor in Utrecht. He is currently director of the
Pacific Institute of Theoretical Physics in Vancouver. He works
on theoretical quantum gravity and condensed matter theory.
6
Spin Waves
Sergej O. Demokritov and Andrei N. Slavin
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin Waves in 3D and 2D Systems: Theory and Experiment . . . . . . . . . . . . . . . . . . . . . . . . . .
Theory of Spin Waves in 3D and 2D Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Brillouin Light Scattering a Powerful Tool for Investigation of Spin Waves . . . . . . . . . . . .
Spin Waves in 1D Magnetic Elements: Standing and Propagating Waves . . . . . . . . . . . . . . . .
BLS in Laterally Confined Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lateral Quantization of Spin Waves in Magnetic Stripes . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin Wave Wells and Edge Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Implementation of Micro-Focus BLS for Laterally Patterned Magnetic Systems . . . . . . . .
Propagating Waves in 1D Magnetic Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Control and Conversion of the Propagating Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Inductive Excitation of Spin Waves in 1D Waveguides . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin-Torque Transfer Effect and Spin Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin Waves in 0D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin-Torque Nano-Oscillator (STNO) and Emitted Spin Waves . . . . . . . . . . . . . . . . . . . . .
Spin-Hall Nano-Oscillator (SHNO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Nature of Spin Wave Modes Excited in 0D Magnetic Nanocontacts . . . . . . . . . . . . . . . . . .
Coupling of a STNO and 1D Spin-Wave Waveguide to Each Other . . . . . . . . . . . . . . . . . .
Conclusion and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
282
284
284
288
291
291
291
296
300
302
307
311
315
320
321
324
329
336
339
341
S. O. Demokritov ()
Institute for Applied Physics and Center for Nanotechnology, University of Muenster, Muenster,
Germany
e-mail: [email protected]
A. N. Slavin
Department of Physics, Oakland University, Rochester, MI, USA
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_6
281
282
S. O. Demokritov and A. N. Slavin
Abstract
Spin waves are the dynamic eigen-excitations of a magnetic system. They
provide the basis for the description of spatial and temporal evolution of the
magnetization distribution of a magnetic object. The unique features of spin
waves such as the possibility to carry spin information over relatively long
distances, the possibility to achieve submicrometer wavelength at microwave
frequencies, and controllability by electronic signal via magnetic fields make
these waves uniquely suited for implementation of novel integrated electronic
devices characterized by high speed, low power consumption, and extended
functionalities. The history of spin waves clearly shows a progressively increasing interest for the spin waves in magnetic samples of reduced dimensionality.
Since 1950s the focus of the researchers has moved from 3D to 2D objects –
thin films and magnetic multilayers, resulting in the discovery of the surface
Damon-Eshbach spin-wave mode. Later in 1990s 1D stripes became the most
actively studied magnetic systems, bringing about the discovery of lateral
quantization and edge modes. Finally, theoretical prediction of the spin-torque
effect and development of novel techniques for nanofabrication allowed for the
investigation of magnetic 0D objects such as spin-torque nano-oscillator. In this
chapter we follow this historical trend and describe the recent development of
spin-wave studies.
Introduction
Felix Bloch introduced the concept of spin waves (SPW), as the lowest-energy magnetic states above the ground state of a magnetic medium [1]. Bloch theoretically
considered quantum states of magnetic systems with spins slightly deviating from
their equilibrium orientations, and found that these disturbances were dynamic: they
propagate as waves through the medium. It is interesting to note that the concept
of dynamic spin waves was introduced to explain experimental data obtained
in static measurements. From spin-wave theory Bloch was able to predict that
the magnetization of a three-dimensional (3D) ferromagnet at low temperatures
should deviate from its zero-temperature value with a T3/2 dependence (the famous
Bloch law), instead of the exponential dependence given by the mean field theory.
Albeit the Bloch theory has restricted itself by assuming the dominating exchange
interaction, now we know that the relativistic magnetic dipole interaction plays
a decisive role in the properties of spin waves having the wavelengths that are
much smaller than the interatomic distance in the magnetic medium. Moreover, phenomenologically, spin waves in a wide interval of wavevectors (30 < k < 106 cm−1 )
that is most important for the practical applications are, on one hand, almost entirely
determined by the magnetic dipole-dipole interaction, and, on the other hand, can
be correctly described when the retardation effects are neglected. Such spin waves
are usually called dipolar magnetostatic waves or magnetostatic modes. Due to the
anisotropic properties of the magnetic dipole interaction, the frequency of a spin
6 Spin Waves
283
wave depends on the orientation of its wavevector relative to the orientation of
the static magnetization. For higher values of the wavevectors, when the exchange
interaction cannot be neglected, one speaks about dipole-exchange spin waves.
Spin waves are the dynamic eigen-excitations of a magnetic system. They
provide the basis for the description of spatial and temporal evolution of the
magnetization distribution of a magnetic object under the general assumption that
locally the length of the magnetization vector is constant. This is correct, if, first,
the temperature is far below the Curie temperature of the medium, as is assumed
throughout this chapter, and, second, if no topological anomalies such as vortices or
domain walls are present. The latter is fulfilled for samples in a single-domain state,
i.e., magnetized to saturation by an external bias magnetic field.
The unique features of spin waves such as the possibility to carry spin information over relatively long distances, the possibility to achieve submicrometer
wavelength at microwave frequencies, and controllability by electronic signal via
magnetic fields make these waves uniquely suited for implementation of novel
integrated electronic devices characterized by high speed, low power consumption,
and extended functionalities. The utilization of spin waves for integrated electronic
applications is addressed within the emerging field of magnonics [2–5]. Although
the application of spin waves for microwave signal processing has been intensively
explored since many decades, recent advances in spintronics and nanomagnetism, as
well as the development of novel techniques for nanofabrication and measurements
of high-frequency magnetization dynamics created essentially new possibilities for
magnonics and brought it onto a new development stage. Of particular importance
here is the recent discovery of the spin-transfer torque (STT) [6–8] and the spin-Hall
effect (SHE) [9–11], both of which have already been demonstrated to enable novel
device geometries and functionalities [12–16].
The history of spin waves clearly shows a progressively increasing interest
to the spin waves in magnetic samples of reduced dimensionality. Although the
original theory of Bloch was developed for 3D magnets, since 1950s the focus
of the researchers has moved to two-dimensional (2D) objects – thin films and
magnetic multilayers, resulting in the discovery of quantized standing spin-wave
resonances [17, 18], surface Damon-Eshbach spin-wave mode [19, 20], and the
interlayer coupling [21]. Later in 1990s quasi-one-dimensional (1D) stripes became
the most actively studied magnetic systems, bringing about the discovery of
lateral quantization [22] and edge modes [23, 24]. Finally, as mentioned above,
theoretical prediction of the spin-torque effect (STE) [6, 7] and development of
novel techniques for nanofabrication allowed for investigation of magnetic zerodimensional (0D) objects such as spin-torque nano-oscillator (STNO) [25–31]. In
this chapter we follow this historical trend. In the Sect. “Spin Waves in 3D and 2D
Systems: Theory and Experiment” we describe the theory of spin waves in 3D and
2D magnetic systems, as well as the main experimental techniques for spin-wave
studies in such systems. In Sect. “Spin Waves in 1D Magnetic Elements: Standing
and Propagating Waves” we focus on spin waves in quasi-1D systems: laterally
quantized and localized spin-wave edge modes in magnetic stripes. We also discuss
the propagating wave modes in magnetic waveguides. The experimental data are
284
S. O. Demokritov and A. N. Slavin
complemented by a short theoretical description of the spin-torque effect and its
role in the damping compensation for propagating spin waves. In Sect. “Spin Waves
in 0D” we describe the spin-wave dynamics of 0D systems on the example of an inplane magnetized STNO structure supporting a self-localized solitonic spin-wave
“bullet” mode. Conclusions are given in Sect. “Conclusion and Outlook”.
Spin Waves in 3D and 2D Systems: Theory and Experiment
This section comprises two subsections. Subsection “Theory of Spin Waves in 3D
and 2D Systems” is devoted to the general theory of spin waves in both 3D and
2D systems. It also demonstrates the modification of the spin-wave properties due
to the dimensionality reduction. Subsection “Brillouin Light Scattering a Powerful
Tool for Investigation of Spin Waves” describes the Brillouin light scattering – the
main experimental techniques for the investigation of spin waves. Since spin waves
in 3D and 2D systems have been intensively studied in the past, the sections devoted
to their description are short, and the main attention is paid to the spin waves in the
quasi-1D and 0D geometries.
Theory of Spin Waves in 3D and 2D Systems
The dynamics of the magnetization vector is described by the Landau-Lifshitz
torque equation [32]:
1 dM
= M × H eff
−
(1)
γ dt
where M = MS + m(R, t) is the total magnetization, MS and m(R , t) are the vectors
of the saturation and the variable magnetization correspondingly, γ is the modulus
of the gyromagnetic ratio for the electron spin (γ/2π = 2.8 MHz/Oe), and
H eff = −
δW
δM
(1a)
is the effective magnetic field calculated as a variational derivative of the energy
function W, where all the relevant interactions in the magnetic substance have
been taken into account (see, e.g., [33–35]). For the case of an unbounded 3D
ferromagnetic medium the variable magnetization m(R, t) depends on time t and
on the three-dimensional radius-vector R. In the spin-wave analysis it is usually
assumed that the variable magnetization m(R, t) is small compared to the saturation
magnetization MS , i.e., the angle of magnetization precession is small. In this case
the variable magnetization can be expanded in a series of plane spin waves (having
a 3D wavevector q):
m (R, t) =
q
mq exp (iqR) .
(2)
6 Spin Waves
285
The spectrum of dipole-exchange spin waves is in an unbounded ferromagnetic
medium which is given by the Herrings-Kittel formula [36].
ω = 2πf = γ
1/2
2A 2
2A 2
H+
H+
q
q + 4π Ms sin2 θq
Ms
Ms
(3)
where A is the exchange stiffness constant, H is the applied magnetic field, and θ q
is the angle between the directions of the wavevector q and the static magnetization
MS with sin2 θ q being the matrix element of the dipole-dipole (magnetostatic)
interaction. Analyzing Eq. (3), one concludes that if the exchange can be neglected
q2 < < HMS /2A and q 2 << 2π MS2 /A, the spin-wave frequency is independent of q.
It depends solely on θ q demonstrating that the nonexchange (magnetostatic) spinwave spectrum is anisotropic and nondispersive. In contrast, for q2 > > HMS /2A
and q 2 >> 2π MS2 /A the spectrum of purely exchange spin waves is isotropic since
the spin wave frequency solely depends on q = |q|.
The transition from 3D to 2D can be made if one considers a magnetic film
with a finite thickness d. In the following, we assume a Cartesian coordinate system
oriented in such a way that the film normal is along the x-axis, and axes y and z are
in the film plane, with the external field H and the static magnetization MS being
aligned along the z-axis. Correspondingly, because the translational invariance along
the direction normal to the film surfaces (axis x) is broken, the three-dimensional
spin wave wavevector is represented as a sum of a two-dimensional continuous inplane wavevector q and quantized wavevector κ p ex (p = 0,1,2 . . . ) along the film
thickness: q = q + κ p ex , while the three-dimensional radius vector is represented
as R = R + x ex . Then, the distribution of the variable magnetization along the
film thickness (axis x) can be represented as a Fourier series in a complete set of
orthogonal functions mp (x) [37]:
m (R, t) =
mp (x) exp iq R .
(2a)
q ,p
These functions mp (x) are chosen in such a way that they satisfy the exchange
differential operator of the second order and the exchange boundary conditions at
the film boundaries [38]:
∂m
∂x
+ D m|x=±d/2 = 0,
(4)
x=±d/2
where D is the so-called “pinning” parameter determined by the ratio of the
effective surface anisotropy ks and the exchange stiffness constant A: D = ks /A.
The modes mp (x) can be interpreted as the modes of the spin-wave resonance
[17] in a particular geometry, and are sometimes called perpendicular standing
spin waves (PSSW). The discrete transverse wavenumbers κ p for these modes
are determined from the eigenvalue problem for the exchange differential operator
286
S. O. Demokritov and A. N. Slavin
with the boundary conditions Eq. (4). Note that in an finite-in-plane nonellipsoidal
magnetic film samples, the “pinning” of the dynamic magnetization at the lateral
edges of the sample could be determined by the local inhomogeneity of internal
dynamic magnetic field, and a different set of the “in-plane” eigenfunctions for the
expansion of the in-plane components of the variable magnetization can be obtained
in that case [39, 40].
Assuming that the thickness spin-wave modes mp (x) do not hybridize, it is
possible to obtain an approximate “diagonal” dispersion equation for the spin-wave
modes in a magnetic film of a finite thickness that is similar to the classical Kittel
equation (3) [37]:
2A 2
2
ωp = 2πfp = γ H +
q + κp
Ms 1/2
2A 2
2
H+
q + κp + 4π Ms Fpp κp , q , d
,
Ms (5)
where Fpp (κ p , q , d) is the matrix element of the dipole-dipole interaction in a film
defined by Eq. (46) in [37].
In the case of “unpinned” spins at the film surfaces
∂m
∂x
= 0,
(6)
x=±d/2
it is possible to obtain a simple explicit expression for the transverse spin-wave
wavenumber κ p = pπ /d, p = 0, 1, 2, . . . , and the expression for the matrix
element Fpp (κ p , q , d) for an arbitrary angle between q|| and MS can be written
in the form:
Fpp = 1 + Ppp (q) 1 − Ppp (q)
4π MS
H + (2A/MS ) q 2
qy2
q2
q2
− Ppp (q) z2
q
,
(7)
where
q 2 = qy2 + qz2 +
pπ
d
2
= q2 +
pπ
d
2
,
(8)
and the function Ppp (q , p) is defined in [37].
We present here only the expression for this function for the lowest (quasiuniform) thickness mode (p = 0) [37]:
P00
1 − exp −q d
.
= P00 q d = 1 +
q d
(9)
6 Spin Waves
287
If the lowest thickness spin wave mode (p = 0) is propagating perpendicular to
the bias magnetic field, q = (qy , 0), its dispersion equation obtained from (5) using
(6), (7), and (8) in the nonexchange limit (A = 0) has the form:
ω0 qy d = 2πf0 qy d
1/2 (10)
,
= γ H (H + 4π MS ) + (4π MS )2 P00 qy d 1 − P00 qy d
which for qy d < < 1 is similar to the dispersion equation obtained by Damon and
Eshbach for the dipolar surface mode [19]:
ωDE qy d = 2πfDE qy d
1/2
.
= γ H (H + 4π MS ) + (2π MS )2 1 − exp −2qy d
(11)
Thus, the spectrum of spin wave modes propagating perpendicular to the
direction of the bias magnetic field in an in-plane magnetized magnetic film obtained
from (5), (7), and (8) contains the lowest dipole-dominated mode (p = 0) with a
quasi-uniform thickness profile, which is analogous to the spin waves in 3D bulk
samples, and higher exchange-dominated spin-wave modes (p > 0), whose thickness
profiles are approximately described by the PSSWs. The higher spin-wave modes
are created because of the broken translational invariance along the x-direction. The
frequencies of these modes (p > 0) in the long wave limit qy d 1 can be calculated
from the following expression:
2A
pπ 2
2
qy +
ωp = 2πfp = γ H +
Ms
d
2 1/2 (12)
2
/H
pπ
4π
M
2A
s
qy2 +
+ 4π Ms + H
qy2
,
H+
Ms
d
pπ/d
which is obtained from Eq. (5) using the expressions for the dipole-dipole matrix
elements Fpp (qy d) calculated in [37].
Figure 1 illustrates the typology of the lowest quasi-uniform (p = 0) dipoledominated spin-wave modes in the quasi-2D case of a magnetic film for different
mutual orientations between the in-plane wavevector, q , and the static magnetization, MS . Three different geometries are shown. If MS is in the film plane, and
if q is perpendicular to MS , the surface or Damon-Eshbach (DE) mode exists (see
Eqs. (10) and (11)). If q and MS are collinear in the film plane, a mode with a
negative dispersion, or, so-called, the backward volume magnetostatic mode (BV)
exists and has the group velocity that is antiparallel to the wavevector. Finally, if
the magnetization MS is perpendicular to the film plane, the existing mode is the,
so-called, forward volume magnetostatic mode. (FV) In the latter case the dipoledipole matrix element F00 (q) in Eq. (5) can be expressed as:
288
S. O. Demokritov and A. N. Slavin
Fig. 1 Typology of spin wave modes in a magnetic film for different orientations of the
magnetization, MS , and the in-plane wavevector, q
F00 q d = P00 q d ,
(13)
with P00 (q d) given by Eq. (9).
Brillouin Light Scattering a Powerful Tool for Investigation of Spin
Waves
The spin-wave spectrum of a magnetic system can be investigated by various
techniques: ferromagnetic resonance [41], time-resolved Kerr magnetometry
[24, 42–44], and Brillouin light scattering spectroscopy (BLS) [45–47]. The
BLS experimental technique has a number of advantages for the investigations of
magnetic structures. It combines the possibility to study the dynamics of magnetic
systems in the frequency range of up to 500 GHz (corresponding time resolution is
2 ps) with a high lateral resolution of 20–40 μm for the regular setup and down to
220 nm for the, so-called, micro-focus BLS. In both cases the spatial resolution is
defined by the size of the laser beam focus.
Another important advantage of BLS is its very high sensitivity, which allows
us to register thermally excited spin wave modes, so the coherent excitation of
the magnetic element by an external signal is not necessary. This property of the
6 Spin Waves
Fig. 2 Scattering process of
a photon from a spin wave
(magnon). θ is the scattering
angle
289
hqI
hω I
θ
hqS
hωS
hq, hω
BLS is especially useful for the experimental investigations of complicated, strongly
confined spin-wave modes in patterned magnetic elements, which will be considered
in the next sections.
The BLS process can be considered as follows (see Fig. 2): monoenergetic
photons (visible light, usually green line 532 nm or blue line 473 nm) with the wave
vector qI and frequency ωI = cqI interact with the elementary quanta of spin waves
(magnons), characterized by the magnon wave vector q and frequency ω. Due to the
conservation laws resulting from the time- and space-translation invariance of a 3D
system the scattered photon increases or decreases its energy and momentum if a
magnon is annihilated or created:
ωS = (ωI ± ω)
(14)
q S = q I ± q
(15)
Measuring the frequency shift of the scattered light one obtains the frequency
of the spin wave participating in the BLS process. From Eq. (15) it is evident that
the wave vector qS − qI , transferred in the scattering process, is equal to the wave
vector q of the spin wave. Changing the scattering geometry one can sweep the value
of q and measure the corresponding ω(q). Thus, the spin-wave dispersion ω(q) can
be studied. Note here that for the 3D scattering process the maximum accessible
wavevector q = 2qI , the double value of the light wave vector, corresponds to the
backscattering geometry with the scattering angle θ = 180◦ .
The electromagnetic field of the scattered wave is proportional to the product
of the Faraday/Kerr and other magneto-optical constants of the medium and the
amplitude of the dynamic magnetization, corresponding to the spin wave [45]. Thus,
the BLS intensity, determined by the squared field, is directly proportional to the
dynamic magnetization squared. Magneto-optical effects relate the dielectric tensor
of the medium with Cartesian components of its magnetization. Usually the nondiagonal elements of the tensor, the magneto-optical effects (magnetic birefringence
and the Faraday effect and the corresponding dichroisms) are responsible for the
scattering. Therefore, the plane of polarization of the scattered light is rotated by
90◦ with respect to that of the incident light.
The conservation laws, given by Eqs. (14) and (15), follow from the time
invariance of the problem and the translation invariance of an infinite medium,
290
S. O. Demokritov and A. N. Slavin
correspondingly. However, if the scattering volume is finite, the selection rule for
the momentum is broken. For the scattering volume with a size less or comparable
with the wavelength of the light, any spin wave with a wavevector comparable
with that of light contributes to the scattering process. The confinement of the
scattering volume can come from the finite size of the element under consideration
and/or from a small scattering volume of the laser beam. For a thin film or for
nontransparent bulk materials, the thickness of the scattering volume is strongly
confined along the direction normal to the surfaces. Therefore, only the in-plane
wave vector is conserved in the light scattering experiments. As shown in Fig. 3, in
the backscattering geometry the transferred in-plane wavevector q|| is determined
by the angle of incidence q|| = 2qsinα, with q being the absolute value of the wave
vector of the incident light. For green laser light q|| varies in the typical range of
(0–2.5) × 105 cm−1 . This approach is illustrated in Fig. 4 showing the spin-wave
dispersion of a permalloy (Ni80 Fe20 ) film with a thickness of 20 nm, measured in
the DE geometry at the applied magnetic field H = 500 Oe. The experimental data
presented in Fig. 4 were obtained by varying α . The solid line in the figure is the
result of calculation based on Eq. (11) with the value of the permalloy magnetization
4πMS = 9.8 kG.
Fig. 3 Backscattering
process from a thin film. qi is
the wavevector of the incident
light; qs is the wavevector of
the incident light; α is the
angle between the
wavevectors and the film
normal
Spin-wave frequency (GHz)
Fig. 4 The spin-wave
dispersion of a permalloy
(Ni80 Fe20 ) film with
parameters listed in the text
measured using BLS. The
solid line is the result of
calculation based on Eq. (11)
qs
α
qi
14
12
10
8
6
0.0
0.5
1.0
1.5
5
-1
q|| (10 cm )
2.0
2.5
6 Spin Waves
291
Spin Waves in 1D Magnetic Elements: Standing and Propagating
Waves
In this section, we will consider spin-wave modes in arrays of micron-size quasi-1D
magnetic elements (stripes and waveguides). We will discuss lateral quantization of
DE modes in a longitudinally magnetized stripe due to its finite width as well as
localization of spin-wave modes near the edges of the stripes. After that, we review
recent experimental investigations of spin-wave propagation, excitation, and control
in microscopic waveguides.
BLS in Laterally Confined Systems
Before we consider new effects resulting from the lateral confinement in stripes, let
us first focus on BLS from laterally confined excitations.
It has been already mentioned in the previous sections that the form of the
conservation laws, which determine the BLS process, depends on the dimensionality
of the studied system: due to lack of translational invariance of a thin magnetic film
along the normal to the film surface, the corresponding component of the wavevector
is not conserved. Instead, the scattering angle (see Fig. 3) determines the 2D inplane vector, q|| only, whereas all the thickness modes possessing this q|| contribute
to BLS, albeit with different intensity according to their thickness profiles. If now
the in-plane translational invariance of the magnetic film is broken by patterning,
the in-plane wavevector is no longer fully conserved in the BLS process. In the case
of a spin-wave mode localized in a long stripe, the only conserved component is
the component of along the stripe axis. In analogy with the films, all the laterally
confined modes contribute to BLS, whereas the mode profile along the stripe width
(more specific: the corresponding Fourier component of the dynamic magnetization)
defines the contribution of the mode to BLS. Finally, if the confinement takes
place in all three dimensions, no conservation laws for wavevectors can be applied.
One should perform a Fourier analysis of the 3D distribution of the dynamic
magnetization of a particular mode to calculate its contribution to the BLS intensity.
Lateral Quantization of Spin Waves in Magnetic Stripes
Mathieu et al. [22] and Jorzick et al. [48] investigated spin-wave excitations by BLS
in arrays of permalloy stripes. They have found the effect of lateral quantization of
spin waves due to a finite width of the stripes and observed several dispersionless
spin-wave modes. Since these experiments provide the first account for spin-wave
modes in 1D magnetic systems and heavily contribute to quantitative understanding
of spin wave quantization effects in systems with reduced dimensions, we consider
them in detail.
The samples were made of 20–40 nm thick permalloy films deposited in
UHV onto a Si(111) substrate by means of e-beam evaporation using using
292
S. O. Demokritov and A. N. Slavin
Fig. 5 Scanning electron micrographs of permalloy stripes with a width of 1.8 μm and a
separation of 0.7 μm. (Reprinted with permission from [45], © 2001 by Elsevier)
X-ray lithography with a following lift-off process. Several types of periodic arrays
of stripes with stripe widths w = 1.7 and 1.8 μm and distances between the
stripes above 0.5 μm were prepared. Thus, the interaction between the stripes were
negligible. The length L of the stripes was 500 μm, ensuring 1D-properties of the
stripes. The patterned area was 500 × 500 μm2 , allowing BLS investigation with
a large beam diameter, providing a good wavevector resolution. One of the studied
arrays is shown in Fig. 5. In agreement with the shape-anisotropy arguments, the
magnetic easy axis of the array was along of the stripe axis.
Let us consider a magnetic stripe magnetized in plane along the z-direction
and having a finite width w along the y-direction as shown in Fig. 5. A boundary
condition similar to Eq. (6) at the lateral edges of the stripe should be imposed:
m|x=±d/2 = 0
(16)
One should emphasize that the internal field in the stripe is strongly nonhomogeneous due to the nonellipsoidal shape. This nonhomogeneity is of particular
importance close to the edges. In the considered geometry, only the dynamic
internal field is nonhomogeneous, since the static magnetization is along the
edge and does not contribute to the demagnetizing effects. Nevertheless, this
nonhomogeneous dynamic internal field results in specific mechanism for pinning
of the magnetization [39]. Therefore, the Eq. (16) differs from Eq. (6) written for
6 Spin Waves
293
an unconfined film possessing a homogeneous internal field. The corresponding
quantization of qy is then obtained as:
qyn =
nπ
w
(17)
where n = 1,2, . . . . Using Eqs. (4), (11), and (12) and the quantization expression
(17) one can calculate the frequencies of these so-called width (or laterally
quantized) modes. The profile of the dynamic part of the magnetization m in the
nth mode can be written as follows:
mn (y) = An sin
w
nπ
y+
w
2
(18)
Equation (18) describes a standing mode consisting of two counter-propagating
waves with quantized wavenumbers, nπ /w. Note here that due to the truncation of
the sin-function at the stripe boundaries the modes are no more infinite plane waves
and the quantized values are not true wavevectors.
In BLS experiments with backscattering geometry the in-plane wavevector q|| ,
transferred in the light scattering process, was oriented perpendicular to the stripes,
and its value was varied by changing the angle of light incidence, α, measured
from the surface normal: as illustrated in Fig. 3. Figure 6 shows a typical BLS
spectrum for the sample with a stripe width of 1.8 μm and as transferred wavevector
q|| = 0.3 × 105 cm−1 , while an external field of 500 Oe was applied along the stripe
axis. As it is seen in Fig. 6, the spectrum contains four distinct modes near 7.8,
9.3, 10.4, and 14.0 GHz. By varying the applied field, the spin wave frequency
for each mode was measured as a function of the field, as displayed in Fig. 7.
The observed dependence of all frequencies on the field confirms that all detected
modes are magnetic excitations. The dispersion law of the modes was obtained by
varying the angle of light incidence, α, and, thus the magnitude of the transferred
wavevector, qy . The obtained results are displayed in Fig. 8 for two with the same
stripe thickness of 40 nm and width of 1.8 pm, but with different stripe separations
of 0.7 pm (open symbols) and 2.2 pm (solid symbols). The dispersion measured
on the arrays with the same lateral layout but with a stripe thickness of 20 nm is
presented in Fig. 9. It is clear from Fig. 8 that one of the detected modes presented
by circles (near 14 GHz) is the PSSW mode, corresponding to p = 1 in Eq. (12).
This mode is not seen in Fig. 9 due to its much higher frequency caused by the
smaller stripe thickness. In the region of low wavevectors the spin wave modes show
a disintegration of the continuous dispersion of the DE mode of an infinite film into
several discrete, resonance-like modes with a frequency spacing between the lowest
lying modes of approximately 0.9 GHz for d = 20 nm and 1.5 GHz for d = 40 nm.
As it is clear from Figs. 8 and 9, there is no significant difference between the
data for the stripes with a separation of 0.7 and 2 μm. This fact indicates that the
mode splitting is purely caused by the quantization of the spin waves in a single
stripe due to its finite width. In other words, the studied stripes can be considered as
independent 1D elements.
294
S. O. Demokritov and A. N. Slavin
PSSW
BLS Intensity (a.u.)
3000
2000
1000
0
6
10
14
Frequency Shift (GHz)
Fig. 6 Experimental BLS spectrum obtained from the stripe array with a stripe thickness of 40 nm,
a width of 1.8 μm, and a separation of 0.7 μm. The applied field is 500 Oe orientated along the
stripe axis. The transferred wavevector of q|| = 0.3 × 105 cm−1 is oriented perpendicular to the
wires. The discrete spin-wave modes are indicated by arrows. PSSW stands for perpendicular
standing spin-wave mode. (Reprinted with permission from [45], © 2001 by Elsevier)
Spin Wave Frequency [GHz]
24
n=3
n=2
n=1
PSSW
20
16
12
Q-DE
8
4
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
Magnetic Field [kOe]
Fig. 7 Frequencies of the in-plane quantized Damon-Eshbach modes (Q-DE) as a function of the
applied field, obtained from the BLS spectra similar to that shown in Fig. 6. The lines are calculated
using Eq. (11) with quantized wavevectors, determined by the quantization numbers n = 1,2,3.
The line labeled PSSW shows the frequency of the first perpendicular standing spin-wave mode.
(Reprinted with permission from [45], © 2001 by Elsevier)
6 Spin Waves
295
Spin Wave Frequency (GHz)
17
16
d = 40 nm
15
14
13
n=5
12
n=4
11
n=3
10
9
n=2
8
n=1
7
6
0.0
0.5
1.0
5
1.5
-1
2.0
2.5
q|| (10 cm )
Spin Wave Frequency (GHz)
Fig. 8 Obtained spin-wave dispersion curves for an array of stripes with a stripe thickness of
40 nm, with a width of 1.8 μm and a separation of 0.7 μm (open symbols) and 2.2 μm (solid
symbols). The external field applied along the stripe axis is 500 Oe. The solid horizontal lines
indicate the results of a calculation using Eq. (11) with the quantized wavevectors, determined by
the quantization numbers n = 1,2,3,4,5. The dashed lines showing the hybridized dispersion of the
Damon-Eshbach mode and the first PSSW mode were calculated numerically for a continuous film
with a thickness of 40 nm. (Reprinted with permission from [45], © 2001 by Elsevier)
14
13
d = 20 nm
12
11
n=5
10
n=4
9
n=3
8
n=2
n=1
7
6
0.0
0.5
1.0
5
1.5
-1
2.0
2.5
q|| (10 cm )
Fig. 9 Obtained spin-wave dispersion curves for an array of stripes for the same conditions as in
Fig. 8, but with a stripe thickness of 20 nm. (Reprinted with permission from [45], © 2001 by
Elsevier)
296
S. O. Demokritov and A. N. Slavin
The main features of the observed spin wave modes in magnetic stripes can be
summarized as follows: (i) For low wavevector values (0–0.8 × 105 cm−1 ) the
discrete modes do not show any noticeable dispersion, behaving like standing wave
resonances. (ii) Each discrete mode is observed over a continuous range of the
transferred wavevector qy . (iii) The lowest two modes appear very close to zero
wavevector, the higher modes appear at higher values. (iv) The frequency splitting
between two neighbored modes is decreasing for increasing mode number. (v)
There is a transition regime (qy = 0.8–1.0 × 105 cm−1 ) where the well-resolved
dispersionless modes converge towards the dispersion of the continuous film (see
dashed lines in Figs. 8 and 9).
All these features can be explained if one implies that the observed discrete,
dispersionless spin wave modes result from a confinement of the DE modes in
the stripes. The confinement causes width-dependent quantization of the in-plane
wavevector of the mode, as discussed above. Considering the corresponding Fourier
component of the dynamic magnetization one can reproduce the measured BLS
intensity of each modes [48].
The frequency of the observed modes can be derived by substituting the
obtained quantized values of the wavevector, qyn , determined by Eq. (17) into the
dispersion equation of the DE mode, Eq. (11). The results of these calculations
are shown in Figs. 8 and 9 by the solid horizontal lines. For the calculation
the geometrical parameters (stripe thickness d = 20 or 40 nm, stripe width
w = 1.8 pm) and the independently measured material parameters 4πMs = 10.2 kG
and γ = 2.95 GHz/kOe were used. Without any fit parameters the calculation
reproduces all mode frequencies very well.
Spin Wave Wells and Edge Modes
In the previous subsection, the simplest geometry with the applied magnetic field
aligned along the stripe axis was considered. In this case, the static field is
homogeneous, while the dynamic field is not homogeneous resulting in the pinned
boundary conditions for the dynamic magnetization. If, however, the applied field is
directed along the width of a thin stripe, both the static and the dynamic internal field
are strongly inhomogeneous. Spin waves propagating along the field are affected in
this case not only by the confinement effects but also by the above inhomogeneities.
Since in unconfined media a wave with q||MS is called the backward volume (BV)
magnetostatic wave, we refer to this experimental geometry as the BV geometry,
contrary to the geometry described in the previous subsection, which we call the
DE geometry. Figure 10 shows two typical BLS spectra obtained from an array of
stripes for the external in-plane magnetic field He = 500 Oe for different orientations
of the field. The spectrum (a) is recorded with both transferred wavevector and He
aligned along the width of the stripe (the BV geometry), whereas the spectrum (b)
is obtained for He oriented along the stripe axis, thus presenting the DE geometry
6 Spin Waves
297
LM
BLS Intensity (a.u.)
Band
(a)
Quantized DE - Modes
PSSW
(b)
0
5
10
15
Frequency shift (GHz)
20
Fig. 10 BLS spectra obtained on the stripe array described in the text, q = 0.3 × 105 cm−1 and
He = 500 Oe for (a) the DE geometry and (b) the BV geometry. LM indicates the localized mode.
(Reprinted with permission from [23], © 2002 by the American Physical Society)
discussed in detail above. As it is seen in Fig. 10, both spectra contain several
distinct peaks corresponding to spin-wave modes. The high-frequency peaks can
easily be identified as exchange-dominated PSSW modes. Thus, a narrow PSSW
peak in spectrum (b) confirms the homogeneity of the static internal field in the DE
geometry. On the contrary, a broad PSSW peak in spectrum (a) clearly indicates
a strong inhomogeneous distribution of the internal field across the stripe in the
BV geometry. For a large He the magnetization is parallel to the applied field within
almost the entire stripe. Therefore, poles are created at the edges of the stripe, which
decrease the internal magnetic field in those regions. A detailed analysis shows that
the internal static field, Hi , has a broad maximum in the center of the stripe while it
is vanishing completely near the edges of the stripe [49, 50].
To get further inside into the physics of the observed spin-wave modes, their
dispersion was measured by varying q. It is displayed in Fig. 11 for both orientations
of He . Figure 11a representing the DE geometry is very similar to Figs. 8 and 9: it
clearly demonstrates the lateral quantization of the DE spin waves, resembling a
typical “staircase” dispersion. The frequency of the PSSW mode coincides with
that of the PSSW mode for the unpatterned film and corresponds to an internal
field of Hi = He = 500 Oe, thus corroborating a negligible static demagnetizing
field in the stripes magnetized along their long axes. The dispersion presented in
Fig. 11b and representing the BV geometry differs completely from that shown
in Fig. 11a. First, the PSSW mode is split into two modes, with frequencies
298
S. O. Demokritov and A. N. Slavin
Fig. 11 Spin wave dispersion of the stripe array measured at He = 500 Oe for (a) the DE geometry
with the quantization numbers of the quantized modes as indicated and (b) the BV geometry. In
the latter case, the shadowed region represents the band of non-localized spin modes, whereas LM
indicates the localized mode. The solid lines represent the results of calculation. (Reprinted with
permission from [23], © 2002 by the American Physical Society)
corresponding to internal fields of Hi = 300 Oe and Hi = 0 Oe, respectively, in
agreement with the above qualitative discussion. Second, a broad peak is seen in
the spectra in the frequency range 5.5–7.5 GHz over the entire accessible interval
of q. The shape of the peak varies with q, thus indicating different contributions of
unresolved modes to the scattering cross-section at different q. Third, a separate,
low-frequency, dispersionless mode with a frequency near 4.6 GHz (indicated as
“LM” in Figs. 10 and 11b) is observed over the entire accessible wave vector
range (qmax = 2.5 × 105 cm−1 ) with almost constant intensity. This is a direct
confirmation of a strong lateral localization of the mode within a region with the
width z = 2π /qmax = 250 nm. From the low frequency of the mode, one can
conclude that it is localized near the edges of the elements, where the internal field
vanishes.
A quantitative analytical description of the spin-wave modes observed in the BV
geometry is made as follows. The frequency of the spin wave as a function of q
and H is given by Eq. (4). Contrary to the DE geometry, where H = He = Hi ,
the demagnetizing effects in the BV geometry are very large: Hi is strongly
inhomogeneous and differs from He . To evaluate Hi :
6 Spin Waves
299
Hi (x, y, z) = He − N (x, y, z) · 4π MS
(19)
where Nzz (x, y, z) is the demagnetizing factor. Here we assume a Cartesian
coordinate system, in which the x-axis is perpendicular to the plane of the elements,
the y-axis is along the long axes of the stripes, and the z-axis is along the width of
the stripe: He ||ez .
The field profile of the internal field H(z) obtained from Eq. (19) for He = 500 Oe
is shown in the inset of Fig. 12. For Hi > 0 the magnetization is parallel to He .
Near the edges, however, regions with Hi = 0 and with continuously rotating
magnetization are formed [49, 50], which reflect spin waves propagating from the
middle of the stripe towards these regions. Moreover, a spin wave propagating
in an inhomogeneous field might encounter the second turning point even if the
magnetization is uniform. In fact, for large enough values of the internal field there
are no allowed real values of q, consistent with the spin-wave dispersion (Eq. 4) – a
potential well for propagating spin waves is created. Similar to the potential well in
quantum mechanics, the conditions determining the frequencies fr of possible spinwave states in the well created by the inhomogeneous internal field are as follows:
2
q (H (z), f ) dz +
ψ1 +
ψ2 = 2rπ
(20)
where r = 1,2,3, . . . and q(H(z),f ) is found from the spin-wave dispersion Eqs. (4)
and (5), and ψ l , ψ 2 are the phase jumps at the left and right turning points,
between which Eq. (4) has a real solution q(z) for a fixed frequency f.
We will illustrate these ideas in the following. The dispersion curves for spin
waves with q||He and p = 0 calculated using Eqs. (4) and (5) for different constant
values of the field are presented in Fig. 12. A dashed horizontal line shows the
frequency of the lowest spin-wave mode f1 = 4.5 GHz obtained from Eq. (20) for
the lowest value r = 1 in good agreement with the experiment. It can be seen from
Fig. 12 that for H > 237 Oe there are no spin waves with the frequency f1 = 4.5 GHz.
Therefore, the lowest mode can exist only in the spatial regions in the magnetic
stripe where 0 Oe < H < 237 Oe. The corresponding turning points are indicated in
the inset of Fig. 12 by the vertical dashed lines. Thus, the lowest mode is localized
in the narrow region z near the lateral edges of the stripe where 0.26 < |z/w| < 0.39.
The mode is composed of exchange-dominated plane waves with qmin < q < qmax ,
as indicated in Fig. 12.
The higher-order spin-wave modes with r > 1 having their frequencies above
5.3 GHz are not strongly localized under the used experimental conditions and exist
everywhere in the stripe where the internal field is positive (0 < |z/w| < 0.39). In
the experiment, they show a band, since the frequency difference between the fr and
fr + 1 modes is below the frequency resolution of the BLS technique. Note here that
several localized modes can be observed at higher values of He [51].
300
S. O. Demokritov and A. N. Slavin
Fig. 12 Dispersion of plane spin waves in the BV geometry at constant internal fields as indicated.
Inset: the profile of the internal field in a stripe. z shows the region of the lowest mode
localization. (Reprinted with permission from [23], © 2002 by the American Physical Society)
Implementation of Micro-Focus BLS for Laterally Patterned
Magnetic Systems
All the above BLS data were obtained in the so-called Fourier microscope mode
[48]. In this case, the diameter of the laser beam was kept quite large (typically
30–50 μm), allowing fulfilling the wavevector conservation law Eq. (15), and the
frequencies and the intensities of the studied spin-wave modes were investigated as
a function of transferred wavevector.
A complimentary approach is the micro-focus BLS [46]. Here the coherent laser
light focused onto the surface of the magnetic system into a diffraction-limited spot
by using a high-quality microscope objective lens with a large numerical aperture.
While the frequency shift of the scattered light is equal to the frequency of the
magnetization oscillations, its intensity (referred to as BLS intensity) is proportional
to the intensity of magnetization oscillations at the position of the probing spot.
The latter fact enables direct spatial imaging of the spin-wave intensity by twodimensional rastering of the probing spot over the sample surface (Fig. 13a). The
acquired intensity maps, such as that shown in Fig. 13b, allow one to obtain
information about the spatial characteristics of spin waves. In the case of spin-wave
6 Spin Waves
301
a)
Microwave
current
Probing
light
Antenna
H0
Magnonic
waveguide
Substrate
b)
1
-1
10
BLS
intensity,
a.u.
c)
-2
10
1
cos (ϕ)
0
500 nm
-1
Fig. 13 (a) Sketch of a typical micro-focus BLS experiment on spin-wave propagation in a
microscopic waveguide. (b) and (c) Representative examples of the two-dimensional maps of the
spin-wave intensity (b) and phase (c) recorded by micro-focus BLS. (© 2015 IEEE. Reprinted,
with permission, from [47])
beams propagating in waveguides, they also provide important information about
the damping and the spatial characteristics of the spin-wave beam.
For reliable two-dimensional imaging of spin waves, the spatial resolution of the
micro-focus BLS apparatus is of crucial importance, which is found to be about
250 nm for the wavelength of 532 nm in agreement with classical optics. This
resolution can be further improved to about 50 nm [52] by utilizing the principles of
near-field optical microscopy. However, the use of this approach inevitably leads to
a reduction of the sensitivity, which noticeably increases the time needed for BLS
measurements. Therefore, the experiments described here were performed by using
free-space-optics micro-focus BLS apparatus with the resolution of 250 nm, which
is sufficient to address most of the spin-wave propagation phenomena.
In agreement with the discussion in Sect. “BLS in Laterally Confined Systems”,
high-spatial resolution of the micro-focus BLS technique is incompatible with the
wavevector resolution due to the large uncertainty of the light-scattering angle
associated with the tight focusing of the probing light. Thus, the information about
the wavevector of the wavelength of the studied spin waves is lost, and the spinwave dispersion cannot be obtained on the usual way. However, this drawback can
be eliminated by making use of the time invariance of the BLS-process, which
results in frequency/phase conservation in the light-scattering process. It means
302
S. O. Demokritov and A. N. Slavin
that the phase of the scattered light is directly correlated to the phase of the
magnetization oscillations in the spin-wave mode. This phase information can be
acquired by utilizing interference of the scattered light with the light modulated
by the signal used to excite the magnetization oscillations. This approach enables
direct measurements of the phase difference between the excitation signal and
the phase of a propagating spin wave at the given spatial location, providing, for
example, direct information about the spin-wave wavelength. The phase-resolution
technique was first demonstrated for standard macro-BLS apparatus [53] and was
subsequently adapted for micro-focus BLS measurements [54, 55]. Figure 13c
shows a representative example of the measured spatial phase map for spin waves
propagating in a submicrometer-width magnonic waveguide, excited by microwave
current in the antenna. The plotted value is cos(ϕ), where ϕ is the phase difference
between the microwave current and the magnetization oscillations in the spin wave.
The spatial period of cos(ϕ)-function is equal to the spin-wave wavelength at a given
excitation frequency. Therefore, by varying the latter and measuring the spatial
period, one can obtain the complete information about the spin-wave dispersion
characteristics of the studied waveguide.
Propagating Waves in 1D Magnetic Structures
By analyzing the lateral quantization of spin waves in stripes in the previous subsections, we have considered standing waves, i.e., we implied that the component of
the wavevector along the stripe axis is zero. To extend this approach to propagating
waves in such a waveguide we should consider two components of the wavevector:
one component quantized due to finite width of the stripe and another component
continuously varying as illustrated in Fig. 14. In fact, if the quantized components
qzn are known, the spectrum of normal waveguide modes can be obtained from the
two-dimensional dispersion surface described by Eq. (4) and by cutting it along qy
at the fixed qzn as illustrated in Fig. 14a by the curves labeled as DE1 − DE3 . For
the sake of clearness we project these curves onto the frequency- qy plane, as shown
in Fig. 14b, keeping in mind that the different curves correspond to different qzn .
As seen from Fig. 14b, the considered modes propagate perpendicular to He , i.e.,
they are analogues of the DE mode in an extended film, their dispersion curves are
shifted to lower frequencies with respect to that of the unconfined DE mode, and this
shift increases with the increase of the mode number. This is not surprising, since all
these modes are characterized by a nonzero component of the wavevector qz parallel
to He . Since the dipolar magnetic energy is known to decrease with the increase
of this component causing the backward dispersion of the BV modes (Fig. 14a),
the dispersion curves of the waveguide modes shift to lower frequencies with the
increase of the mode number, which corresponds to the increase in qz .
We emphasize that the above-described approach to calculation of the dispersion
curves of the normal waveguide modes is a rough approximation, since they do not
take into account the reduction of the static magnetic field inside the waveguide
303
a)
z
y
DE3
DE
DE1
f0
DE2
DE1
DE2
BV
qy
DE3
H0
qz
b)
Frequency, GHz
Fig. 14 (a) Two-dimensional
dispersion spectrum of spin
waves in an extended in-plane
magnetized ferromagnetic
film. Inset shows the
geometry of the stripe
waveguide and transverse
profiles of the dynamic
magnetization for normal
waveguide modes. (b)
Calculated (solid lines) and
measured (symbols)
dispersion curves for a
waveguide with the width
w = 800 nm and the
thickness d = 20 nm
magnetized by the static field
He = 900 Oe. Dashed line
shows the dispersion curve
for Damon-Eshbach mode in
an extended film. (© 2015
IEEE. Reprinted, with
permission, from [47])
Frequency
6 Spin Waves
10
DE1
DE
DE2
f1
9
DE3
8
q2
q1
0
1
2
3
-1
4
q3
5
6
qy, μm
caused by the demagnetization effects, which can be further taken in account using
Eq. (19).
The data of Fig. 14b show that the dispersion spectrum of waveguide modes
supports multimode propagation of spin waves at all frequencies above f0 . For example, by exciting spin waves at the frequency f1 (see Fig. 14b), one simultaneously
excites a number of modes with different longitudinal wavevectors qy . Neglecting
attenuation of spin waves, the spatial distribution of the intensity of the dynamic
magnetization in these patterns can be described as (cf. Eq. 18):
I (y, z) =
n
w
nπ
z+
An sin
w
2
exp −iq n y
2
(21)
where An are the amplitudes of the modes and qn are their longitudinal wavevectors
at the given excitation frequency (see Fig. 14b). Figure 15 shows the results of
calculations based on Eq. (21) performed for different ratios between the amplitudes
An and the dispersion data taken from Fig. 14b. In the simplest case, where the
only present mode is the fundamental mode with n = 1 (Fig. 15a), the intensity
distribution is uniform in the longitudinal direction and shows a half-sine profile
in the transverse direction. The co-propagation of the fundamental mode and the
mode with n = 2 (Fig. 15b) results in an appearance of a “snake”-like pattern,
which becomes more pronounced with the increase of A2 . We note that, because
of the symmetry reasons, the mode n = 2 possessing an antisymmetric distribution
of the dynamic magnetization across the waveguide width (inset in Fig. 14a) is
304
z, μm
a)
A1=1, A2=0, A3=0
0.4
0
-0.4
z, μm
b)
A1=1, A2=0.15, A3=0
0.4
0
-0.4
z, μm
Fig. 15 Interference patterns
for the three lowest-order
waveguide modes calculated
for different ratios between
their amplitudes, as labeled.
Calculations were performed
for the waveguide with the
width of 800 nm and the
thickness of 20 nm
magnetized by the field
He = 900 Oe. (© 2015 IEEE.
Reprinted, with permission,
from [47])
S. O. Demokritov and A. N. Slavin
A1=1, A2=0.3, A3=0
0.4
0
-0.4
z, μm
c)
A1=1, A2=0, A3=0.15
0.4
0
z, μm
-0.4
A1=1, A2=0, A3=0.3
0.4
0
-0.4
0
1
2
y, μm
3
4
5
normally not excited in axially symmetric guiding systems and can only be observed
if this symmetry is broken. In contrast, a significant contribution of the mode
with n = 3 is detected in most of the experiments. As seen from Fig. 15c, the
co-propagation of this mode and the fundamental mode of the waveguide results
in a periodic spatial beating pattern, where the spin-wave energy is periodically
concentrated in the middle of the waveguide, while the transverse width of the spinwave beam shows a periodic modulation. By analogy with the light focusing in
optics, this effect was given a name of “spin-wave focusing” [56]. Figure 16a shows
a typical measured spin-wave intensity map for a 2.4 μm wide and 36 nm thick
Py waveguide clearly demonstrating this effect (compare with Fig. 15c). In order
to highlight the details of the interference pattern in Fig. 16a, the spatial decay
of spin waves is numerically compensated by multiplying the experimental data
by exp.(2y/ξ), where ξ is the spin-wave decay length – the distance over which
the wave amplitude decreases by a factor of e. The latter is determined from the
dependence of the BLS intensity integrated across the transverse waveguide section
versus the propagation coordinate (solid symbols Fig. 16b). As seen from Fig. 16b,
spin waves in the studied waveguide exhibit clear exponential decay characterized
by ξ = 6.4 μm. Figure 16b also shows by open symbols the transverse width
of the spin-wave beam versus the propagation coordinate, which can be used to
quantitatively characterize the strength of the focusing effect. In particular, for the
data of Fig. 16b, the modulation of the beam width caused by the focusing is equal to
about 70%, while the smallest width observed at the focal point is equal to 0.65 μm.
305
Propagation
z, μm
a)
1
1
0
0.5
-1
0
0
1
2
3
4
5
6
7
BLS intensity, a.u.
6 Spin Waves
y, μm
1.5
Integral
intensity, a.u.
1
1.0
0.1
Beam width, μm
b)
0.5
0
1
2
3
4
5
6
7
y, μm
Fig. 16 (a) Measured map of the spin-wave intensity for a waveguide with the width of 2.4 μm
and the thickness of 36 nm magnetized by the field of 900 Oe. Excitation frequency is 9.4 GHz.
Spatial decay of spin waves is numerically compensated. (b) Solid symbols – BLS intensity
integrated across the transverse waveguide section versus the propagation coordinate in the loglinear scale. Line is the exponential fit to the experimental data. Open symbols – transverse width
of the spin-wave beam measured at one half of the maximum intensity versus the propagation
coordinate. (© 2015 IEEE. Reprinted, with permission, from [47])
In the above discussion on the normal waveguide modes, we neglected the
nonuniformity of the magnetic field inside the waveguide caused by the demagnetization effects. On one side, this nonuniformity does not qualitatively affect
the structure of the modes evolving from plane spin waves due their geometrical
confinement. As discussed above, the nonuniformity is known [49] to result in the
appearance of the regions of strongly reduced internal field close to the edges of the
stripe [50], which gives rise to additional spin-wave modes having no analogue in
the case of extended magnetic films [23]. According to [49], the distribution of the
internal field across the width of a magnetic stripe magnetized perpendicular to its
axis can be approximated as (cf. Eq. 18):
d
d
4π MS
atan
− atan
Hi (z) = He −
π
2z + w
2z − w
(22)
Figure 17a shows this distribution calculated for the waveguide with the width
of w = 2.1 μm and the thickness of d = 20 nm magnetized by the static field
He = 1100 Oe. As seen from this data, close to the edges of the waveguide, the
internal field is drastically reduced resulting in the appearance of field-induced
channels, where spin waves can be localized, as discussed in previous subsection.
306
S. O. Demokritov and A. N. Slavin
Internal
field, Oe
a)
1000
500
Field-induced
channels
0
-1.0
-0.5
0.0
0.5
1.0
z, μm
z, μm
1
Propagation
9.8 GHz
0
1
-1
z, μm
1
0.5
9.0 GHz
0
0
BLS intensity, a.u.
b)
-1
Distance between
the beams, μm
c)
4
y, μm
6
8
10
1.5
0.6
1.0
0.5
Beam width, μm
2
0
0.4
0.5
9.0
9.2
9.4
9.6
9.8
Frequency, GHz
Fig. 17 (a) Calculated distribution of the internal static magnetic field across the width of a
waveguide with the width of 2.1 μm and the thickness of 20 nm magnetized by the static field of
1100 Oe. Horizontal dashed line marks the value of the external magnetic field. (b) Measured maps
of the spin-wave intensity for two excitation frequencies, as labeled. Spatial decay of spin waves
is numerically compensated. (c) Distance between the centers of the spin-wave beams and their
transverse width measured at one half of the maximum intensity versus the spin-wave frequency.
(© 2015 IEEE. Reprinted, with permission, from [47])
Since these modes are mostly concentrated in the areas of the reduced field,
their typical frequencies are lower than the frequencies of the “center” modes as
discussed above. Therefore, in order to address them experimentally, one needs
to excite the waveguide at frequencies below the frequency of the uniform ferromagnetic resonance f0 (see Fig. 14a). Figure 17b shows two decay-compensated
spin-wave intensity maps measured for a waveguide with the parameters given
above by applying the excitation signal at the frequencies of 9.0 and 9.8 GHz, which
are smaller than f0 equal to about 10 GHz for the used experimental conditions.
6 Spin Waves
307
The data of Fig. 17b show that, in agreement with the simple qualitative model,
at these frequencies, the spin waves do not occupy the entire cross-section of
the waveguide. Instead, they form two narrow beams with the submicrometer
width whose spatial positions depend on the spin-wave frequency. The quantitative
analysis (see Fig. 17c) shows that, in the wide frequency interval, the widths of
the beams vary moderately staying in the range 400–500 nm, while the distance
between their centers monotonously increases with the decrease of the frequency
from 0.8 μm to 1.4 μm, i.e., by more than 70%.
Control and Conversion of the Propagating Waves
One of the great advantages of spin waves for implementation of signal-processing
devices is their controllability by the static magnetic field, which allows one to
efficiently manipulate the spin-wave propagation. Although this control mechanism
is straightforward, its implementation on the macroscopic scale requires the use of
electromagnets making the resulting devices extremely space and power consuming.
The downscaling of spin-wave devices provides a route for overcoming this
drawback, since, in microscopic systems, the control magnetic field has to be created
over small distances and sufficiently large local magnetic fields can be created
by using relatively small electric currents [55]. This approach is schematically
illustrated in Fig. 18a. Instead of using external electromagnets, the control magnetic
field is created by the electric current flowing in a control line, which is directly
integrated into the waveguide. The composite waveguide consists of two layers:
the upper 20 nm thick Py layer guiding spin waves and the bottom 100 nm thick
Cu layer used as a current-carrying line to generate controlling magnetic fields.
Because of the large difference in conductivities of Cu and Py, the shunting of the
control current through the Py layer is negligible, which makes electrical isolation
between the two layers unnecessary.
The Oersted field in the Py layer produced by the current in the Cu layer can be
approximated as H = I/(2w + 2d), where I is the current strength and w = 0.8 μm
and d = 0.1 μm are the width and the thickness of the Cu line, respectively. Due to
the small cross-section of the control line, one achieves sufficiently high efficiency
of the magnetic field generation of about 7 Oe/mA. The effect of the current on the
dispersion characteristics of spin waves in the waveguide is illustrated by Fig. 18b.
By applying I = ± 12 mA one creates controlling magnetic field H = ± 83 Oe,
which adds or subtracts from the static magnetic field He = 520 Oe resulting in
the shift of the dispersion curve by more than ±500 MHz. This shift leads to
a significant variation of the longitudinal wavevector qy at the given spin-wave
frequency. Figure 18c demonstrates the controllability of the wavevector and the
wavelength of sin waves with the frequency of 7 GHz. These data show that by
applying I = ±12 mA one can change these parameters by more than 50%. Note
that the variation of qy with the current is nearly linear. Since the phase accumulated
by the spin wave over a propagation distance L is proportional to qy : ϕ = Lqy , this
also implies a linear controllability of the phase accumulation, which is attractive
308
z
a)
y
He
Py
(20 nm)
DH
dc
Frequency, GHz
8
I=12 mA
7
6
I=-12 mA
5
I=0
0
c)
1
2
3
4
-1
qy, μm
5
6
5
2.5
qy, μm
2.0
4
1.5
3
Wavelength, μm
b)
Cu
(100 nm)
-1
Fig. 18 (a) Schematic of a
composite waveguide with
the integrated control Cu line.
(b) Dispersion curves of the
fundamental waveguide mode
for different currents in the
control line, as labeled. Lines
show the calculated
dispersion curves, and
symbols show the results of
measurements by
phase-resolved BLS
technique. (c) Longitudinal
wavevector qy and the
wavelength of the spin wave
at the frequency of 7 GHz
versus the control current.
Symbols – experimental data,
lines – guides for the eye. (©
2015 IEEE. Reprinted, with
permission, from [47])
S. O. Demokritov and A. N. Slavin
1.0
-15
-10
-5
0
5
10
15
Current, mA
for technical applications. Based on the data of Fig. 18c, one can conclude that by
applying the control current of ±12 mA, the phase accumulated by the spin wave
can be changed by ±π radians at the propagation distance of about 3.2 μm which is
smaller than the spin-wave propagation length. This makes the proposed mechanism
well suited for implementation of magnonic logic devices [57], where the digital
information is coded into the phase of propagating spin waves.
In addition to the use of magnetic fields created by electric currents, the control
of spin-wave propagation can also be realized by using demagnetizing fields. Since
the demagnetizing field depends on the ratio between the width and the thickness of
the waveguide (Eq. 22), simple variation of one of these parameters enables efficient
manipulation of spin waves [58, 59].
A waveguide with the varying width is schematically shown in Fig. 19a: while
the thickness of the waveguide d = 36 nm remains constant, its width w varies from
1.3 to 2.4 μm over a transition region with the length L. According to Eq. (22),
such a variation results in a spatial variation of the internal field, which changes
from 870 Oe in the narrow part of the waveguide to 950 Oe in the wide part
(Fig. 19b). Due to this variation, the dispersion curves of the fundamental center
waveguide mode are shifted in the two parts by about 500 MHz in the frequency
domain (Fig. 19c). If the excitation frequency is chosen to be located between the
6 Spin Waves
309
b)
y
Internal
field, Oe
1.3 μm
He
L
8.7 GHz
e) L=3 μm
Frequency, GHz
0
1
11
2
3 4
y, μm
5
6
Wide
waveguide
f2
10
Narrow
waveguide
f1
9
0
d)
900
850
2.4 μm
c)
950
Wide
waveguide
z
Narrow
waveguide
a)
1
2
-1
qy, μm
Propagation
3
9.1 GHz
1 μm
L=1 μm
Fig. 19 (a) Schematic of a spin-wave waveguide with a varying width. (b) Calculated distribution
of the internal static magnetic field in the section along the axis of the waveguide with the
thickness of 36 nm and the geometrical parameters given in (a). The external static magnetic field
He = 1000 Oe. (c) Calculated dispersion curves for the fundamental center waveguide mode in the
wide and the narrow parts of the waveguide. (d) Maps of the spin-wave intensity measured at the
excitation frequencies of 8.7 and 9.1 GHz, as labeled. Width of the transition region L = 2 μm. (e)
Maps of the spin-wave intensity measured at the excitation frequency of 9.7 GHz in waveguides
with L = 3 and 1 μm, as labeled. In (d) and (e) the spatial decay of spin waves is numerically
compensated. (© 2015 IEEE. Reprinted, with permission, from [47])
cut-off frequencies of the two dispersion curves, i.e., between 8.7 and 9.2 GHz (f1 in
Fig. 19c), the center mode propagating in the narrow part of the waveguide cannot
pass into the wide part. Instead, it should be transformed into the edge mode, whose
frequency range is located below that of the center modes. This case is illustrated
in Fig. 19d showing two spin-wave intensity maps measured for the excitation
frequencies of 8.7 and 9.1 GHz. These maps clearly demonstrate the conversion
of the center mode into the edge mode characterized by two narrow spin-wave
beams with frequency-dependent spatial separation (see Sect. “Implementation of
Micro-Focus BLS for Laterally Patterned Magnetic Systems”). Since the two beams
310
S. O. Demokritov and A. N. Slavin
propagate in the field-induced channels and are independent from each other, the
observed transformation can be used for implementation of a spin-wave splitter.
One also observes interesting behaviors in the case, when the frequency of
spin waves is larger than cut-off frequencies in both parts of the waveguide (f2 in
Fig. 19c). In the waveguides with the relatively long transition region (L = 3 μm
in Fig. 19d) the propagation of spin waves from the narrow to the wide part
is quasi-adiabatic. It is only accompanied by the increase in the wavelength,
while the spatial structure of the spin-wave beam remains unchanged. However,
in systems with shorter transitions (e.g., L = 1 μm in Fig. 19e), the propagation
is accompanied by an appearance of a complex intensity pattern, which can be
recognized as an interference pattern created by several waveguide modes with
comparable amplitudes (see Fig. 15c). This is due to the strong coupling of the
waveguide modes mediated by the spatial nonuniformity in the waveguide, which
results in the efficient energy transfer from the fundamental mode to the higherorder modes. As seen from Fig. 19e, this effect causes a strong concentration of
the spin-wave energy in the middle of the waveguide at a certain distance from the
transition region, which can be treated as an enhanced spin-wave focusing.
A particularly interesting case is a junction between a 1D waveguide and 2D
film [60, 61] (Fig. 20a). Because of the abrupt transition in such a system, the
wavevector of spin waves is not conserved during the conversion of the waveguide
modes into the modes of the extended film. In other words, being radiated from
the open end of the waveguide, the waveguide mode excites spin waves within
a large range of wavevectors. Because of the temporal translation symmetry, the
frequency of radiated spin waves is equal to that of the waveguide mode. Therefore,
the characteristics of the radiated waves can be obtained by considering constantfrequency contours of the two-dimensional dispersion surface (Fig. 14a), as shown
in Fig. 20b. Figure 20c shows such contours projected onto the qz − qy plane,
calculated for the conditions used in the experiment: d = 36 nm, He = 690 Oe.
In this representation the vector of the group velocity of spin waves Vg is directed
along the normal to the constant-frequency contour: Vg = 2π ∇ f (qy , qz ). As
seen from Fig. 20c, except for the region of small qz , the direction of the group
velocity is practically constant and builds a well-defined angle with the direction
of the static magnetic field He . This indicates that a large group of spin waves with
different wavevectors transmits energy in the same direction. Therefore, one expects
predominant radiation of the spin-wave energy from the waveguide along this
direction. Figure 20d shows an experimental spin-wave intensity map corresponding
to the frequency of 8.2 GHz. The shown experimental data confirm the conclusions
of the above analysis: the spin waves are radiated in a form of two narrow beams and
their directions coincide well with the direction of the group velocity obtained from
the analysis of the 2D dispersion surface (arrows in Fig. 20d). Since the direction
of the beams is determined by He , it can be electronically steered. In general, the
phenomenon enables a directional, 1D transmission of spin waves in an extended
2D magnetic film without utilization of the geometrical confinement. Recently, this
phenomenon was also observed for spin waves radiated by a spin-torque nanooscillator [62], which, due to its small size, also emits spin waves within large
interval of wavevectors.
311
a)
b)
z
y
Waveguide
Frequency
6 Spin Waves
DE
f2
f1
f0
He
BV
f0
qy
Extended film
qz
8.2 GHz
qz, μm
-1
c)
10.2 GHz
d)
4
Waveguide
Vg
2
0
He
-2
-4
0
2
4
6
1 μm
-1
qy, μm
Fig. 20 (a) Schematic of a junction between 1D waveguide and 2D extended film. (b) Twodimensional dispersion surface of spin waves in an extended magnetic film with marked constantfrequency contours. (c) Constant-frequency contours projected onto the qz -qy plane calculated for
t = 36 nm and He = 690 Oe. (d) Measured map of the intensity of spin waves with the frequency
of 8.2 GHz radiated from a waveguide with the width of 2 μm into an extended magnetic film.
The spatial decay of spin waves is numerically compensated. (© 2015 IEEE. Reprinted, with
permission, from [47])
Inductive Excitation of Spin Waves in 1D Waveguides
In spite of the recent progress in studies of spin-transfer torque excitation of
propagating spin waves (see Sect. “Spin Waves in 0D”), the inductive excitation mechanism shortly introduced above still remains the most widely used in
experimental investigations of spin-wave phenomena in both 2D and 1D magnetic
systems, because its implementation does not require complex nanolithography
techniques. This method is also characterized by the full control over the frequency
of excited spin waves, which makes it attractive for research purposes. The inductive
excitation by means of spin-wave antennae was widely used in the past for implementation of macroscopic-scale devices (see, e.g., [63, 64]) and was subsequently
transferred onto the microscopic scale without significant modifications.
Figure 21a shows schematics of a spin-wave waveguide with a spin-wave antenna
on top. A microwave-frequency electric current transmitted through the antenna
creates the dynamic magnetic field h, which couples to the dynamic magnetization
in the waveguide and excites propagating spin waves. Experimentally, the excitation
312
S. O. Demokritov and A. N. Slavin
process can be efficiently characterized by micro-focus BLS by placing the probing
laser spot onto the waveguide in the vicinity of the antenna and recording the BLS
intensity as a function of the frequency of the excitation current. Such an excitation
curve is shown in Fig. 21b. The curve exhibits a fast drop of the spin-wave intensity
at frequencies below the frequency of the ferromagnetic resonance f0 . However, the
intensity of excited spin waves still remains noticeable in this region and shows a tail
extending far into the low-frequency spectral interval. Within this frequency region,
propagating edge modes are excited. The part of the excitation curve at frequencies
above f0 corresponding to the center waveguide modes shows a nonmonotonous
behavior with several oscillations. This oscillatory behavior can be understood
based on the spin-wave excitation theory adapted for the case of microscopic
waveguides [65]. According to this theory, the amplitudes of the waveguide modes
An excited by the antenna are determined by the spatial overlap of the dynamic
magnetic field h(x,y,z) created by the antenna with the dynamic magnetization in
the waveguide m(x,y,z). As seen from Fig. 21a, the former has x- and y-components.
Both these components are perpendicular to the direction of the static magnetization
and, therefore, can linearly couple to the dynamic magnetization. However, because
of the strong dynamic demagnetization in the thin-film waveguides, the effect of
the out-of-plane component hx is relatively small and can be neglected in the first
approximation. Then the excitation problem is reduced to the consideration of
the spatial overlap of the hy component with the corresponding component of the
magnetization. Neglecting the variations of the field and the dynamic magnetization
across the waveguide thickness, the amplitudes of excited spin-wave modes can be
expressed as:
w/2
An ∝
∞
hy (z)mny (z)dz
−w/2
·
hy (y)mny (y)dy
(23)
−∞
where w is the width of the waveguide. The corresponding profiles of components
of the field and the dynamic magnetization are schematically shown in the inset
in Fig. 21a: hy (z) = const, and hy (y) can be approximated by a rectangular
pulse function with the width equal to that of the antenna d. As discussed in
Sect. “Propagating Waves in 1D Magnetic Structures” (Eq. 21), the transverse
profiles of the dynamic magnetization corresponding to the center modes can
approximated as mny (z) ∝ sin (nπ (z + w/2)), while the longitudinal profile
represents a propagating wave: mny (y) ∝ exp −iq ny y , where qyn are the
longitudinal wavevectors of the modes at the given spin-wave frequency.
Substituting these expressions into Eq. (22) one obtains:
n
1 − (−1)n sin qy b
An ∝
n
qyn
(24)
6 Spin Waves
313
x
a)
Antenna
Microwave
current y
hy(y)
z
b
h
He
Magnetic
waveguide
n=2
n=1
hy(z)
my(y)
y
n=3
n
z
my (z)
b)
BLS Intensity,
a.u.
Edge
Modes
DE
Modes
f0
8
qy, μm
-1
c)
6
n=3
2π/d
4
n=1
2
0
An, a.u.
d)
1.0
n=1
0.5
n=3
0.0
7
8
9
10 11
Frequency, GHz
12
13
Fig. 21 (a) Schematic of a waveguide with the spin-wave antenna on top. Inset schematically
shows the y- and z-profiles of the dynamic field of the antenna and those of the dynamic
magnetization in the spin-wave modes. (b) Spin-wave excitation curve measured in a waveguide
with the width of 2 μm and the thickness of 36 nm magnetized by the static field of 900 Oe. (c)
Calculated dispersion curves for the first two symmetric waveguide modes. Horizontal dashed line
shows the value of qy , at which the excitation efficiency vanishes. Arrows show the corresponding
frequencies for the two modes. (d) Calculated amplitudes of the first two symmetric waveguide
modes versus the excitation frequency. The vertical dashed line in (b)–(d) marks the frequency of
the uniform ferromagnetic resonance f0 . (© 2015 IEEE. Reprinted, with permission, from [47])
314
S. O. Demokritov and A. N. Slavin
The first term in Eq. (24) shows that the amplitudes of the modes decrease
with the increase of the mode number as 1/n and that only modes with symmetric
transverse profiles can be excited. The second term shows that the excitation
efficiency has a maximum at qy = 0, exhibits an oscillatory behavior in agreement
with the experimental data of Fig. 21a, and vanishes for qy = 2π m/b or b = mλ,
where m = 1,2,3 . . . and λ is the wavelength of the spin wave.
Figure 21d presents the frequency dependences of the amplitudes of the first
two symmetric waveguide modes calculated based on Eq. (23) and the dispersion
curves obtained by using Eqs. (4) and (5) (shown in Fig. 21c). The calculations
were performed for the conditions used to acquire the experimental excitation curve
shown in Fig. 21b: w = 2 μm, d = 36 nm, He = 900 Oe, and b = 1.5 μm. As
seen from Fig. 21d, the fundamental mode with n = 1 clearly dominates over
the mode with n = 3 and the nodes of the corresponding curve match well with
those seen in the experimental data. We note that, due to the shift of the dispersion
curves corresponding to different modes (Fig. 21c), the frequencies, at which the
excitation efficiency vanishes, are different for the modes with n = 1 and 3. Since
the contribution of the mode with n = 3 is relatively small, vanishings of its
amplitude cannot be seen in the experimental curve in Fig. 21b. Nevertheless, this
vanishing can be proven by the spatial imaging of the spin-wave propagation. These
measurements show a strong reduction in the spatial transverse modulation of the
spin-wave beam at the frequency of about 10 GHz in agreement with the data
of Fig. 21d. We emphasize that the dependence of the relative amplitudes of the
waveguide modes on the excitation frequency allows one to control the strength of
the spin-wave focusing effect. If the transverse modulation of the spin-wave beam
is not desired, one can choose the spin-wave frequency in the vicinity of the point
of the vanishing excitation efficiency of the mode with n = 3, while by choosing
the frequency in the region, where the excitation efficiencies of both modes are
approximately equal, one obtains the strongest mode interference.
The above-described simple model can be extended by additionally taking
into account the out-of-plane component of the dynamic magnetic field of the
antenna [65]. This extension does not significantly modify the discussed frequency
dependence of the excitation efficiency. However, it results in different excitation
efficiencies of spin waves propagating in the positive and the negative direction of
the y-axis. This excitation non-reciprocity is often confused with the intrinsic nonreciprocity of DE modes [19], which has no effect for spin waves with qy < <1/d
addressed in most of the experiments utilizing the inductive spin-wave excitation.
For typical parameters of the Py micro-waveguides, the excitation non-reciprocity
results in the factor of 4–5 difference between the intensities of waves propagating
in the opposite directions from the antenna.
The above analysis shows that by using an inductive spin-wave antenna one
can only efficiently excite spin waves within a certain interval of wavevectors
corresponding to the first oscillation lobe of the excitation efficiency function
(Fig. 21d), whose spectral width is determined by the width of the spin-wave
antenna b. This limits the applicability of standard inductive excitation, since, in
order efficiently to excite short-wavelength spin waves necessary for most of the
magnonic applications, one has to reduce the width of the antenna, which inevitably
6 Spin Waves
315
causes the microwave impedance matching problems and increases the microwave
losses in magnonic devices. This drawback can be partially overcome [59] by
utilizing the spin-wave controllability by the demagnetizing fields discussed in
Sect. “Propagating Waves in 1D Magnetic Structures”.
Figure 22a schematically shows a system enabling inductive excitation of spin
waves with the wavelength smaller than the width of the antenna: the width of
the waveguide equal to 2 μm in the excitation area gradually reduces to 0.5 μm,
while the thickness of the waveguide d = 40 nm stays unchanged. As discussed
in detail in Sect. “Propagating Waves in 1D Magnetic Structures”, the variation of
the waveguide width leads to the variation of the internal field along the waveguide
axis, which causes the shift of the dispersion curves of the waveguide modes in
the frequency domain. As a result, spin waves excited in the 2 μm-wide part of
the waveguide continuously decrease their wavelength propagating in the tapered
part. This process is illustrated by the two-dimensional phase map (Fig. 22b) and its
section along the waveguide axis (Fig. 22c) obtained by using the phase-resolved
micro-focus BLS technique. The experimental data clearly show a significant
reduction of the wavelength, which, after passing the tapered part, becomes smaller
than the width of the antenna equal to 2.2 μm. Figure 22d further characterizes
this wavelength conversion process. It shows the dependence of the spin-wave
wavelength at the output of the tapered part on the wavelength in the excitation
area. These data show that the discussed conversion process enables the reduction
of the wavelength by up to a factor of 15 and allows one to efficiently excite
spin waves with the wavelength below 1 μm, which is more than by a factor of 2
smaller than the antenna width. We also note that, as shown in [59], the conversion
process is not accompanied by noticeable additional losses associated with the spinwave reflection in the transition region, which makes it favorable for technical
applications.
Spin-Torque Transfer Effect and Spin Waves
Since the first demonstration [25, 27–29, 66, 67] of the possibility to excite
magnetization dynamics by spin-polarized electric currents due to the spin-transfer
torque (STT) effect [6, 7], dynamic spin-torque phenomena have become the subject
of intense research. The ability to control high-frequency magnetization dynamics
by dc currents is promising for the generation of microwave signals [68–72] and
propagating spin waves [16, 30, 73–75] in magnetic nanocircuits. Electronically
controlled local generation of spin waves is particularly important for the emerging
field of nanomagnonics, which utilizes propagating spin waves as the medium for
the transmission and processing of signals, logic operations, and pattern recognition
on nanoscale [3–5].
Initially, STT phenomena have been studied in 0D-nanodevices based on the
tunneling or giant magnetoresistance spin-valve structures, where STT is induced
by the electric current flowing through a multilayer that consists of a “fixed”
magnetic spin-polarizer and the active magnetic layer, separated by a nonmagnetic
316
S. O. Demokritov and A. N. Slavin
a)
2 μm
z
b)
y
He
Pr
op
ag
ati
on
L
0.5 μm
c)
cos (ϕ)
1
2.2 μm
0
-1
cos (ϕ)
1
0
-1
0
2
4
6
8
10
d)
Converted
wavelength, μm
y, μm
1.5
1.0
0
5
10
15
20
25
Excited wavelength, μm
Fig. 22 (a) Schematic of a tapered waveguide with the spin-wave antenna located in the wide
part. (b) Measured phase map of spin waves propagating in a 40 nm thick tapered waveguide
magnetized by the static field He = 900 Oe. The length of the tapered part L = 5 μm. Excitation
frequency is 9.8 GHz. (c) Section of the phase map along the waveguide axis. (d) Dependence of
the spin-wave wavelength at the output of the tapered part on the wavelength in the excitation area.
(© 2015 IEEE. Reprinted, with permission, from [47])
metallic or insulating spacer. In these structures, the electric charges must cross
the active magnetic layer to excite its magnetization dynamics. To enable current
flow through the active magnetic layers, STT devices operating with spin-polarized
electric current require that current-carrying electrodes are placed both on top and
on the bottom of the spin valve. To keep the electric current below a reasonable
limit, the devices should have sub-100-nm dimensions in both lateral directions.
An alternative approach to the implementation of STT devices that avoids
these shortcomings utilizes pure spin currents – flows of spin not accompanied
by directional transfer of electrical charge. This approach does not require the
flow of electrical current through the active magnetic layer, resulting in reduced
Joule heating and electromigration effects. One can also eliminate the electrical
leads attached to the magnetic layer to drain the electrical current, enabling novel
geometries and functionalities of the STT devices. Moreover, it becomes possible to
use insulating magnetic materials such as Yttrium Iron Garnet (YIG) [12, 76].
6 Spin Waves
317
Among the physical mechanisms for creation of pure spin currents, the spinHall effect (SHE) [9–11, 77] plays the most important role so far. The effect is
generally significant in nonmagnetic materials with strong spin-orbit interaction,
such as Pt and Ta. An electrical current in these materials produces a spin current
in the direction perpendicular to the charge flow, due to a combination of spinorbit splitting of the band structure (intrinsic SHE), and the spin dependence of the
electron scattering on phonons and impurities (extrinsic SHE) [11, 77]. When a SHE
layer is brought in contact with a ferromagnetic film, the spin current flows through
the interface into the ferromagnet and exerts STT on its magnetization [78]. The
ability to exert STT on ferromagnets over extended areas is a significant benefit of
SHE. Indeed, when an in-plane current flows through an extended bilayer formed by
a SHE material and a magnetic film, the spin current produced by SHE is injected
over the entire area of the sample, which can be as large as several millimeters
[78]. This feature makes SHE uniquely suited for the control of the spatial decay
of propagating spin waves in 1D waveguides, when STT partially compensates the
natural magnetic damping.
The effect of pure spin current on the magnetization is similar to that of
spin-polarized electric currents. Both can be described by the Slonczewski-LandauLifshitz-Gilbert equation [8]:
dM
β
α
dM
= −γ M × H eff +
+ 2 M × M × ŝ
M×
dt
Ms
dt
Ms
(25)
where α is the Gilbert damping parameter, β is the strength of the spin-transfer
torque proportional to the spin current density, and ŝ is the unit vector in the
direction of the spin-current polarization. All other notations are similar to those
of Eqs. (1) and (2). In fact, Eq. (25) is an extension of Eq. (1), which takes into
account magnetic damping and the STT effect. Within this model, the third term
is mathematically very similar to the second one. Correspondingly, one expects
that spin current with an appropriate polarization can reduce magnetic damping of
propagating spin waves [78].
The above spin-wave damping compensation has been recently demonstrated
experimentally [79]. Figure 23 shows the schematic of the test devices. They
are based on a 20 nm thick YIG film grown by the pulsed laser deposition on
Gadolinium Gallium Garnet (GGG) (111) substrate. The film is covered by an
8 nm thick layer of Pt deposited using dc magnetron sputtering and the YIG/Pt
bilayer is patterned by e-beam lithography into a stripe waveguide with the width of
1 μm. The system is insulated by a 300 nm thick SiO2 layer, and a broadband
3 μm wide microwave antenna made of 250 nm thick Au is defined on top of
the system by the optical lithography. The waveguide is magnetized by the static
magnetic field He = 1000 Oe applied in its plane perpendicular to the long axis.
A dc electrical current I flowing in the plane of the Pt film is converted by the
SHE into the transverse spin accumulation (see inset in Fig. 23). The associated
pure spin current IS is injected into the YIG film resulting in a spin-transfer torque
on its magnetization. Depending on the relative orientation of the current and the
318
S. O. Demokritov and A. N. Slavin
Pt
Microwave
current
Spin-wave
antenna
YIG
I
IS
He
M
Spin wave
I
z
x
y
Probing
laser light
Pt (8 nm) /
YIG (20 nm)
waveguide
GGG substrate
Fig. 23 Schematic of the experiment on compensation of the spin-wave damping by pure spin
current. Inset illustrates the generation of the pure spin current by the spin-Hall effect. (Reprinted
from [79], with the permission of AIP Publishing)
static magnetic field, the spin current either compensates or enhances the effective
magnetic damping in the YIG film.
The effects of spin current on the decay length of propagating spin waves were
performed by applying a microwave signal at the frequency corresponding, for the
given conditions, to a spin wave with the wavelength of about 5 μm, which can be
efficiently excited by the used 3 μm wide inductive antenna and possess sufficiently
large group velocity. The propagation of spin waves was mapped by rastering
the probing laser spot over the surface of the YIG waveguide with step sizes of
200 and 250 nm in the transverse and the longitudinal directions, respectively.
Figure 24a shows a representative map of the BLS intensity, proportional to the
local intensity in the spin wave, obtained for I = 2.55 mA. As seen from these data,
the spin wave propagates along the waveguide nearly uniformly without changing
its transverse profile (inset in Fig. 24a), which is a clear signature of the singlemode propagation regime caused by the strong separation of the transverse modes
in a narrow waveguide. The intensity of the wave decreases by only 60% over the
propagation path of 10 μm. To characterize the decay length of spin waves and its
dependence on the current, we plot in Fig. 24b the dependences of the spin-wave
intensity on the propagation coordinate obtained for different dc currents in the Pt
layer. These data show that spin waves in the waveguide experience well-defined
exponential decay (note the logarithmic vertical scale) ∼ exp(−2y/ξ), where ξ is the
decay length defined as a distance over which the wave amplitude decreases by a
factor of e. As seen from the figure, the decay length strongly increases with the
increase of the dc current, as expected for the effect of spin current on the effective
magnetic damping.
6 Spin Waves
319
Fig. 24 (a) Normalized spatial intensity map of the propagating spin wave excited by the antenna.
The map was recorded for I = 2.55 mA. The mapping was performed by rastering the probing spot
over the area 1.6 by 10 μm, which is larger than the waveguide width of 1 μm. Dashed lines
show the edges of the waveguide. Inset shows the transverse profile of the spin-wave intensity. (b)
Dependences of the spin-wave intensity on the propagation coordinate for different currents, as
labeled, in the log-linear scale. Lines show the exponential fit of the experimental data. (c) Current
dependences of the decay length and the decay constant. Vertical dashed line marks IC . Solid line is
the linear fit of the experimental data at I < IC . The data were obtained at He = 1000 Oe. (Reprinted
from [79], with the permission of AIP Publishing)
320
S. O. Demokritov and A. N. Slavin
Figure 24c summarizes the results of the spatially resolved measurements. The
decay length (up-triangles) monotonously increases with the increase of I < IC
and then shows an abrupt decrease at I > IC in contradiction to naive expectations
that for large values of I the magnetic damping should be overcompensated by the
spin current, and the propagating spin wave should be amplified. This experimental
observation can be attributed to the strong nonlinear scattering of the propagating
spin waves from large-amplitude current-induced magnetic fluctuations, which have
been observed independently. To characterize the variation of the decay length
with current in detail, we plot in Fig. 24c its inverse value – the decay constant
(down-triangles), which is proportional to effective Gilbert damping constant αeff .
In agreement with the simple theoretical model assuming the linear variation of αeff
with current, the decay constant shows a linear dependence on I. By extrapolating
this dependence to I = 0, we obtain the propagation length at zero current
ξ0 = 2.4 μm. Additionally, one expects the linear dependence in Fig. 24c to cross
zero at I = IC , which corresponds to an infinitely large decay length under conditions
of the complete damping compensation. The data of Fig. 24c show, however, that
the linear fit yields the intercept value larger than IC . This disagreement can be
attributed to the Joule heating of the waveguide by the electric current in Pt resulting
in the significant reduction of the effective magnetization. Since the decay length is
proportional to the group velocity, which is known to decrease with the decrease
in Meff , the effects of the heating on the propagation length counteract those of the
spin current and do not allow one to achieve the decay-free propagation regime. We
note, however, that the maximum achieved propagation length of 22.5 μm is nearly
by a factor of two larger compared to the value of 12 μm estimated for a waveguide
made of a bare YIG film without Pt on top (α = 5 × 10−4 ).
Spin Waves in 0D
The STT effect discussed above is of a particular importance for magnetic systems
fully confined in all three directions. It is now well established that a spin-polarized
electric current or, alternatively, a pure spin current, injected into a ferromagnetic
layer through a nanocontact exerts a torque on the magnetization, leading to a
strongly localized microwave-frequency precession of magnetization, which can be
considered as a 0D spin-wave mode. This phenomenon can serve as a basis for the
development of tunable nanometer-size microwave oscillators, the so-called spintorque nano-oscillators (STNO) [25, 27–29, 66, 67]. The density of magnetic energy
in auto-oscillations excited by STT in a magnetic nanocontact could be very high.
Therefore, the STT-induced precession modes are, usually, strongly nonlinear. Also,
since the spin precession excited in a magnetic nanocontact is, usually, surrounded
by a 2D film or is coupled to a 1D waveguide, it may radiate propagating spin waves.
All these makes the phenomena connected with the STT-driven magnetization
dynamics multifarious and intriguing. This section is devoted to the 0D spin-wave
modes driven by spin-polarized electric or pure spin currents.
6 Spin Waves
321
Probing laser light
Top electrode
Current
flow
200
nm
Insulator
500
nm
Py film
Nanopillar
Bottom electrode
Fig. 25 Schematic of the studied STNO with an AFM image superimposed. The devices consist
of an extended 6 nm thick Permalloy free layer and an elliptical nanopillar formed by a 9 nm thick
Co70 Fe30 polarizing layer and a 3 nm thick Cu spacer. The nanopillar is located close to the edge of
the top electrode enabling optical access to the free layer for BLS microscopy. Magnetic precession
in the device is induced by dc current flowing from the polarizer to the free layer. The spatially
resolved detection of spin waves is accomplished by focusing the probing laser light into a 250 nm
spot, which is scanned over the surface of the Py film. (Reprinted from [31], with the permission
of Springer Nature)
Spin-Torque Nano-Oscillator (STNO) and Emitted Spin Waves
Let us consider an STNO shown in Fig. 25 [30]. The device is formed by a
nanocontact on an extended Permalloy (Py) film. The nanocontact is shaped as
an elliptical nanopillar formed by the nanopattered polarizing Co70 Fe30 layer and
a Cu spacer. A dc current I flowing from the polarizer to the Py film induces
local magnetization oscillations in this film. The nanocontact is located within 200
nanometers from the edge of the top device electrode, enabling optical access to the
Py film at larger distances. The spatially resolved detection of spin waves emitted by
STNO was performed by micro-focus BLS spectroscopy, as described above. The
probing laser light was focused onto the surface of the Py film and scanned in plane
to record two-dimensional maps of the spin-wave intensity.
The oscillation characteristics of STNOs were determined from the microwave
signals generated due to the magnetoresistance effect as shown in Fig. 26. The plots
of the power spectral density (PSD) illustrate the dependence of the oscillation
frequency on the bias current I for three different angles ϕ between the in-plane
bias magnetic field He = 900 Oe and the easy axis of the nanostructured polarizer.
The microwave generation starts at an onset current I = 2.5–3.5 mA that depends
on ϕ. The dependence of the generation frequency on current above the onset is
caused by the nonlinear frequency shift, due to a combination of the demagnetizing
322
S. O. Demokritov and A. N. Slavin
Frequency, GHz
He
He
He 5°
8.2
8.2
8.2
8.0
8.0
8.0
7.8
7.8
7.8
7.6
7.6
7.6
7.4
7.4
7.4
7.2
7.2
2
3 4
I, mA
5
45°
25°
7.2
2
3
PSD,
-5
pW/MHz 10
4
2
5
-4
10
10
-3
3
10
-2
4
10
5
-1
Fig. 26 Pseudo-color logarithmic maps of the power spectral density (PSD) of the signal
generated by the device due to the magnetoresistance effect at different angles ϕ between the
in-plane magnetic field and the easy axis of the nanopillar, as labeled. (Reprinted from [31], with
the permission of Springer Nature)
effects in Py and the dipolar field of the structured Co70 Fe30 polarizer. For small ϕ,
the nonlinear shift is strongly negative. It becomes less pronounced with increasing
ϕ and changes to positive at small I and ϕ > 20◦ . The region of positive nonlinear
frequency shift is reduced at larger He and eventually disappears for He > 1200 Oe,
suggesting its origin from the dipolar field of the polarizer. The possibility to control
the nonlinear behaviors by varying the angle ϕ makes the studied STNOs uniquely
suited for the analysis of the effects of the nonlinearity on the spin-wave emission.
Figure 27 shows two-dimensional intensity maps of spin waves emitted by
STNO at I = 5 mA, measured for different in-plane directions of the applied field
He = 900 Oe. As seen in Fig. 27, the emission mainly occurs in the direction
perpendicular to the in-plane field, regardless of its orientation, the generation
frequency, or the magnitude of the nonlinear frequency shift. We note that although
the sign of the nonlinear shift is expected to be important for the efficiency of
spin-wave emission, the maps of Fig. 27 corresponding to significantly different
nonlinear behaviors of the STNO (see Fig. 26) differ predominantly by the direction
of emission, which rotates together with the field.
Figure 28 illustrates the spin-wave characteristics determined at the location of
the maximum spin wave intensity. Figure 28a–c show the BLS spectra of the emitted
spin waves for I increasing from 3 to 5 mA, together with the spectrum of the
thermal spin waves. The spectra exhibit a small nonlinear frequency shift at ϕ = 45◦ ,
which increases as ϕ is reduced, in agreement with the electrical measurements
shown in Fig. 26. Figure 28d summarizes the dependences of the frequencyintegrated spin-wave intensity on the current I. As seen from these data, the intensity
of the emitted spin waves increases linearly with current for the angle ϕ = 45◦
characterized by a small nonlinear frequency shift. In contrast, the data for ϕ = 25◦
6 Spin Waves
323
a
b
He
He
He
He
500 nm
c
d
He
He
He
He
Normalized
intensity 0.0
0.2
0.4
0.6
0.8
1.0
Fig. 27 Normalized color-coded intensity maps of spin waves emitted by the STNO, recorded
at different angles ϕ between the in-plane magnetic field He = 900 Oe and the easy axis of the
elliptical nanopillar: (a) ϕ = 5◦ o, (b) ϕ = 25◦ o, (c) ϕ = 45◦ o, (d) ϕ = − 45◦ o. The bias current is
I = 5 mA. The schematic of the top electrode is superimposed on each map, with a cross indicating
the location of the nanocontact. The intensity maps acquired at I = 0 were subtracted to eliminate
the contribution from the thermal spin waves. Arrows show the direction of the static magnetic
field, and the dashed lines indicate the direction of the spin-wave emission. (Reprinted from [31],
with the permission of Springer Nature)
and ϕ = 5◦ exhibit a decrease of the spin-wave intensity starting from a certain
value of current that decreases with decreasing ϕ. These findings are correlated
with a larger nonlinear frequency shift, resulting in more significant reduction of the
emission frequency far below FMR. In contrast, magnetoresistance measurements
(Fig. 26) showed similar monotonic increases of generated power for all three
configurations. Therefore, the decrease in the BLS intensity is associated with a
decreased emission efficiency rather than a reduced amplitude of the oscillation in
the nano-contact area.
However, one should admit that decay length of the emitted waves was rather
small, below 500 nm. Further studies [16, 73, 80] have shown that the spin waves
emitted in these experiments have an evanescence nature, since their frequency were
slightly below the spin-wave spectrum of the surrounding Py film. As demonstrated
in Fig. 29, microwave parametric pumping can be used as a mechanism for the
transfer of the generated microwave energy into the desirable spectral range above
the FMR frequency [80]. This effect enables an increase of the propagation length of
a
c
7
45°
8
9 10 11
Frequency, GHz
He
25°
7
8
9 10 11
Frequency, GHz
He
b
8
9 10 11
Frequency, GHz
d
Integral intensity
BLS intensity, a.u.
7
He 5°
BLS intensity, a.u.
S. O. Demokritov and A. N. Slavin
BLS intensity, a.u.
324
3
4
I, mA
5
Fig. 28 (a–c), Dependence of BLS spectra on the current for different in-plane directions of
He = 900 Oe, as indicated. Shadowed regions show the spectrum of the thermally excited spin
waves determined by measurement at I = 0. Color lines show the spectra acquired at the currents
of 3.0 mA (black), 3.5 mA (blue), 4.0 mA (green), 4.5 mA (red), and 5.0 mA (pink). Dashed
vertical lines mark the frequency of the ferromagnetic resonance (FMR). (d) Dependences of the
integrated intensity of emitted spin waves on current for ϕ = 5◦ (triangles), 25◦ (squares), and 45◦
(dots). (Reprinted from [31], with the permission of Springer Nature)
spin waves emitted by STNOs: the decay length of 540 nm for the auto-emission was
increased to 940 nm for the pumping-induced emission. Moreover, the phenomenon
of the pumping-induced emission does not disturb the unique directionality found
for the emission in the auto-oscillation regime, as illustrated by Fig. 29.
Spin-Hall Nano-Oscillator (SHNO)
In the previous section, a STNO driven by spin-polarized electric current was
considered. Another possibility to inject angular momentum into a magnetic system
is utilization of pure spin current. As it has been already mentioned above, the
application of pure spin current has numerous advantages compared to the spinpolarized electric current when the excitation of a large-amplitude 0D spin-wave
modes is discussed. A complete compensation of damping by the spin current
6 Spin Waves
325
Fig. 29 Pseudocolor spatial intensity maps of the emitted spin waves, acquired at I = 5 mA.
A schematic of the top electrode and a cross indicating the location of the nanocontact is
superimposed on each map. (a) Spin-wave auto-emission, in the absence of the external pumping
microwaves. (b) Spin-wave emission under influence of parametric pumping. Note an extended
spin-wave propagation area for (b). (Reprinted with permission from [80], © 2011 by the American
Physical Society)
appears to be a straightforward extension of the damping reduction, described in
Sect. “Spin-Torque Transfer Effect and Spin Waves”. However, as the compensation
point is approached, additional nonlinear damping emerges due to the nonlinear
interactions among different dynamical modes enhanced simultaneously by the spin
current, preventing the onset of auto-oscillation. Since magnon-magnon scattering
rates are proportional to the populations of the corresponding modes, detrimental
effects of nonlinear damping can be avoided by selectively suppressing all the
modes, except for the ones that can be expected to auto-oscillate. To achieve
selective suppression, the frequency-dependent damping caused by the spin-wave
radiation was used. To take advantage of this radiative damping, the spin current
was locally injected into an extended magnetic film, in contrast with the geometry
described in Sect. “Spin-Torque Transfer Effect and Spin Waves”. In fact, the local
spin current enhances a large number of dynamical modes, but those having higher
frequencies, and, consequently, higher group velocities, quickly escape from the
active region, which results in their efficient suppression by the radiation losses.
Meanwhile, the modes at frequencies close to the bottom of the spin-wave spectrum
have a much smaller group velocity, and, therefore, minimum radiation losses.
The scheme of our experiment with pure spin current is shown in Fig. 30a [31].
The studied device is formed by a bilayer of a 8 nm thick film of Pt and a 5 nm thick
film of Py patterned into a disk with a diameter of 4 μm. Two 150 nm thick Au
electrodes with sharp points separated by a 100 nm wide gap are placed on top of
the bilayer, forming an in-plane point contact. The sheet resistance of the bilayer is
nearly two orders of magnitude larger than that of the Au electrodes. Consequently,
the electrical current induced by voltage between the electrodes should be strongly
localized in the gap. Indeed, a calculation of the current distribution through a
326
a
DC
Current
Probing laser
light
Py(5)Pt(8)
Disk
He
1μ
b
Normalized
current density
Fig. 30 (a)
Scanning-electron
microscopy image of the test
spin-Hall nano-oscillator. The
device consists of a 4 μm
diameter disk formed by a
8 nm thick Pt on the bottom
and a 5 nm thick Py layer on
top, covered by two pointed
Au(150 nm) electrodes
separated by a 100 nm gap.
(b) Normalized calculated
distribution of current
through the section of the
device shown in the inset by a
dashed line. (Reprinted from
[30], with the permission of
Springer Nature)
S. O. Demokritov and A. N. Slavin
Au(150)
Top electrodes
m
1
z
0.5
250 nm
0
-2
-1
0
z, μm
1
2
section across the middle of the gap (Fig. 30b) shows that most of the current flows
through a 250 nm wide Pt strip. This electric current creates a pure spin current
flowing into Py, due to the spin-Hall effects. The spin current injected into Py exerts
spin-transfer torque on its magnetization. As a result, the damping is compensated,
and the dynamic magnetic modes are enhanced.
Figure 31 shows the BLS spectra obtained with the probing spot positioned in
the center of the gap between the electrodes, at different values of the dc current
I. At I = 0, the BLS spectrum exhibits a broad peak corresponding to incoherent
thermal magnetization fluctuations in the Py film (Fig. 31a). As this thermal peak
grows with increasing current, its rising front becomes increasingly sharper than
the trailing front, consistent with the preferential enhancement of the low-frequency
modes. Analysis of the dependence of the frequency-integrated BLS intensity on
current (Fig. 31b) shows that the intensity of magnetic fluctuations diverges as the
current approaches a critical value of Ic ≈ 16.1 mA. In contrast to confined systems
driven by spatially uniform spin currents [81], the intensity of fluctuations does
not saturate as the current approaches Ic , indicating that the nonlinear processes
preventing the onset of auto-oscillations are avoided.
At I ≥ Ic , a new peak appears in the BLS spectrum below the thermal peak, as
indicated in Fig. 31a by an arrow. The calculated current density in the center of the
gap at the onset is 3 × 108 A/cm2 , which is only slightly larger than the extrapolated
value 1 × 108 A/cm2 obtained for a similar system without radiation losses [81].
Since this peak is not present in the thermal fluctuation spectrum, we can conclude
that it corresponds to a new auto-oscillation mode that does not exist at I < Ic . The
peak rapidly grows and then saturates above 16.3 mA (Fig. 31c–d). Comparing the
Fig. 31 (a) BLS spectra of thermal fluctuation amplified by the spin current at currents below the onset of auto-oscillation. (b) Integral intensity of amplified
thermal fluctuations and its inverse versus current. Both dependencies are normalized by their values at I = 0. (c) BLS pectra of the magnetization autooscillation driven by the spin current. Filled areas are the results of fitting by the Gaussian function. Note, that the spectral widths are determined by the
resolution of the BLS setup. (d) The intensity and the center frequency of the auto-oscillation peak versus current. Curves are guides for the eye. Reprinted
from [30], with the permission of Springer Nature
6 Spin Waves
327
328
S. O. Demokritov and A. N. Slavin
spectra for I = 16.1 mA and 16.3 mA, we see that the onset of auto-oscillations is
accompanied by a decrease in the intensity of thermal fluctuations, suggesting that
the energy of the spin current is mainly channelled into the auto-oscillation mode.
The spectral width of the auto-oscillation peak characterizing the coherence
of auto-oscillations decreases just above Ic , and stabilizes above 16.3 mA. Note
that the linewidth in the spectra shown in Fig. 31a and c is determined by the
spectral resolution of our optical technique under usual conditions. Additional
measurements at our instrument’s ultimate spectral resolution of 60 MHz show that
the actual linewidth in the saturated regime is below this value, suggesting a high
degree of coherence of the observed auto-oscillation mode.
The frequency of the auto-oscillation peak monotonically decreases with increasing I (Fig. 31d). We note that the generated frequency is significantly below the
frequencies of magnetic fluctuations even at the onset of auto-oscillations. We draw
three important conclusions based on this observation. First, the auto-oscillation
mode does not belong to the thermal spin-wave spectrum. Second, this mode is
formed abruptly at the onset current, and not by gradual reduction of frequency
from the spin-wave spectrum due to the red nonlinear frequency shift. Third, since
the energy can be radiated only by propagating spin waves and there are no available
spin-wave spectral states at the auto-oscillation frequency, the auto-oscillation mode
is not influenced by the radiation losses.
To determine the spatial profile of the auto-oscillation mode, we performed
two-dimensional mapping of the dynamic magnetization at the frequency of autooscillations, by rastering the probing laser spot in the two lateral directions and
simultaneously recording the BLS intensity. An example of the obtained maps
is presented in Fig. 32. These data show that the auto-oscillations are localized
in a very small area in the gap between the electrodes. The spatial distribution
Fig. 32 Normalized
color-coded map of the
measured BLS intensity over
the auto-oscillation area, and
two orthogonal sections
through its center. Symbols
are the experimental data, and
filled areas under solid curves
are the results of fitting by a
Gaussian function. Dashed
lines on the map show the
contours of the top electrodes.
The data were recorded at
I = 16.2 mA. (Reprinted from
[30], with the permission of
Springer Nature)
6 Spin Waves
329
of the BLS intensity is well described by a Gaussian function with the width
of 250 ± 10 nm, close to the diameter of the probing laser spot. The measured
spatial distribution is a convolution of the actual spatial profile with the instrumental
function determined by the shape of the laser spot. Therefore, we estimate that
the size of the auto-oscillation region is less than 100 nm, significantly smaller
than the characteristic size of the current localization (Fig. 30b). Therefore, we
conclude that the auto-oscillation area is determined not by the spatial localization
of the driving current, but by the nonlinear self-localization processes defining the
geometry of a standing spin-wave “bullet” [82]. We emphasize that the observed
quick saturation of the intensity of the auto-oscillation peak above the onset and
its monotonic red frequency shift are the intrinsic characteristics of the “bullet”
mode. Only one “bullet” mode exists at the frequency of the auto-oscillations and
this frequency is well separated from the continuous spectrum of non-localized spin
waves, Therefore, our findings provide strong evidence for that auto-oscillations
involve only a single mode in the studied system.
Nature of Spin Wave Modes Excited in 0D Magnetic Nanocontacts
The nature of the auto-oscillation spin wave mode excited by either spin-polarized
or pure spin current in magnetic nanocontacts (0D objects) is of a fundamental
importance for the current-induced magnetization dynamics.
The first theoretical analysis of the nature of the spin-wave eigen-mode excited
by spin-polarized current in a nano-contact geometry was performed by J. Slonczewski [8]. He developed a spatially nonuniform linear theory of spin wave
excitations in a nano-contact, where the “free” ferromagnetic layer is infinite in
plane, while the spin-polarized current traversing this layer has a finite cross-section
S = π Rc2 , where Rc is the contact radius. Considering a perpendicularly magnetized
nano-contact Slonczewski showed that in the linear case the lowest threshold
of excitation by spin-polarized current is achieved for an exchange-dominated
propagating cylindrical spin wave mode having wave number q0 = 1.2/Rc and
frequency [8]:
ω (q0 ) = ω0 + Dex q02 .
(26)
Here ω0 is the ferromagnetic resonance (FMR) frequency in the magnetic film,
2 , ω ≡ 4π γ M , γ is the gyromagnetic ratio for electron spin, l
Dex = ωM lex
M
S
ex =
1/2
2
A/2π MS
is the exchange length, A is the exchange constant, and MS is the
value of the saturation magnetization.
It was also shown that the threshold current Ith in such a geometry consists of
two additive terms: the first one arises from the radiative loss of energy carried by
the propagating spin wave out of the region of current localization, while the second
one is caused by the usual energy dissipation in the current-carrying region:
330
S. O. Demokritov and A. N. Slavin
lin
Ith
= 1.86
D
(H )
.
+
σ
σ Rc2
(27)
Here σ = εgμB /2eMS dS whereε is the spin-polarization efficiency defined in
[8], g is the spectroscopic Lande factor, μB is the Bohr magneton, e is the modulus
of the electron charge, d is the thickness of the “free” magnetic layer, S is the crosssection area of the nano-contact), and (H) is the spin wave damping dependent
on the bias magnetic field H. It turns out that for a typical nano-contact of the
radius Rc = 20 − 30 nm the radiative losses are about one order of magnitude
larger than the direct energy dissipation, and should give the main contribution into
the threshold current. This result, however, contradicts experimental observations
(see, e.g., [83]): the experimentally measured magnitude of the threshold current
in an in-plane magnetized nano-contact is much smaller than the value predicted
by Eq. (27), although the dependence of this current on the magnetic field H is
satisfactory described by this equation.
In this section we present a spatially nonuniform nonlinear theory of spin wave
excitation by spin-polarized current in a nano-contact geometry for the case of the
in-plane magnetization [82]. We show that in an in-plane magnetized magnetic film
the competition between the nonlinearity and exchange-related dispersion leads to
the formation of a stationary two-dimensional self-localized nonpropagating spin
wave mode. Such nonlinear self-localized wave modes in two- or three-dimensional
cases are conventionally called wave “bullets” [84]. The frequency of this spin wave
“bullet” is shifted by the nonlinearity below the spectrum of linear spin waves
and, therefore, this nonlinear mode has an evanescent character with vanishing
radiative losses, which leads to a substantial decrease of its threshold current Ith
in comparison to the linear propagating mode shown in Eq. (27).
To describe the generation of a spin wave bullet by the spin-polarized current
we consider a “free” ferromagnetic layer, infinite in y − z plane and having finite
thickness d in the x direction (d is assumed to be sufficiently small for us to consider
that the magnetization M is constant along the film thickness, and that the dipoledipole interaction can be described by a simple demagnetization field). We assume
that the internal magnetic field H = Happ + Hex , consisting of the applied Happ and
interlayer exchange Hex fields, is applied in the z direction in the film plane. Using
the standard Hamiltonian spin-wave formalism [33], which has been successfully
used to develop a spatially uniform nonlinear model of spin wave generation by
spin-polarized current [85, 86], one can derive an approximate equation for the
dimensionless complex spin wave amplitude b ≡ b(t, r):
∂b
= −i ω0 b − DD b + N |b|2 b − b + f (r/Rc ) σ I b − f (r/Rc ) σ I |b|2 b.
∂t
(28)
√
Here ω0 ≡
ωH (ωH + ωM ) is the linear FMR frequency, (ωH ≡ γ H,
DD ≡ (2A/MS ) ∂ω0 /∂H = (2γ A/MS )(ωH + ωM /2)/ω0 is the dispersion coefficient
6 Spin Waves
331
for spin waves,
is the two-dimensional Laplace operator in the film plane,
N = − ωH ωM (ωH + ωM /4)/ω0 (ωH + ωM /2) is the coefficient describing nonlinear
frequency shift, and ≡ α G (ωH + ωM /2) is the spin wave damping rate (α G is
the dimensionless Gilbert damping parameter). The dimensionless function f (x)
describes the spatial distribution of the spin-polarized current. The dimensionless
spin wave amplitude b is connected with the z-component of the magnetization by
the equation |b|2 = (MS − Mz )/2MS .
Equation (28) differs from the Eq. (9) in [8] (which resulted in the solution
(27)) by the presence of two additional nonlinear terms: the term containing the
coefficient N and describing a nonlinear frequency shift of the excited mode, and
the last term describing the current-induced positive nonlinear damping that stops
the increase of the amplitude of the excited mode at relatively large currents. Also,
since the Eq. (28) was obtained as a Taylor expansion it is exactly correct only for
sufficiently small spin wave amplitudes |b| < 1.
Without damping and current terms ( = 0, I = 0) Eq. (28) coincides with
the well-known (2 + 1)-dimensional nonlinear Schrödinger equation (NSE) [87].
In the considered case of an in-plane magnetized film the nonlinear coefficient N
is negative, and the nonlinearity and dispersion satisfy the well-known Lighthill
criterion ND < 0 (i.e., they act in opposite directions), and the NSE has a nonlinear
self-localized radially symmetric standing solitonic solution (or the solution in the
form of a standing spin wave bullet)
b (t, r) = B0 ψ (r/) e−iωt ,
(29)
where dimensionless function ψ(x), having maximum value of 2.2 at x = 0,
describes the profile of the bullet. This function is the localized solution of the
equation
ψ +
1
ψ + ψ 3 − ψ = 0,
x
(30)
which has to be found numerically (see e.g., [84]).
In Eq. (29) B0 , , and ω are the characteristic amplitude, characteristic size, and
frequency of the bullet, respectively. Among these three parameters only one is
independent. Taking the amplitude B0 as an independent parameter, we can express
the two other parameters as
√
ω=
ω0 + NB 20 ,
=
|D/N|
.
B0
(31)
We would like to stress that the frequency of the spin wave bullet lies below
the linear frequency ω0 of the ferromagnetic resonance (see Eq. (31), and note that
N < 0), i.e., outside the spectrum of linear spin waves. This is the main reason for
the self-localization of the spin wave bullet, as the effective wave number of the spin
332
S. O. Demokritov and A. N. Slavin
wave mode with frequency (6) is purely imaginary. It also follows from Eq. (29) and
the expansion condition |b| < 1 that the maximum magnitude of B0 for which our
perturbative approach is still correct is B0 = 0.46.
It is well known [87] that the bullet-like solutions of (2 + 1)-dimensional NSE
are unstable with respect to the small perturbations: the wave packets having the
bullet shape Eq. (29), but amplitudes smaller than B0 , decay due to the dispersion
spreading, while the wave packets having amplitudes higher than B0 collapse due
to the nonlinearity. At the same time, Eq. (28) with both Gilbert dissipation and
current I is a two-dimensional analog of a Ginzburg-Landau equation that is known
to have stable localized solutions (see, e.g., review [88]). One can assume that for
a small damping rate and current I the full nonconservative eq. (28) will have a
bullet-like solution, only slightly different from the exact solution Eq. (29) of the
conservative NSE equation. It is clear, however, that not all of such solutions can be
supported in our case. For example, small-amplitude bullets, for which > > Rc ,
practically do not interact with the spatially localized current and will decay due to
the linear dissipation. The large-amplitude (B0 ≥ 1) bullets, on the other hand, will
also decay because the effective damping − σ I(1 − |b|2 ) for them changes sign
and becomes positive.
The excitation threshold of the spin wave bullet mode was calculated in [82] and
the minimum value of this threshold turned out to be equal to the second term in Eq.
(27), i.e., sobstantially lower than the threshold of excitation of the propagating
spin-wave mode in the perpendicularly magnetized magnetic nanocontact
Eq. (27).
To find the spatial profile of the spin-wave bullet mode Eq. (28) was solved
numerically. The results of comparison of the spin-wave excitation profiles at
the threshold obtained for a typical set of experimental parameters [83] from the
analytical solution Eq. (29) (solid black line) and numerical solution of Eq. (28)
(black dots) are shown in Fig. 33. One clearly sees that the numerical profile of
the nonlinear eigen-mode is practically indistinguishable from the approximate
“bullet-like” profile, so the “bullet” model works exceptionally well in this case.
For comparison we present in Fig. 33 the spatial profile of the Slonczhewski-like [8]
linear mode, that is obtained from the solution of Eq. (28) where the nonlinear terms
(terms containing |b|2 ) are omitted (red line). The amplitude of this linear mode at
the threshold is vanishingly small, |b(r)|2 → 0. We also present the normalized
spatial profile of a bullet mode above the thershold numerically calculated form Eq.
(28), to show that with the increase of the bullet amplitude its width decreases in
accordance with the experssion shown in Eq. (31) (blue line).
As it was mentioned above, the bullet mode is excited in an in-plane magetized
nanocontact, while the linear propagating spin wave mode (Slonczewski mode [8])
could be excited in a perpendicularly magnetized nanocotact. It is also well-known
that in the case when the direction of the external bias magnetic field varies from inplane to the perpendicular the coefficient of the nonlinear frequency shift N changes
its sign from negative to positive [89, 90]. Therefore, it is interesting to see the nature
of the spin-wave mode excited by a spin-polarized (or pure spin) current under the
oblique magnetization of a nanocontact.
333
Normalized power |b(r)|2/|b(0)|2
6 Spin Waves
1.0
Linear mode
0.8
Nonlinear "bullet"
0.6
Nonlinear "bullet"
above the threshold
0.4
0.2
0.0
0.0
0.5
1.0
1.5
2.0
2.5
3.0
Normalized distance r/Rc
Fig. 33 Normalized profiles of the spin wave mode generated by spin-polarized current at (the
threshold: solid black line – bullet profile (29), circles – result of the numerical solution of Eq.
(28), red line – profile of the linear eigen-mode calculated from the linearized Eq. (28). Blue line
demonstrates the numerically calculated profile of the bullet mode above the thershold. Vertical
dash-dotted line shows the region of the current localization. The parameters are: 4π M0 = 16.6 kG,
Happ = Hex = 5 kOe, A = 2.85 · 10−6 erg/cm, α G = 0.015, d = 1.2 nm, Rc = 25 nm, ε = 0.3
The analytic [91], numerical [92], and experimental [93] studies of the spin wave
excitation in current-driven magnetic nanocontacts were successfully performed,
and confirmed the conclusions of the above-presented theory.
In particular, the experimental study [93] was performed on a nanocontact
2Rc = 40 nm to the thin film tri-layer Co81 Fe19 (20 nm)/Cu(6 nm)/Ni80 Fe20 (4.5 nm),
patterned into a 8 μ m × 26μ m mesa. On top of this mesa, a circular Al nanocontact
was defined through SiO2 using e-beam lithography and an external magnetic field
of a constant magnitude (μ0 He = 1.1 T) was applied to the sample at an angle
θ e with respect to the film plane. Microwave excitations were only observed for a
single current polarity, corresponding to electrons flowing from the “free” (thinner)
to the “fixed” (thicker) magnetic layer. All measurements were performed at room
temperature.
Figure 34 shows the detailed angular dependence of the microwave frequencies
generated at a constant current of I = 14 mA and a constant magnetic field amplitude
of μ0 He = 1.1 T. The generated frequencies are approximately independent of
the magnetic field angle up to about θ e = 35◦ , and then decrease from about
35 GHz to 10 GHz with the increasing angle. The most important feature of the
results presented in Fig. 34 is the existence of two distinct and different modes for
sufficiently small values of θ e – linear propagating spin wave Slonczewski mode,
having a higher frequency, and a nonlinear self-localized spin-wave bullet mode
having a lower frequency. The frequencies of these two modes differ by about
2.5 GHz at angles up to θ e = 40◦ , and then they start to approach each other
up to θ e ≈ 55◦ where the mode having lower frequency completely disappears.
334
S. O. Demokritov and A. N. Slavin
40
Propagating mode
Spin wave bullet
40
20
f [GHz]
f [GHz]
30
30
20
10
10
0
0
0
0
15
30
45
60
75
Applied field agle θe [deg]
15
30
45
90
60
75
90
Applied field agle θe [deg]
Fig. 34 Experimental frequencies of the current-induced spin wave modes as a function of the
applied field angle θ e between the external bias magnetic field and the nanocontavt plane at I = 14
mA and μ0 He = 1.1 T (symols). Inset: theoretically calculated frequencies of the propagating
(upper curve) and “bullet” (lower curve) modes at the threshold of their excitation for nominal
parameters of the nanocontact sample. (Reprinted with permission from [82], © 2005 by the
American Physical Society)
The behavior of the frequencies of the current-induced spin wave modes shown in
Fig. 34 remains the same for any magnitude of the bias current that is larger than
I ≈ 10 mA. This behavior is also qualitatively similar to the behavior of the mode
frequencies derived from analytic theory (see Fig. 2 in [91] and the inset in Fig. 34)
and from the numerical simulation (see Fig. 4 in [92]).
The experimental threshold currents for the two excited spin wave modes shown
in Fig. 35 as functions of θ e were determined using the method proposed in [90]. The
graph in Fig. 35 only shows the threshold currents determined from experiment for
the magnetization angles 20◦ < θ e < 80◦ , since outside this range the signal was too
weak to allow reliable determination of the excitation threshold. It is clear that the
lower-frequency (bullet) mode has a lower threshold current at low magnetization
angles. As the angle increases, the threshold currents for the two modes gradually
approach each other and become essentially equal close to the critical angle
θ e ≈ 50◦ , where the low-frequency mode disappears. These experimental data are
also qualitatively similar to the threshold curves calculated analytically in [91] (see
solid lines in Fig. 35) and the similar curves simulated numerically (see Fig. 3 in
[92]). The inset in Fig. 35 shows the numerically calculated profile of the both
modes: nonlinear self-localized bullet mode (left frame) and linear propagating
mode (right frame). Thus, the results of the laboratory experiment [93] and the
results of the numerical simulations [92] fully confirmed the theoretical ideas [8,
82] about the possibility of current-induced excitation of two qualitatively different
spin wave modes in magnetic nanocontacs.
6 Spin Waves
335
Threshold current lth [mA]
30
25
100
100
y [nm]
20
0
−100 −100
0
x [nm]
200
y [nm]
0
−200
200
0 x [nm]
−200
15
10
Propagating mode
Spin wave bullet
5
0
0
15
30
45
60
75
90
Applied field agle qe [deg]
Fig. 35 Measured threshold current for the propagating (empty triangles) and bullet (filled circles)
modes as a function of the applied field angle θ e . Solid lines: thresholds for the same modes versus
applied field angle theoretically calculated using the formalism of [91] (red line – bullet mode,
blue line- propagating mode). Upper inset demonstrates the numerically calculated spatial profiles
of the bullet mode (left) and propagating mode (right). (Reprinted with permission from [82], ©
2005 by the American Physical Society)
It is important to note that the excitation of self-localized evanescent spin wave
bullet modes by pure spin current was subsequently observed in several independent
experiments [31, 94, 95].
Another example of an interesting and unusual solitonic spin wave mode that
can be excited by spin-polarized current is given by the so-called spin-wave droplet
(or spin-wave droplet soliton) existing in perpendicularly magnetized nanocontacts
having a large perpendicular magnetic anisotropy (PMA) [96–98]. As it was first
demonstrated analytically in [99, 100], the Landau-Lifshitz equation for magnetic
films can sustain a family of so-called magnon drop solitons, provided there is no
spin wave damping.
While any realistic magnetic system always exhibits some spin wave damping,
hence making magnon drops unrealistic, it was demonstrated theoretically in [96]
that in a current-driven magnetic nanocontact with PMA, where the spin wave
damping is completely compensated by the STT effect [8], it would be possible
to excite a magnon-drop-like excitations. In contrast to the conservative magnon
drops, these so-called magnetic droplets are strongly dissipative, relying not only
on the zero balance between the exchange and anisotropy, but also on the balance
between the negative damping created by the STT effect and the positive nonlinear
damping in the current-driven magnetic material. As a consequence, out of a large
family of magnon drops, the additional net zero damping condition singles out a
particular magnetic droplet with both a well-defined frequency and a well-defined
direction of the dynamic magnetization at the center of the excited droplet. Note
336
S. O. Demokritov and A. N. Slavin
that at large droplet amplitudes the direction of magnetization at the droplet center
could be almost completely opposite to the similar direction at the droplet periphery
(see Fig. 2 in [96]). These rather exotic nonlinear dynamical magnetic modes
were observed experimentally in [97], and a more detailed description of magnetic
droplets is presented in [98].
Coupling of a STNO and 1D Spin-Wave Waveguide to Each Other
In the previous sections, we have demonstrated that STNO devices, employing 0D
spin-wave modes can convert the energy of direct electrical current into propagating
spin waves. We have also noticed that it is difficult to achieve frequency matching
of STNO with the propagating spin waves, since the large-amplitude spin wave
modes in STNOs are frequency shifted due to nonlinear properties of spin waves
with respect to characteristic frequencies of 1D and 2D propagating spin waves.
However, if one uses a spin-wave waveguide of a particular geometry as described
below, efficient matching between such waveguides and STNOs can be achieved.
This matching is realized by taking advantage of the dipolar magnetic field within
the waveguide, which acts on 1D propagating spin-wave modes [16].
Figure 36a shows the layout of the studied device. A point-contact STNO
is comprised of a multilayer Cu(4)/Co70 Fe30 (4)/Au(150) shaped into an elliptic
nanopillar with dimensions of 120 nm × 40 nm fabricated on top of an extended
5 nm thick Permalloy (Py) film. Additionally, the device incorporates a 5 nm thick
and 200 nm wide Co70 Fe30 nanostripe below the Py film. The device is magnetized
by a static magnetic field He = 800 – 1200 Oe applied in the plane of the Py film
perpendicular to the CoFe nanostripe.
Figure 36b shows the characteristics of the oscillation of STNO determined
by the standard electronic spectroscopy measurements. Above the onset current
of about 3.5 mA, both the amplitude and the frequency of the auto-oscillations
exhibit a smooth dependence on current, indicating a single-mode operation of
the STNO. Correspondingly, Fig. 36c shows representative BLS spectra recorded
with the probing laser spot positioned above the CoFe nanostripe. Note here
that by comparing Fig. 36b and c, one can conclude that the frequency of
the microwave signal is twice the frequency of the magnetization oscillation
measured by BLS, since the former is due to the quadratic magnetoresistance
effect.
While the BLS spectra acquired above the CoFe nanostripe clearly show the
signals resulting from the STNO oscillation, no such signals were detected away
from the nanostripe. This observation indicates that the STNO can efficiently
generate 1D spin waves, propagating along the CoFe nanostripe, but a radiation
of 2D spin waves into the free Py film is inefficient.
To understand this phenomenon, one has to consider the effects of the dipolar
field of the CoFe nanostripe on the internal field in the magnetic layers. Both
micromagnetic simulations and studies of spin-wave spectra of thermal fluctuations
6 Spin Waves
337
a
Top
electrode
Current
200 nm
Py (5) film
z
CoFe (5)
nano-wire
Nanopillar
y
He
Bottom Cu(40)
electrode
c
b
2.5
1.0
BLS intensity, a.u.
PSD, pW/MHz
7 mA
2.0
1.5
6 mA
1.0
5 mA
0.5
4 mA
0
7 mA
0.8
0.6
5.5 mA
0.4
0.2
4 mA
0.0
13
14
Frequency, GHz
15
Frequency, GHz
Fig. 36 (a) Layout of the studied STNO with an incorporated waveguide. Inset: SEM micrograph
of the device. He is the static magnetic field. Numbers in parentheses indicate the thicknesses of
the layers in nanometres. (b) Spectra of the current-induced oscillations of the STNO measured
by a spectrum analyzer at different driving dc currents, as indicated. (c) BLS spectra recorded at
different driving currents measured by positioning the probing laser spot on the nano-waveguide.
Note that the spectral widths are determined by the resolution of the BLS setup. (Reprinted from
[16], with the permission of Springer Nature)
show that the internal field is significantly reduced in the magnetic film in the region
of the CoFe nanostripe, as compared to the magnitude away from the nanostripe.
The reduction of the internal field results in lowering of the local spin-wave
spectrum, creating a one-dimensional channel with allowed spin-wave frequencies
below the bottom of the spectrum in the free Py film. Low-frequency magnons
excited by STNO are directionally guided along the CoFe nanostripe, since there
are no states available at these frequencies in the free Py film. Thus, the staticfield channel induced by the CoFe nanostripe plays the role of a compound dipolar
spin-wave waveguide formed by the strongly exchange-coupled bilayer of the CoFe
nanostripe and the Py film on top of it.
338
S. O. Demokritov and A. N. Slavin
Fig. 37 (a) Normalized decay-compensated spatial map of the spin-wave intensity. The positions
of the top device electrode and the CoFe nanostripe are schematically shown. (b) Measured
dependence of the integral spin-wave intensity on the propagation coordinate (symbols), on
the log-linear scale. The line shows the result of the fitting of the experimental data by the
exponential function. (c) Distribution of the spin-wave intensity in the section transverse to the
nano-waveguide. Symbols are experimental data, curve is a fit by the Gaussian function. w is
the full width at half maximum of the transverse intensity profile. (d) Dependence of w on
the propagation coordinate. Symbols are experimental data, horizontal line is the mean value.
(Reprinted from [16], with the permission of Springer Nature)
6 Spin Waves
339
The measured propagation characteristics of spin waves in the nano-waveguide
are illustrated in Fig. 37. Figure 37a shows the normalized spatial map of the
BLS intensity, which is proportional to the local spin wave intensity. The map was
recorded at a constant dc current of 5 mA by rastering the probing laser spot over
a 1.6 μm by 1.6 μm area with the step size of 100 nm. To highlight the transverse
profile of the propagating wave, the spatial decay in the direction of propagation was
compensated by normalizing the signal with the integral over the transverse section
of the map (along the z-coordinate). The map of Fig. 37a clearly shows that the spin
wave energy is concentrated entirely in the nano-waveguide, i.e., spin waves are
guided by the field-induced channel without noticeable losses associated with the
radiation of energy into the surrounding free Py film.
The BLS intensity integrated over the transverse section of the map exhibits a
simple exponential spatial decay in the direction of propagation (shown on the log
scale in Fig. 37b). We define the decay length ξ as the distance over which the wave
amplitude decreases by a factor of e. By fitting the data of Fig. 37b with the function
exp(−2y/ξ), we obtain ξ = 1.3 μm. We note that this value is close to the best spinwave propagation characteristics obtained in low-loss Py films with comparable
thickness, despite the higher dynamical losses expected due to the stronger damping
in CoFe.
By analyzing transverse cross-sections of the BLS intensity map (Fig. 37c),
we determine the transverse full width at half maximum w of the spin wave
intensity distribution for different positions along the waveguide. The obtained value
w = 320 nm is independent of the propagation coordinate (Fig. 37d), which confirms
that the spin wave is efficiently localized in the waveguide without spreading out.
We note that the measured spatial profile (Fig. 37c) represents a convolution of
the actual profile of the spin wave intensity with the distribution of intensity in
the diffraction-limited probing light spot whose estimated diameter is 250 nm.
The value w = 320 nm is therefore in a reasonable agreement with the measured
waveguide width of 200 nm (inset in Fig. 36a).
Conclusion and Outlook
The post-CMOS information technology will require radically new solutions for
digital and analog information processing. One promising approach is to employ the
spin degree of freedom of electron for information storage and computing, which
is the main focus of the rapidly growing field of spintronics [101–104]. In this
new signal-processing paradigm signals will be codes in terms of a spin angular
momentum that can be carried by either polarized electrons or spin waves.
The use of spin waves (magnons) as carriers of spin angular momentum is
preferable to the use of spin-polarized electrons, because Gilbert magnetic damping,
associated with the transport of spin waves, is, usually, lower than the Ohmic losses
associated with the transport of electrons. The typical medium for the spin wave
propagation in the existing nano-scale magnonics is a soft magnetic metal such
as permalloy (Py). This material choice is mainly dictated by the relative ease of
340
S. O. Demokritov and A. N. Slavin
magnetic information readout via various types of magneto-resistance observed in
metallic ferromagnetic heterostructures. A substantial progress has been made in
this field during the last two decades. In particular, generation of self-sustained
microwave magnetic oscillations by STT effect from spin-polarized currents
[27, 29] as well as pure spin currents arising from spin-Hall effect [31, 94, 95, 105]
have been demonstrated. Novel techniques for precise characterization of magnetization dynamics in nano-scale metallic magnetic systems, such as the technique of
spin-torque ferromagnetic resonance (ST-FMR) [106, 107], have been developed.
In spite of the rapid research progress in the field of metal-based magnonics,
several significant limitations of the metal-based magnonic systems are very evident.
One of them is the relatively large magnetic damping of ferromagnetic metals,
which translates into large spin current densities needed to induce magnetization
switching or self-generation of spin waves in ferromagnetic metals. The large
magnetic damping also results in short propagation lengths (typically ∼1 μm)
of magnons in metallic magnetic nanostructures, which critically hinders the
transition from single magnonic elements to large-scale spintronic circuits based on
propagating magnons. In addition, high electric conductivity of metallic magnets
and the corresponding short charge screening length do not allow to employ
magneto-electric effects such as the recently predicted flexoelectric effect [108–110]
for manipulation of spin waves using electric field.
These drawbacks are absent in magnetic dielectrics, the most common of
which is yttrium iron garnet (YIG) – a ferrimagnetic insulator with very low
magnetic damping (magnon lifetimes reaching 1 μs and magnon propagation length
exceeding 1 cm) [35]. However, the technique of liquid phase epitaxy typically
employed for the growth of high-quality YIG crystals does not allow for deposition
of films sufficiently thin for observation of interfacial spin-dependent phenomena,
which will determine the future of manipulation of spin waves at nanoscale.
The pioneering experiments in YIG-based spintronics performed on the relatively
thick (∼1–3 μm) epitaxial YIG films revealed some weak effects, but failed to
demonstrate reproducible excitation or/and manipulation of propagating magnons
by interfacial spintronic effects [12, 13, 111, 112].
The recently developed methods for growth of ultra-thin (∼ 10 nm) high-quality
YIG films (ferromagnetic resonance (FMR) linewidth ∼3–5 Oe) by pulsed laser
deposition (PLD) [113–115] and patterning of thin YIG films [116] remove major
roadblocks for using magnetic dielectrics in nano-scale spintronic devices and
open a new field of magnon-based spintronics of magnetic dielectrics. Pioneering
experiments in this field performed in the last two years demonstrated excitation
of magnonic signals in magnetic dielectrics by interfacial spin orbit torques,
compensation of magnetic damping in magnetic dielectrics by pure spin currents,
and resulting self-sustained generation of microwave magnetization oscillations in
YIG film samples [117, 118].
We firmly believe that spin waves propagating in magnetic dielectrics and
antiferromagnets will determine the future of the microwave signal processing at
nanoscale, and will form a basis for a new generation of energy-efficient microwave
signal processing devices which will use spin-orbital effects, like the spin-Hall
6 Spin Waves
341
effect [9–11] and inverse spin-Hall effect [119], and electric fields (e.g., through
the flexoelectric effect [109, 110]) for generation, reception, and manipulation of
signals coded in terms of the spin angular momentum and carried by different spinwave modes in magnetic nanostructures.
References
1. Bloch, F.: Zur Theorie des Ferromagnetismus. Z. Phys. 61, 206–219 (1930)
2. Neusser, S., Grundler, D.: Magnonics: spin waves on the nanoscale. Adv. Mater. 21, 2927–
2932 (2009)
3. Kruglyak, V.V., Demokritov, S.O., Grundler, D.: Magnonics. J. Phys. D. Appl. Phys. 43,
264001 (2010)
4. Lenk, B., Ulrichs, H., Garbs, F., Münzenberg, M.: The building blocks of magnonics. Phys.
Rep. 507, 107 (2011)
5. Chumak, A.V., Vasyuchka, V.I., Serga, A.A., Hillebrands, B.: Nat. Phys. 11, 453–461 (2015)
6. Slonczewski, J.C.: Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater.
159, L1–L7 (1996)
7. Berger, L.: Emission of spin waves by a magnetic multilayer traversed by a current. Phys.
Rev. B. 54, 9353–9358 (1996)
8. Slonczewski, J.C.: Excitation of spin waves by an electric current. J. Magn. Magn. Mater.
195, L261–L268 (1999)
9. Dyakonov, M.I., Perel, V.I.: Possibility of orienting electron spins with current. Sov. Phys.
JETP Lett. 13, 467–469 (1971)
10. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834–1837 (1999)
11. Hoffmann, A.: Spin Hall effects in metals. IEEE Trans. Magn. 49, 5172–5193 (2013)
12. Kajiwara, Y., et al.: Transmission of electrical signals by spin-wave interconversion in a
magnetic insulator. Nature. 464, 262–266 (2010)
13. Wang, Z., Sun, Y., Wu, M., Tiberkevich, V., Slavin, A.: Control of spin waves in a
ferromagnetic insulator through interfacial spin scattering. Phys. Rev. Lett. 107, 146602
(2011)
14. Demidov, V.E., Urazhdin, S., Rinkevich, A.B., Reiss, G., Demokritov, S.O.: Spin Hall
controlled magnonic microwaveguides. Appl. Phys. Lett. 104, 152402 (2014)
15. An, K., et al.: Control of propagating spin waves via spin transfer torque in a metallic bilayer
waveguide. Phys. Rev. B. 89, 140405(R) (2014)
16. Urazhdin, S., Demidov, V.E., Ulrichs, H., Kendziorczyk, T., Kuhn, T., Leuthold, J., Wilde, G.,
Demokritov, S.O.: Nanomagnonic devices based on the spin-transfer torque. Nat Nanotechnol. 9, 509–513 (2014)
17. Kittel, C.: Excitation of spin waves in a ferromagnet by uniform RF field. Phys. Rev. 110,
1295 (1958)
18. Seavey Jr., M.H., Tannenwald, E.: Direct observation of spin wave resonance. Phys. Rev. Lett.
1, 168 (1958)
19. Damon, R.W., Eshbach, J.R.: Magnetostatic modes of a ferromagnet slab. J. Phys. Chem.
Solids. 19, 308–320 (1961)
20. Grünberg, P., Metawe, F.: Light scattering from bulk and surface spin waves in EuO. Phys.
Rev. Lett. 39, 1561 (1977)
21. Grünberg, P., Schreiber, R., Pang, Y., Brodsky, M.B., Sowers, H.: Layered magnetic structures: evidence for antiferromagnetic coupling of Fe layers across Cr interlayers. Phys. Rev.
Lett. 57, 2442 (1986)
22. Mathieu, C., Jorzick, J., Frank, A., Demokritov, S.O., Slavin, A.N., Hillebrands, B., Bartenlian, B., Chappert, C., Decanini, D., Rousseaux, F., Cambrill, E.: Lateral quantization of spin
waves in micron size magnetic wires. Phys. Rev. Lett. 81, 3968 (1998)
342
S. O. Demokritov and A. N. Slavin
23. Jorzick, J., Demokritov, S.O., Hillebrands, B., Berkov, D., Gorn, N.L., Guslienko, K., Slavin,
A.N.: Spin wave wells in nonellipsoidal micrometer size magnetic elements. Phys. Rev. Lett.
88, 047204 (2002)
24. Park, J., Eames, D.M., Engebretson, J., Berezovsky, A., Crowell, P.: Phys. Rev. Lett. 89,
277201 (2002)
25. Tsoi, M., et al.: Excitation of a magnetic multilayer by an electric current. Phys. Rev. Lett.
80, 4281–4284 (1998)
26. Tsoi, M., Jansen, A.G.M., Bass, J., Chiang, W.C., Tsoi, V., Wyder: Generation and detection
of phase-coherent current-driven magnons in magnetic multilayers. Nature. 406, 46 (2000)
27. Kiselev, S.I., et al.: Microwave oscillations of a nanomagnet driven by a spin-polarized
current. Nature. 425, 380–383 (2003)
28. Rippard, W.H., Pufall, M.R., Kaka, S., Russek, S.E., Silva, T.J.: Direct-current induced
dynamics in Co90Fe10/Ni80Fe20 point contacts. Phys. Rev. Lett. 92, 027201 (2004)
29. Krivorotov, I.N., Emley, N.C., Sankey, J.C., Kiselev, S.I., Ralph, D.C., Buhrman, R.A.: Timedomain measurements of nanomagnet dynamics driven by spin-transfer torques. Science. 307,
228 (2005)
30. Demidov, V.E., Urazhdin, S., Demokritov, S.O.: Direct observation and mapping of spin
waves emitted by spin-torque nano-oscillators. Nat. Mater. 9, 984–988 (2010)
31. Demidov, V.E., et al.: Magnetic nano-oscillator driven by pure spin current. Nat. Mater. 11,
1028–1031 (2012)
32. Landau, L.D., Lifshitz, E.M.: Theory of the dispersion of magnetic permeability in ferromagnetic bodies. Phys. Z. Sowjet. 8, 153 (1935)
33. L’vov, V.S.: Wave Turbulence Under Parametric Excitation. Springer, New York (1994)
34. Cottam, M.G. (ed.): Linear and Nonlinear Spin Waves in Magnetic Films and Superlattices.
World Scientific, Singapore (1994)
35. Gurevich, A.G., Melkov, G.A.: Magnetization Oscillations and Waves. CRC, New York
(1996)
36. Herring, C., Kittel, C.: On the theory of spin waves in ferromagnetic media. Phys. Rev. 81,
869 (1951)
37. Kalinikos, B.A., Slavin, A.N.: Theory of dipole-exchange spin wave spectrum for ferromagnetic films with mixed exchange boundary conditions. J. Phys. C. 19, 7013 (1986)
38. Rado, G.T., Weertman, J.R.: Spin-wave resonance in a ferromagnetic metal. J. Phys. Chem.
Solids. 11, 315–333 (1959)
39. Guslienko, K.Y., Demokritov, S.O., Hillebrands, B., Slavin, A.N.: Dynamic pinning of dipolar
origin in nonellipspoidal magnetic stripes. Phys. Rev. B. 66, 132402 (2002)
40. Guslienko, K.Y., Slavin, A.N.: Boundary conditions for magnetization in magnetic nanoelements. Phys. Rev. B. 72, 014463 (2005)
41. Heinrich, B., Urban, R., Woltersdorf, G.: J. Appl. Phys. 91, 7523 (2002)
42. Hiebert, W.K., Stankiewicz, A., Freeman, M.R.: Phys. Rev. Lett. 79, 1134 (1997)
43. Barman, A., Kruglyak, V.V., Hicken, R.J., Rowe, J.M., Kundrotaite, A., Scott, J., Rahman,
M.: Imaging the dephasing of spin wave modes in a square thin film magnetic element. Phys.
Rev. B. 69, 174426 (2004)
44. Acremann, Y., Back, C.H., Buess, M., Portmann, O., Vaterlaus, A., Pescia, D., Melchior, H.:
Imaging precessional motion of the magnetization vector. Science. 290, 492 (2000)
45. Demokritov, S.O., Hillebrands, B., Slavin, A.N.: Brillouin light scattering studies of confined
spin waves: linear and nonlinear confinement. Phys. Rep. 348, 441 (2001)
46. Demokritov, S.O., Demidov, V.E.: Micro-Brillouin light scattering spectroscopy of magnetic
nanostructures. IEEE Trans. Magn. Adv. Magn. 44, 6 (2008)
47. Demidov, V.E., Demokritov, S.O.: Magnonic waveguides studied by microfocus Brillouin
light scattering. IEEE Trans. Magn. 51, 8578 (2015)
48. Jorzick, J., Demokritov, S.O., Mathieu, C., Hillebrands, B., Bartenlian, B., Chappert, C.,
Rousseaux, F., Slavin, A.: Brillouin light scattering from quantized spin waves in micronsize magnetic wires. Phys. Rev. B. 60, 15194 (1999)
49. Joseph, R.I., Schlomann, E.: Demagnetizing field in nonellipsoidal bodies. J. Appl. Phys. 36,
1579–1593 (1965)
6 Spin Waves
343
50. Bryant, H.: Suhl: thin-film magnetic patterns in an external field. Appl. Phys. Lett. 54, 2224
(1989)
51. Bayer, C., Demokritov, S.O., Hillebrands, B., Slavin, A.N.: Spin wave wells with multiple
states created in small magnetic elements. Appl. Phys. Lett. 82, 607 (2003)
52. Jersch, J., Demidov, V.E., Fuchs, H., Rott, K., Krzysteczko, P., Munchenberger, J., Reiss, G.,
Demokritov, S.O.: Mapping of localized spin-wave excitations by near-field Brillouin light
scattering. Appl. Phys. Lett. 97, 152502 (2010)
53. Serga, A.A., Schneider, T., Hillebrands, B., Demokritov, S.O., Kostylev, M.: Phase-sensitive
Brillouin light scattering spectroscopy from spin-wave packets. Appl. Phys. Lett. 89, 063506
(2006)
54. Vogt, K., Schultheiss, H., Hermsdoerfer, S.J., Pirro, Serga, A.A., Hillebrands, B.: All-optical
detection of phase fronts of propagating spin waves in a Ni81Fe19 microstripe. Appl. Phys.
Lett. 95, 182508 (2009)
55. Demidov, V.E., Urazhdin, S., Demokritov, S.O.: Control of spin-wave phase and wavelength
by electric current on the microscopic scale. Appl. Phys. Lett. 95, 262509 (2009)
56. Demidov, V.E., Demokritov, S.O., Rott, K., Krzysteczko, J., Reiss, G.: Self-focusing of spin
waves in Permalloy microstripes. Appl. Phys. Lett. 91, 252504 (2007)
57. Khitun, A., Bao, M., Wang, K.L.: Magnonic logic circuits. J. Phys. D. Appl. Phys. 43, 264005
(2010)
58. Demidov, V.E., Jersch, J., Demokritov, S.O., Rott, K., Krzysteczko, J., Reiss, G.: Transformation of propagating spin-wave modes in microscopic waveguides with variable width. Phys.
Rev. B. 79, 054417 (2009)
59. Demidov, V.E., Kostylev, M., Rott, K., Münchenberger, J., Reiss, G., Demokritov, S.O.:
Excitation of short-wavelength spin waves in magnonic waveguides. Appl. Phys. Lett. 99,
082507 (2011)
60. Demidov, V.E., Demokritov, S.O., Birt, D., O’Gorman, B., Tsoi, M., Li, X.: Radiation of
spin waves from the open end of a microscopic magnetic-film waveguide. Phys. Rev. B. 80,
014429 (2009)
61. Schneider, T., et al.: Nondiffractive subwavelength wave beams in a medium with externally
controlled anisotropy. Phys. Rev. Lett. 104, 197203 (2010)
62. Ulrichs, H., Demidov, V.E., Demokritov, S.O., Urazhdin, S.: Spin-torque nano-emitters for
magnonic applications. Appl. Phys. Lett. 100, 162406 (2012)
63. Adam, J.D.: Analog signal-processing with microwave magnetic. Proc. IEEE. 76, 159–170
(1988)
64. Ishak, W.S.: Magnetostatic wave technology: a review. Proc. IEEE. 76, 171–187 (1988)
65. Demidov, V.E., Kostylev, M., Rott, K., Krzysteczko, J., Reiss, G., Demokritov, S.O.:
Excitation of microwaveguide modes by a stripe antenna. Appl. Phys. Lett. 95, 112509
(2009)
66. Myers, E., Ralph, D.C., Katine, J.A., Louie, R.N., Buhrman, R.A.: Current-induced switching
of domains in magnetic multilayer devices. Science. 285, 867–870 (1999)
67. Katine, J.A., Albert, F.J., Buhrman, R.A., Myers, E.B., Ralph, D.C.: Current-driven magnetization reversal and spin-wave excitations in Co/Cu/Co pillars. Phys. Rev. Lett. 84, 3149–3152
(2000)
68. Ralph, D.C., Stiles, M.D.: Spin transfer torques. J. Magn. Magn. Mater. 320, 1190–1216
(2008)
69. Silva, T.J., Rippard, W.H.: Developments in nano-oscillators based upon spin-transfer pointcontact devices. J. Magn. Magn. Mater. 320, 1260–1271 (2008)
70. Brataas, A., Kent, A.D., Ohno, H.: Current-induced torques in magnetic materials. Nat. Mater.
11, 372 (2012)
71. Locatelli, N., Cros, V., Grollier, J.: Spin-torque building blocks. Nat. Mater. 13, 11 (2014)
72. Chen, T., Dumas, R.K., Eklund, A., Muduli, K., Houshang, A., Awad, A.A., Duerrenfeld,
Malm, B.G., Rusu, A., Akerman, J.: Spin-torque and spin-Hall nano-oscillators. IEEE Trans.
Magn. 99, 1 (2016)
73. Madami, M., et al.: Direct observation of a propagating spin wave induced by spin-transfer
torque. Nat. Nanotechnol. 6, 635–638 (2011)
344
S. O. Demokritov and A. N. Slavin
74. Houshang, A., Iacocca, E., Duerrenfeld, P., Sani, S.R., Akerman, J., Dumas, R.K.: Spin-wavebeam driven synchronization of nanocontact spin-torque oscillators. Nat. Nanotechnol. 11,
280–286 (2016)
75. Demidov, V.E., Urazhdin, S., Liu, R., Divinskiy, B., Telegin, A., Demokritov, S.O.: Excitation
of coherent propagating spin waves by pure spin currents. Nat. Commun. 7, 10446 (2016)
76. Xiao, J., Bauer, G.E.W.: Spin-wave excitation in magnetic insulators by spin-transfer torque.
Phys. Rev. Lett. 108, 217204 (2012)
77. Jungwirth, T., Wunderlich, J., Olejnik, K.: Spin Hall effect devices. Nat. Mater. 11, 382 (2012)
78. Ando, K., Takahashi, S., Harii, K., Sasage, K., Ieda, J., Maekawa, S., Saitoh, E.: Electric
manipulation of spin relaxation using the spin Hall effect. Phys. Rev. Lett. 101, 036601 (2008)
79. Evelt, M., et al.: High-efficiency control of spin-wave propagation in ultra-thin Yttrium Iron
Garnet by the spin-orbit torque. Appl. Phys. Lett. 108, 172406 (2016)
80. Demidov, V.E., Urazhdin, S., Tiberkevich, V., Slavin, A., Demokritov, S.O.: Control of spinwave emission from spin-torque nano-oscillators by microwave pumping. Phys. Rev. B. 83,
060406 (R) (2011)
81. Demidov, V.E., et al.: Control of magnetic fluctuations by spin current. Phys. Rev. Lett. 107,
107204 (2011)
82. Slavin, A., Tiberkevich, V.: Spin wave mode excited by spin-polarized current in a magnetic
nanocontact is a standing self-localized wave bullet. Phys. Rev. Lett. 95, 237201 (2005)
83. Rippard, W.H., Pufall, M.R., Silva, T.J.: Quantitative studies of spin-momentum-transferinduced excitations in Co/Cu multilayer films using point-contact spectroscopy. Appl. Phys.
Lett. 82, 1260 (2003)
84. Silberberg, Y.: Collapse of optical pulses. Opt. Lett. 15, 1282 (1990)
85. Rezende, S.M., de Aguiar, F.M., Azevedo, A.: Spin-wave theory for the dynamics induced by
direct currents in magnetic multilayers. Phys. Rev. Lett. 94, 037202 (2005)
86. Slavin, A.N., Kabos: Approximate theory of microwave generation in a current-driven
magnetic nanocontact magnetized in an arbitrary direction. IEEE Trans. Magn. 41, 1264
(2005)
87. Akhmediev, N.N., Ankiewicz, A.: Solitons. Nonlinear Pulses and Beams. Chapman & Hall,
London (1997)
88. Aranson, I.S., Kramer, L.: The world of the complex Ginzburg-Landau equation. Rev. Mod.
Phys. 74, 99 (2002)
89. Slavin, A.N., Tiberkevich, V.S.: Excitation of spin waves by spin-polarized current in
magnetic nano-structures. IEEE Trans. Magn. 44, 1916–1927 (2008)
90. Slavin, A.N., Tiberkevich, V.S.: Nonlinear auto-oscillator theory of microwave generation by
spin-polarized current. IEEE Trans. Magn. 45, 1875 (2009)
91. Gerhart, G., Bankowski, E., Melkov, G.A., Tiberkevich, V.S., Slavin, A.N.: Angular dependence of the microwave-generation threshold in a nanoscale spin-torque oscillator. Phys. Rev.
B. 76, 024437 (2007)
92. Consolo, G., Azzerboni, B., Lopez-Diaz, L., Gerhart, G., Bankowski, E., Tiberkevich, V.,
Slavin, A.N.: Micromagnetic study of the above-threshold generation regime in a spin-torque
oscillator based on a magnetic nanocontact magnetized at an arbitrary angle. Phys. Rev. B.
78, 014420 (2008)
93. Bonetti, S., et al.: Experimental evidence of self-localized and propagating spin wave modes
in obliquely magnetized current-driven nanocontacts. Phys. Rev. Lett. 105, 217204 (2010)
94. Liu, R.H., Lim, W.L., Urazhdin, S.: Spectral characteristics of the microwave emission by the
spin Hall nano-oscillator. Phys. Rev. Lett. 110, 147601 (2013)
95. Duan, Z., et al.: Nanowire spin torque oscillator driven by spin orbit torques. Nat. Commun.
5, 5616 (2014)
96. Hoefer, M.A., Silva, T.J., Keller, M.W.: Theory for a dissipative droplet soliton excited by a
spin torque nanocontact. Phys. Rev. B. 82, 054432 (2010)
97. Mohseni, S.M., Sani, S.R., Persson, J., Nguyen, T.N.A., Chung, S., Pogoryelov, Y., Muduli,
K., Iacocca, E., Eklund, A., Dumas, R.K., Bonetti, S., Deac, A., Hoefer, M.A., Akerman, J.:
Spin torque–generated magnetic droplet solitons. Science. 339, 1295 (2013)
6 Spin Waves
345
98. Chung, S., Mohseni, M., Sani, S.R., Iacocca, E., Dumas, R.K., Anh Nguyen, T.N., Pogoryelov, Y., Muduli, K., Eklund, A., Hoefer, M., Akerman, J.: Spin transfer torque generated
magnetic droplet solitons. J. Appl. Phys. 115, 172612 (2014)
99. Ivanov, B.A., Kosevich, A.M.: Zh. Eksp. Teor. Fiz. 72, 2000 (1977)
100. Kosevich, A.M., Ivanov, B.A., Kovalev, A.S.: Magnetic solitons. Phys. Rep. 194, 117–238
(1990)
101. Wolf, S.A., Awschalom, D.D., Buhrman, R.A., Daughton, J.M., von Molnar, S., Roukes,
M.L., Chtchelkanova, A.Y., Treger, D.M.: Spintronics: a spin-based electronics vision for
the future. Science. 294, 1488 (2001)
102. Zutic, I., Fabian, J., Das Sarma, D.: Spintronics: fundamentals and applications. Rev. Mod.
Phys. 76, 323 (2004)
103. Kang, W., Zhang, Y., Wang, Z.H., Klein, J.O., Chappert, C., Ravelosona, D., Wang, G.F.,
Zhang, Y.G., Zhao, W.S.: Spintronics: emerging ultra-low-power circuits and systems beyond
MOS technology. ACM J. Emerg. Technol. Comput. Syst. 12, 16 (2015)
104. Hoffmann, A., Bader, S.D.: Opportunities at the frontiers of spintronics. Phys. Rev. Appl. 4,
047001 (2015)
105. Liu, L.Q., Pai, C.F., Ralph, D.C., Buhrman, R.A.: Magnetic oscillations driven by the spin
Hall effect in 3-terminal magnetic tunnel junction devices. Phys. Rev. Lett. 109, 186602
(2012)
106. Tulapurkar, A.A., Suzuki, Y., Fukushima, A., Kubota, H., Maehara, H., Tsunekawa, K.,
Djayaprawira, D.D., Watanabe, N., Yuasa, S.: Spin-torque diode effect in magnetic tunnel
junctions. Nature. 438, 339 (2005)
107. Sankey, J.C., Braganca, M., Garcia, A.G.F., Krivorotov, I.N., Buhrman, R.A., Ralph, D.C.:
Spin-transfer-driven ferromagnetic resonance of individual nanomagnets. Phys. Rev. Lett. 96,
227601 (2006)
108. Bar’yakhtar, V.G., L’vov, V.A., Yablonskii, D.A.: Inhomogeneous magnetoelectric effect.
JETP Lett. 37, 673 (1983)
109. Dzyaloshinskii, I.: Magnetoelectricity in ferromagnets. Europhys. Lett. 83, 67001 (2008)
110. Mills, D.L., Dzyaloshinskii, I.E.: Influence of electric fields on spin waves in simple
ferromagnets: role of the flexoelectric interaction. Phys. Rev. B. 78, 184422 (2008)
111. Sandweg, C.W., Kajiwara, Y., Chumak, A.V., Serga, A.A., Vasyuchka, V.I., Jungfleisch, M.B.,
Saitoh, E., Hillebrands, B.: Spin pumping by parametrically excited exchange magnons. Phys.
Rev. Lett. 106, 216601 (2011)
112. Chumak, A.V., Serga, A.A., Jungfleisch, M.B., Neb, R., Bozhko, D.A., Tiberkevich, V.S.,
Hillebrands, B.: Direct detection of magnon spin transport by the inverse spin Hall effect.
Appl. Phys. Lett. 100, 082405 (2012)
113. Heinrich, B., Burrowes, C., Montoya, E., Kardasz, B., Girt, E., Song, Y.-Y., Sun, Y., Wu, M.:
Spin pumping at the magnetic insulator (YIG)/normal metal (Au) interfaces. Phys. Rev. Lett.
107, 066604 (2011)
114. Sun, Y., Song, Y.-Y., Chang, H., Kabatek, M., Jantz, M., Schneider, W., Wu, M., Schultheiss,
H., Hoffmann, A.: Growth and ferromagnetic resonance properties of nanometer-thick yttrium
iron garnet films. Appl. Phys. Lett. 101, 152405 (2012)
115. Chang, H., Li, Zhang, W., Liu, T., Hoffmann, A., Deng, L., Wu, M.: Nanometer-thick yttrium
iron garnet films with extremely low damping. IEEE Magn. Lett. 5, 6700104 (2014)
116. Hahn, C., Naletov, V.V., de Loubens, G., Klein, O., d’Allivy Kelly, O., Anane, A., Bernard,
R., Jacquet, E., Bortolotti, Cros, V., Prieto, J.L., Munoz, M.: Measurement of the intrinsic
damping constant in individual nanodisks of Y3 Fe5 O12 and Y3 Fe5 O12 |Pt. Appl. Phys. Lett.
104, 152410 (2014)
117. Hamadeh, A., et al.: Electronic control of the spin-wave damping in a magnetic insulator.
Phys. Rev. Lett. 113, 197203 (2014)
118. Collet, M., et al.: Generation of coherent spin-wave modes in Yttrium Iron Garnet microdiscs
by spin-orbit torque. Nat. Commun. 7, 10377 (2016)
119. Saitoh, E., Ueda, M., Miyajima, H., Tatara, G.: Conversion of spin current into charge current
at room temperature: inverse spin-Hall effect. Appl. Phys. Lett. 88, 182509 (2006)
346
S. O. Demokritov and A. N. Slavin
Sergej O. Demokritov received his Ph.D at Kapitsa Institute for
Physical Problems, Moscow, Russia. In the 1990s, he moved to
Germany to start to work with P. Grünberg at Research Center
Jülich. Since 2004, he is a Professor at Münster University, Germany. His main directions of research are dynamics and quantum
thermodynamics of magnetic structures, spin-wave research, and
magnonics.
Andrei N. Slavin is a Distinguished Professor and Chair of the
Physics Department, Oakland University, Michigan, USA. He
received his Ph.D from the St. Petersburg Technical University,
Russia. Andrei is Fellow of the American Physical Society and
Fellow of the IEEE. He is a specialist in magnetization dynamics
and spin waves and published over 280 research papers in this
field.
7
Micromagnetism
Lukas Exl , Dieter Suess , and Thomas Schrefl
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Micromagnetics Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Gibbs Free Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spin, Magnetic Moment, and Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Exchange Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Zeeman Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetostatic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Crystal Anisotropy Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetoelastic and Magnetostrictive Energy Terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Characteristic Length Scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Exchange Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Critical Diameter for Uniform Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Wall Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Mesh Size in Micromagnetic Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
348
349
350
350
351
356
357
357
363
366
372
372
373
375
376
L. Exl ()
University of Vienna Research Platform MMM Mathematics – Magnetism – Materials,
University of Vienna, and Wolfgang Pauli Institute, Wien, Austria
e-mail: [email protected]
D. Suess
University of Vienna Research Platform MMM Mathematics – Magnetism – Materials, and
Physics of Functional Materials, Faculty of Physics, University of Vienna, Wien, Austria
e-mail: [email protected]
T. Schrefl
Christian Doppler Laboratory for Magnet Design Through Physics Informed Machine Learning,
Department of Integrated Sensor Systems, Danube University Krems, Wiener Neustadt, Austria
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_7
347
348
L. Exl et al.
Brown’s Micromagnetic Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Euler Method: Finite Differences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Ritz Method: Finite Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetization Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
376
377
380
382
386
Abstract
Computational micromagnetics is widely used for the design and development
of magnetic devices. The theoretical background of these simulations is the
continuum theory of micromagnetism. It treats magnetization processes on a
significant length scale which is small enough to resolve magnetic domain walls
and large enough to replace atomic spins by a continuous function of position.
The continuous expression for the micromagnetic energy terms are either
derived from their atomistic counterpart or result from symmetry arguments.
The equilibrium conditions for the magnetization and the equation of motion
are introduced. The focus of the discussion lies on the equations that form the
basic building blocks of micromagnetic solvers. Numerical examples illustrate
the micromagnetic concepts. An open-source simulation environment was used
to address the ground state of thin film magnetic elements, initial magnetization
curves, stress-driven switching of magnetic storage elements, the grain size
dependence of the coercivity of permanent magnets, and damped oscillations
in magnetization dynamics.
Introduction
Computer simulations are essential tools for product design in modern society. This
is also true for magnetic materials and their applications. The design of magnetic
data storage systems such as hard disk devices [1, 2, 3, 4, 5] and random access
memories [6, 7] relies heavily on computer simulations. Similarly, the computer
models assist the development of magnetic sensors [8, 9] used as biosensors or
position and speed sensors in automotive applications [10]. Computer simulations
give guidance for the advance of high performance permanent magnet materials
[11, 12, 13] and devices. In storage and sensor applications, the selection of
magnetic materials, the geometry of the magnetically active layers, and the layout
of current lines are key design questions that can be answered by computations. In
addition to the intrinsic magnetic properties, the microstructure including grain size,
grain shape, and grain boundary phases is decisive for the magnet’s performance.
Computer simulations can quantify the influence of microstructural features on the
remanence and the coercive field of permanent magnets.
The characteristic length scale of the abovementioned computer models is
in the range of nanometers to micrometers. The length scale is too big for a
description by spin polarized density functional theory. Efficient simulations by
atomistic spin dynamics [14] are possible for nano-scale devices only. On the other
hand, macroscopic simulations using Maxwell’s equations hide the magnetization
7 Micromagnetism
349
processes that are relevant for the specific functions of the material or device under
consideration. Micromagnetism is a continuum theory that describes magnetization
processes on significant length scales that are
• large enough to replace discrete atomic spins by a continuous function of position
(the magnetization), but
• small enough to resolve the transition of the magnetization between magnetic
domains
For most ferromagnetic materials, this length scale is in the range of a few
nanometers to micrometers. The first aspect leads to a mathematical formulation
which makes it possible to simulate materials and devices in a reasonable time.
Instead of billions of atomic spins, only millions of finite elements have to be taken
into account. The second aspect keeps all relevant physics so that the influence of
structure and geometry on the formation of reversed domains and the motion of
domain walls can be computed.
The theory of micromagnetism was developed well before the advance of modern
computing technology. Key properties of magnetic materials can be understood by
analytic or semi-analytic solutions of the underlying equations. However, the future
use of powerful computers for the calculation of magnetic properties by solving the
micromagnetic equations numerically was already proposed by Brown [15] in the
late 1950s. The purpose of micromagnetics is the calculation of the magnetization
distribution as function of the applied field or the applied current taking into account
the structure of the material and the mutual interactions between the different
magnetic parts of a device.
Micromagnetics Basics
The key assumption of micromagnetism is that the spin direction changes only by
a small angle from one lattice point to the next [16]. The direction angles of the
spins can be approximated by a continuous function of position. Then the state
of a ferromagnet can be described by a continuous vector field, the magnetization
M(x). The magnetization is the magnetic moment per unit volume. The direction of
M(x) varies continuously with the coordinates x, y, and z. Here we introduced the
position vector x = (x, y, z). Starting from the Heisenberg model [17, 18] which
describes a ferromagnet by interacting spins associated with each atom, we derive
the micromagnetic equations whereby several assumptions are made:
1. Micromagnetism is a quasi-classical theory. The spin operators of the Heisenberg
model are replaced by classical vectors.
2. The The length of the magnetization vector is a constant that is uniform over
each material of the ferromagnetic body and only depends on temperature.
3. The temperature is constant in time and in space.
4. The Gibbs free energy of the ferromagnetic body is expressed in terms of the
direction cosines of the magnetization.
350
L. Exl et al.
5. The energy terms are derived either by the transition from an atomistic model to
a continuum model or phenomenologically.
In classical micromagnetism, the magnetization can only rotate. A change of the
length of M is forbidden. Thus, a ferromagnet is in thermodynamic equilibrium,
when the torque on the magnetic moment MdV in any volume element dV is zero.
The torque on the magnetic moment MdV caused by a magnetic field H is
T = μ0 MdV × H ,
(1)
where μ0 is the permeability of vacuum (μ0 = 4π × 10−7 Tm/A). The equilibrium
condition (1) follows from the direct variation of the Gibbs free energy. If only the
Zeeman energy of the magnet in an external field is considered, H is the external
field, H ext . In general additional energy terms will be relevant. Then H has to be
replaced by the effective field, H eff . Each energy term contributes to the effective
field.
In section “Magnetic Gibbs Free Energy”, we will derive continuum expressions
for the various contributions to the Gibbs free energy functional using the direction
cosines of the magnetization as unknown functions. In section “Characteristic
Length Scales”, we discuss the different characteristic length scales used to describe
magnetic phenomena. In section “Brown’s Micromagnetic Equation”, we show
how the equilibrium condition can be obtained by direct variation of the Gibbs free
energy functional.
Magnetic Gibbs Free Energy
We describe the state of the magnet in terms of the magnetization M(x). In the
following we will show how the continuous vector field M(x) is related to the
magnetic moments located at the atom positions of the magnet.
Spin, Magnetic Moment, and Magnetization
The local magnetic moment of an atom or ion at position x i is associated with the
spin angular momentum, h̄S,
μ(x i ) = −g
|e|
h̄S(x i ) = −gμB S(x i ).
2m
(2)
Here e is the charge of the electron, m is the electron mass, and g is the Landé factor.
The Landé factor is g ≈ 2 for metal systems with quenched orbital moment. The
constant μB = 9.274 × 10−24 Am2 = 9.274 × 10−24 J/T is the Bohr magneton.
The constant h̄ is the reduced Planck constant, h̄ = h/(2π ), where h is the Planck
constant. The magnetization of a magnetic material with N atoms per unit volume is
7 Micromagnetism
351
M = Nμ.
(3)
The magnetic moment is often given in Bohr magnetons per atom or Bohr
magnetons per formula unit. The magnetization is
M = Nfu μfu ,
(4)
where μfu is the magnetic moment per formula unit and Nfu is the number of
formula units per unit volume.
The length of the magnetization vector is assumed to be a function of temperature
only and does not depend on the strength of the magnetic field:
|M| = Ms (T ) = Ms ,
(5)
where Ms is the saturation magnetization. In classical micromagnetism the temperature, T, is assumed to be constant over the ferromagnetic body and independent of
time t. Therefore Ms is fixed and time evolution of the magnetization vector can be
expressed in terms of the unit vector m = M/|M|
M(x, t) = m(x, t)Ms .
(6)
The saturation magnetization of a material is frequently given as μ0 Ms in units of
Tesla.
Example. The saturation magnetization is an input parameter for micromagnetic
simulations. In a multiscale simulation approach of the hysteresis properties of a
magnetic material, it may be derived from the ab initio calculation of magnetic
moment per formula unit. In NdFe11 TiN, the calculated magnetic moment per
formula unit is 26.84 μB per formula unit [19]. The computed lattice constants
were a = 8.537 × 10−10 m, b = 8.618 × 10−10 m, and c = 4.880 × 10−10 m [19]
which give a volume of the unit cell of v = 359.0×10−30 m3 . There are two formula
units per unit cell and Nfu = 2/v = 5.571 × 1027 . With (4) and (5), the saturation
magnetization of NdFe11 TiN is Ms = 1.387 × 106 A/m (μ0 Ms = 1.743 T).
Exchange Energy
The exchange energy is of quantum mechanical nature. The energy of two ferromagnetic electrons depends on the relative orientation of their spins. When the two spins
are parallel, the energy is lower than the energy of the antiparallel state. Qualitatively
this behavior can be explained by the Pauli exclusion principle and the electrostatic
Coulomb interaction. Owing to the Pauli exclusion principle, two electrons can only
be at the same place if they have opposite spins. If the spins are parallel, the electrons
tend to move apart which lowers the electrostatic energy. The corresponding gain in
energy can be large enough so that the parallel state is preferred.
352
L. Exl et al.
The exchange energy, Eij , between two localized spins is [18]
Eij = −2Jij S i · S j ,
(7)
where Jij is the exchange integral between atoms i and j and h̄S i is the angular
momentum of the spin at atom i. For cubic metals and hexagonal closed packed
metals with ideal c over a ratio there holds Jij = J . Treating the exchange energy
for a large number of coupled spins, we regard Eij as a classical potential energy
and replace S i by a classical vector. Let mi be the unit vector in direction −S i . Then
mi is the unit vector of the magnetic moment at atom i. If ϕij is the angle between
the vectors mi and mj , the exchange energy is
Eij = −2J S 2 cos(ϕij ),
(8)
where S = |S i | = |S j | is the spin quantum number.
Now, we introduce a continuous unit vector m(x) and assume that the angle
ϕij between the vectors mi and mj is small. We set m(x i ) = mi and expand m
around x i
m(x i + a j ) =m(x i )+
∂m
∂m
∂m
aj +
bj +
cj +
∂x
∂y
∂z
1 ∂ 2m 2 ∂ 2m 2 ∂ 2m 2
+ ....
a
+
b
+
c
2 ∂x 2 j
∂y 2 j
∂z2 j
(9)
Here a j = (aj , bj , cj )T is the vector connecting points x i and x j = x i + a j .
We can replace cos(ϕij ) by cos(ϕij ) = m(x i ) · m(x j ) in (8). Summing up
over the six nearest neighbors of a spin in a simple cubic lattice gives (see
Fig. 1) the exchange energy of the unit cell. The vectors a j take the values
(±a, 0, 0)T , (0, ±a, 0)T , (0, 0, ±a)T . For every vector a, there is the corresponding vector −a. Thus the linear terms in the variable a in (9) vanish in the summation.
The same holds for mixed second derivatives in the expansion (9). The constant
term, m · m = 1, plays no role for the variation of the energy and will be neglected.
The exchange energy of a unit cell in a simple cubic lattice is
6
j =1
Eij = − J S 2
6 ∂ 2 mi 2
∂ 2 mi 2
∂ 2 mi 2
mi ·
a
+
m
·
b
+
m
·
c
i
i
∂x 2 j
∂y 2 j
∂z2 j
j =1
∂ 2 mi
∂ 2 mi
∂ 2 mi
+
m
·
+
m
·
= − 2J S 2 a 2 mi ·
i
i
∂x 2
∂y 2
∂z2
(10)
To get the exchange energy of the crystal, we sum over all atoms i and divide by 2
to avoid counting each pair of atoms twice. We also use the relations
7 Micromagnetism
353
Fig. 1 Nearest neighbors of
spin i for the calculation of
the exchange energy in a
simple cubic lattice
m·
∂ 2m
∂m 2
=
−
,
∂x
∂x 2
(11)
which follows from differentiating m · m = 1 twice with respect to x. Thus we can
write
Eex
∂mi 2
J S2 3
∂mi 2
∂mi 2
=
a
+
+
.
a
∂x
∂y
∂z
(12)
i
The sum in (12) is over the unit cells of the crystal with volume V . In the continuum
limit, we replace the sum with an integral. The exchange energy is
Eex =
A
V
∂m
∂x
2
+
∂m
∂y
2
+
∂m
∂z
2 dV .
(13)
Expanding and rearranging the terms in the bracket and introducing the nabla
operator, ∇, we obtain
A (∇mx )2 + ∇my
Eex =
V
2
+ (∇mz )2 dV .
(14)
354
L. Exl et al.
In equations (13) and (14), we introduced the exchange constant
A=
J S2
n.
a
(15)
In cubic lattices, n is the number of atoms per unit cell (n = 1, 2, and 4 for
simple cube, body-centered cubic, and face-centered cubic lattices, respectively).
In a hexagonal
closed packed structures, n is the ideal nearest neighbor distance
√
(n = 2 2). The number N of atoms per unit volume is n/a 3 . At non-zero
temperature, the exchange constant may be expressed in terms of the saturation
magnetization, Ms (T ). Formally we replace S by its thermal average. Using
equations (2) and (3), we rewrite
A(T ) =
J [Ms (T )]2
n.
(NgμB )2 a
(16)
The calculation of the exchange constant by (15) requires a value for the
exchange integral, J . Experimentally, one can measure a quantity that strongly
depends on J such as the Curie temperature, TC ; the temperature dependence of
the saturation magnetization, Ms (T ); or the spin wave stiffness parameter, in order
to determine J. According to the molecular field theory [20], the exchange integral
is related to the Curie temperature given by
J =
3 k B TC S n
3 k B TC 1
or A =
.
2 S(S + 1) z
2 a(S + 1) z
(17)
The second equation follows from the first one by replacing J with the relation (15).
Here z is the number of nearest neighbors (z = 6, 8, 12, and 12 for simple cubic,
body-centered cubic, face-centered cubic, and hexagonal closed packed lattices,
respectively) and kB =1.3807×10−23 J/K is Boltzmann’s constant. The use of (17)
together with (15) underestimates the exchange constant by more than a factor of 2
[21]. Alternatively one can use the temperature dependence of the magnetization as
arising from the spin wave theory
Ms (T ) = Ms (0)(1 − CT 3/2 ).
(18)
Equation (18) is valid for low temperatures. From the measured temperature
dependence Ms (T ), the constant C can be determined. Then the exchange integral
[21, 22] and the exchange constant can be calculated from C as follows:
J =
0.0587
nSC
2/3
kB
or A =
2S
0.0587
n2 S 2 C
2/3
kB
.
2a
(19)
This method was used by Talagala and co-workers [23]. They measured the
temperature dependence of the saturation magnetization in NiCo films to determine
7 Micromagnetism
355
the exchange constant as function of the Co content. The exchange constant can also
be evaluated from the spin wave dispersion relation (see Chapter SPW) which can
be measured by inelastic neutron scattering, ferromagnetic resonance, or Brillouin
light scattering [24]. The exchange integral [22] and the exchange constant are
related to the spin wave stiffness constant, D, via the following relations:
J =
D
D 1
or A = NS.
2
2 Sa
2
(20)
For the evaluation of the exchange constant, we can use S = Ms (0)/(NgμB ) [25]
for the spin quantum number in equations (17), (19), and (20). This gives the relation
between the exchange constant, A, and the spin wave stiffness constant, D,
A=
DMs (0)
,
2gμB
(21)
when applied to (20). Using neutron Brillouin Scattering, Ono and co-workers [26]
measured the spin wave dispersion in a polycrystalline Nd2 Fe14 B magnet, in order
to determine its exchange constant.
Ferromagnetic exchange interactions keep the magnetization uniform. Depending on the sample, geometry external fields may lead to a locally confined
non-uniform magnetization. Probing the magnetization twist experimentally and
comparing the result with the computed equilibrium magnetic state (see section “Brown’s Micromagnetic Equation”) is an alternative method to determine the
exchange constant. The measured data is fitted to the theoretical model whereby
the exchange constant is a free parameter. Smith and co-workers [27] measured
the anisotropic magnetoresistance to probe the fanning of the magnetization in a
thin permalloy film from which its exchange constant was calculated. Eyrich and
co-workers [24] measured the field-dependent magnetization, M(H ), of a trilayer
structure in which two ferromagnetic films are coupled antiferromagnetically. The
M(H ) curve probes the magnetization twist within the two ferromagnets. Using
this method the exchange constant of Co alloyed with various other elements was
measured [24].
The interplay between the effects of ferromagnetic exchange coupling, magnetostatic interactions, and the magnetocrystalline anisotropy leads to the formation of
domain patterns (for details on domain structures, see Chapter Domains). With magnetic imaging techniques, the domain width, the orientation of the magnetization,
and the domain wall width can be measured. These values can be also calculated
using a micromagnetic model of the domain structure. By comparing the predicted
values for the domain width with measured data, Livingston [28, 29] estimated
the exchange constant of the hard magnetic materials SmCo5 and Nd2 Fe14 B. This
method can be improved by comparing more than one predicted quantity with
measured data. Newnham and co-workers [30] measured the domain width, the
orientation of the magnetization in the domain, and the domain wall width in foils
356
L. Exl et al.
of Nd2 Fe14 B. By comparing the measured values with the theoretical predictions,
they estimated the exchange constant of Nd2 Fe14 B.
Input for micromagnetic simulations: The high temperature behavior of permanent magnets is of utmost importance for the applications of permanent magnets
in the hybrid or electric vehicles. For computation of the coercive field by
micromagnetic simulations, the exchange constant is needed as input parameter.
Values for A(T ) may be obtained from the room temperature value of A(300 K)
and Ms (T ). Applying (16) gives A(T ) = A(300 K) × [Ms (T )/Ms (300 K)]2 .
Magnetostatics
We now consider the energy of the magnet in an external field produced by
stationary currents and the energy of the magnet in the field produced by the
magnetization of the magnet itself. The latter field is called demagnetizing field.
In micromagnetics, these fields are treated statically if eddy currents are neglected.
In magnetostatics, we have no time-dependent quantities. In the presence of a
stationary magnetic current, Maxwell’s equations reduce to [31]
∇ ×H = j
(22)
∇ ·B = 0
(23)
Here B is the magnetic induction or magnetic flux density, H is the magnetic field,
and j is the current density. The charge density fulfills ∇ · j = 0 which expresses
the conservation of electric charge. We now have the freedom to split the magnetic
field into its solenoidal and nonrotational part
H = H ext + H demag .
(24)
By definition, we have
∇ · H ext = 0,
(25)
∇ × H demag = 0.
(26)
Using (22) and (24), we see that the external field, H ext , results from the current
density (Ampere’s law)
∇ × H ext = j .
(27)
On a macroscopic length scale, the relation between the magnetic induction and the
magnetic field is expressed by
B = μH ,
(28)
7 Micromagnetism
357
where μ is the permeability of the material. Equation (28) is used in magnetostatic
field solvers [32] for the design of magnetic circuits. In these simulations, the
permeability describes the response of the material to the magnetic field. Micromagnetics describes the material on a much finer length scale. In micromagnetics, we
compute the local distribution of the magnetization as function of the magnetic field.
This is the response of the system to (an external) field. Indeed, the permeability
can be derived from micromagnetic simulations [33]. For the calculation of the
demagnetizing field, we can treat the magnetization as fixed function of space.
Instead of (28), we use
B = μ0 (H + M) .
(29)
The energy of the magnet in the external field, H ext , is the Zeeman energy. The
energy of the magnet in the demagnetizing field, H demag , is called magnetostatic
energy.
Zeeman Energy
The energy of a magnetic dipole moment, μ, in an external magnetic induction Bext
is −μ · Bext . We use B ext = μ0 H ext and sum over all local magnetic moments at
positions x i of the ferromagnet. The sum,
Eext = −μ0
μi · H ext ,
(30)
i
is the interaction energy of the magnet with the external field. To obtain the Zeeman
energy in a continuum model, we introduce the magnetization M = Nμ, define the
volume per atom, Vi = 1/N, and replace the sum with an integral. We obtain
Eext = −μ0
(M · H ext )Vi → −μ0
(M · H ext )dV .
(31)
V
i
Using (6), we express the Zeeman energy in terms of the unit vector of the
magnetization
Eext = −
μ0 Ms (m · H ext )dV.
(32)
V
Magnetostatic Energy
The magnetostatic energy is also called dipolar interaction energy. In a crystal each
moment creates a dipole field, and each moment is exposed to the magnetic field
created by all other dipoles. Therefore magnetostatic interactions are long range.
The magnetostatic energy cannot be represented as a volume integral over the
magnet of an energy density dependent on only local quantities.
358
L. Exl et al.
Demagnetizing Field as Sum of Dipolar Fields
The total magnetic field at point x i , which is created by all the other magnetic
dipoles, is the sum over the dipole fields from all moments μj = μ(x j )
μj
1 (μj · r ij )r ij
− 3 .
3
H dip (x i ) =
4π
rij5
rij
j =i
(33)
The vectors r ij = x i − x j connect the source points with the field point. The
distance between a source point and a field point is rij = |r ij |. In order to obtain
a continuum expression for the field, we split the sum (33) into two parts. The
contribution to the field from moments that are far from x i will not depend strongly
on their exact position at the atomic level. Therefore we can describe them by a
continuous magnetization and replace the sum with an integral. For moments μj
which are located within a small sphere with radius R around x i , we keep the sum.
Thus we split the dipole field into two parts [34]:
H dip (x i ) = H near (x i ) + H demag (x i ).
(34)
Here
1 H near (x i ) =
4π
rij <R
3
(μj · r ij )r ij
rij5
−
μj
rij3
(35)
is the contribution of the sum of the dipoles within the sphere (see Fig. 2). For
the dipoles outside, the sphere we use a continuum approximation. Introducing the
magnetic dipole element MdV , we can replace the sum in (33) with an integral for
rij ≥ R
Fig. 2 Computation of the
total magnetostatic field at
point atom i. The near field is
evaluated by a direct sum
over all dipoles in the small
sphere. The atomic moments
outside the sphere are
replaced by a continuous
magnetization which
produces the far field acting
on i
7 Micromagnetism
359
1
H demag (x i ) =
4π
V*
M(x ) · (x i − x ) (x i − x )
M(x )
−
3
dV .
|x i − x |5
|x i − x |3
(36)
The integral is over V *, the volume of the magnet without the small sphere around
the field point x i .
The sum in (35) is the contribution of the dipoles inside the sphere to the total
magnetostatic field. The corresponding energy term is local. It can be expressed
as an integral of an energy density that depends only on local quantities [34].
The term depends on the symmetry of the lattice and has the same form as the
crystalline anisotropy. Therefore it is included in the anisotropy energy. When the
anisotropy constants in (58) are determined experimentally, they already include the
contribution owing to dipolar interactions.
Magnetic Scalar Potential
The demagnetizing field is nonrotational. Therefore we can write the demagnetizing
field as gradient of a scalar potential
H demag = −∇U.
(37)
Applying −∇ to
U (x) = −
1
4π
M(x ) ·
V*
x − x
dV |x − x |3
(38)
gives (36). In computational micromagnetics, it is beneficial to work with effective
magnetic volume charges, ρm = −∇ · M(x ), and effective magnetic surface
charges, σm = M(x ) · n. Using
1
x − x
1
1
and ∇
= −∇ = −∇
|x − x |
|x − x |
|x − x |
|x − x |3
we obtain
1
U (x) =
4π
V*
M(x ) · ∇ 1
dV .
|x − x |
(39)
(40)
Now we shift the ∇ operator from 1/|x − x | to M. We use
∇ ·
∇ · M(x )
1
M(x )
=
+ M(x ) · ∇ ,
|x − x |
|x − x |
|x − x |
apply Gauss’ theorem, and obtain [31]
1
ρm (x )
1
U (x) =
dV +
4π V * |x − x |
4π
∂V *
σm (x )
dS .
|x − x |
(41)
(42)
360
L. Exl et al.
Magnetostatic Energy
For computing the magnetostatic energy, there is no need to take into account
(35). The near field is already included in the crystal anisotropy energy. We now
compute the energy of each magnetic moment μi in the field H demag (x i ) from the
surrounding magnetization. The sum over all atoms is the magnetostatic energy of
the magnet
Edemag = −
μ0 μi · H demag (x i ).
2
(43)
i
The factor 1/2 avoids counting each pair of atoms twice. Similar to the procedure
for the exchange and Zeeman energy, we replace the sum with an integral
Edemag = −
μ0
2
M · H demag = −
V
1
2
μ0 Ms m · H demag dV .
(44)
V
Alternatively, the magnetostatic energy can be expressed in terms of a magnetic
scalar potential and effective magnetic charges. We start from (44), replace H demag
by −∇U , and apply Gauss’ theorem on ∇ · (MU ) to obtain
Edemag =
μ0
2
ρm U dV +
V
μ0
2
σm U dS.
(45)
∂V
Equation (45) is widely used in numerical micromagnetics. Its direct variation
(see section “Brown’s Micromagnetic Equation”) with respect to M gives the
cell averaged demagnetizing field. This method was introduced in numerical
micromagnetics by LaBonte [35] and Schabes and Aharoni [36]. For discretization
with piecewise constant magnetization only, the surface integrals remain.
In a uniformly magnetized spheroid, the demagnetizing field is antiparallel to the
magnetization. The demagnetizing field is
H demag = −N M,
(46)
where N is the demagnetizing factor. For a sphere the demagnetizing factor is 1/3.
Using (44), we find
μ0
NMs2 V
(47)
Edemag =
2
for the magnetostatic energy of a uniformly magnetized sphereoid with volume V .
In a cuboid or polyhedral particle, the demagnetizing field is nonuniform. However
we still can apply (47) when we use a volume averaged demagnetizing factor
which is obtained from a volume-averaged demagnetizing field. Interestingly the
volume-averaged demagnetizing factor for a cube is 1/3 the same value as for
the sphere. For a general rectangular prism, Aharoni [37] calculated the volume
averaged demagnetizing factor. A convenient calculation tool for the demagnetizing
7 Micromagnetism
361
factor, which uses Aharoni’s equation, is given on the Magpar website [38]. A
simple approximate equation for the demagnetizing factor of a square prism with
dimensions l × l × pl is [39]
N=
1
,
2p + 1
(48)
where p is the aspect ratio and N is the demagnetizing factor along the edge with
length pl.
Magnetostatic Boundary Value Problem
Equation (42) is the solution of the magnetostatic boundary value problem, which
can be derived from Maxwell’s equations as follows. From (23) and (29), the
following equation holds for the demagnetizing field
∇ · H demag = −∇ · M.
(49)
Plugging (37) into (49), we obtain a partial differential equation for the scalar
potential
∇ 2 U = ∇ · M.
(50)
Equation (50) holds inside the magnet. Outside the magnet M = 0 and we have
∇ 2 U = 0.
(51)
At the magnet’s boundary, the following interface conditions [31] hold
U (in) = U (out) ,
∇U (in) − ∇U (out) · n = M · n,
(52)
(53)
where n denotes the surface normal. The first condition follows from the continuity
of the component of H parallel to the surface (or ∇ ×H = 0). The second condition
follows from the continuity of the component of B normal to the surface (or ∇ ·B =
0). Assuming that the scalar potential is regular at infinity,
|U (x)| ≤ C
1
for |x| large enough and constant C > 0
|x|
(54)
the solution of equations (50) to (53) is given by (42). Formally the integrals in (42)
are over the volume, V *, and the surface, ∂V *, of the magnet without a small sphere
surrounding the field point. The transition from V *→ V adds a term −M/3 to the
field and thus shifts the energy by a constant which is proportional to Ms2 . This is
usually done in micromagnetics [34].
362
L. Exl et al.
The above set of equations for the magnetic scalar potential can also be derived
from a variational principle. Brown [16] introduced an approximate expression
= μ0
Edemag
V
M · ∇U dV −
μ0
2
(∇U )2 dV
(55)
for the magnetostatic energy, Edemag . For any magnetization distribution M(x), the
following equation holds
Edemag (M) ≥ Edemag
(M, U ),
(56)
where U is an arbitrary function which is continuous in space and regular at infinity
[16]. A proof of (56) is given by Asselin and Thiele [40]. The inequality (56) is sharp
in the sense that if maximized with respect to the variable U , equality holds in (56)
and U is the scalar potential owing to M. Then equality holds and Edemag
reduces
to the usual magnetostatic energy Edemag . Equation (55) is used in finite element
micromagnetics for the computation of the magnetic scalar potential. The EulerLagrange equation of (55) with respect to U gives the magnetostatic boundary value
problem (50) to (53) [40].
Examples
Magnetostatic energy in micromagnetic software: For physicists and software engineers developing micromagnetic software, there are several options to implement
magnetostatic field computation. The choice depends on the discretization scheme,
the numerical methods used, and the hardware. Finite difference solvers including
OOMMF [41], MuMax3 [42], and FIDIMAG [43] use (45) to compute the magnetostatic energy and the cell-averaged demagnetizing field. For piecewise constant
magnetization only, the surface integrals over the surfaces of the computational cells
remain. MicroMagnum [44] uses (42) to evaluate the magnetic scalar potential.
The demagnetizing field is computed from the potential by a finite difference
approximation. This method shows a higher speed up on Graphics Processor Units
[45] though its accuracy is slightly less. Finite element solvers compute the magnetic
scalar potential and build its gradient. Magpar [46], Nmag [47], and magnum.fe
[48] solve the partial differential equations (50) to (53). FastMag [49], a finite
element solver, directly integrates (42). Finite difference solvers apply the Fast
Fourier Transforms for the efficient evaluation of the involved convolutions. Finite
element solvers often use hierarchical clustering techniques for the evaluation of
integrals [50].
Magnetic state of nano-elements: From (45), we see that the magnetostatic
energy tends to zero if the effective magnetic charges vanish. This is known as
pole avoidance principle [34]. In large samples where the magnetostatic energy
dominates over the exchange energy, the lowest energy configurations are such that
∇ · M in the volume and M · n on the surface tend to zero. The magnetization
is aligned parallel to the boundary and may show a vortex. These patterns are
known as flux closure states. In small samples, the expense of exchange energy
7 Micromagnetism
363
Fig. 3 Computed magnetization patterns for a soft magnetic square element
(K1 = 0, μ0 Ms = 1 T, A = 10 pJ/m, mesh size h = 0.56 A/(μ0 Ms2 ) = 2 nm) as function of
element size L. The dimensions are L × L × 6 nm3 . The system was relaxed multiple times from
an initial state with random magnetization. The lowest energy states are the leaf state, the C-state,
and the vortex state for L = 80 nm, L = 150 nm, and L = 200 nm, respectively. For each state,
the relative contributions of the exchange energy and the magnetostatic energy to the total energy
are given
for the formation of a closure state is too high. As a compromise the magnetization
bends towards the surface near the edges of the sample. Depending on the size,
the leaf state [51] or the C-state [52] or the vortex state has the lowest energy.
Figure 3 shows the different magnetization patterns that can form in thin film
square elements. The results show that with increasing element size the relative
contribution of the magnetostatic energy, Fdemag /(Fex + Fdemag ) decreases. All
micromagnetic examples in this chapter are simulated using FIDIMAG [43]. Code
snippets are given in the appendix.
Crystal Anisotropy Energy
The magnetic properties of a ferromagnetic crystal are anisotropic. Depending on
the orientation of the magnetic field with respect to the symmetry axes of the
crystal, the M(H ) curve reaches the saturation magnetization, Ms , at low or high
field values. Thus easy directions in which saturation is reached in a low field and
hard directions in which high saturation requires a high field are defined. Figure 4
shows the magnetization curve, measured parallel to the easy and hard direction, of
a uniaxial material with strong crystal anisotropy. The initial state is a two domain
state with the magnetization of the domains parallel to the easy axis. The snapshots
364
L. Exl et al.
Fig. 4 Initial magnetization curves with the field applied in the easy direction (dashed line) and
the hard direction (solid line) computed for a uniaxial hard magnetic material (Nd2 Fe14√
B at room
temperature: K1 = 4.9 MJ/m3 , μ0 Ms = 1.61 T, A = 8 pJ/m, the mesh size is h = 0.86 A/K1 =
1.1 nm). The magnetization component parallel to the field direction is plotted as a function of the
external field. The field is given in units of HK . The sample shape is thin platelet with the easy
axis in the plane of the film. The sample dimensions are 200 × 200 × 10 nm3 . The insets show
snapshots of the magnetization configuration along the curves. The initial state is the two domain
state shown at the lower left of the figure
of the magnetic states show that domain wall motion occurs along the easy axis and
rotation of the magnetization occurs along the hard axis.
The crystal anisotropy energy is the work done by the external field to move
the magnetization away from a direction parallel to the easy axis. The functional
form of the energy term can be obtained phenomenologically. The energy density,
eani (m), is expanded in a power series in terms of the direction cosines of the
magnetization. Crystal symmetry is used to decrease the number of coefficients.
The series is truncated after the first two non-constant terms.
Cubic Anisotropy
Let a, b, and c be the unit vectors along the axes of a cubic crystal. The crystal
anisotropy energy density of a cubic crystal is
eani (m) = K0 + K1 (a · m)2 (b · m)2 + (b · m)2 (c · m)2 + (c · m)2 (a · m)2
+ K2 (a · m)2 (b · m)2 (c · m)2 + . . . .
(57)
7 Micromagnetism
365
The anisotropy constants K0 , K1 , and K2 are functions of temperature. The first
term is independent of m and thus can be dropped since only the change of the
energy with respect to the direction of the magnetization is of interest.
Uniaxial Anisotropy
In hexagonal or tetragonal crystals, the crystal anisotropy energy density is usually
expressed in terms of sin θ , where θ is the angle between the c-axis and the
magnetization. The crystal anisotropy energy of a hexagonal or tetragonal crystal
is
eani (m) = K0 + K1 sin2 (θ ) + K2 sin4 (θ ) + . . . .
(58)
In numerical micromagnetics, it is often more convenient to use
eani
(m) = −K1 (c · m)2 + . . . .
(59)
as expression for a uniaxial crystal anisotropy energy density. Here we used the
identity sin2 (θ ) = 1 − (c · m)2 , dropped two constant terms, namely, K0 and K1 ,
and truncated the series. When keeping only the terms which are quadratic in m,
the crystal anisotropy energy can be discretized as quadratic form involving only a
geometry-dependent matrix.
The crystalline anisotropy energy is
Eani =
eani (m)dV ,
(60)
V
whereby the integral is over the volume, V , of the magnetic body.
Anisotropy Field
An important material parameter, which is commonly used, is the anisotropy
field, HK . The anisotropy field is a fictitious field that mimics the effect of the
crystalline anisotropy. If the magnetization vector rotates out of the easy axis, the
crystalline anisotropy creates a torque that brings M back into the easy direction.
The anisotropy field is parallel to the easy direction, and its magnitude is such that
for deviations from the easy axis, the torque on M is the same as the torque by the
crystalline anisotropy. If the energy depends on the angle, θ , of the magnetization
with respect to an axis, the torque, T , on the magnetization is the derivative of the
energy density, e, with respect to the angle [20]
T =
∂e
.
∂θ
(61)
Let θ be a small angular deviation of M from the easy direction. The energy density
of the magnetization in the anisotropy field is
366
L. Exl et al.
eK = −μ0 Ms HK cos(θ )
(62)
TK = μ0 Ms HK sin(θ ) ≈ μ0 Ms HK θ.
(63)
the associated torque is
For the crystalline anisotropy energy density
eani = K1 sin2 (θ )
(64)
the torque towards the easy axis is
Tani = 2K1 sin(θ ) cos(θ ) = K1 sin(2θ ) ≈ 2K1 θ.
(65)
From the definition of the anisotropy field, namely, TK = Tani , we get
HK =
2K1
μ0 Ms
(66)
Anisotropy field, easy and hard axis loops: K1 and – depending on the material
to be studied – K2 are input parameters for micromagnetic simulation. The
anisotropy constants can be measured by fitting a calculated magnetization curve to
experimental data. Figure 4 shows the magnetization curves of a uniaxial material
computed by micromagnetic simulations. For simplicity we neglected K2 and
described the crystalline anisotropy with (59). The M(H ) along the hard direction
is almost a straight line until saturation where M(H ) = Ms . Saturation is reached
when H = HK .
The above numerical result can be found theoretically. A field is applied
perpendicular to the easy direction. The torque created by the field tends to increase
the angle, θ , between the magnetization and the easy axis. The torque asserted by
the crystalline anisotropy returns the magnetization towards the easy direction. We
set the total torque to zero to get the equilibrium condition
−μ0 Ms H cos(θ ) + 2K1 sin(θ ) cos(θ ) = 0.
The value of H that makes M parallel to the field is reached when sin(θ ) = 1. This
gives H = 2K1 /(μ0 Ms ). If higher anisotropy constants are taken into account the
field that brings M into the hard axis is H = (2K1 + 4K2 )/(μ0 Ms ).
Magnetoelastic and Magnetostrictive Energy Terms
When the atom positions of a magnet are changed relative to each other the
crystalline anisotropy varies. Owing to magnetoelastic coupling a deformation
produced by an external stress makes certain directions to be energetically more
7 Micromagnetism
367
favorable for the magnetization. Reversely, the magnet will deform in order to
minimize its total free energy when magnetized in certain directions.
Spontaneous Magnetostrictive Deformation
Most generally the spontaneous magnetostrictive deformation is expressed by the
0 as
symmetric tensor strain εij
0
εij
=
(67)
λij kl αk αl ,
kl
where λij kl is the tensor of magnetostriction constants. Measurements of the relative
change of length along certain directions owing to saturation of the crystal in
direction α = (α1 , α2 , α3 ) give the magnetostriction constants. For a cubic material,
the following relation holds
εii0
3
1
2
,
= λ100 αi −
2
3
0
εij
=
(68)
3
λ111 αi αj for i = j.
2
(69)
The magnetostriction constants λ100 and λ111 are defined as follows: λ100 is the
relative change in length measured along [100] owing to saturation of the crystal in
[100]; similarly λ111 is the relative change in length measured along [111] owing
to saturation of the crystal in [111]. The term with 1/3 in (68) results from the
definition of the spontaneous deformation with respect to a demagnetized state with
the averages αi2 = 1/3 and αi αj = 0.
Magnetoelastic Coupling Energy
All energy terms discussed in the previous sections can depend on deformations.
The most important change of energy with strain arises from the crystal anisotropy
energy. Thus the crystal anisotropy energy is a function of the magnetization and the
deformation of the lattice. We express the magnetization direction in terms of the
direction cosines of the magnetization α1 = a · m, α2 = b · m, and α3 = c · m (a, b,
and c are the unit lattice vectors) and the deformation in terms of the symmetric
strain tensor εij to obtain
eani = eani (αi , εij ).
(70)
A Taylor expansion of (70)
eani = eani (αi , 0) +
∂eani (αi , 0)
ij
∂εij
εij
(71)
368
L. Exl et al.
gives the change of the energy density owing to the strain εij . Owing to symmetry,
the expansion coefficients ∂eani (αi , 0)/∂εij do not dependent on the sign of the
magnetization vector and thus are proportional to αi αj . The second term on the
right-hand side of (71) is the change of the crystal anisotropy energy density
with
deformation. This term is the magnetoelastic coupling energy density. Using
ij Bij kl αi αj as expansion coefficients, we obtain
eme =
ij
Bij kl αi αj εkl ,
(72)
kl
where Bij kl is the tensor of the magnetoelastic coupling constants. For cubic
symmetry, the magnetoelastic coupling energy density is
eme,cubic = B1 (ε11 α12 + ε22 α22 + ε33 α32 )+ 2B2 (ε23 α2 α3 + ε13 α1 α3 + ε12 α1 α2 ) + . . .
(73)
with the magnetoelastic coupling constants B1 = B1111 and B2 = B2323 .
Equation (72) describes change of the energy density owing to the interaction of
magnetization direction and deformation. The magnetoelastic coupling constants
can be derived from the ab initio computation of the crystal anisotropy energy as
function of strain [53]. Experimentally the magnetoelastic coupling constants can
be obtained from the measured magnetostriction constants.
When magnetized in a certain direction, the magnet tends to deform in a way
that minimizes the sum of the magnetoelastic energy density, eme , and of the elastic
energy density of the crystal, eel . The elastic energy density is a quadratic function
of the strain
eel =
1 cij kl εij εkl ,
2
ij
(74)
kl
where cij kl is the elastic stiffness tensor. For cubic crystals the elastic energy is
1
2
2
2
+ ε22
+ ε33
)+
eel,cubic = c1111 (ε11
2
c1122 (ε11 ε22 + ε22 ε33 + ε33 ε11 )+
(75)
2
2
2
2c2323 (ε12
+ ε23
+ ε31
).
Minimizing eme + eel with respect to εij under fixed αi gives the equilibrium strain
or spontaneous magnetostrictive deformation
0
0
εij
= εij
(Bij kl , cij kl ).
(76)
7 Micromagnetism
369
in terms of the magnetoelastic coupling constants and the elastic stiffness constants.
Comparison of the coefficients in (76) and the experimental relation (67) allows to
express the magnetoelastic coupling coefficients in terms of the elastic stiffness constants and the magnetostriction constants. For cubic symmetry the magnetoelastic
coupling constants are
3
B1 = − λ100 (c1111 − c1122 )
2
B2 = −3λ111 c1212.
(77)
(78)
External Stress
A mechanical stress of nonmagnetic origin will have an effect on the magnetization
owing to a change of magnetoelastic coupling energy. The magnetoelastic coupling
energy density owing to an external stress σ ext is [54]
eme = −
0
σijext εij
.
(79)
ij
For cubic symmetry, this gives [20]
3
eme,cubic = − λ100 (σ11 α12 + σ22 α12 + σ33 α22 )
2
(80)
− 3λ111 (σ12 α1 α2 + σ23 α2 α3 + σ31 α3 α1 )
The above results can be derived from the strain induced by the external stress
which is
ext
εij
=
sij kl σklext ,
(81)
kl
where sij kl is the compliance tensor. Inserting (81) into (72) gives the magnetoelastic
energy density owing to external stress. For an isotropic material, for example, an
amorphous alloy, we have only a single magnetostriction constant λs = λ100 =
λ111 . For a stress σ along an axis of a unit vector a, the magnetoelastic coupling
energy reduces to
3
eme,isotropic = − λs σ (a · α)2 .
2
(82)
This equation has a similar form as that for the uniaxial anisotropy energy density
(59) with an anisotropy constant Kme = 3λs σ/2.
Magnetostrictive Self-Energy
A nonuniform magnetization causes a nonuniform spontaneous deformation owing
to (67). As a consequence, different parts of the magnet do not fit together. To
370
L. Exl et al.
el , will occur. The
compensate this misfit, an additional elastic deformation, εij
associated magnetostrictive self-energy density is
emagstr =
1 el el
cij kl εij
εkl .
2
ij
(83)
kl
el we have to solve an elasticity problem. The total strain,
To compute εij
el
0
+ εij
,
εij = εij
(84)
can be derived from a displacement field, u = (u1 , u2 , u3 ), according to [55]
1
εij =
2
∂uj
∂ui
+
∂xj
∂xi
.
(85)
We start from a hypothetically undeformed, nonmagnetic body. If magnetism is
0 causes a stress which we treat as virtual body forces. Once these
switched on, εij
forces are known, the displacement field can be calculated as usual by linear
elasticity theory. The situation is similar to magnetostatics where the demagnetizing
field is calculated from effective magnetic charges. The procedure is as follows [56].
First we compute the spontaneous magnetostrictive strain for a given magnetization
distribution with (67) or in case of cubic symmetry with (68) and (69). Then we
apply Hooke’s law to compute the stress
σij0 =
0
cij kl εkl
(86)
kl
owing to the spontaneous magnetostrictive strain. The stress is interpreted as virtual
body force
fi = −
∂
σ0.
∂xj ij
(87)
j
The forces enter the condition for mechanical equilibrium
∂
σij = fi with σij =
cij kl εkl .
∂xj
j
(88)
kl
Equations (85) to (88) lead to a system of partial differential equations for the
displacement field u(x). This is an auxiliary problem similar to the magnetostatic
boundary value problem (see section “Magnetostatic Boundary Value Problem”)
which is to be solved for a given magnetization distribution.
7 Micromagnetism
371
Based on the above discussion, we can identify two contributions to the total
magnetic Gibbs free energy: The magnetoelastic coupling energy with an external
stress
0
Eme = −
σijext εij
dV
(89)
V
ij
and the magnetostrictive self-energy
Emagstr =
1
2
V ij
0
0
cij kl (εij − εij
)(εkl − εkl
)dV .
(90)
kl
Artificial multiferroics: The magnetoelastic coupling becomes important in artificial
multiferroic structures where ferromagnetic and piezoelectric elements are combined to achieve a voltage controlled manipulation of the magnetic state [57]. For
example, piezoelectric elements can create a strain on a magnetic tunnel junction
of about 10−3 causing the magnetization to rotate by 90 degrees [58]. Breaking
the symmetry by a stress-induced uniaxial anisotropy, which can be created by a
piezoelectric element, the deterministic switching between two metastable states in
square nano-element is possible as shown in Fig. 5.
3
Fig. 5 Simulation of the stress-driven switching of a CoFeB nano-element
(Ku = 1.32 kJ/m ,
μ0 Ms = 1.29 T, A = 15 pJ/m, λs = 3 × 10−5 , mesh size h = 0.59 A/(μ0 Ms2 ) = 2 nm,
the magnetostrictive self-energy is neglected). The sample is a thin film element with dimensions
120 × 120 × 2 nm3 . The system switches from 0 to 1 by a compressive stress (−0.164 GPa) and
from 1 to 0 by a tensile stress (0.164 GPa)
372
L. Exl et al.
Characteristic Length Scales
To obtain a qualitative understanding of equilibrium states, it is helpful to consider
the relative weight of the different energy terms towards the total Gibbs free energy.
As shown in Fig. 3, the relative importance of the different energy terms changes
with the size of the magnetic sample. We can see this most easily when we write the
total Gibbs free energy
Etot = Eex + Eext + Edemag + Eani + Eme + Emagstr ,
(91)
in dimensionless form. From the relative weight of the energy contributions in
dimensionless form, we will derive characteristic length scales which will provide
useful insight into possible magnetization processes depending on the magnet’s size.
Let us assume that Ms is constant over the magnetic body (conditions 2 and 3
in section “Micromagnetics Basics”). We introduce the external and demagnetizing
field in dimensionless form hext = H ext /Ms and hdemag = H demag /Ms and rescale
the length x̃ = x/L, where L is the sample extension. Let us choose L so that
tot = Etot /(μ0 Ms2 V ). The
L3 = V . We also normalize the Gibbs free energy E
2
normalization factor, μ0 Ms V , is proportional to the magnetostatic self-energy of
the fully magnetized sample. The energy contributions in dimensionless form are
ext
E
2 lex
x 2 + ∇m
y
∇m
2
L
V
,
=−
m · hext d V
ex =
E
demag = − 1
E
2
ani = −
E
V
z
+ ∇m
2
,
dV
(92)
(93)
,
m · hdemag d V
(94)
K1
,
(c · m)2 d V
μ0 Ms2
(95)
V
V
2
is the domain after transformation of the length. Further, we assumed
where V
uniaxial magnetic anisotropy and neglected magnetoelastic coupling and magnetostriction. The constant lex in (92) is defined in the following section.
Exchange Length
In (92) we introduced the exchange length
lex =
A
.
μ0 Ms2
(96)
It describes the relative importance of the exchange energy with respect to the
magnetostatic energy. Inspecting the factor (lex /L)2 in front of the brackets in (92),
7 Micromagnetism
373
we see that the exchange energy contribution increases with decreasing sample
size L. The smaller the sample, the higher is the expense of exchange energy for
nonuniform magnetization. Therefore small samples show a uniform magnetization.
If the magnetization remains parallel during switching, the Stoner-Wohlfarth [59]
model can be applied.
In the literature, the exchange length is either defined by (96)
= 2A/(μ M 2 ) [61].
[60] or by lex
0 s
Critical Diameter for Uniform Rotation
In a sphere with uniaxial anisotropy, the magnetization reverses uniformly if its
diameter is below D ≤ Dcrit = 10.2lex [60]. During uniform rotation of the
magnetization, the exchange energy is zero, and the magnetostatic energy remains
constant. It is possible to lower the magnetostatic energy during reversal by
magnetization curling. Then the magnetization becomes nonuniform at the expense
of exchange energy. The total energy will be smaller than for uniform rotation if
the sphere diameter, D, is larger than Dcrit . Nonuniform reversal decreases the
switching field as compared to uniform rotation. The switching fields of a sphere
are [60]
Hc =
2K1
for D ≤ Dcrit .
μ0 Ms
(97)
Hc =
2K1
1
34.66A
− Ms +
for D > Dcrit .
μ0 Ms
3
μ0 Ms D 2
(98)
In cuboids and particles with polyhedral shape, the nonuniform demagnetizing field
causes a twist of the magnetization near edges or corners [62]. As a consequence
nonuniform reversal occurs for particle sizes smaller than Dcrit . The interplay
between exchange energy and magnetostatic energy also causes a size dependence
of the switching field [63, 64].
Grain size dependence of the coercive field. The coercive field of permanent
magnets decreases with increasing grain size. This can be explained by the different
scaling of the energy terms [64, 65]. The smaller the magnet, the more dominant is
the exchange term. Thus it costs more energy to form a domain wall. To achieve
magnetization reversal, the Zeeman energy of the reversed magnetization in the
nucleus needs to be higher. This can be accomplished by a larger external field.
Figure 6 shows the switching field a Nd2 Fe14 B cube as a function of its edge
length. In addition we give the theoretical switching field for a sphere with the same
volume according to (97) and (98). Magnetization reversal occurs by nucleation
and expansion of reversed domains unless the hard magnetic cube is smaller
than 6lex .
3
Fig. 6 Computed grain size
√ dependence of the coercive field of a perfect Nd2 Fe14 B cube at room temperature (K1 = 4.9 MJ/m , μ0 Ms = 1.61 T, A = 8 pJ/m,
the mesh size is h = 0.86 A/K1 = 1.1 nm, the external field is applied at an angle of 10−4 rad with respect to the easy axis). The sample dimensions are
L × L × L nm3 . Left: Switching field as function of L in units of HK . The squares give the switching field of the cube. The dashed line is the theoretical
switching field of a sphere with the same volume. A switching field smaller than HK indicates nonuniform reversal. Right: Snapshots of the magnetic states
during switching for L = 10 nm and L = 80 nm
374
L. Exl et al.
7 Micromagnetism
375
Wall Parameter
The square root of the ratio of the exchange length and the prefactor of the crystal
anisotropy energy gives another critical length. The Bloch wall parameter
δ0 =
A
K
(99)
denotes the relative importance of the exchange energy versus crystalline anisotropy
energy. It determines the width of the transition of the magnetization between two
magnetic domains. In a Bloch wall, the magnetization rotates in a way so that no
magnetic volume charges are created. The mutual competition between exchange
and anisotropy determines the domain wall width: Minimizing the exchange energy
favors wide transition regions, whereas minimizing the crystal anisotropy energy
favors narrow transition regions. In a bulk uniaxial material the wall width is
δB = π δ0 .
Single Domain Size
With increasing particle, the prefactor (lex /L)2 for the exchange energy in (92)
becomes smaller. A large particle can break up into magnetic domains because the
expense of exchange energy is smaller than the gain in magnetostatic energy. In
addition to the exchange energy, the transition of the magnetization in the domain
wall
√ also increases the crystal anisotropy energy. The wall energy per unit area is
4 AK1 . The energy of uniformly magnetized cube is its magnetostatic energy,
Edemag1 = μ0 Ms2 L3 /6. In the two domain states, the magnetostatic energy is
roughly one half
value, Edemag2 = μ0 Ms2 L3 /12. The energy of the wall
√ of this
2
is Ewall2 = 4 AK1 L . Equating the energy of the single domain state, Edemag1 ,
with the energy of the two domain state, Edemag2 + Ewall2 , and solving for L give
the single domain size of a cube
LSD ≈
√
48 AK1
.
μ0 Ms2
(100)
The above equation simply means that the energy of a ferromagnetic cube with a
size L > LSD is lower in a the two domain state than in the uniformly magnetized
state. A thermally demagnetized sample with L > LSD most likely will be in a
multidomain state.
We have to keep in mind that the magnetic state of a magnet depends on its
history and whether local or global minima can be accessed over the energy barriers
that separate the different minima. The following situations may arise:
(1) A particle in its thermally demagnetized state is multidomain although L <
LSD [66]. When cooling from the Curie temperature, a particle with L < LSD
may end up in a multidomain state. Although the single domain state has a
lower energy, it cannot be accessed because it is separated from the multidomain
376
L. Exl et al.
state by a high energy barrier. This behavior is observed in small Nd2 Fe14 B
particles [66].
(2) An initially saturated cube with L > LSD will not break up into domains
spontaneously if its anisotropy field is larger than the demagnetizing field.
The sample will remain in an almost uniform state until a reversed domain is
nucleated.
(3) Magnetization reversal of a cube with L < LSD will be nonuniform. Switching
occurs by the nucleation and expansion of a reversed domain for a particle size
down to about 5lex . For example in Nd2 Fe14 B, the single domain limit is LSD ≈
146 nm, and the exchange length is lex = 1.97 nm. The simulation presented in
Fig. 6 shows the transition from uniform to nonuniform reversal which occurs
at L ≈ 6lex .
Mesh Size in Micromagnetic Simulations
The required minimum mesh size in micromagnetic simulations depends on the
process that should be described by the simulations. Here are a few examples:
(1) For computing the switching field of a magnetic particle, we need to describe
the formation of a reversed nucleus. A reversed nucleus is formed near edges
or corners where the demagnetizing field is high. We have to resolve the
rotations of the magnetization that eventually form the reversed nucleus. For
the computation of the nucleation field the required minimum mesh size has to
be smaller than the exchange length [61] at the place where the initial nucleus
is formed.
(2) For the simulation of domain wall motion, the transition of the magnetization
between the domains needs to be resolved. A failure to do so will lead to an
artificial pinning of the domain wall on the computational grid [67]. For the
study of domain wall motion in hard magnetic materials, the required minimum
mesh size has to be smaller than the Bloch wall parameter.
(3) In soft magnetic elements with vanishing crystal or stress-induced ansisotropy,
the magnetization varies continuously [68]. The smooth transitions of the magnetization transitions can be resolved with a grid size larger than the exchange
length. Care has to be taken if vortices play a role in the magnetization process
to be studied. Then artificial pinning of vortex cores on the computational grid
[67] has to be avoided.
Brown’s Micromagnetic Equation
In the following, we will derive the equilibrium equations for the magnetization.
The total Gibbs free energy of a magnet is a functional of m(x). To compute an
equilibrium state, we have to find the function m(x) that minimizes Etot taking into
account |m(x)| = 1. In addition the boundary conditions
7 Micromagnetism
377
∇mx · n = 0, ∇my · n = 0, and ∇mz · n = 0
(101)
hold, where n is the surface normal. The boundary conditions follow from (11) and
the respective equations for y and z and applying Green’s first identity to each term
of (14). The boundary conditions (101) can also be understood intuitively [15].
To be in equilibrium, a magnetic moment at the surface has to be parallel with its
neighbor inside when there is no surface anisotropy. Otherwise there is an exchange
torque on the surface spin.
Most problems in micromagnetics can only be solved numerically. Instead
of solving the Euler-Lagrange equation that results from the variation of (91)
numerically, we directly solve the variational problem. Direct methods [69, 70]
represent the unknown function by a set of discrete variables. The minimization
of the energy with respect to these variables gives an approximate solution to the
variational problem. Two well-known techniques are the Euler method and the Ritz
method. Both are used in numerical micromagnetics.
Euler Method: Finite Differences
In finite difference micromagnetics, the solution m(x) is sampled on points
(xi , yj , zk ) so that mij k = m(xi , yj , zk ). On a regular grid with spacing h, the
positions of the grid points are xi = x0 + ih, yj = y0 + j h, and zk = z0 + kh. The
points (xi , yj , zk ) are the cell centers of the computational grid. The magnetization
is assumed to be constant within each cell. To obtain an approximation of the energy
functional, we apply the trapezoidal rule; more precisely, we replace m(x) by the
values at the cell centers mij k and the spatial derivatives of m(x) with the finite
difference quotients. The approximated solution values mij k are the unknowns of an
algebraic minimization problem. The indices i, j , and k run from 1 to the number
of grid points Nx , Ny , Nz in x, y, and z direction, respectively. In the following, we
will derive the equilibrium equations whereby for simplicity we will not take into
account the magnetoelastic coupling energy and the magnetostrictive self-energy.
We can approximate the exchange energy (14) on the finite difference grid as [71]
Eex
2Ai+1j k Aij k mx,i+1j k − mx,ij k 2
≈ h3
+ ··· ,
Ai+1j k + Aij k
h
(102)
ij k
where we introduced the notation Aij k = A(xi , yj , zk ). The prefactor in (102) is
the harmonic mean of the values for the exhange constants in cells i + 1j k and ij k.
This follows from the interface condition Ai+1j k (mx,i+1j k − mx,interface )/(h/2) =
Aij k (mx,interface − mx,ij k )/(h/2), where the minterface is the magnetization at the
interface between the two cells.
Eext ≈ −μ0 h3
ij k
Ms,ij k (mij k · H ext,ij k ).
(103)
378
L. Exl et al.
To approximate the magnetostatic energy, we use (42) and (45). Replacing the
integrals with sums over the computational cell, we obtain
Edemag ≈
μ0 Ms,ij k Ms,i j k 8π
∂Vij k
ij k i j k
∂Vi j k (mij k · n)(mi j k · n )
dSdS .
|x − x |
(104)
The volume integrals in (42) and (45) vanish when we assume that m(x) is constant
within each computational cell ij k. The magnetostatic energy is often expressed in
terms of the demagnetizing tensor Nij k,i j k Edemag ≈
μ0 3 h
Ms,ij k mTij k Nij k,i j k mi j k Ms,i j k 2
(105)
ij k i j k
We approximate the anisotropy energy (60) by
Eani ≈ h3
(106)
eani (mij k ).
ij k
The total energy is now a function of the unknowns mij k . The constraint (5) is
approximated by
|mij k | = 1
(107)
where ij k runs over all computational cells. We obtain the equilibrium equations
from differentiation
⎤
⎡
Lij k
∂ ⎣
(mij k · mij k − 1)⎦ = 0,
Etot (. . . , mij k , . . . ) +
∂mij k
2
⎡
∂ ⎣
Etot (. . . , mij k , . . . ) +
∂Lij k
ij k
Lij k
ij k
2
(108)
⎤
(mij k · mij k − 1)⎦ = 0.
(109)
In the brackets we added a Lagrange function to take care of the constraints (107).
Lij k are Lagrange multipliers. From (108) we obtain the following set of equations
for the unknowns mij k
−2Aij k h3
2Ai−1j k
mi+1j k − mij k
mi−1j k − mij k
2Ai+1j k
+
+
·
·
·
Ai+1j k + Aij k
Ai−1j k + Aij k
h2
h2
−μ0 Ms,ij k h3 H ext,ij k
(110)
7 Micromagnetism
+μ0 Ms,ij k h3
379
N ij k,i j k mi j k Ms,i j k i j k
+h3
∂eani
= −Lij k mij k .
∂mij k
The term in brackets is the Laplacian discretized on a regular grid. First-order
equilibrium conditions require also zero derivative with respect to the Lagrange
multipliers. This gives back the constraints (107). It is convenient to collect all terms
with the dimensions of A/m to the effective field
H eff,ij k = H ex,ij k + H ext,ij k + H demag,ij k + H ani,ij k.
(111)
The exchange field, the magnetostatic field, and the anisotropy field at the computational cell ij k are
H ex,ij k =
2Aij k
μ0 Ms,ij k
H demag,ij k = −
2Ai−1j k
mi+1j k − mij k
2Ai+1j k
+
Ai+1j k + Aij k
Ai−1j k + Aij k
h2
mi−1j k − mij k
+
·
·
·
(112)
h2
Nij k,i j k mi j k Ms,i j k (113)
i j k
H ani,ij k = −
∂eani
1
,
μ0 Ms,ij k ∂mij k
(114)
respectively. The evaluation of the exchange field (112) requires values of mij k
outside the index range [1, Nx ] × [1, Ny ] × [1, Nz ]. These values are obtained by
mirroring the values of the surface cell at the boundary. This method of evaluating
the exchange field takes into account the boundary conditions (101).
Using the effective field, we can rewrite the equilibrium equations
μ0 Ms,ij k h3 H eff,ij k = Lij k mij k .
(115)
Equation (115) states that the effective field is parallel to the magnetization at each
computational cell. Instead of (115) we can also write
μ0 Ms,ij k h3 mij k × H eff,ij k = 0.
(116)
The expression Ms,ij k h3 mij k is the magnetic moment of computational cell ij k.
Comparison with (1) shows that in equilibrium the torque for each small volume
element h3 (or computational cell) has to be zero. The constraints (107) also have
to be fulfilled in equilibrium.
380
L. Exl et al.
Ritz Method: Finite Elements
Within, the framework of the Ritz method the solution is assumed to depend on a
few adjustable parameters. The minimization of the total Gibbs free energy with
respect to these parameters gives an approximate solution [15, 16]. In the following
we describe a famous and computationally efficient Ritz method, namely, the finite
element ansatz.
Most finite element solvers for micromagnetics use a magnetic scalar potential
for the computation of the magnetostatic energy. This goes back to Brown [16] who
introduced an expression for the magnetostatic energy, Edemag
(m, U ), in terms of
the scalar potential for the computation of equilibrium magnetic states using the
Ritz method. We replace Edemag (m) with Edemag
(m, U ), as introduced in (55), in
the expression for the total energy. The vector m(x) is expanded by means of basis
functions ϕi with local support around node x i
mfe (x) =
(117)
ϕi (x)mi .
i
Similarly, we expand the magnetic scalar potential
U fe (x) =
(118)
ϕi (x)Ui .
i
The index i runs over all nodes of the finite element mesh. The expansion
coefficients mi and Ui are the nodal values of the unit magnetization vector and
the magnetic scalar potential, respectively. We assume that the constraint |m| = 1
is fulfilled only at the nodes of the finite element mesh. We introduce a Lagrange
function; Li are the Lagrange multipliers at the nodes of the finite element mesh.
By differentiation with respect to mi , Ui , and Li , we obtain the equilibrium
conditions
∂
∂mi
∂
∂Ui
∂
∂Li
Etot (. . . , mi , Ui . . . ) +
Li
i
Etot (. . . , mi , Ui . . . ) +
Li
i
Etot (. . . , mi , Ui . . . ) +
2
2
Li
i
2
(mi · mi − 1) = 0,
(119)
(mi · mi − 1) = 0,
(120)
(mi · mi − 1) = 0.
From (119) we obtain the following set of equations for the unknowns mi
(121)
7 Micromagnetism
381
2
A∇ϕi · ∇ϕj dV mj
V
j
−
μ0 Ms H ext ϕi dV
V
+
(122)
μ0 Ms ∇U ϕi dV
V
∂eani (
+
V
j
ϕj mj )
∂mi
dV = −Li mi .
Equation (120) is the discretized form of the partial differential equation (50) for the
magnetic scalar potential. Equation (121) gives back the constraint |m| = 1.
In the following, we introduce the effective field at the nodes of the finite element
mesh
H eff,i = −
2 μ0 M
j
A∇ϕi · ∇ϕj dV mj
V
+ H ext,i + H demag,i
1
−
μ0 M
V
(123)
∂eani
dV ,
∂mi
where M = V Ms ϕi dV . H demag,i is the demagnetizing field averaged over the
finite elements surrounding node i. This average can be computed by plugging (118)
into the third line of (122) and dividing the resulting expression by −μ0 M.
The equilibrium equations are
μ0 MH eff,i = Li mi .
(124)
We can write the equilibrium conditions in terms of a cross product of the magnetic
moment, Mmi , and the effective field at node i
μ0 Mmi × H eff,i = 0.
(125)
The system is in equilibrium if the torque equals zero and the constraint |mi | = 1 is
fulfilled on all nodes of the finite element mesh.
Instead of a Lagrange function for keeping the constraint |m| = 1, projection
methods [72] are commonly used in fast micromagnetic solvers [73]. In the iterative
scheme for solving (125), the search direction d k+1
is projected onto a plane
i
perpendicular to mki , corresponding to first-order approximation of the constraint
is normalized.
at node i. After each iteration k, the vector mk+1
i
382
L. Exl et al.
Magnetization Dynamics
Brown’s equations describes the conditions for equilibrium. In many applications,
the response of the system to a time varying external field is important. The
equations by Landau-Lifshitz [74] or Gilbert [75] describes the time evolution of
the magnetization. The Gilbert equation in Landau-Lifshitz form
|γ |μ0
∂m
|γ |μ0 α
=−
m × H eff −
m × (m × H eff )
∂t
1 + α2
1 + α2
(126)
is widely used in numerical micromagnetics. Here |γ | = 1.76086 × 1011 s−1 T−1 is
the gyromagnetic ratio and α is the Gilbert damping constant. In (126) the unit
vector of the magnetization and the effective field at the grid point of a finite
difference grid or finite element mesh may be used for m and H eff . The first term of
(126) describes the precession of the magnetization around the effective field. The
last term of (126) describes the damping. The double cross product gives the motion
of the magnetization towards the effective field.
The interplay between the precession and the damping terms leads to damped
oscillations of the magnetization around its equilibrium state. In the limiting case of
small deviations from equilibrium and uniform magnetization, the amplitude of the
oscillations decay as [76]
a(t) = Ce−t/t0 .
(127)
For small damping, the oscillations decay time is [76]
t0 =
2
.
αγ μ0 Ms
(128)
Switching of magnetic nano-elements. Small thin film nano-elements are key
building blocks of magnetic sensor and storage applications. By application of a
short field pulse, a thin film nano-element can be switched. After reversal, the system
relaxes to its equilibrium state by damped oscillations. Figure 7 shows the switching
dynamics of a NiFe film with a length of 100 nm, a width of 20 nm, and a thickness
of 2 nm. In equilibrium the magnetization is parallel to the long axis of the particle
(x axis). A Gaussian field pulse (dotted line in Fig. 7) is applied in the (-1,-1,-1)
direction. After the field is switched off the magnetization oscillates towards the
long axis of the film. From an exponential fit to the envelope of the magnetization
component, My (t), parallel to the short axes, we derived the characteristic decay
times of the oscillation which are t0 ≈ 0.613 ns and t0 ≈ 0.204 ns for a damping
constant of α = 0.02 and α = 0.06, respectively. According to (128), the difference
between the two relaxations times is a factor of 3, given by the ratio of the damping
constants.
Fig. 7 Switching of a thin film nano-element by
a short field pulse in the (-1,-1,-1) direction for α = 0.06 (top row) and α = 0.02 (bottom row). (K1 = 0,
μ0 Ms = 1 T, A = 10 pJ/m, mesh size h = 0.56 A/(μ0 Ms2 ) = 2 nm). The sample dimensions are 100 × 20 × 2 nm3 . The sample is originally magnetized in
the +x direction. Left: Magnetization as function of time. The thin dotted line gives the field pulse, Hext (t). Once the field is switched off damped oscillations
occur which are clearly seen in My (t). The bold grey line is a fit to the envelope of the magnetization component parallel to the short axis. Right: Transient
magnetic states. The numbers correspond to the black dots in the plot of My (t) on the left
7 Micromagnetism
383
384
L. Exl et al.
Acknowledgments The authors thank the Austrian Science Fund (FWF) under grant No. F4112
SFB ViCoM and grant No. P31140-N32 for financial support. The financial support by the
Austrian Federal Ministry for Digital and Economic Affairs, the National Foundation for Research,
Technology and Development and the Christian Doppler Research Association is gratefully
acknowledged.
Appendix
The intrinsic material properties listed in Table 1 are taken from [77]. The exchange
lengths
and
the
wall
parameter
are
calculated
as
follows:
l
=
A/(μ0 Ms2 ), δ0 =
ex
√
A/|K1 |.
Table 1 Intrinsic magnetic properties and characteristic lengths of selected magnetic materials
Material
Fe
Co
Ni
Ni0.8 Fe0.2
CoPt
Nd2 Fe14 B
SmCo5
Sm2 Co17
Fe3 O4
TC (K)
1044
1360
628
843
840
588
1020
1190
860
μ0 Ms (T)
2.15
1.82
0.61
1.04
1.01
1.61
1.08
1.25
0.6
A(pJ/m)
22
31
8
10
10
8
12
16
7
K1 (kJ/m3 )
48
410
-5
-1
4900
4900
17200
4200
-13
lex (nm)
2.4
3.4
5.2
3.4
3.5
2.0
3.6
3.6
4.9
δ0 (nm)
21
8.7
40
100
1.4
1.3
0.8
2.0
23
The examples given in Figs. 3 to 7 were computed using the micromagnetic
simulation environment FIDIMAG [43]. FIDIMAG solves finite difference micromagnetic problems using a Python interface. The reader is encouraged to run
computer experiments for further exploration of micromagnetism. In the following
we illustrate the use of the Python interface for simulating the switching dynamics
of a magnetic nano-element (see Fig. 7). The function relax_system computes
the initial magnetic state. The function apply_field computes the response of the
magnetization under the influence of a time varying external field.
import numpy a s np
from f i d i m a g . m i c r o import Sim
from f i d i m a g . common import CuboidMesh
from f i d i m a g . m i c r o import UniformExchange , Demag
from f i d i m a g . m i c r o import TimeZeeman
mu0 = 4 ∗ np . p i ∗ 1 e−7
7 Micromagnetism
A
Ms
385
= 1 . 0 e −11
= 1 . / mu0
d e f r e l a x _ s y s t e m ( mesh ) :
sim = Sim ( mesh , name= ’ r e l a x ’ )
sim . d r i v e r . s e t _ t o l s ( r t o l =1e −10 , a t o l =1e −10)
sim . d r i v e r . a l p h a = 0 . 5
sim . d r i v e r . gamma = 2 . 2 1 1 e5
sim . Ms = Ms
sim . d o _ p r e c e s s i o n = F a l s e
sim . s e t _ m ( ( 0 . 5 7 7 3 5 0 2 6 9 , 0 . 5 7 7 3 5 0 2 6 9 , 0 . 5 7 7 3 5 0 2 6 9 ) )
sim . add ( U n i f o r m E x c h a n g e (A=A) )
sim . add ( Demag ( ) )
sim . r e l a x ( )
np . s a v e ( ’m0 . npy ’ , sim . s p i n )
d e f a p p l y _ f i e l d ( mesh ) :
sim = Sim ( mesh , name= ’ dyn ’ )
sim . d r i v e r . s e t _ t o l s ( r t o l =1e −10 , a t o l =1e −10)
sim . d r i v e r . a l p h a = 0 . 0 2
sim . d r i v e r . gamma = 2 . 2 1 1 e5
sim . Ms = Ms
sim . s e t _ m ( np . l o a d ( ’m0 . npy ’ ) )
sim . add ( U n i f o r m E x c h a n g e (A=A) )
sim . add ( Demag ( ) )
s i g m a = 0 . 1 e−9
def gaussian_fun ( t ) :
r e t u r n np . exp ( −0.5 ∗ ( ( t −3∗ s i g m a ) / s i g m a ) ∗ ∗ 2 )
mT = 0 . 0 0 1 / mu0
zeeman = TimeZeeman ([ −100 ∗ mT, −100 ∗ mT, −100 ∗
mT] , t i m e _ f u n = g a u s s i a n _ f u n , name= ’H ’ )
sim . add ( zeeman , s a v e _ f i e l d = T r u e )
sim . r e l a x ( d t = 1 . e −12 , m a x _ s t e p s = 10000)
i f __name__ == ’ __main__ ’ :
mesh = CuboidMesh ( nx =50 , ny =10 , nz =1 , dx =2 , dy =2 ,
dz =2 , u n i t _ l e n g t h =1e −9)
r e l a x _ s y s t e m ( mesh )
a p p l y _ f i e l d ( mesh )
386
L. Exl et al.
References
1. Fukuda, H., Nakatani, Y.: Recording density limitation explored by head/media cooptimization using genetic algorithm and GPU-accelerated LLG. IEEE Trans. Magn. 48(11),
3895–3898 (2012)
2. Greaves, S., Katayama, T., Kanai, Y., Muraoka, H.: The dynamics of microwave-assisted
magnetic recording. IEEE Trans. Magn. 51(4), 1–7 (2015)
3. Wang, H., Katayama, T., Chan, K.S., Kanai, Y., Yuan, Z., Shafidah, S.: Optimal write head
design for perpendicular magnetic recording. IEEE Trans. Magn. 51(11), 1–4 (2015)
4. Kovacs, A., Oezelt, H., Schabes, M.E., Schrefl, T.: Numerical optimization of writer and media
for bit patterned magnetic recording. J. Appl. Phys. 120(1), 013902 (2016)
5. Vogler, C., Abert, C., Bruckner, F., Suess, D., Praetorius, D.: Areal density optimizations for
heat-assisted magnetic recording of high-density media. J. Appl. Phys. 119(22), 223903 (2016)
6. Makarov, A., Sverdlov, V., Osintsev, D., Selberherr, S.: Fast switching in magnetic tunnel
junctions with two pinned layers: micromagnetic modeling, IEEE Trans. Magn. 48(4), 1289–
1292 (2012)
7. Lacoste, B., de Castro, M.M., Devolder, T., Sousa, R., Buda-Prejbeanu, L., Auffret, S., Ebels,
U., Ducruet, C., Prejbeanu, I., Rodmacq, B., Dieny, B., Vila, L.: Modulating spin transfer
torque switching dynamics with two orthogonal spin-polarizers by varying the cell aspect ratio.
Phys. Rev. B 90(22), 224404 (2014)
8. Hou, Z., Silva, A., Leitao, D., Ferreira, R., Cardoso, S., Freitas, P.: Micromagnetic and
magneto-transport simulations of nanodevices based on mgo tunnel junctions for memory and
sensing applications. Phys. B Condens. Matter 435, 163–167 (2014)
9. Leitao, D.C., Silva, A.V., Paz, E., Ferreira, R., Cardoso, S., Freitas, P.P.: Magnetoresistive
nanosensors: controlling magnetism at the nanoscale. Nanotechnology 27(4), 045501 (2015)
10. Ennen, I., Kappe, D., Rempel, T., Glenske, C., Hütten, A.: Giant magnetoresistance: Basic
concepts, microstructure, magnetic interactions and applications. Sensors 16(6), 904 (2016)
11. Sepehri-Amin, H., Ohkubo, T., Nagashima, S., Yano, M., Shoji, T., Kato, A., Schrefl, T.,
Hono, K.: High-coercivity ultrafine-grained anisotropic Nd–Fe–B magnets processed by hot
deformation and the nd–cu grain boundary diffusion process. Acta Mater. 61(17), 6622–6634
(2013)
12. Bance, S., Oezelt, H., Schrefl, T., Winklhofer, M., Hrkac, G., Zimanyi, G., Gutfleisch, O.,
Evans, R., Chantrell, R., Shoji, T., Yano, M., Sakuma, N., Kato, A., Manabe, A.: High energy
product in Battenberg structured magnets. Appl. Phys. Lett. 105(19), 192401 (2014)
13. Fukunaga, H., Hori, R., Nakano, M., Yanai, T., Kato, R., Nakazawa, Y.: Computer simulation
of coercivity improvement due to microstructural refinement. J. Appl. Phys. 117(17), 17A729
(2015)
14. Evans, R.F., Fan, W.J., Chureemart, P., Ostler, T.A., Ellis, M.O., Chantrell, R.W.: Atomistic
spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter 26(10), 103202
(2014)
15. Brown Jr, W.F.: Micromagnetics, domains, and resonance. J. Appl. Phys. 30(4), S62–S69
(1959)
16. Brown, W.F.: Micromagnetics. Interscience Publishers, New York/London (1963)
17. Heisenberg, W.: Zur Theorie des Ferromagnetismus. Zeitschrift für Physik 49(9), 619–636
(1928)
18. Vleck, V.: The Theory of Electric and Magnetic Susceptibilities. Oxford University Press,
London (1932)
19. Harashima, Y., Terakura, K., Kino, H., Ishibashi, S., Miyake, T.: Nitrogen as the best interstitial
dopant among x = B, C, N, O, and F for strong permanent magnet NdFe11 TiX: First-principles
study. Phys. Rev. B 92, 184426 (2015)
20. Cullity, B.D., Graham, C.D.: Introduction to Magnetic Materials, 2nd edn. IEEE Press,
Hoboken (2009)
21. Kneller, E.: Ferromagnetismus. Springer, Berlin/Göttingen/Heidelberg (1962)
7 Micromagnetism
387
22. Kittel, C.: Introduction to Solid State Physics, 8th edn. Wiley, New York (2005)
23. Talagala, P., Fodor, P., Haddad, D., Naik, R., Wenger, L., Vaishnava, P., Naik, V.: Determination
of magnetic exchange stiffness and surface anisotropy constants in epitaxial Ni1−x Cox (001)
films. Phys. Rev. B 66(14), 144426 (2002)
24. Eyrich, C., Zamani, A., Huttema, W., Arora, M., Harrison, D., Rashidi, F., Broun, D., Heinrich,
B., Mryasov, O., Ahlberg, M., Karis, O., Jönsson, P., From, M., Zhu, X., Girt, E.: Effects of
substitution on the exchange stiffness and magnetization of co films. Phys. Rev. B 90(23),
235408 (2014)
25. Weissmüller, J., McMichael, R.D., Michels, A., Shull, R.: Small-angle neutron scattering by
the magnetic microstructure of nanocrystalline ferromagnets near saturation. J. Res. Natl. Inst.
Stand. Technol. 104(3), 261 (1999)
26. Ono, K., Inami, N., Saito, K., Takeichi, Y., Yano, M., Shoji, T., Manabe, A., Kato, A., Kaneko,
Y., Kawana, D., Yokoo, T., Itoh, S.: Observation of spin-wave dispersion in nd-fe-b magnets
using neutron brillouin scattering. J. Appl. Phys. 115(17), 17A714 (2014)
27. Smith, N., Markham, D., LaTourette, D.: Magnetoresistive measurement of the exchange
constant in varied-thickness permalloy films. J. Appl. Phys. 65(11), 4362–4365 (1989)
28. Livingston, J., McConnell, M.: Domain-wall energy in cobalt-rare-earth compounds. J. Appl.
Phys. 43(11), 4756–4762 (1972)
29. Livingston, J.: Magnetic domains in sintered Fe-Nd-B magnets. J. Appl. Phys. 57(8), 4137–
4139 (1985)
30. Newnham, S., Jakubovics, J., Daykin, A.: Domain structure of thin NdFeB foils. J. Magn.
Magn. Mater. 157, 39–40 (1996)
31. Jackson, J.D.: Classical Electrodynamics, 3rd edn. Wiley, Singapore (1999)
32. Steele, C.: Iterative algorithm for magnetostatic problems with saturable media. IEEE Trans.
Magn. 18(2), 393–396 (1982)
33. Senanan, K., Victora, R.: Effect of medium permeability on the perpendicular recording
process. Appl. Phys. Lett. 81(20), 3822–3824 (2002)
34. Brown Jr, W.F.: Domains, micromagnetics, and beyond: Reminiscences and assessments. J.
Appl. Phys. 49(3), 1937–1942 (1978)
35. LaBonte, A.: Two-dimensional bloch-type domain walls in ferromagnetic films. J. Appl. Phys.
40(6), 2450–2458 (1969)
36. Schabes, M., Aharoni, A.: Magnetostatic interaction fields for a three-dimensional array of
ferromagnetic cubes. IEEE Trans. Magn. 23(6), 3882–3888 (1987)
37. Aharoni, A.: Demagnetizing factors for rectangular ferromagnetic prisms. J. Appl. Phys. 83(6),
3432–3434 (1998)
38. Dittrich, R., Scholz, W.: Calculator for magnetostatic energy and demagnetizing factor. http://
www.magpar.net/static/magpar/doc/html/demagcalc.html (2016) Accessed 11 Aug 2016
39. Sato, M., Ishii, Y.: Simple and approximate expressions of demagnetizing factors of uniformly
magnetized rectangular rod and cylinder. J. Appl. Phys. 66(2), 983–985 (1989)
40. Asselin, P., Thiele, A.: On the field lagrangians in micromagnetics. IEEE Trans. Magn. 22(6),
1876–1880 (1986)
41. Donahue, M., Porter, D., McMichael, R., Eicke, J.: Behavior of μmag standard problem no. 2
in the small particle limit. J. Appl. Phys. 87, 5520–5522 (2000)
42. Vansteenkiste, A., Leliaert, J., Dvornik, M., Helsen, M., Garcia-Sanchez, F., Van Waeyenberge,
B.: The design and verification of MuMax3. AIP Adv. 4(10), 107133 (2014)
43. Bisotti, M.-A., Cortés-Ortuño, D., Pepper, R., Wang, W., Beg, M., Kluyver, T., Fangohr, H.:
Fidimag–a finite difference atomistic and micromagnetic simulation package. J. Open Res.
Softw. 6, 1 (2018)
44. Abert, C., Selke, G., Kruger, B., Drews, A.: A fast finite-difference method for micromagnetics
using the magnetic scalar potential. IEEE Trans. Magn. 48(3), 1105–1109 (2012)
45. Fu, S., Cui, W., Hu, M., Chang, R., Donahue, M.J., Lomakin, V.: Finite-difference micromagnetic solvers with the object-oriented micromagnetic framework on graphics processing units.
IEEE Trans. Magn. 52(4), 1–9 (2016)
388
L. Exl et al.
46. Scholz, W., Fidler, J., Schrefl, T., Suess, D., Dittrich, R., Forster, H., Tsiantos, V.: Scalable
parallel micromagnetic solvers for magnetic nanostructures. Comput. Mater. Sci. 28(2), 366–
383 (2003)
47. Fischbacher, T., Franchin, M., Bordignon, G., Fangohr, H.: A systematic approach to multiphysics extensions of finite-element-based micromagnetic simulations: Nmag. IEEE Trans.
Magn. 43(6), 2896–2898 (2007)
48. Abert, C., Exl, L., Bruckner, F., Drews, A., Suess, D.: magnum.fe: A micromagnetic finiteelement simulation code based on FEniCS. J. Magn. Magn. Mater. 345, 29–35 (2013)
49. Chang, R., Li, S., Lubarda, M., Livshitz, B., Lomakin, V.: Fastmag: Fast micromagnetic
simulator for complex magnetic structures. J. Appl. Phys. 109(7), 07D358 (2011)
50. Forster, H., Schrefl, T., Dittrich, R., Scholz, W., Fidler, J.: Fast boundary methods for
magnetostatic interactions in micromagnetics. IEEE Trans. Magn. 39(5), 2513–2515 (2003)
51. Cowburn, R., Welland, M.: Micromagnetics of the single-domain state of square ferromagnetic
nanostructures. Phys. Rev. B 58(14), 9217 (1998)
52. Goll, D., Schütz, G., Kronmüller, H.: Critical thickness for high-remanent single-domain
configurations in square ferromagnetic thin platelets. Phys. Rev. B 67(9), 094414 (2003)
53. James, P., Eriksson, O., Hjortstam, O., Johansson, B., Nordström, L.: Calculated trends of the
magnetostriction coefficient of 3D alloys from first principles. Appl. Phys. Lett. 76(7), 915–917
(2000)
54. Hubert, A., Schäfer, R.: Magnetic Domains, corrected, 3rd printing ed. Springer, Berlin/New
York (2009)
55. Rand, O., Rovenski, V.: Analytical Methods in Anisotropic Elasticity. Birkhäuser,
Boston/Basel/Berlin (2005)
56. Shu, Y., Lin, M., Wu, K.: Micromagnetic modeling of magnetostrictive materials under
intrinsic stress. Mech. Mater. 36(10), 975–997 (2004)
57. Peng, R.-C., Hu, J.-M., Momeni, K., Wang, J.-J., Chen, L.-Q., Nan, C.-W.: Fast 180o
magnetization switching in a strain-mediated multiferroic heterostructure driven by a voltage.
Sci. Rep. 6, 27561 (2016)
58. Zhao, Z., Jamali, M., D’Souza, N., Zhang, D., Bandyopadhyay, S., Atulasimha, J., Wang, J.-P.:
Giant voltage manipulation of mgo-based magnetic tunnel junctions via localized anisotropic
strain: A potential pathway to ultra-energy-efficient memory technology. Appl. Phys. Lett.
109(9), 092403 (2016)
59. Stoner, E.C., Wohlfarth, E.: A mechanism of magnetic hysteresis in heterogeneous alloys.
Philos. Trans. R. Soc. Lond. A Math. Phys. Eng. Sci. 240(826), 599–642 (1948)
60. Skomski, R.: Nanomagnetics. J. Phys. Condens. Matter 15(20), R841 (2003)
61. Rave, W., Ramstöck, K., Hubert, A.: Corners and nucleation in micromagnetics. J. Magn.
Magn. Mater. 183(3), 329–333 (1998)
62. Schabes, M.E., Bertram, H.N.: Magnetization processes in ferromagnetic cubes. J. Appl. Phys.
64(3), 1347–1357 (1988)
63. Schmidts, H., Kronmüller, H.: Size dependence of the nucleation field of rectangular ferromagnetic parallelepipeds. J. Magn. Magn. Mater. 94(1), 220–234 (1991)
64. Bance, S., Seebacher, B., Schrefl, T., Exl, L., Winklhofer, M., Hrkac, G., Zimanyi, G., Shoji,
T., Yano, M., Sakuma, N., Ito, M., Kato, A., Manabe, A.: Grain-size dependent demagnetizing
factors in permanent magnets. J. Appl. Phys. 116(23), 233903 (2014)
65. Thielsch, J., Suess, D., Schultz, L., Gutfleisch, O.: Dependence of coercivity on length ratios in
sub-micron Nd2 Fe14 B particles with rectangular prism shape. J. Appl. Phys. 114(22), 223909
(2013)
66. Crew, D., Girt, E., Suess, D., Schrefl, T., Krishnan, K., Thomas, G., Guilot, M.: Magnetic
interactions and reversal behavior of Nd2 Fe14 B particles diluted in a Nd matrix. Phys. Rev. B
66(18), 184418 (2002)
67. Donahue, M., McMichael, R.: Exchange energy representations in computational micromagnetics. Phys. B Condens. Matter 233(4), 272–278 (1997)
68. Schäfer, R.: Domains in extremely soft magnetic materials. J. Magn. Magn. Mater. 215, 652–
663 (2000)
7 Micromagnetism
389
69. Courant, R., Hilbert, D.: Methods of Mathematical Physics. Interscience Publishers, New York
(1953)
70. Komzsik, L.: Applied Calculus of Variations for Engineers. CRC Press, Boca Raton/London/New York (2009)
71. Victora, R.: Micromagnetic predictions for magnetization reversal in coni films. J. Appl. Phys.
62(10), 4220–4225 (1987)
72. Cohen, R., Lin, S.-Y., Luskin, M.: Relaxation and gradient methods for molecular orientation
in liquid crystals. Comput. Phys. Commun. 53(1–3), 455–465 (1989)
73. Exl, L., Bance, S., Reichel, F., Schrefl, T., Stimming, H.P., Mauser, N.J.: LaBonte’s method
revisited: An effective steepest descent method for micromagnetic energy minimization. J.
Appl. Phys. 115(17), 17D118 (2014)
74. Landau, L.D., Lifshitz, E.: On the theory of the dispersion of magnetic permeability in
ferromagnetic bodies. Phys. Z. Sowjetunion 8(153), 101–114 (1935)
75. Gilbert, T.: A lagrangian formulation of the gyromagnetic equation of the magnetization field.
Phys. Rev. 100, 1243 (1955)
76. Miltat, J., Albuquerque, G., Thiaville, A.: An introduction to micromagnetics in the dynamic
regime. In: Hillebrands, B., Ounadjela, K. (eds.) Spin Dynamics in Confined Magnetic
Structures, pp. 1–34. Springer, Berlin (2002)
77. Coey, M.D.J.: Magnetism and Magnetic Materials:. Cambridge University Press, Cambridge
001 (2001)
Lukas Exl studied mathematics and computational physics and
received his PhD from TU-Wien in 2014. He is currently running
the project “Reduced Order Approaches in Micromagnetism” at
WPI. He works on computational methods in magnetism and
quantum mechanics with emphasis on (data-driven) PDEs and
model reduction. He is Senior Scientist at the University of
Vienna and lectures numerical methods.
Dieter Suess received his PhD from the TU-Wien in 2002 where
he completed his Habilitation in 2007 in “Computational Material
Science.” In 2006 he proposed “Exchange Spring Media” for
recording. Since 2018 he is assoc. Prof. and Group Speaker of
the “Physics of Functional Materials” group at the University of
Vienna.
390
L. Exl et al.
Thomas Schrefl received his PhD from the TU-Wien in 2002
where he completed his Habilitation in 1999 in “Computational
Physics.” He worked on the development of numerical micromagnetic solvers for application in magnetic recording and permanent
magnet. He his head of the Center for Modelling and Simulation
at Danube University Krems, Austria
8
Magnetic Domains
Rudolf Schäfer
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Relevance of Domains and Domain Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Domain Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Driving Forces for Domain Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Interplay of Energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Domain Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Domain Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Domain Wall Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Domain Wall Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Current-Driven Domain Wall Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
392
393
398
399
404
406
409
414
415
420
426
432
Abstract
Magnetic domains are the basic elements of the magnetic microstructure of
magnetically ordered materials. They are formed to minimize the total energy,
with the stray field energy being the most significant contribution. The reordering
of domains in magnetic fields determines the magnetization curve, domains can
be engineered on purpose, and they can be applied in devices. In this chapter
a review of the basics of magnetic domains is presented. It will be shown how
the magnetic energies act together to determine the domain character and how
R. Schäfer ()
Institute for Metallic Materials, Leibniz Institute for Solid State and Materials Research (IFW)
Dresden, Dresden, Germany
Institute for Materials Science, Dresden University of Technology, Dresden, Germany
e-mail: [email protected]
© Springer Nature Switzerland AG 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_8
391
392
R. Schäfer
domains can be classified. Domain walls and their dynamics, both field- and
current-driven, will be addressed.
Introduction
According to Fig. 1, magnetically ordered materials may be described by five scaledependent hierarchic levels [1]: atomic level theory, level (1), explains the origin and
magnitude of magnetic moments, crystal anisotropy, or magnetoelastic interactions,
and it deals with the arrangement of spins on the crystal lattice sites. The theory
works on a microscopic, sub-nanometer scale level. The other extreme level, the
magnetization Curve in level (5), describes the average magnetization of a specimen
as a function of applied magnetic field and may be seen as a macroscopic descriptive
level.
These two extreme levels are interlinked by magnetic microstructure analysis,
levels (2) to (4) in Fig. 1. Here the individual atomic magnetic moments, defined in
5. Magnetization Curve
(independent of scale)
M/Ms 1
H/Ha
1
sis
c
eti
gn
Ma
eA
tur
2. Micromagnetic
Analysis
(1 - 1000 nm)
uc
str
ly
na
cro
Mi
H
3. Domain Analysis
(1 - 1000 µm)
1. Atomic Level Theory
(< 1 nm)
4. Phase Analysis
(> 100 µm)
Fig. 1 Five descriptive levels of magnetically ordered materials, illustrating the link-function
of magnetic microstructure between atomic foundations and technical applications of magnetic
materials. The anisotropy field Ha , used in the magnetization curve, is defined as Ha = 2K1 /μ0 Ms
with K1 and Ms being the first-order cubic anisotropy constant and saturation magnetization,
respectively. Indicated are the sample dimensions for which the five concepts are applicable. The
domain image in level (3) was obtained by magneto-optical Kerr microscopy on the (100) surface
of an Fe3wt%Si-sheet of 0.5 mm thickness, and the M(H )-loop in level (5) was calculated by
magnetic phase theory for the configuration in level (4). (The domain image at level (3) is adapted
by permission from Ref. [1] (c) Springer 1998)
8 Magnetic Domains
393
level (1), are no longer considered. One rather sums them in a certain neighborhood
and represents the local average over a small volume of magnetic moments by a
classical magnetization vector M – the discrete, quantum mechanical, and statistical
properties of the spins and elementary moments are thus ignored. For a constant
temperature this vector has a constant length, the technical saturation magnetization
Ms . Then one can divide the M-vector by the saturation magnetization, leading
to the unit vector of magnetization m(r) = M(r)/Ms with m2 = 1. For this
mesoscopic approach in the description of magnetic materials, it does not matter
whether a material is ferromagnetic or ferrimagnetic, as the latter is also characterized by a net magnetization vector. The purpose of magnetic microstructure analysis
is the determination of the vector field m(r) and its response to a magnetic field.
The magnetization vectors are typically arranged as magnetic domains, so Domain
Analysis – level (3) – is in the center of magnetic microstructure analysis. Level
(2), the continuum theory of micromagnetics, deals with the connecting elements
between the domains, the domain walls, and their substructures. Level (4), phase
theory, ignores the specific arrangement of the domains and rather focuses on their
volume distribution by collecting all domains that are magnetized along a specific
direction in a domain phase. The rearrangement of the phases in a magnetic field
finally leads to the (idealized) magnetization curve. So the domains are at the end
responsible for magnetization curves.
In a conventional definition, magnetic domains are uniformly magnetized regions
that appear spontaneously in otherwise unstructured ferro- or ferrimagnetic samples
[1]. In the example presented in the center of Fig. 1, the domains are well-ordered
owing to the facts that the surface of the crystal is “well”-oriented, meaning
that it contains two easy anisotropy axes for the magnetization, that the crystal
is largely free from internal mechanical stress, and that no magnetic field is
applied to the specimen. In general, however, magnetic domains do not have to be
uniformly magnetized, and they can be more or less complex depending on many
circumstances. To give an impression of the variability of domain patterns, a collage
of selected domain images of various magnetic materials is shown in Fig. 2.
In this chapter a review of some basic aspects of magnetic domains is presented
that is based on an earlier textbook on magnetic domains [1]. The magneto-optical
Kerr images were partly taken and adapted from the book.
Relevance of Domains and Domain Analysis
The magnetization curve in Fig. 1 is calculated under the assumption of an infinitely
extended specimen, the arrangement of domain phases rather than individual
domains is considered, and the properties of domain walls are completely ignored.
In this phase-theoretical approach, it is further assumed that the domain phase
volumes can freely reach their optimum equilibrium values, thus ignoring coercivity and irreversibility effects. The result is a reversible, vectorial magnetization
394
R. Schäfer
NiFe film
NiFe
film
dCo
/Si/G
Co
Co crysta
l
Fe
lass
cg
talli
film
me
et
he
s
Si
Fe
NiFe
FeS
i sh
Ga
rne
t fi
eet
lm
FeSi
shee
t
Fig. 2 Collection of domain images, obtained by Kerr microscopy on various magnetic materials.
(Most domain images are adapted by permission from Ref. [1] (c) Springer 1998)
curve that may be used as a theoretical reference. In practice, however, those
conditions are rarely met. Applied materials have a polycrystalline-, amorphous-,
or nanocrystalline microstructure; they are of finite size in the shape of particles,
films, ribbons, sheets, or bulk magnets; domain walls may be pinned at defects
and grain boundaries; or they may interact with each other in case of thin films;
mechanical stress can influence the local preference of the magnetization direction;
and the domains and effective magnetic field are strongly influenced by surfaces and
the sample geometry. All these features may have strong effects on the measured
magnetization curve leading to coercivity and irreversibility. Therefore the correct
interpretation of measured hysteresis curves often requires the experimental analysis
of the domains that are responsible for the loop and of their (in general) irreversible
response to magnetic fields, known as magnetization process.
Figure 3 demonstrates such processes for four different magnetic films with
strong perpendicular anisotropy, i.e., they all have an easy axis for the magnetization
that is aligned perpendicular to the film plane. In the course of the experiments, a
strong positive magnetic field was applied along the anisotropy axis, and then the
field was inverted and successively increased in the opposite direction, thus following the upper branch of the shown hysteresis loops. The magnetization process is
initiated by domain nucleation, followed by domain wall motion (upper three rows
8 Magnetic Domains
1
395
-15.5 mT
M/Ms
0
-16.2 mT
-17.5 mT
-17.9 mT
-7.8 mT
-7.9 mT
-8.1 mT
-300 mT
-370 mT
-400 mT
-20 mT
-67 mT
-106 mT
µ0H
in mT
20 µm
-1
-40
-20
0
20
40
1
-7.0 mT
0
100 µm
-1
-40
-20
0
20
40
1
-200 mT
0
-1
10 µm
-1000 -500 0 500 1000
1
-1 mT
0
10 µm
-1
-1000 -500 0 500 1000
Fig. 3 Domain nucleation and growth in magnetic films with strong perpendicular anisotropy,
together with the hysteresis loops along which the domains were imaged. First row: [Co
(0.3 nm)/Pt (0.7 nm)]3 multilayer (sample courtesy D. Makarov, Dresden). Second row: Pt
(3 nm)/Co (1 nm)/Pt (1.5 nm) trilayer (sample courtesy P.M. Shepley and T.A. Moore, Leeds).
Third row: FePt film, 16 nm thick (sample courtesy P. He and S.M. Zhou, Tongji [4]). Fourth
row: FePd(11 nm)/FePt (24 nm) double layer (Sample courtesy L. Ma and S.M. Zhou, Tongji [3])
in the figure) or proceeding nucleation (lower row). For such polycrystalline thin
films, the domain character is more determined by domain wall coercivity effects
due to defects and roughness, randomness of the sample nanostructure, etc. rather
than by an equilibrium of magnetic energies [2, 3]. Typical for such films is a slow
creeping of the nucleated reversed domains into the still saturated area, indicating
thermally activated processes. The domains shown can therefore hardly be predicted
by domain theory – the only way to interpret the hysteresis curves is by observation
of the domains that are responsible for the loops.
This is also true for the M(H )-loop of the amorphous ribbon in Fig. 4. Rather
than revealing a rectangular loop as expected for such material, a predominantly
flat curve is measured inductively with two distinct steps at small field. Domain
observation immediately discloses the reason for such behavior: perpendicularly
magnetized domains across most of the ribbon’s cross section are seen from the side,
whereas longitudinal 180◦ domains are found on the surface. The perpendicular
domains are magnetized by reversible rotational processes in an applied magnetic
396
R. Schäfer
Volume M(H)curve (inductively
measured)
–8
20
–1
µm
M/Ms
Surface M(H)curve
(by MOKE
magnetometry)
M
ag
fie net
ld ic
1
–6
–4
–2
0
2
4
6
H in kA/m
8
Fig. 4 Hysteresis curves and domains of a surface-crystallized, 20-μm-thick amorphous ribbon
of composition Fe84.3 Cu0.7 Si4 B8 P3 . Together with E. Lopatina, IFW Dresden (Adapted and
reprinted from Ref. [5] with permission from Elsevier)
field along the ribbon axis, leading to the flat portions of the loop, while the surfaces
are magnetized by the coercive motion of 180◦ domain walls, which is responsible
for the low-field behavior. In fact, a rectangular surface loop is measured by
magneto-optical Kerr (MOKE) magnetometry with the same coercivities as found
inductively. From the relative induction amplitude at the switching fields of the
surface, a depth of about one micrometer for the longitudinally magnetized surface
regions can be derived. The reason for this behavior is a crystallized surface that
sets the volume under planar compression stress, whereas the surface itself is under
tension. By magnetoelastic interaction, this inhomogeneous stress state leads to a
perpendicular anisotropy in most of the volume and longitudinal anisotropy at the
surfaces.
There are also cases where the interpretation of a hysteresis loop is apparently
simple: the rectangular loop measured on a permalloy film that is in direct exchange
contact with an antiferromagnetic NiO film (Fig. 5) suggests 180◦ wall motion as
dominating magnetization process. In fact this can be proven by domain imaging.
However, the domain walls surprisingly change character between symmetric Néel
and cross-tie wall (see section “Domain Walls”) on the descending and ascending
branches of the loop, respectively. This subtle difference can by no means be derived
from the magnetization loop. These reversible wall alterations indicate the existence
of bi-modal coupling strengths due to pinned and unpinned spins at the interface
between ferro- and antiferromagnet [6].
Surprising domain phenomena may also be found when magnetic samples are
excited by high-frequency magnetic fields. The ground domain state of a nanocrystalline FeNbSiCuB tape wound core with a week circumferential anisotropy, for
example, consists of 180◦ domains that are running along the circumferential
direction (Fig. 6a). Excited quasistatically along the easy axis, i.e., in a slowly
changing magnetic field well below the Hertz regime, 180◦ walls are nucleated,
8 Magnetic Domains
397
M/Ms
1
10 µm
0
Heb
–1
–40
–20
0
20
µ0H in mT
40
Fig. 5 Field-shifted hysteresis loop in a Ni81 Fe7 (30 nm) / NiO (30 nm) exchange-bias double film
together with high-resolution domain wall images along the forward and backward branch of the
loop. A symmetric Néel wall (upper image) and an asymmetrically distorted cross-tie wall (lower
image) are observed. Obtained at IFW Dresden together with J. McCord. (Adapted and reproduced
from Ref. [6] by permission from IOP Publishing)
At 1 kHz
At 0.04 Hz
Ground state
H
M/Ms
1
Ku
M/Ms
1
–10
H in A/m
5
–5
–5
5
H in A/m
10
–1
0.2 mm
a)
b)
Fig. 6 Difference between quasistatic and dynamic magnetization process, demonstrated for
a nanocrystalline Fe73 Cu1 Nb3 Si16 B7 tape wound core with circumferential anisotropy. The
quasistatic hysteresis loop in (a) is governed by the motion of 180◦ walls along the easy axis,
whereas dynamic excitation with a sinusoidal ac field of 1 kHz frequency results in a domain
nucleation-dominated reversal (b). Together with S. Flohrer, IFW Dresden. (Adapted and reprinted
from Ref. [7] with permission from Elsevier)
shifted and annihilated, leading to a square-type hysteresis loop. If excited at high
frequency, however, the domain character surprisingly changes to a patch-type
pattern (b) and the area of the hysteresis loop, and thus the energy loss increases
due to eddy current effects.
398
R. Schäfer
After laser scribing
Without domain refinement
2 mm
2 µm
a)
b)
Fig. 7 Engineering and application of domains: (a) Domain refinement in a grain-oriented
transformer steel sheet by laser scribing for the purpose of energy loss reduction. (b) Currentinduced motion of domain walls in a CoNi multilayer microwire with perpendicular anisotropy, an
experimental test structure for the race-track memory [8]. (Courtesy S.S.P. Parkin, IBM and Halle)
From these examples we see that the correct interpretation of hysteresis curves
may indeed be strongly supported by the analysis of the domains that are responsible
for the loop. But magnetic domains can also be engineered on purpose to obtain
favorable properties in devices, and they can be actively used in devices as
functional units. The best-known example for domain engineering is the intended
domain refinement in transformer steel sheets by scratching or laser scribing
(Fig. 7a). The stress introduced locally in this way interrupts the basic domains,
acting like an artificial grain boundary. This mechanism works for well-textured
material, in which the basic domains would otherwise be wide thus causing large
anomalous eddy current losses when operated at power frequency.
A classical example for the application of magnetic domains is the bubble
memory which was developed back in the 1970s. In this device, cylindrical domains
as carriers of information were shifted in a magnetic garnet film along deposited
ferromagnetic structures on top of the film by applying a rotating magnetic field.
A modern concept of such a domain shift register device is the race track memory
(Fig. 7b), in which perpendicularly magnetized domains are shifted in a magnetic
nanowire by electrical current.
Domain Formation
Magnetic domains are formed to minimize the total free energy [9]. Under ideal
conditions (i.e., ignoring coercivity), a vector field of magnetization directions m(r)
arises in a ferro- or ferrimagnetic specimen so that the total energy reaches an
absolute or relative minimum under the constraint of a constant magnetization, i.e.,
m2 = 1. In this section the energies are reviewed, and it is demonstrated how they
act together to define the domain character.
8 Magnetic Domains
399
Magnetic Energies
The relevant magnetic energies are summarized in an illustrative way in Fig. 8. They
can be classified in local terms, which depend on the local magnetization direction
(anisotropy, applied field, and magnetoelastic coupling energy) and nonlocal terms
that give rise to torques on the magnetization vector, which depend at any point on
the magnetization direction at every other point. The stray field and magnetostrictive
self-energies belong to this class. The exchange energy may be seen as local, but it
depends on the derivatives of the magnetization direction. In the following the most
important aspects of the energy terms are listed.
Given are energy densities Ex ,
which by integration lead to the energies εx = Ex dV with V being the sample
volume.
Exchange energy The alignment of magnetic moments occurs via exchange
coupling. Deviations from a constant magnetization direction therefore invoke the
penalty of exchange energy
Exchange energy
Anisotropy energy
E ex = 0
M
M
M
Easy
axis
H ext
M
Ea > 0
S
N
S
N
N
M
Hd
S
Magnetostrictive
self energy
m
divH d
M
H ext
Magneto-elastic
coupling energy
[100] [100]
E ms = 0
N
100
>0
0]
[10
N
H ext = 0
Ea = 0
E ex > 0
Stray field energy
External field energy
0]
[01
0]
[10
Tensile
stress
E ms > 0
Fig. 8 Summary of magnetic energies that are relevant for the formation and character of magnetic
domains
400
R. Schäfer
Eex = A(grad m)2 ,
(1)
where A is the exchange stiffness constant, a temperature-dependent material
constant in units J/m.
Anisotropy energy Most magnetic materials are anisotropic, i.e., there are easy
axes along which the magnetization vector is preferably aligned and along which
the saturated state can be “more easily” obtained than along other directions. These
can be preferred crystal axes (magnetocrystalline anisotropy) or an axis that arises
from some short-range ordering of atoms like Ni-Ni and Fe-Fe atomic pairs in NiFe
alloys. The driving force for this induced anisotropy is the magnetization of the
material that is present at an annealing temperature below the Curie temperature
and which can intentionally be aligned by an applied external magnetic field.
Also annealing under mechanical stress may result in a uniaxial anisotropy, called
creep-induced anisotropy. In amorphous and nanocrystalline ribbons, this type of
anisotropy may be dominating, and often an easy plane of magnetization transverse
to the stress axis is created by stress annealing. Shape effects are part of the stray
field energy, and they do not belong to the anisotropy terms.
Deflecting the magnetization out of an easy axis requires additional energy, called
anisotropy energy. In the most simple case of a uniaxial anisotropy (as it occurs
in crystal lattices with hexagonal or tetragonal symmetry or in case of an induced
anisotropy), the energy density is written as
Ea = Ku1 sin2 ϑ + Ku2 sin4 ϑ ,
(2)
where ϑ is the angle between anisotropy axis and magnetization direction and Ku1
and Ku2 are the anisotropy constants of first and second order – higher orders can
usually be neglected. An easy axis is described by a large positive Ku1 , whereas
‘planar’ and ‘conical’ anisotropies are found for large negative Ku1 and intermediate
Ku2 values. The anisotropy constant Ku1 corresponds to the energy needed to
saturate the sample in the so-called “hard” direction (ϑ = 90◦ ). In multiaxial
materials such as iron, all 100 directions are easy, whereas in nickel the 111
axes are the preferred crystal axes.
External field energy, also called Zeeman energy, is added to a magnet if an
external magnetic field H ext is applied. It is given by
EZ = −μ0 H ext · M = −μ0 Hext · M · cos(ϕ) ,
(3)
where ϕ is the angle between magnetization and field. This interaction energy of
external field and magnetization vector field m(r) causes domain wall motion and
rotational processes and finally leads to saturation along the field direction if the
field is strong enough. The minimum of the Zeeman energy is achieved when the
magnetization is aligned to the magnetic field (ϕ = 0).
8 Magnetic Domains
401
Stray field energy Sinks and sources of the magnetization vector field (div m)
lead to magnetic poles, which act as sources and sinks for a magnetic stray field.
Magnetic poles can be present as volume or as surface poles if the magnetization
vector M is not parallel to the surface. The stray field H d , arising from the poles, is
illustrated in Fig. 8 for a finite ellipsoidal magnet that is homogeneously magnetized
to the right. This leads to north (N) and south (S) poles at the edges and a stray field
from (N) to (S). Within the magnet the stray field is called demagnetizing field as it
opposes the magnetization. The presence of poles and stray fields causes the stray
field energy
1
Ed = − μ0 H d · M with H d = −N · M .
2
(4)
The demagnetizing factor N (a tensor in general) is zero for infinitely extended
bodies and becomes the larger the closer the specimen edges along M. The stray
field energy thus scales with N and with the average magnetization. A particularly
unfavorable case is an infinitely extended plate that is magnetically saturated
perpendicular to its surface. The demagnetizing factor along M is 1 then, and the
demagnetizing or stray field energy is written
Ed =
1
μ0 Ms2 = Kd .
2
(5)
The stray field energy coefficient Kd is a measure for the maximum energy densities
which may be connected with stray fields. Independent of the complexity of real
stray fields, their energy always scales with the material parameter Kd . As the
demagnetizing field of a body along a short axis is stronger than along a long axis,
the applied magnetic field along the short axis has to be stronger to produce the same
field inside the specimen. The shape of the magnet is thus the source of magnetic
anisotropy (shape anisotropy). For an infinitely extended body that is magnetized
along an infinite direction, the demagnetizing factor N is zero, and there will be no
stray field energy at all.
Magnetostrictive self-energy In magnetostrictive material, the crystal lattice is
spontaneously elongated or contracted along the magnetization direction if the magnetostriction constant λ is positive or negative, respectively. As magnetostriction
is quadratic in the magnetization vector, this lattice distortion is not important for
180◦ domains as all domains will lead to the same deformation. However, for
the 90◦ domain configuration shown in Fig. 8, it will cause elastic energy (called
magnetostrictive self-energy) as the spontaneous deformations of various parts of
the domain pattern, indicated by ellipses in the figure, do not fit together elastically.
The energy density of this incompatible domain configuration is
9
Ems = − Cλ2100 ,
8
(6)
402
R. Schäfer
λs > 0
λs = +35·10–6
λs = +24·10–6
λs = 0
100 µm
λs = +8·10
a)
λs < 0.2·10–6
–6
b)
Fig. 9 Illustration of magnetostrictive and magnetoelastic energies: (a) Closure domains,
observed on the surfaces of two amorphous ribbons with positive and zero magnetostriction as
indicated. A backward pointing magnetic field along the ribbon axis was applied. (b) Typical
domains in the as-quenched state of amorphous ribbons with different magnetostriction constants.
Frozen-in internal mechanical stress, which is present in all four materials after quenching, leads to
different complexity in domains depending on the magnetostriction constant. The ribbon thickness
is about 20 μm in each case
where λ100 is the magnetostriction constant for the 100 directions and C is the
relevant shear modulus for the given configuration. If not enforced for topological
reasons or by magnetic fields, nature tries to avoid such incompatible domain
arrangements. Figure 9a demonstrates this effect for amorphous ribbons with some
induced anisotropy perpendicular to the ribbon surface. A moderate magnetic field
along the ribbon axes was applied that enforces a certain longitudinal magnetization.
If magnetostriction is positive, domains running transverse to the field direction are
more favorable because here the sample elongation in the neighboring domains fits
together elastically. For a material without magnetostriction, a longitudinal domain
arrangement will cause no problem as the elastic distortion, indicated in Fig. 8, will
not occur. In fact this arrangement is even more favorable because here the specific
wall energy of the basic domain walls is lower than for the transverse case.
Magnetoelastic interaction energy There is also an inverse effect: applied
mechanical stress of nonmagnetic origin can act on the magnetization direction
in materials with non-zero magnetostriction by adding magnetoelastic interaction
energy. The stress can be an external stress or some nonmagnetic internal stress
resulting from dislocations or inhomogeneities in composition, structure, and
temperature. In the example shown in Fig. 8, a horizontal tensile stress favors
the horizontal anisotropy axes in a material with otherwise dominating positive
cubic crystal anisotropy. If magnetocrystalline anisotropy is low or absent like in
8 Magnetic Domains
403
amorphous materials, the magnetoelastic coupling may lead to dominating stressinduced anisotropies (Fig. 9b). This can be seen by writing the magnetoelastic
coupling energy as
Eme =
3
λs σ sin2 ϑ
2
(7)
where σ is the uniaxial mechanical stress, λs the isotropic magnetostriction constant,
and ϑ the angle between magnetization vector and stress axis. This energy term
describes a uniaxial anisotropy along the stress axis with an anisotropy constant of
Ku = 32 λs σ , compare Eq. (2). Although magnetostriction is a relatively weak effect
with induced strains of typically 10−5 only, the examples in Fig. 9 demonstrate that
its effect on domain patterns can be significant.
Domain wall energy The specific energy of domain walls is not an independent
term, but rather consists of exchange and anisotropy energy, owing to the deviation
of magnetization from the anisotropy axes and non-parallel magnetic moments
across the wall. In the most simple case of a Bloch√wall in an infinitely extended
uniaxial material the specific wall energy is γw0 = 4 A/Ku . See section “Domain
Walls” for more information on domain walls.
For summary, the characteristic coefficients of the magnetic energy terms and
their order of magnitude in typical magnetic materials are collected in Fig. 10.
Magnetic Energy
Energy Coefficient
Range
Exchange energy
Exchange stiffness constant A
Material constant
10
Anisotropy energy
Anisotropy constant K
Material constant
K1: Crystal anisotropy
Ku: Induced (uniaxial) anisotropy
±(102
External field energy
µ0HextMs
Hext = external field
Ms = saturation magnetization
Depends on field
magnitude Hext,
unit J/m3
Stray field energy
Kd = 1/2 µ0Ms
Magnetoelastic
interaction energy
J/m
7
2
λ
= mechanical stress
λ = magnetostriction constant
6
) J/m3
J/m3
Depends on stress
magnitude ext,
unit J/m3
2
Magnetostrictive self
energy
Cλ
C = shear modulus
λ = magnetostriction constant
Fig. 10 Summary of energy coefficients and their range of magnitude
3
J/m3
404
R. Schäfer
Driving Forces for Domain Formation
Primarily it is the stray field energy that is responsible for the development of
magnetic domains: domains are formed to reduce or avoid stray field energy. This
fact is illustrated in Fig. 11 by comparing an infinitely extended with a finite sample.
The NiFe (permalloy) film was sputter-deposited in the presence of a magnetic
field, so it has a weak uniaxial anisotropy along the vertical direction in the images.
For Fig. 11a, a piece of the film was broken from the wafer which extends 30 mm
along the easy axis. In view of the thickness of just 240 nm, the specimen may be
considered as infinitely extended so that stray field energy does not play a role. In
fact, domains are not visible in that case: when a magnetic field along the easy axis is
applied, the film switches from magnetization up to down and vice versa, leading to
a magnetization curve with two steep, discontinuous steps at the coercivity field. The
switching occurs by the fast and abrupt motion of a 180◦ domain wall as indicated
schematically in the figure. The multidomain states at the discontinuity fields cannot
be captured in the experiment because the expense of domain wall energy makes
them statically unfavorable.
The situation changes if an open sample, prone to a demagnetizing field, is
considered. For Fig. 11b the infinite permalloy film was replaced by a 100×100 μm2
film element. For this finite specimen, the in-plane demagnetizing factor N has
raised to 0.0015, and a demagnetizing energy of N Kd m2 is added. A saturated state
at remanence, like in Fig. 11a, would thus be highly unfavorable as it would cost
30 x 5 mm2
M/Ms
1
1
Finite sample
(
= 0)
Hext = Hint +
µ0Hext in mT
–5
a)
5
–1
M
µ0Hext in mT
–5
5
–1
b)
Fig. 11 Comparison between infinitely extended and finite samples. Shown are the magnetooptically measured magnetization curves along the (vertical) induced anisotropy axis in a Ni80 Fe20
Permalloy film of 240 nm thickness, which is infinitely extend in (a) and of finite size in (b). The
domain images in (b) show the full patterned element, whereas in (a, upper inset) only a part of
the extended film is shown. The lower inset in (a) is a schematics
8 Magnetic Domains
405
maximum stray field energy (| m |= 1). At zero field, the film element is rather
in a demagnetized, multidomain state that reveals a flux-closed, pole-free nature.
In the magnetization curve, this is expressed by the shearing transformation: for
a given magnetization value m, the external field Hext must be enlarged by the
demagnetizing field −Hd = N Ms m to reach the same magnetization state. Thus
the discontinuous magnetization curve is transformed into a finite-slope curve with
a well-defined magnetization value for every field value. For the finite element, the
domains can thus be followed along the sheared M(H )-loop.
So for the existence of magnetic domains, a finite sample size along the
magnetization direction (easy axis) is required as the stray field energy is the driving
force for domain observation. Infinitely extended films will consequently also have
domains if the easy axis is perpendicular to the film plane, compare Fig. 3.
There are two further cases in which magnetic domains or related objects can
exist even in the absence of demagnetizing effects:
• Consider a sample that is embedded in an ideal soft magnetic yoke to obtain fluxclosure. With a coil, wrapped around the yoke, a certain average magnetization
can be enforced in the yoke by some feedback mechanism. If a magnetization
value is then enforced in the sample that lies within the range of a discontinuous
jump in the magnetization curve, a multidomain state will be enforced in which
the two states at the endpoints of the jump will be mixed in a certain volume
ratio so that the enforced magnetization is achieved. In case of the previously
discussed uniaxial material, this would be a 180◦ domain state in which the 180◦
domain walls move as the enforced magnetization is varied. Such circuits are
realized in machines with an inductive load on a rigid voltage, such as an idling
transformer.
• The second case may be found in magnetic materials with broken inversion
symmetry in the atomic lattice in which the crystallographic handedness induces
a quantum-mechanical Dzyaloshinskii-Moriya interaction (DMI) [10] by spinorbit scattering. Unlike direct Heisenberg or superexchange, which favor parallel
or antiparallel alignment of neighboring magnetic moments according to a
Hamiltonian that is proportional to S i · S j , the DMI is proportional to S i × S j
thus favoring perpendicularly aligned neighboring spins S. In competition with
collinear coupling, the DMI can lead to nanoscale, homochiral magnetization
modulations like long-period helical spin-spiral phases. Most prominent is the
topologically stable skyrmion spin structure that was predicted theoretically by
A. Bogdanov [11, 12] and which has been directly observed in nanolayers of
cubic helimagnets with intrinsic DMI [13] and in Fe/Ir bilayers [14] with surface/interface-induced chiral interactions [15]. Magnetic skyrmions are axisymmetric
vortex patterns with a homochiral rotation of spins that can exist as isolated
entities in the saturated states of chiral magnets [14,16] or in form of skyrmionic
condensates (two-dimensional lattices and other mesophases) [12, 17].
In modern literature, intrinsically caused magnetic modulations (e.g., chiral
helicoids and skyrmions) are often classified as magnetic micro- or spintextures,
406
R. Schäfer
whereas the modulated elements of multidomain states (domain walls, Bloch lines,
Bloch points, magnetic swirls etc. – compare Fig. 18) are commonly addressed as
being part of magnetic microstructure [1]. This different classification, however, is
questionable: spatially inhomogeneous spin structures arising in magnetic nanolayers are formed under the mutual influence of intrinsic and dipolar forces [18]. This
levels out the difference between the terms magnetic microstructure and magnetic
micro- or spintextures [19].
Interplay of Energies
Once the precondition for domain formation is fulfilled, the domain character is
finally determined by an interplay of the magnetic energies. For demonstration of
this principle, let us have a look at the prominent example of domain formation in
grain-oriented Fe3wt%Si steel that is used as core material in transformers. Like
for pure iron, the easy directions of magnetization are the 100 directions for this
material. Transformer sheets are typically 0.3 mm thick, consist of wide grains in the
centimeter regime, and are Goss-textured. In this [001](110) texture the [001] easy
direction is oriented, within a few degrees deviation, along the rolling axis during the
manufacturing process of the sheets. The grain surfaces are (110) oriented within the
same accuracy, and the other two easy axes are oriented at angles of ±45◦ relative
to the surface. As shown in Fig. 12, the domain structure of such sheets consists
of simple slab domains that are separated by 180◦ walls in case of ideally oriented
grains. Their existence may, e.g., be enforced by some demagnetization effects at
the grain boundaries. For increasing out-of-plane misorientation of the [001] easy
direction, fine lancet-shaped domains of increasing density are superimposed on the
basic domains.
The formation of those so-called supplementary domains is a consequence of
energy optimization. Let us firstly assume an infinitely extended grain with ideal
(110) orientation. It will be homogeneously magnetized along the surface-parallel
easy axis (Fig. 13a), thus completely avoiding magnetic poles. As the grain is
infinitely extended, domains are not to be expected. A different situation arises if
the [001]-axis is misoriented by some degrees relative to the surface (b). Assuming
that the magnetization strictly follows the [001] axis, magnetic surface poles will
arise. The associated stray field energy can be reduced by forming ±180◦ basis
domains (c) which leads to the presence of opposite poles on the same surface, thus
allowing the field lines of the stray field to run along the surface. A further reduction
of stray field energy could be achieved by reducing the basic domain spacing as
this would bring the opposite poles in closer distance. However, the narrower the
domain width, the higher the expense of domain wall energy associated with the
rising wall area of the basic domain walls that extend all through the thickness.
Nature finds a more economic way to keep the overall energy low by adding
supplementary domains to the basic domains (d). The shallow lancet domains at
the surface collect the net flux that is transported toward the surface in the basic
domain. The lancets are oppositely magnetized to the basic domains, thus leading to
8 Magnetic Domains
407
(110) surface
Goss texture
100 easy axes
Out-of-plane
misorientation
0.1 mm
2° miso
riented
4° miso
riented
idealy
oriente
d
8° mis
30 mm
oriente
d
Fig. 12 Domains on a Goss-textured transformer sheet. Shown are the domains of four grains with
increasing out-of-plane misorientation as indicated. The ceramic insulation coating, by which such
sheets are usually covered to avoid eddy currents between the sheets, was removed for domain
imaging by Kerr microscopy. (The domain images are adapted by permission from Ref. [1] (c)
Springer 1998)
a narrow spacing of opposite surface poles as required for stray field reduction. The
flux is then transported to a surface of opposite polarity and distributed again. This
is achieved by internal domains that are magnetized along the internal, transverse
easy axes. Those transverse domains can extend all through the volume, or they
can be connected to a basic domain wall so that the neighboring basic domain
is used to lead flux downwards. Because this system of compensating domains is
superimposed on the basic domains that would be present without misorientation,
these domains are called supplementary domains.
If a (moderate) magnetic field is applied along the surface-parallel easy axis,
Zeeman energy is added and those basic domains with magnetization along the field
direction will grow on expense of the opposite basic domains by 180◦ wall motion.
The rise in stray field energy, caused by the absence of oppositely magnetized
basic domains, is then compensated by an increasing number of supplementary
domains. At the same time those internal transverse domains, which are connected
to the basic domain walls, have to extend across the whole sheet thickness. So the
transverse domain volume is larger compared to the demagnetized state. This change
in relative domain volumes has consequences for the stress state of the sheet: as the
magnetostriction constant is positive for FeSi, the cubic crystal lattice is tetragonally
distorted along the magnetization direction. The basic domains thus cause an
408
R. Schäfer
N
Easy
axes
S
a)
N
N
N
N
S
S
N
S
N
S
S
N
S
N
N
d)
N
N
N
S
S
S
S
N
S
S
N
N
S
S
S
N
c)
N
N
N
S
S
N
N
N
N
b)
S
S
N
N
e)
N
Tensile stress
µ*-corrected
N
N
N
N
Without tensile stress
g)
With tensile stress
f)
Fig. 13 Interplay of magnetic energies, illustrated on the example of domain formation in
FeSi transformer sheets with (110)-related surfaces. Shown is the introduction of basic and
supplementary domains (b–d) in case of a slightly misoriented surface, starting from an ideally
oriented surface in (a). Tensile stress leads to domain refinement (e, f). In (g) the μ*-effect is
illustrated. (Image (d) is adapted by permission from Ref. [1] (c) Springer 1998)
elongation of the sheet along the rolling direction, while in the transverse domains
the sheet is transversely expanded. A change in the transverse domain volume will
thus result in a magnetostrictive change in length during remagnetization along the
[001] easy axis. Driven in a magnetic field at power frequency, the sheet will be
set in mechanical vibration leading to acoustic transformer noise. Furthermore, the
repeated destroying and rebuilding of supplementary domains forms an important
part of hysteresis loss as the energy bound in the supplementary domains is lost in
every cycle.
Magnetostrictive interaction can, however, also be favorably used in transformer
sheets. The supplementary domains are suppressed under tensile stress applied
along the preferred axis, because tensile stress magnetostrictively disfavors the
transverse domains that are attached to the supplementary domains. A domain state
as in Fig. 13c would thus result. Rather than superimposing supplementary domains
8 Magnetic Domains
409
to lower the stray field energy, which is forbidden now, a similar effect is achieved
by lowering the basic domain width (e). The domain images in (f) demonstrate this
effect. Obviously even ideally oriented grains assume a small domain width if they
are coupled to less well-oriented grains to achieve flux continuity. A narrow domain
with is favorable if the domains are excited by AC magnetic fields. The larger the
density of the walls, the smaller the velocity of every wall for a given induction level
which lowers domain wall-related eddy current effects (so-called anomalous eddy
current losses). In practice the tensile stress is created by the insulation coating that
is at the same time stress-effective. The planar stress exerted by the coating is for the
Goss texture equivalent to a uniaxial stress and will thus suppress the supplementary
domains.
Two further, energy-related aspects are worth to be noted: (i) so far it was
assumed that the domains are strictly magnetized along the easy crystal axes in the
demagnetized state and up to moderate applied magnetic fields. This is in fact true
for most of the volume domains. By approaching the (110) surface, however, the
magnetization bends toward the surface (Fig. 13g). So the surface poles are spread
over a certain volume and not just at the surface which helps to reduce the stray
field energy at the expense of some anisotropy energy, though. The phenomenon
is known as μ*-effect. (ii) The basic ±180◦ walls are zigzag folded across the
thickness as indicated in Fig. 13. Although the total wall area is larger than in case
of straight, perpendicular (110) walls that would have the smallest area, the total
wall energy is reduced by the folding. The reason is the specific wall energy, which
is lower for {100} wall orientations. The (110) wall therefore tends to rotate toward
these orientations, forming tilted or zigzag walls with a lower overall energy.
Domain Classification
The magnetic energy coefficients, listed in Fig. 10, can be combined in several ways
to obtain dimensionless parameters that reflect the interplay of energies and thus the
domain character. The ratio between anisotropy and stray field energy is the most
important. This ratio is called the quality factor, defined by
Q=
Keff
.
Kd
(8)
Here Keff is the effective anisotropy constant and Kd the stray field energy
coefficient defined in Eq. (5). If the anisotropy energy dominates over the stray field
energy (Q > 1), domains are formed that avoid an expense of anisotropy energy
while keeping the stray field energy as low as possible. If the stray field energy is
dominant (Q 1), stray fields are avoided by flux-closed domain patterns that
adapt to keep the anisotropy energy as low as possible. In the following discussion
we use the quality factor as primary criterion as it leads to the most fundamental
way of classifying domains and magnetic materials. In Fig. 14 a number of typical
materials are listed in the order of decreasing quality factor. Further criteria are the
410
R. Schäfer
Material
µ0 Ms
in Tesla
SmCo5
1.05
CoPt (L10)
0
w
0
w
K1 , Ku
in J/m3
Q
12
1.7 107
hexagonal
39
0.84
57
1.0
10
4.9 106
tetragonal
12
1.5
28
Sm2Co17
1.29
14
4.2 106
rhombohed.
6.3
1.83
31
Nd2Fe14B
1.61
7
4.5 106
tetragonal
4.4
1.25
23
BaFe12O19
0.48
7
3.2 105
hexagonal
3.5
4.68
6
Cobalt (Co)
1.79
31
4.5 105
hexagonal
0.35
= –45
= –260
8.3
15
= +22
–21
20.9
4
–55
–23
42
0.8
1
300
A
J/m
in 10
10
11
44
in 10
m in mJ/m2
Iron (Fe)
2.15
21
4.8 104
bcc
0.03
Nickel (Ni)
0.60
8
–4.5 103
fcc
0.03
Permalloy film
(Ni81Fe19wt%)
1.00
13
50 - 200
Ku induced
3 10
Fe74Cu1Nb3Si15B7
nanocryst. ribbon
1.24
6
~20
Ku induced
4 10
s
0.2
550
0.04
Amorph. ribbon,
Co-based
0.6
2.5
~3
Ku induced
2 10
s
0.1
900
0.01
100
111 =
100 =
111 =
s
Fig. 14 Material parameters that are important for domain analysis. The listed materials are
ordered in terms of decreasing quality factor Q. Listed are furthermore saturation polarization
μ0 Ms , exchange stiffness constant A, first order anisotropy constant K1,u , magnetostriction
constant λ, wall width parameter Δ0w [see Eq. (14)], and specific wall energy of a 180◦ Bloch
wall γw0 [see Eq. (13)]. (Data are taken from Refs. [20, 21])
manifold of easy directions and the surface orientation of the investigated specimen,
which we treat as secondary criteria to classify the wide variability of domain
phenomena.
In Fig. 15 the interplay of stray field and anisotropy energy is illustrated by
comparing three material classes with uniaxial anisotropy but highly different
Q-factors. In all cases the easy axis is perpendicular to the plate surface, on
which domain observation was performed by Kerr microscopy, i.e., the specimens
are extremely misoriented with respect to the imaged surface. Compared are the
domains of a NdFeB single crystal (left column) with those in amorphous films and
ribbons (right column). The strong magnetocrystalline anisotropy of 4.5 · 106 J/m3
8 Magnetic Domains
411
D
Ku
Dominating
anisotropy energy (Q >1)
D =
5 µm
Film:
W
1
7 µm
N
S
N
D =
1 µm
N
S
N
S
Q=
0.01
e)
Towards bulk:
N
7
N
S
S
N
N
Surface
5 µm
S
a)
14 µm
Dominating
stray-field energy (Q<<1)
D =
25 µm
Q=
0.0003
b)
f)
D =
25 µm
40 µm
Q=
0.001
c)
g)
D =
25 µm
120 µm
Q=
0.002
20 µm
h)
d)
(Q
polar magnetization
1)
in-plane magnetization
D =
1 mm
5 µm
Fig. 15 Classification of magnetic domains for three extrem cases of the quality factor. (a–d)
NdFeB single crystal, Q = 4.4. (e–h) FeBSi-based amorphous film and ribbons with stressinduced perpendicular anisotropy and Q-values as indicated. (i) Cobalt single crystal, Q = 0.35.
In each case the domain images are taken on top surface, while the sketches show side views. The
Kerr images in (a–h) are adapted by permission from Ref. [1] (c) Springer 1998, while (i) was
obtained together with I. Soldatov, Dresden, and reproduced from Ref. [22] with the permission of
AIP Publishing
412
R. Schäfer
makes NdFeB a material with Q = 4.4, while Q 1 for amorphous material
owed to some weak (stress-)induced anisotropy around the order of 10–100 J/m3
(compare Fig. 14). An intermediate anisotropy of 4.5 · 105 J/m3 results in a quality
factor of 0.35 for hexagonal cobalt, shown on the bottom of the figure.
In case of the high-anisotropy material, the domains are strictly magnetized along
the easy axis to avoid anisotropy energy, even though this causes magnetic poles at
the surface thus costing stray field energy. This rule is strictly followed for films and
bulk specimens, though the domain character changes as function of thickness. For
films (Fig. 15a) a simple plate domain state with up-and-down domains is observed.
For rising
√ sample thickness D, the domain width W increases according to
W ∼ D. Beyond a critical thickness (about 5 μm in the example), the domain
walls get corrugated close to the surface (b), which leads to a better intermixing
of poles. This lowers the stray field energy compared to the hypothetical case of
straight domain walls. For further rising thickness (c), domain branching sets in:
close to the surface a fine domain pattern is enforced to minimize the stray field
energy by bringing opposite poles close together, whereas in the bulk, wide domains
are favored to save wall energy. With increasing thickness (d) further iterated
generations of domains are added, leading to a progressive domain refinement
toward the surface in a fractal way. Theory [1] yields a characteristic D 2/3
dependence of the basic domain width and a constant surface domain width within
the branching regime. As only up-and-down domains are involved in high Q
uniaxial material, the branching scheme is called two-phase branching.
Different arguments apply to the low-anisotropy material presented in the right
column of Fig. 15. In the limit of small thickness (not shown), a thin film with a weak
perpendicular anisotropy would be in-plane magnetized, because the anisotropy
energy density of this state would be smaller than the stray field energy density of a
uniformly perpendicularly magnetized state. Beyond a critical thickness, however,
the magnetization starts to oscillate out of the plane in a periodic manner to save
part of the anisotropy energy (Fig. 15e). The oscillation modulation assumes the
character of a two-dimensional flux-closed pattern, called dense stripe domains,
the half period of which typically equals the film thickness. Due to flux closure,
stray field energy is completely avoided as required for a Q 1 material,
whereas anisotropy energy and exchange energy are spent due to the deviations
from the easy axis and the non-parallel alignment of magnetization, respectively. For
larger thickness, an oscillating magnetization would consume increasing anisotropy
energy so that nature prefers a Landau pattern (f). Here the bulk is strictly
magnetized along the easy axis, and the expenses of exchange and anisotropy energy
are concentrated in regular domain walls and closure domains (being magnetized
perpendicular to the easy axis), respectively. When the perpendicular anisotropy
increases slightly, still within the Q 1 regime, the anisotropy energy in the
closure domains would rise so that beyond a critical anisotropy level (or beyond a
critical thickness
– note that the closure domain volume increases with thickness
√
due to a D-increase of the basic domain width) a three-dimensional branching
scheme takes over (g): the closure domains become themselves modulated in a
similar continuous manner as seen for films in (e), thus lowering their anisotropy
energy density. These stripe oscillations are connected with the basic domains by
8 Magnetic Domains
413
assuming the corrugated shape visible in the photograph and model. With a further
increase in anisotropy or sample thickness, the stripe domains do grow into regular
domains, the closure domains of which now decay into a further generation of
stripe pattern as shown in (h). This type of branching may be considered as multiphase branching which occurs despite the fact of having a uniaxial material rather
than cubic or other multiaxial materials. In any case, the overall domain patterns
(e - h) are completely free from stray field energy as required for such low-Q
material.
If anisotropy and stray field energy are competing with about equal magnitude
(Q ≈ 1), a hybrid structure is formed as shown in Fig. 15i for a thick cobalt
crystal. The character of the domain pattern is similar to that of NdFeB in the
macroscopic aspects, i.e., the branching mode in the bulk agrees with that of the
high-anisotropy material. The surface domain width, however, is smaller because of
the smaller wall energy of cobalt and theory predicts closure domains with a tilted
magnetization, owing to a lower anisotropy compared to NdFeB. By comparing
the two domain photographs in Fig. 15i, which show the out-of-plane and in-plane
magnetization components separately, it seems that the fine surface pattern of the
in-plane component resembles the branched domains of the amorphous ribbon in
image (h). Obviously the closure domains are modulated in a dense stripe domain
pattern. According to these findings, cobalt with its intermediate anisotropy forms
a kind of hybrid, following the high-anisotropy two-phase branching scheme in the
bulk and a low-anisotropy multiaxial branching scheme at the surface.
The classification principle discussed so far is generally valid – different crystal
symmetries or sample orientations just add modifications. Consider the case of
Fe3wt%Si material in which the three 100 axes are easy. The quality factor of 0.03
implies flux-closure domain configurations at zero field. Their character depends
on the surface orientation, and by having two further easy axis compared to the
uniaxial materials in Fig. 15, more degrees of freedom for the domain formation
are available. This becomes immediately apparent by looking at the domains of
the non-oriented sheet presented in Fig. 16a. The variety of flux-closed domain
patterns can be sub-classified according to the surface orientation: on an ideally
oriented surface, the principle of flux closure is immediately seen by wide domains
that are separated by well-oriented domain walls. Two grains with (110) and (100)
orientations are marked in the figure. A pole-free wall orientation requires that the
component of the magnetization perpendicular to the wall is the same on both sides
of the wall as indicated in Fig. 16b. For slight misorientation of a few degrees
flux collection is achieved by supplementary domains as seen for the (100)-related
surface in Fig. 16c (see Fig. 13 for a thorough explanation of this phenomenon for
a (110)-related surface). For stronger misorientation a domain branching scheme is
energetically preferred. In Fig. 16d this is illustrated for a (111) surface, i.e., the case
of extreme misorientation. Here in most of the volume a domain structure is formed
that occupies easy directions only, and these domains are joined so that no magnetic
stray fields are generated as required by Q 1. Near the surface zones, however,
the two requirements of using only easy directions and avoiding stray fields are
incompatible as the surface does not contain an easy direction. By branching nature
finds a compromise: the domains get finer toward the surface by adding several
414
R. Schäfer
~(111)
~(110)
(100)
10 µm
1)(11
(110)
100 µm
a)
mn
m2
m1
ace
f
sur
~(100)
mn
e.a.
(100)-cut
d)
Pole-free
wall orientation
requires:
(m1 m2) n = 0
n
b)
a)
c)
Fig. 16 The domains on a non-oriented Fe3wt%Si sheet (0.5 mm thick) reveal the influence
of the surface orientation on the domain character. (a) Low-resolution Kerr image giving an
overview. (b) Schematics of a 90◦ wall illustrating the condition for a pole-free wall orientation. (c)
Sketch of the fir tree structure, a supplementary pattern that appears on slightly misoriented (100)
surfaces of iron-like material. (d) High-resolution Kerr image on a (111) surface, together with
a schematics showing the phenomenon of multiaxial branching in iron-like material with cubic
crystal anisotropy. (The images in (a) are adapted by permission from Ref. [23] (c) Springer 2009)
generations of echelon domains in a fractal way (the number of generations depends
on the thickness). By the wide volume domains, wall energy is saved, and by the fine
surface domains, the volume of the outermost closure domains, which cannot be
magnetized along an easy axis, is minimized. Those closure domains are actually
magnetized in a continuously varying way similar to the dense stripe domains in
films (see Fig. 15e). So the unavoidable anisotropy energy of the surface zone is
reduced and right at the surface the magnetization lies parallel to the surface as
required by the low-quality factor.
Domain Walls
Domain walls form a continuous transition between neighboring domains. The
domain wall structure and character primarily depends on the Q-factor and thickness of the specimen. Furthermore the wall angle (i.e., the relative angle of the
8 Magnetic Domains
415
Left domain
(x)
Bloch path
1
x
S
Néel
path
Right domain
a)
x
sin (x)
0
–4
–2
0
2
4
x / A/Ku
N
b)
W180
Fig. 17 (a) Bloch and Néel wall paths in an infinite uniaxial material. The magnetic poles for the
Néel wall are indicated in the vector plot. (b) Wall profiles of a 180◦ Bloch wall. The indicated
wall width, W180 , is defined on basis of the slope of the magnetization angle ϕ(x). (Adapted by
permission from Ref. [1] (c) Springer 1998)
neighboring domain magnetizations) and magnetostriction may have an influence.
Here some basic aspects of domain walls are reviewed; for details and a thorough
review, we refer to Ref. [1].
Domain Wall Types
Two principle modes of magnetization rotation across a domain wall can be
distinguished: the Bloch and Néel wall. In Fig. 17a the two paths are illustrated
for the simplest of all domain walls, a planar 180◦ wall in a (hypothetic) infinitely
extended medium with uniaxial anisotropy that separates two domains of opposite
magnetization. If the domain magnetizations are parallel to the wall, there will be no
global magnetic poles, meaning that the component of magnetization perpendicular
to the wall is the same on both sides of the wall (compare Fig. 16b). In the Bloch
wall the magnetization rotates parallel to the wall plane, so there are no poles inside
the wall either (divm = dmx /dx = 0), and the stray field energy is zero. For the
Néel path the stray field energy would be maximum, making this wall path inferior
to the Bloch path in bulk material. Néel-type walls can nevertheless be favorable in
magnetic thin films as shown below.
Classical Bloch wall If magnetostriction and higher-order anisotropy constants are
neglected, the specific wall energy γw0 of a Bloch wall is written as an integral over
exchange energy (considering the non-parallel moments in the wall) and anisotropy
energy (due to deviations from the easy axis):
γw0 =
∞
−∞
[A(dϕ/dx)2 + Ku cos2 ϕ] dx , ϕ(−∞) =
π
π
, ϕ(∞) = − .
2
2
(9)
Here x is the coordinate perpendicular to the wall, the angle ϕ rotates in the
wall from 90◦ to −90◦ , and ϕ(±∞) are the boundary conditions given by the
416
R. Schäfer
neighboring domains (see Fig. 17a). The solution of this ansatz is obtained by
variational calculus, which leads to a function ϕ(x) that minimizes γw0 under the
boundary conditions. Starting with Euler’s equation
2A(dϕ/dx) = −2Ku sin ϕ cos ϕ ,
(10)
multiplying it with (dϕ/dx) and integrating with respect to x leads to the first
integral:
A(dϕ/dx) 2 = Ku cos2 ϕ .
(11)
According to this equation, the exchange and anisotropy energy densities are equal
at every point in the wall: at positions where the anisotropy energy is high, the
magnetization rotates rapidly leading to a large exchange energy. From Eq. (11) we
obtain
dx =
A/Ku dϕ/ cos ϕ ,
(12)
which, inserted together with (11) into the total wall energy (9), yields
γw0 = 2
∞
−∞
Ku cos2 ϕ dx = 2 AKu
π/2
−π/2
cos ϕ dϕ = 4 AKu .
(13)
√
Integration of (12) leads to the functional dependence sin ϕ = tanh(x/ A/Ku ),
which is plotted in Fig. 17b. From the indicated definition, the classical Bloch wall
width is derived as
W180 = π Δ0w , with Δ0w =
A/K
(14)
being the wall width parameter (see Fig. 14 for examples).
Although calculated for infinite samples, the classical Bloch wall also occurs in
“real” specimens if the quality factor of the material is larger than one. Then the
magnetic surface poles, which are caused by the out-of-plane magnetic moments
of the wall in case of in-plane magnetized domains, can be tolerated. For low
anisotropy material (Q 1), however, the requirement of pole avoidance enforces
different wall types. In Fig. 18 they are collected for materials with a (weak)
uniaxial, in-plane anisotropy Ku . There is a difference between thin films, thick
films, and bulk specimens.
Walls in thin films with Q < 1 From the micromagnetic point of view, magnetic
films are defined as “thin” if their thickness is below the classical Bloch wall width.
Then wall modes using a predominantly in-plane rotation of magnetization, known
as symmetric Néel wall and cross-tie wall, have a lower energy than the Bloch mode,
although magnetic poles cannot be avoided in those walls. The characteristics of
8 Magnetic Domains
417
D = 60 nm
D = 460 nm
Cross-tie wall
D = 10 nm
Asymmetric
Bloch wall
Symmetric
wall
20 µm
Permalloy thickness in nm
100
200
375
Wall angle
50
Symmetric
S
S
S
S
S
Tail
Asymmetric
N
N
N
N
N
Tail
Core
Cross-tie Asymmetric
wall
Bloch wall
Symmetric Néel wall
1
2
3 4 5 6 7 10 12 15 20 25 30 40 50 60
//
Vortex wall
5
D / A/K d
Side view
D / A/K u
Cross Bloch line
Circular Bloch line
Swirl
Fig. 18 Phase diagram for various types of domain walls that exist in low-Q thin and thick films
at zero applied field and in a hard-axis field that causes magnetization rotation in the domains,
thus reducing the wall angle. The corresponding thicknesses for permalloy are indicated. Shown
are high-resolution Kerr images of permalloy together with sketches (symmetric Néel and cross-tie
wall, Bloch lines and swirl) and micromagnetically simulated vector plots (asymmetric Bloch and
Néel wall, calculated for permalloy). The contour lines in the calculated wall profiles indicate the
center of the walls, i.e., the surfaces on which the magnetization is strictly aligned in the drawing
plane. The pictures are taken and adapted from Ref. [1]. Since the anisotropy has only a moderate
influence on the wall energy in films, the diagram is valid for a wider range of low-Q materials.
(Adapted by permission from Ref. [1] (c) Springer 1998)
the symmetric Néel wall is its decomposition in a sharply localized core and two
extremely wide tails that take over a large part of the total rotation. A dipolar pole
pattern appears at the core, which carries about half of the pole density, and the other
half of the poles are displaced in the tails, both of them adding stray field energy.
The two characteristic lengths of a 180◦ Néel wall in a film of thickness D are then
given by
Wcore = 2 A/(Ku + Kd ) and Wtail ≈ 0.56DKd /Ku .
(15)
Because Kd is much√larger than Ku for materials with Q 1, the core width
roughly scales with A/Kd , i.e., in the core the exchange energy (A) is primarily
balanced by the stray field energy (Kd ) that is connected with the magnetic poles.
The tail is rather determined by a balance between stray field (Kd ) and anisotropy
418
R. Schäfer
energy (Ku ). The stray field energy thus has an opposite effect on both parts of
the wall. The poles in the extended tail