Handbook of Magnetism and Magnetic Materials 3030632083, 9783030632083

This handbook presents a comprehensive survey of magnetism and magnetic materials. The dramatic advances in information

959 140 55MB

English Pages 1735 [1714] Year 2021

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Handbook of Magnetism and Magnetic Materials
 3030632083, 9783030632083

Table of contents :
Preface
Contents
About the Editors
Contributors
Part I Fundamentals
1 History of Magnetism and Basic Concepts
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
2 Magnetic Exchange Interactions
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
3 Anisotropy and Crystal Field
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
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
4 Electronic Structure: Metals and Insulators
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
5 Quantum Magnetism
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
6 Spin Waves
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
7 Micromagnetism
Contents
Introduction
Micromagnetics Basics
Magnetic Gibbs Free Energy
Spin, Magnetic Moment, and Magnetization
Exchange Energy
Magnetostatics
Zeeman Energy
Magnetostatic Energy
Demagnetizing Field as Sum of Dipolar Fields
Magnetic Scalar Potential
Magnetostatic Energy
Magnetostatic Boundary Value Problem
Examples
Crystal Anisotropy Energy
Cubic Anisotropy
Uniaxial Anisotropy
Anisotropy Field
Magnetoelastic and Magnetostrictive Energy Terms
Spontaneous Magnetostrictive Deformation
Magnetoelastic Coupling Energy
External Stress
Magnetostrictive Self-Energy
Characteristic Length Scales
Exchange Length
Critical Diameter for Uniform Rotation
Wall Parameter
Single Domain Size
Mesh Size in Micromagnetic Simulations
Brown's Micromagnetic Equation
Euler Method: Finite Differences
Ritz Method: Finite Elements
Magnetization Dynamics
Appendix
References
8 Magnetic Domains
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
9 Magnetotransport
Contents
Basics
Classical Magnetoresistance in Semiconductors and Semimetals
Magnetotransport and Ferromagnetism
Anisotropic Magnetoresistance, Planar Hall Effect, and Two-Current Model
Giant Magnetoresistance (GMR)
Colossal Magnetoresistance (CMR)
Tunneling Magnetoresistance (TMR)
Powder Magnetoresistance (PMR)
Organic Magnetoresistance (OMAR)
Quantum Transport
Exotic Tunneling
Hall Effect
Spin Currents
Spin Hall Effect
Inverse Spin Hall Effect
Quantum Spin Hall Effect
Magnetoimpedance
Measurements
References
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films
Contents
Introduction
Magneto-Optics
Basics of Magneto-Optics from a Macroscopic Perspective
Micropic Understanding of Magneto-Optics
Measuring Magnetism Using MOKE
Different Configurations
Layer-Specific MOKE
MO Spectroscopy
MOKE Microscopy
Measuring Ultrafast Magnetization Dynamics
Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism
Conceptual Introduction
Ultrafast Laser-Induced Loss of Magnetic Order
Experimental Demonstration of Laser-Induced Demagnetization
Key Observations in Laser-Induced fs Demagnetization
Conservation of Angular Momentum Revisited
Theories for Femtosecond Demagnetization
Towards Quantitative Understanding
All-Optical Switching of Magnetization
All-Optical Switching in Ferrimagnetic Alloys
All-Optical Switching in Ferromagnetic Systems
New Directions in All-Optical Switching
Laser-Pulse-Excited Spin Currents
Optically-Induced Spin Transfer
Optical Spin-Transfer Torque
Optical THz Spin Wave Excitation
Conclusions and Outlook
Notes
References
11 Magnetostriction and Magnetoelasticity
Contents
Introduction
Classification of Magnetoelastic Effects
Anisotropic Magnetostriction (Joule Magnetostriction)
Magnetovolume Effects, Spontaneous Magnetostriction, Forced Magnetostriction, and Invar Effects
Villari Effect
E Effect
Magnetomechanical Damping
Wiedemann Effect
Matteucci Effect
Magnetic Field-Induced Strain Phenomena, Which Differ from JouleMagnetostriction
Ferromagnetic Shape-Memory Alloys (FSMA)
Magnetic Field-Driven Spin-State Transition in La(1-x)SrxCoO3-δ (x≥0.3)
Magnetostriction in Superconductors
Magnetoelasticity and Joule Magnetostriction
Derivation of the Magnetostrictive Strain Tensor: Cubic Case
Magnetostriction of Polycrystalline Cubic Materials
Derivation of the Magnetostrictive Strain Tensor: Hexagonal Case
Magnetostriction of Polycrystalline Hexagonal Materials
Magnetostriction and Stress-Induced Magnetic Anisotropy
Magnetoelastic Effects in Films
Experimental Determination of Magnetostriction and Magnetoelastic Coupling
Magnetoelastic Coupling in Films
Magnetostriction and Magnetoelasticity: Physical Origin and Insights from Theory
Compilation of Data
Magnetoelastic and Elasticity Data for Bulk Transition Metals
Theoretical and Experimental Values of Magnetoelastic Coupling Coefficients and Their Strain Dependence (Table 4)
Magnetostriction Data of Amorphous Fe Alloys (Table 5)
Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si, Fe-Ga-Al, and Fe-Ga-Ge Alloys
Zero Magnetostriction Alloys and Soft Magnetic Materials
Magnetostriction Data for Paramagnetic Metals and Alloys
Magnetostriction Data for Bi
Magnetostriction Data for Tb, Dy, and Ho
Magnetostriction Data of TbFe2 (Terfenol) and Tb27Dy73Fe2 (Terfenol-D)
Magnetostriction Data of Oxide Materials
Appendix
Notations for Lattice Strain
Relation Between λ and B for the Hexagonal System
References
12 Magnetoelectrics and Multiferroics
Contents
Magnetoelectrics
Multiferroics
The Evolving Terminology
Ferroelectric Ferromagnets
Ferroelectric Antiferromagnets
Ferroelectric Ferrimagnets
Electric Field Switching of Magnetization in Multiferroic BiFeO3: Status and Perspective
BiFeO3 Bulk Single Crystals
Epitaxial BiFeO3 Thin Films
A Brief Future Perspective
Electric Field Control of Spin Cycloid in BiFeO3 Thin Films: From the Control of Collinear Magnetism to Noncollinear Magnetism
Towards the Control of Magnetism in BiFeO3 Thin Films at the THz Frequency: New Opportunities with Ultrafast Stimuli
Composite Multiferroics and Magnetoelectrics
Terminology and Exiting Reviews
Mechanisms and Application of Magnetoelectric Effects and Experimental Data of Magnetoelectric Coefficients in Composite Magnetoelectrics
Direct Magnetoelectric Effect
Inverse Magnetoelectric Effect
References
13 Magnetism and Superconductivity
Contents
Introduction
Paramagnetic Limit and Nonuniform FFLO Superconducting State
Interplay of Zeeman Field and Spin-Orbit Interaction: Spinless Fermions and a Route Towards Topological Superconductivity
Ising Superconductors: Interplay of Magnetic Field and Spin-Triplet Channels
Superconductivity in the Presence of Antiferromagnetic Order
Phenomenological Description
Microscopic Consideration: Important Aspects
Ferromagnetic Superconductors
Conclusions
References
Part II Magnetic Materials
14 Magnetism of the Elements
Contents
Introduction
The Magnetism of Iron, Cobalt and Nickel
Band Structures
Magnetic Properties
Thin Films of Fe, Co and Ni
Iron, Steels and Other Iron-Based Alloys
Phase Diagram
Manganese and Chromium
Manganese
Chromium
Spin Density Waves
Rare Earths
Magnetism of the Rare Earths
Magnetic Structures and Phase Transitions
Fermi-Level Spin Polarisation of the Magnetic Elements
Spin Polarisation of the 3d Elements
Spin Polarisation of the 4f Elements
Oxygen
Molecular Magnetism
Other Examples of Magnetic Order in the p- and d-Shell Elements
References
15 Metallic Magnetic Materials
Contents
Magnetic Metallic Glasses and Nanocrystalline Alloys
Alnicos
Intermetallic Compounds of d-d and d-p Types
Magnetic Shape Memory Alloys and Compounds
Magnetic Heusler Compounds
Intermetallic Compounds of 3d-4f Type
Sm-Co Permanent Magnets
Nd-Fe-B Permanent Magnets
Magnetocaloric Intermetallics
Heavy-Fermion Compounds
References
16 Metallic Magnetic Thin Films
Contents
Introduction
GaAs (001) Substrates
Structure and Magnetism of Magnetic 3d Transition Metals
Cr/GaAs (001)
Mn/GaAs (001)
Fe/Cu (001)
Co/GaAs (001)
Ni/GaAs (001)
Structure and Magnetism of Magnetic 3d Transition Metal Alloys
Py/GaAs (001)
CoxMn1−x/GaAs (001)
FexCu1−x/GaAs (001)
FexPd1−x/Cu (100)
References
17 Magnetic Oxides and Other Compounds
Contents
Background
Rocks, Solid Solutions, and Percolation
Principles of Oxide Magnetism
Iron Oxides and Hydroxides
Hematite
Magnetite
Maghemite
Wüstite
Iron Hydroxides
Goethite and Other Oxyhydroxides
Ferrihydrite and Ferric Gel
Ferrous Hydroxide
Ferrites
Spinels
Garnets
Orthoferrites
Hexagonal Ferrites
Other Magnetic Oxides
3d Oxides
Monoxides
Sesquioxides
Dioxides
Mixed-Valence Oxides
Perovskites
Rare-Earth Orthoferrites and Related Compounds
Perovskite Solid Solutions
Double Perovskites and Related Materials
Pyrochlores
4d and 5d Oxides
4f Oxides
5f Oxides and Related Compounds
Related Compounds
Halides
Chalcogenides
Pnictides
Silicates and Carbonates
Silicates
Carbonates, Phosphates
Oxide Thin Films
Substrates, Caps, and Buffers
Thin Film Preparation
Thermal Evaporation
Molecular Beam Epitaxy (MBE)
Sputtering
Pulsed Laser Deposition (PLD)
Chemical Vapor Deposition (CVD)
Atomic Layer Deposition (ALD)
Magnetic Oxide Monolayers
Dead Layers
Dilute Magnetic Oxides
Oxide Heterostructures and Interfaces
Oxide Interfaces
Spin Pumping
Magnonics
Conclusion
References
18 Dilute Magnetic Materials
Contents
Introduction
Dilute Magnetic Materials
Dilute Magnetic Metals (DMMs)
p-Type Dilute Ferromagnetic Semiconductors (DFSs)
Other Dilute Ferromagnetic Semiconductors
Dilute Magnetic Semiconductors (DMSs)
Dilute Magnetic Topological Materials
Heterogenous Magnetic Semiconductors and Oxides
Energy States of Magnetic Dopants in Solids
Exchange Interactions Between Band and Localized Spins
Effects of sp-d(f) Exchange Interactions
Spin-Splitting of Extended States: Weak Coupling
Spin-Splitting of Extended States: Strong Coupling
Alloy and Spin-Disorder Scattering: Weak Coupling
Alloy and Spin-Disorder Scattering: Strong Coupling
Bound Magnetic Polarons
Quantum Localization and Mesoscopic Phenomena: Colossal Magnetoresistance
Interplay of sp–d(f) Exchange Interactions and Spin-Orbit Coupling
Dominant Spin-Spin Interactions
Dipole-Dipole Interactions
Direct Spin-Spin Interactions
Superexchange
DMSs with Transition Metals
IV-VI DMSs with Rare-Earth Metals
Carrier-Mediated Spin-Spin Coupling: Intra- and Interband Contributions
RKKY Interaction
Bloembergen-Rowland Mechanism
p–d Zener Model
Theory of Curie Temperature
Magnetic Properties of Dilute Magnetic Materials
Spin-Glass Systems
p-Type Dilute Ferromagnetic Semiconductors
Dilute Ferromagnetic Insulators and Topological Insulators
Heterogenous Magnetic Semiconductors and Oxides
Phase Separation Effects in (Ga,Mn)As
Phase Separation Effects Beyond (Ga,Mn)As
References
19 Single-Molecule Magnets and Molecular Quantum Spintronics
Contents
Introduction
Spin Hamiltonian
Single Ion Spin Hamiltonian
Transition Metal Ions
Lanthanide(III) Ions
Multi-spin Hamiltonian
Giant Spin Hamiltonian
Quantum Tunneling of Magnetization
Landau–Zener–Stückelberg (LZS) Model
Spin Parity and Quantum Phase Interference
Spin Parity
Quantum Phase Interference
Quantum Coherence in Molecular Magnets
Resonant Photon Absorption
Rabi Oscillations
Molecular Quantum Spintronics
Direct Coupling
Read-Out of a Single Nuclear Spin
Indirect Coupling
Quantum Algorithms
Conclusion
References
20 Magnetic Nanoparticles
Contents
Ideal Single-Domain Nanoparticles
How Real Magnetic Nanoparticles Can Be Different
Magnetic Nanoparticle Preparation
General Nanoparticle Formation
Liquid Phase Syntheses
Nonaqueous Syntheses
Aqueous Syntheses
Nanoparticle Coatings and Extra Requirements of Biomedical Applications
Applications of Magnetic Nanoparticles
Important Magnetic Characteristics
Magnetic Separation and Manipulation
Magnetic Hyperthermia
Magnetic Particle Imaging (MPI)
Contrast Agents for Magnetic Resonance Imaging (MRI)
Ferrofluids
Magnetic Recording Media
Nonbiomedical Topics of Recent Interest
L10 FePt Nanoparticles for Magnetic Recording Media
Core-Shell NPs
Surface Effects and Spin Canting
Frontiers and Future Directions
References
21 Artificially Engineered Magnetic Materials
Contents
Nanocomposite materials
Thin-film structures
Bilayers and multilayers with engineered properties
Exchange bias
Interlayer exchange coupling
Perpendicular magnetic anisotropy
Dzyaloshinskii-Moriya interaction
2D electron gases
Nanowires
Template-grown nanowires
Edge states in 2D topological insulators
Lithographically-patterned 2D ferromagnetic arrays
Hysteresis of nano/microdisc arrays
Artificial spin-ice structures
Artificial atoms and the Coulomb blockade
Coulomb blockade
Single electron spintronics
Spin accumulation and spin lifetime
Cotunnelling effects
Kondo effect
Chemical potential effects
Single atom manipulation and measurements
Quantum corrals and wave function imaging
Magnetism at the atomic scale
Anisotropy of individual atoms
References
Part III Methods
22 Magnetic Fields and Measurements
Contents
Introduction
Magnetic Field Generation
Permanent Magnets
Physical and Material Properties of Permanent Magnets
Applications of Permanent Magnets As Flux Sources
Electromagnets
Resistive and Superconducting Solenoids
Helmholtz Coils
Electromagnets with Magnetic Core
High-Field Magnet Facilities
Performance Limitations and Magnet Classification
DC Magnets and Facilities
Pulsed Magnets and Facilities
Megagauss Magnetic Fields
Magnetic Measurements
Magnetic Field Effects
General Technical Principles
Technical Parameters: Accuracy, Precision, Sensitivity, Resolution, and Responsivity
Common Problems: Noise
Common Solutions: Flux Concentration, Modulation, and Null Detection
Magnetic Field Measurements
Induction and Fluxgate Sensors
Hall, Magnetoresistance (MR), and Magnetoimpedance (MI) Sensors
Superconducting Quantum Interference Devices (SQUIDs)
Nuclear Magnetic Resonance (NMR) Sensors
Nonresonant Electromagnetic Probing: Faraday Rotators
Mechanical Sensors: MEMS-Based Magnetic Force Meters and Magnetoelectric Laminates
Bulk Magnetic Measurements
Induction Magnetometers
Torque Magnetometers
Stray Field Mapping
Calibration and Metrology
References
23 Material Preparation and Thin Film Growth
Contents
Thin Film Growth: General Concepts
Phase Transition Out of Equilibrium, Surface Kinetics and Growth Modes
Choice of Substrate and Role of Buffer Layer
Molecular Beam Epitaxy
Introduction
The Ultrahigh Vacuum Environment
Control of Molecular Beam Fluxes and Substrate Temperature During Growth
The In Situ Monitoring of the Film Growth
The Applications of MBE in Novel Functional Materials
Two-Dimensional Magnetic Materials
Topological Insulators
Dirac and Weyl Semimetals
High-Temperature Superconductors
Layered Transition Metal Dichalcogenides
Comparison with Other Deposition Techniques
Pulsed Laser Deposition
Introduction
Target Preparation and Manipulation
Laser
Ablation Process
Optics
Thickness Monitoring
Substrate Heating
Selected Applications
Magnetron Sputtering Deposition
Introduction
Material Targets
Magnetron Sources
Magnetron Operation
Chimneys and Shutters
Reactive Magnetron Sputtering
RF Magnetron Sputtering
Growth of Alloy Films
Ion Beam Sputter Deposition
Ion Beam Deposition Source
Ion Beam Deposition Geometry
Ion Beam Deposition Targets
Ion Beam Deposition Source Operation
Selected Applications
References
24 Magnetic Imaging and Microscopy
Contents
Introduction
Electron Microscopy
Transmission Electron Microscopy
Lorentz Microscopy
Fresnel Imaging
Foucault Imaging
Differential Phase Contrast Microscopy
Electron Holography
Off-Axis Holography
Aberration Correction
Scanning Electron Microscopy
SEMPA
Scanning Probe Microscopy
Spin-Polarized Scanning Tunneling Microscopy
Technical Details
Experimental Examples
Magnetic Force Microscopy
Technical Details
Experimental Examples
X-Ray Imaging
X-Ray Magnetic Circular Dichroism: A Contrast Mechanism
TXM: Quick Full-Field Imaging in Transmission Geometry
STXM: Optimized for Dynamic Imaging
PEEM: Imaging Surfaces of Bulk Samples
CXI: Zero Drift and Femtosecond Temporal Resolution
SP-ARPES: Microscopy in Momentum Space
Medical Magnetic Imaging
Magnetic Resonance Imaging
Nuclear Magnetic Resonance
Imaging and Pulse Sequences
Functional Magnetic Resonance Imaging
Nuclear Quadrupole Resonance Imaging
Magnetoencephalography
The Inverse Problem
Summary
References
25 Magnetic Scattering
Contents
Introduction
Magnetic Neutron Diffraction Technique
Polarized Neutron Techniques
Polarized Neutron Reflectometry
Resonant Magnetic X-ray Diffraction Technique
Dynamics
Inelastic Neutron Scattering Technique
Resonant Inelastic X-ray Scattering Technique
Magnetic Diffraction Examples with Neutrons
Magnetic Diffraction Examples with X-rays
Spin Dynamics with Neutrons
Spin Dynamics with RIXS
Facilities and Online Information
Summary and Future Directions
References
26 Electron Paramagnetic and Ferromagnetic Resonance
Contents
Introduction
Magnetic Resonance of Electrons
Microscopic Equation of Motion
Macroscopic Equation of Motion
On the Difference Between EPR and FMR
Electron Paramagnetic Resonance
Bloch Equations
Effective Spin Hamiltonian
Anisotropic Zeeman Interaction
Fine Structure
Hyperfine Interaction
Ferromagnetic Resonance
Effective Magnetic Fields
Shape Anisotropy
Magnetocrystalline Anisotropy
Exchange Energy and Spin Wave Resonance
FMR in Metals
Experimental Observation of Electron Magnetic Resonance
Frequency Domain Cavity EPR/FMR
Survey of Other Methods
Frequency Domain Techniques
Time-Domain Techniques
Electrically Detected Magnetic Resonance
Optically Detected Magnetic Resonance
Conclusion
References
27 Magnetization Dynamics
Contents
Introduction
Spin-Transfer Torque, Magnetic Switching, and Oscillations
Spin-Transfer Torque in Phenomenological Form
Spin-Torque-Driven Anti-damping Magnetic Switching
Orthogonal Spin-Torque-Driven Magnetic Switching
Spin-Torque Oscillators
Spin-Transfer-Induced Excitation of Spin Waves
Spin-Transfer Vortex Oscillators
Synchrotron and Femtosecond Laser-Based Time-Resolved Spin Dynamics
Switching Schemes
Microscopy Using Visible Light or X-Rays
Ultrafast Spin-Transfer Torques
Summary and Outlook
References
Part IV Applications
28 Permanent Magnet Materials and Applications
Contents
Permanent Magnets and Physics Behind Them
Stray Fields and Demagnetizing Fields
On the Thermodynamics of Magnets
Hysteresis Curves and Magnetization Processes of Permanent Magnets
Permanent Magnet Materials
Hard Magnetic Steels and Alnico
Hard Ferrite
Rare-Earth Magnets Based on SmCo5
Sm2Co17-Based Magnets
Nd2Fe14B-Based Magnets
Other Permanent Magnet Materials
Resource Constraints
Future Permanent Magnet Materials
Permanent Magnet Applications
Permanent Magnet Assemblies and Their Dynamic Application with Mechanical Recoil
Uniform Magnetic Fields
Nonuniform Magnetic Fields
Dynamic Applications with Active Recoil
References
29 Soft Magnetic Materials and Applications
Contents
Losses in Soft Magnets
Hysteresis Loss
Low-Frequency Losses
Steinmetz Model
Rayleigh Loops
Eddy Currents
Classical Losses
Skin Effect
Excess Loss
The Pry and Bean Model
Bertotti's Model
Rotational Losses
Rotational Eddy Current Loss
Rotational Hysteresis Loss
High-Frequency Losses
High-Frequency Losses in Metals
High-Frequency Losses in Insulators
Iron and Low-Carbon Steels
Electrical Steels
Composition, Processing, and Texture
Composition
Processing
Texture
Designation of Magnetic Steels
Non-oriented Silicon Steel Sheets
Cutting
Thin-Gauge Steels
Grain-Oriented Silicon Steel Sheets
Conventional GOSS
High Induction GOSS (HiB)
Cutting
Trends in Fe-Si Products
Improved Texture NOSS
Low Si Alloys
Silicon-Enriched Electrical Steels
Iron-Cobalt Alloys
Equiatomic Fe-Co Alloy
Low-Cobalt Alloys
Iron-Nickel Alloys
Ni-Rich Alloys
Fe-Rich Alloys
Thermal Alloys
Amorphous and Nanocrystalline Alloys
Iron-Based Amorphous Alloys
Cobalt-Based Amorphous Alloys
Nanocrystalline Alloys
Soft Ferrites
Spinel Ferrites
Synthesis of Sintered Ferrites
Mn-Zn Ferrites
Influence of Chemical Composition on Intrinsic Magnetic Properties
Influence of Microstructure on Extrinsic Magnetic Properties
Ni-Zn Ferrites
Other Soft Ferrites
Ni-Zn-Cu Ferrite
Microwave Ferrites
Effect of a Gap on Magnetic Properties
Core with Lumped Gap
Cores with Spread Gap
Iron Powder Cores
Mo-Permalloy Powder Cores
Other Powder Cores
Annex 1: Comparison of Materials for Power Application in the VLF to HF Band
Annex 2: Materials Selection Table
Notes
References
30 Magnetocaloric Materials and Applications
Contents
Introduction: The Magnetocaloric Effect
Cooling at Low or Intermediate Temperatures
A Brief History of the Magnetocaloric Effect
Adiabatic Demagnetization (ADR) Materials
Nuclear Adiabatic Demagnetization (NDR)
The Magnetocaloric Effect Near Room Temperature
Room Temperature Magnetocaloric Materials
The Gd5(Si,Ge)4 System
The (Mn,Fe)2(P,Si) System
The La(Fe,Co,Mn,Si)13Hy System
Fe-Rh
Heusler Compounds and Other Martensitic Materials
Room Temperature Magnetic Cooling Engines
Issues for Deployment of Known Magnetocaloric Materials in Cooling Engines
Power Conversion and Spin Caloritronics
Power Conversion via Pyromagnetic Materials
Spin Seebeck and Spin Nernst Effects and Spin Caloritronics
References
31 Magnetic Sensors
Contents
Magnetic Sensing
Introduction
Relevant Parameters for Choice of a Magnetic Sensor Source Signal Parameters
Sensor Characteristics
External Parameters
Field or Flux Sensing
Sensor Types
Flux Sensors
Coils and Inductive Sensors
Principle
Main Applications
Fluxgates
Principle
Main Applications
SQUIDs
Principle
Main Applications
Field Sensors
Hall Effect Sensors
Principle
Main Applications
AMR Sensors
Principle
Main Applications
Spin Electronic Sensors (GMR, TMR)
Principle
Main Applications
Optically Pumped Magnetometers
Principle
Main Applications
Nitrogen Vacancies Centers
Principle
Main Applications
Giant Magnetoimpedance
Principle
Main Applications
Magnetoelectric Sensors
Principle
Noise
Definitions
Noise Sources [31]
Detection Schemes
Magnetometer
Gradiometer
Closed Loop
Bridge
Applications
Field/Flux Sensing and Flux Transformers
Current Sensing
Position Sensing
References
32 Magnetic Memory and Logic
Contents
Introduction: Magnetic Memory and Logic
Magnetic Hard Drive Memory
Hard Drive and Its Road Map
Giant Magnetoresistance and RKKY Coupling
Tunneling Magnetoresistance
Magnetic Random-Access Memory
Magnetic Random-Access Memory
Spin-Transfer Torque Magnetic Random-Access Memory
Spin-Orbit Torque Magnetic Random-Access Memory
New-Type Magnetic Memories
Domain Wall Racetrack Memory
Skrymion Racetrack Memory
Antiferromagnetic Memory
2D van der Waals Magnets Memory
Magnetic Logic
Nanomagnet Logic
Magnetic Domain Wall Logic
Magnetic Tunnel Junction Logic
Spin Current Logic
All Spin Logic
Graphene Spin Logic
Magnetoelectric Spin-Orbit Logic
Spin Wave Logic
Conclusion and Outlook
References
33 Magnetochemistry and Magnetic Separation
Contents
Introduction
Magnetic Fields
Magnetic Susceptibility
Magnetic Susceptibility of Liquids
Pascal's Constants
Magnetic Field Effects in Chemistry
Magnetoelectrochemistry
Electrochemistry
Magnetic Forces in a Uniform Magnetic Field
In the Absence of Currents
Lorentz Force
Characteristic MHD Numbers
Deposition
Corrosion
Magnetic Forces in a Non-uniform Magnetic Field
Magnetic Structuring and Magnetophoresis
Magnetic Water Treatment
Magnetic Separation
High-Gradient Magnetic Separation (HGMS)
Magnetic Carriers
Wastewater Treatment
Biological Separation
Magnetically Stabilized Bed (MSB) Reactors
Conclusions
References
34 Magnetism and Biology
Contents
Biomagnetism
Magnetic Navigation in Living Organisms
Magnetotactic Bacteria
Avian Magnetic Navigation
Magnetite-Based Magnetoreception
Radical-Pair-Based Magnetoreception
Magnetic Fields Produced by the Body
Iron in the Body
Magnetobiology
Magnetic Nanoparticles for Biomedical Applications
Magnetic Imaging
Magnetic Resonance Imaging (MRI)
Functional Magnetic Resonance Imaging (f-MRI)
Static Magnetic Field Sources Used for MRI and f-MRI
Magnetic Particle Imaging (MPI)
Magnetic Fixation
Magnetic Actuation
Magnetic Trapping (Non-Lab-on-a-Chip)
Magnetic Cell Separation
Magnetic Hemofiltration
Magnetic Targeted Drug Delivery
Magnetic Manipulation in Lab-on-a-Chip
Magnetic Trapping and Separation
Magnetic Transportation
Magnetic Mixing
Magnetic Tweezers
Measurement of Physical Properties
Mechanical Stimulation
Magnetically Induced Cell Destruction
Magnetoporation
Magnetic Orientation
Magnetic Heating
Effects of Magnetic Fields on Biological Organisms
Transcranial Magnetic Stimulation (TMS)
Recommended Magnetic Field Exposure Limits
Conclusions and Prospects
Further Reading
References
Index

Citation preview

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 , , /2 10 i 0 0 −1

(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 ≈ . 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 = =

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 = U (r , r ) to Eq. (4) yields Coulomb energy H12 1 2 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 = 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

 4T 2 +

(15) U2 4

(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 U ), Eq. (17) becomes J =

U + JD − |T | 4

(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 |
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 T ), Eq. (17) becomes J = JD −

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 cˆi↑ cˆj↑ + cˆi↓ cˆ↓ + U i nˆ i↑ nˆ i↓ (28) where nˆ = cˆ+ cˆ [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 ∇ + j Vo (r − r j ) 2me

(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 nˆ i↑ nˆ i↓ |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 + Cij + co

(44)

and neglecting  the thermodynamic correlation term Cij = (S i − ) · S j − and the constant co = · . 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 · − 2μo μB S · H

(45)

This equation amounts to the introduction of a mean field μo μB H =2 z J 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 has the form of a Brillouin function (BS ),

84

R. Skomski

and self-consistently evaluating 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 = j Jij

(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 = j JAA + JAB

(48a)

3 kB T = j JBA + JBB

(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 ∗ 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 j J 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 sˆ · sˆ → J sˆx · sˆ x + sˆy · sˆ y + Jz sˆz · sˆ 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 = 0 and = 0, whereas the idea of the (original) Ising model is to suppress such states, = = 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., Ole´s, 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. Sa¸ ¸ sıo˘glu, 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., Go´scia´nska, 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) = −

N qi e 4πεo | Ri − r |

(13)

i=1

This sum is easily converted into a sum of spherical harmonics by exploiting the identity: 1 4 π rl = Y1 m∗ (Θ, Φ) Y1 m (θ, φ) l (2 l + 1) R l+1 |m| 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 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 = − i ψ* lz ψ dφ yields = 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 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 = ± 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 4 4 4 4 4 2 2 4 4 4 20 x + y + z − 3r /5 = 35z − 30 z r + 3r + 5 (x + iy) + (x–iy) 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 , 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 = 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 = 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,

  • = 0, and therefore
  • = < 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

    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

    Eu −Eo

    (50)

    The numerical determination of the anisotropy constants requires the evaluation of 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 , 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 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: =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 = ±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) = 2i, (b) = i, and (c) = 3 i. The dashed lines are out of the paper plane and visualize the direction of ˆl , 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 , , , , and , whereas that of Fig. 15(c) has two realizations, namely, and . A physical interpretation of matrix elements is that = 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 exp ik · Rj φμ r − Rj j

    (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 sˆ z 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 Sˆx · Sˆx –Jyy Sˆy · Sˆy –Jzz Sˆz · Sˆz

    (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 − S (S + 1) 2

    (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. The parameter J reflects spin-orbit coupling, very similar to singleion 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, ≈ 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

    Nl m =

    (2l + 1) (l − m)! 4π (l + 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: =

     n2 ro 2  2 5n + 1 − 3l (l + 1) 2

    =

     ro  2 3 n − l (l + 1) 2

    =

    =

    =

    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, and the square root of are used to estimate shell radii, gives the electronic energy, and 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, (h¯ = 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 ˆ of the local spin moment m. Accordingly, one can represent the the orientation m spin-dependent part of the effective 2 × 2 potential matrix function by: ˆ B(r), V spin (r) = σ · m

    (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

    ˆ = U † (θ, φ) σz U (θ, φ). means of a transformation matrix U (θ, φ) such that: σ · m This matrix can be obtained from Eq. (80) (see below) by setting θ and φ according ˆ and q = 0. In either case, the form of the spin-dependent to the orientation m 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 nˆ ilmσ nˆ ilm σ  2   il



    mσ m σ

    1   † † Jmm cˆilmσ cˆilm σ¯ cˆilm ˆilmσ  σ¯ c 2  il mσ m  − Δl nˆ ilmσ . il



    (13)

    i=id ,l=d

    Here, c, ˆ cˆ† , and nˆ 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 φ˙ lν (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 φ˙ lν (r): ψj k (r) =

    

    jk jk Alm φlν (r) Ylm (ˆr ) + Blm φ˙ lν (r) Ylm (ˆr ),

    (16)

    lm jk

    jk

    where the expansion coefficients Alm and Blm are given by the eigenvectors α j k and Ylm (ˆr ) 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]: 1 Hˆ D = −icα · ∇ + c2 (β − 1) + V¯ (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 V¯ (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

    pˆ 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 =

    (pˆ + h¯ k)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

    ∂k

    

    uj 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), h∂k h ¯ ¯

    (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 h¯

     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 ) h j

    (37)

    where mn (k) = (e/2h)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)˙r D BZ (2π )  dDk e 1 Ω(k)log(1 + e−β((k)−μ) ). −∇ × β BZ (2π )D h¯

    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

    ˆi ·m ˆ j, Jij m

    (51)

    i=j

    ˆ i(j ) the orientation of the magnetic moment on the lattice site i(j ), for with m 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 ˆ i }) on the rather simple expression in of the complicated energy landscape E({m 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 ˆ i and m ˆ j , one gets a one-to-one mapping of the orientation angle of the moments m 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



    

    ˆ iJ m ˆj + m

    

    ij

    ˆ i) K(m

    (53)

    i

    ˆi ·m ˆj − Jij m

    

    ˆ iJ S m ˆj m ij

    i=j

       ˆi ×m ˆj + ˆ i ), D ij · m Ki ( m

    i=j

    (54)

    i

    with the accompanying single-site magnetic anisotropy represented by the term ˆ i ). In Eq. (53), the coupling tensor J has been decomposed into its isotropic K(m 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 ˆ i }) of a system approximate mapping of the complicated energy landscape E({m 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 ˆ of its magnetization with the anisotropic part a system on the orientation m ˆ of the energy taking a minimum for m ˆ along a so-called easy direction EA (m) ˆ is split into the intrinsic material-specific of the magnetization. Usually, EA (m) ˆ and the extrinsic shape magnetocrystalline anisotropy (MCA) energy EMCA (m) ˆ determined by the shape of the sample: anisotropy energy Eshape (m) ˆ = EMCA (m) ˆ + Eshape (m). ˆ EA (m)

    (57)

    ˆ m ˆ  ), with X=A, MCA or shape, Considering the difference in energy ΔEX (m, ˆ and m ˆ  , respectively, respectively, for the magnetization oriented along directions m   ˆ m ˆ ) = EX (m) ˆ − EX (m ˆ ). one has accordingly ΔEX (m, A convenient phenomenological representation of the magnetocrystalline anisotropy energy can be given by an expansion in terms of spherical harmonics ˆ Ylm (m) ˆ = EMCA (m)

     m=l 

    ˆ κlm Ylm (m)

    (58)

    l even m=−l

    or alternatively by an expansion in powers of the direction cosines (α1 , α2 , α3 ) = ˆ · x, ˆ m ˆ · y, ˆ m ˆ · zˆ ): (m ˆ = b0 + EMCA (m)

     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 magnetizaˆ of the magnetization tion, only terms that are even with respect to the orientation m 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 (m) ˆ are given by: expressions for EMCA hex ˆ = K˜ 0 + K˜ 1 Y20 (θ, φ) + K˜ 2 Y40 (θ, φ) EMCA (m)

    +K˜ 3 Y60 (θ, φ) + K˜ 4 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 K˜ 1 = 21 (7K1 + 8K2 + 8 16 1 ˜ ˜ ˜ 8K3 ), K2 = 385 (11K2 + 18K3 ), K3 = 231 K3 and K4 = 10,395 K4 . ˆ is usually associated with the classical The shape anisotropy energy Eshape (m) 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:

    ˆ = Edip (m)

     mν mν   c2

    νν 

    Rn

     ˆ 2 [R nνν  · m] 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 ˆ has to be ascribed to the spin-orbit coupling. Accordingly, the energy EMCA (m) ˆ m ˆ  ) can be determined by total energy corresponding energy difference ΔESOC (m, ˆ and m ˆ  , respeccalculations with the magnetization oriented along directions m tively, and taking the difference. Obviously, this implies a full SCF calculation for ˆ m ˆ  ). both orientations and taking the difference of large numbers to get ΔESOC (m, The problems connected with this approach can be avoided by making use of the ˆ m ˆ  ) by so-called magnetic force theorem that allows to approximate ΔESOC (m, 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]: ˆ m ˆ ) = − ΔESOC (m,

    

    EFmˆ

     dE N mˆ (E) − N mˆ (E)

       1  − nmˆ (EFmˆ ) (EFmˆ − EFmˆ )2 + O(EFmˆ − EFmˆ )3 . 2

    (63)

    ˆ while nmˆ (E) and Here EFmˆ is the Fermi energy for the magnetization along m  E N mˆ (E) = dE  nmˆ (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 ˆ m ˆ  ) could be achieved [157]. Application of this scheme for the energy ΔESOC (m, ˆ m ˆ  ) shows that the dominating contributions spatial decomposition of ΔESOC (m, 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.

    ˆ to the single-particle energies as a sum of two-site interactions: ESOC (m)  ˆ = ESOC (m)

    EF

    dE δN(E) =

    

    ˆ Eij (m)

    (64)

    ij

    with ˆ =− Eij (m)

    1  2π

    

    EF

    ij

    j

    ji

    i ˆ HSOC G0 (m) ˆ HSOC dE Trace G0 (m) ,

    (65)

    ij

    ˆ are real space structural Green function matrices corresponding to the where G0 (m) 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 ˆ and the spin-orbit-induced the interconnection between the energy correction E(m) 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 ˆ = − C ζ σ · l , ESOC (m) 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 ˆ to m ˆ  to get a component when changing the orientation of the magnetization from m ˆ m ˆ  ). resolved estimate for the corresponding anisotropy energy ΔESOC (m, 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

    

    ∂HSOC

    ψj k , 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

    Tαmˆuˆ = −

    1  π

    

    EF

    dE

      ∂

    −1 0 ˆ ln det t( , m) − G ∂α uˆ

    227

    (68)

    where the torque component with respect to a rotation of the magnetization around ˆ is the single-site t-matrix for an orientation an axis uˆ is considered and where t(m) 0 ˆ and G is the corresponding free electron Green function of the moments along m matrix. On the basis of Eq. (67) or (68), respectively, the anisotropy energy ˆ m ˆ  ) is obtained from the torque by integrating along a path connecting ΔESOC (m,  ˆ and m ˆ . This approach is especially suited when dealing with systems with m 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 . ˆ m ˆ  ) to the total anisotropy energy that is associated The contribution ΔEdip (m, with the dipole-dipole interaction of the individual magnetic moments is usually treated classically by evaluating a corresponding Madelung sum (see Eq. (62)) [155, ˆ m ˆ  ) is much larger than ΔEdip (m, ˆ m ˆ  ), 163, 164]. While for most cases ΔESOC (m, both contributions are often found for layered systems to be in the same order ˆ m ˆ  ) always favors an in-plane orientation of the of magnitude. As ΔEdip (m, ˆ m ˆ  ) in general favors an out-of-plane orientation, magnetization while ΔESOC (m, 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]. ˆ m ˆ  ) and ΔEdip (m, ˆ m ˆ  ) are of the In particular in cases for which ΔESOC (m, 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 ˆ m ˆ  ) is caused by the Breit pointed out already nearly 30 years ago that ΔEdip (m, 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 ˆ m ˆ  ) that are very close to those of a coherent quantumto results for ΔEdip (m, mechanical 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 m Am n θJ n r 4f On 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 ˆ i implying a coherent rotation of the magnetization within the cell during direction m 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 ˆ i that can be written as [144, 180, 181]: for the local magnetic moments mi = mi m 2μB 1 ∂E d ˆi =− ˆ i, m ×m ˆi h¯ mi ∂ m 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]: h¯ 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 zˆ 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 h¯ ω(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 hω(q) ¯ from the expression [191]: 4 ΔE(q, θ ) . θ→0 m sin2 θ

    hω(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 hω(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 h¯ ω(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 Sa¸ ¸ sıo˘glu 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 ) nˆ 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 nˆ is determined by the statistical average over all possible orientations eˆ with the probability distribution for the local magnetic moments given by: P nˆ (ˆe) = 

    ˆ e e−β hW n·ˆ ˆ e d eˆ e−β hW n·ˆ

    ,

    (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 Gm (z), RPA z→0 m k B Tc

    (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. Ležai´c 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

    600

    T (K)

    500

    245

    300

    RPA MFA rRPA Expt.

    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 {ˆei }. Within the DLM theory, this is determined by the corresponding single-site probabilities P nˆ (eˆi ) obtained from Eq. (86) for the Weiss field hW = hnˆ . Following Györffy et al. [230], hnˆ 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 H0nˆ = j hnˆ · eˆ j with hnˆ = hnˆ nˆ and using the Feynman-Peierls inequality [231]: F ≤ F0 + H − H0 ,

    (94)

    where the canonical distribution Eq. (86) is used to calculate the average. Using the Weiss field hnˆ as a variational parameter to minimize the right-hand side of Eq. (94) one is led to [162]: 3 h = 4π nˆ

    

    d eˆ i eˆ i Hnˆ eˆ i ,

    (95)

    where Hnˆ eˆ i denotes the average of Hnˆ with the restriction that the orientation of the moment on site i is fixed to eˆ i [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 )nˆ can be obtained from  M(T ) = mloc (T )

    d eˆ i P nˆ (ˆei ) nˆ · eˆ i ,

    (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. Kováˇcik, 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. Sa¸ ¸ sıo˘glu, 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. Vondráˇcek, 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. Ležai´c, 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. Sa¸ ¸ sıo˘glu, 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., Sa¸ ¸ sıo˘glu, 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., Sa¸ ¸ sıo˘glu, E., Mavropoulos, P., Ležai´c, 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. Ležai´c, 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 /h¯ , exponentiating a phase ϕ = Aμ /h, ¯ 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 h¯ → 0, the spin moment hS ¯ 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 ˆ coordinate n(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 = hS, ¯ 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 (τˆ ) = o τˆx , where τˆ 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 (τˆ ) = o τˆ x + o τˆ 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 τˆi , 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 τˆix + i τˆiz + Vij τˆiz τˆjz 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 = h¯ S, along a path which  has coordinate ˆ n(t) = S(t)/S, and encloses a solid angle . The kinetic phase φB = q/h¯ 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 zˆ -axis, hard x-axis) ˆ 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⊥ ˆ pulls the two states, and the paths between o , applied along the hard x-axis, them, toward x, ˆ 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⊥  ˆ as a function of transverse field o ), for a typical biaxial potential (easy axis zˆ , hard axis x), ⊥ H⊥ o . If Ho is rotated away from xˆ 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 = o /Vo 1 (the “dipolar-dominated” regime in Fig. 2c). As the resonant spins 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 Eo to 0.1 K, we should get “mesoscopic” coherent dynamics (i.e., Qφ > 1). 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 > 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 = 0, ∂x x=±d/2

    (6)

    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 2 d

    = q2 +

    pπ 2 d

    ,

    (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:  

    pπ 2  2A 2 qy + ωp = 2πfp = γ H + Ms d  2 1/2 (12)  

    pπ 2  /H 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:



    w mn (y) = An sin y+ (18) w 2 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): 2 



      w z+ exp −iq n y I (y, z) = An sin n w 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 ∞  n n hy (z)my (z)dz · hy (y)my (y)dy An ∝ −w/2 −∞

    (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:

    nb q sin y 1 − (−1) An ∝ n qyn n

    (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 < 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, h¯ S, μ(x i ) = −g

    |e| hS(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 h¯ is the reduced Planck constant, h¯ = 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 hS ¯ 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



      ∂ 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 

    

    2 A (∇mx )2 + ∇my + (∇mz )2 dV .

    Eex = V

    (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 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 x˜ = 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 2 + ∇m  z 2 dV ∇m  L2 V  , =− m · hext d V

    ex = E

    demag = − 1 E 2  ani = − E

     V

    

    (93)

    , m · hdemag d V

    (94)

    K1 , (c · m)2 d V μ0 Ms2

    (95)

     V

     V

    (92)

     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 ≈

    (mij k · n)(mi  j  k  · n ) μ0   Ms,ij k Ms,i  j  k  dSdS  . | 8π |x − x ∂V ∂V    ij k    i j k ij k i j k

    (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



    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 =

    d, 2

    =0

    H

    Wall velocity

    valid. Then the spin polarization of the current cannot follow the rapidly changing magnetization profile anymore, i.e., the exchange of spin angular momentum between the conduction electron and the local moment is not conservative, leading eventually to mistracking or even reflection (domain wall scattering) of the itinerant spin. These effects are considered in the second, non-adiabatic spin torque term in (16) that was independently introduced in refs. [36, 35] and which describes a torque perpendicular to the adiabatic torque. The deviation from adiabaticity of the STT is designated by the non-adiabatic STT parameter β, the ratio between the non-adiabatic and adiabatic torques. It considers the mentioned spin mistracking and spin relaxation in the wall by spin-flip scattering at impurities, phonons, etc. In general both adiabatic and non-adiabatic terms have to be considered for a given domain wall structure. For current-driven wall motion, similar mechanisms are predicted as for the field torque. Their appearance depends on the relative values of the Gilbert damping parameter α and the β parameter, which both are dimensionless. As schematically shown in Fig. 22a, there is some threshold current density in the absence of the nonadiabatic torque (β = 0), below which the wall does not move at all. This threshold is intrinsic (i.e., not caused by extrinsic pinning effects) because it originates only from magnetization changes in the wall. The relevant torques for this case are illustrated in Fig. 22b for the example of a 180◦ Bloch wall, which is reduced to its central vector like in Fig. 21b. The adiabatic torque τAD on this moment, together with the resulting demagnetizing fields, leads to a Néel-like deformation of the wall structure and to a downward torque on the central vector that effectively compensates τAD . This compensation of torques prevents wall motion up to a critical velocity (i.e., current density) uc = γ Δ0w Ha /2 that is proportional to the anisotropy field Ha . Up to uc , the adiabatic torque thus acts like an applied magnetic field transverse to the wall plane that just distorts the wall without inducing a steady motion. Above the threshold, the internal torque is not sufficient anymore to balance the spin transfer torque, and the wall is set in motion together with a continuous precession of the wall magnetization.

    AD

    m

    H d,1

    H d,1 m

    Hd m

    H d,2 m

    0 a)

    uc

    u

    b)

    Hd NA

    c)

    Fig. 22 (a) Average wall velocity as a function of the effective drift velocity u of the spin current, schematically plotted for three different non-adiabaticity coefficients β (Adapted from Ref. [35]). The adiabatic torque leads to wall distortion and motion (b), while the non-adiabatic torque (c) just moves the wall. Indicated are the primary STT vectors, τAD and τNA , together the following torques due to the generated demagnetizing fields [37]

    428

    R. Schäfer

    The non-adiabatic torque (Fig. 22c), on the other hand, causes a demagnetizing field with a consequent downward torque that sets the wall in steady motion right away by mimicking an easy axis field (compare Fig. 21b). According to Fig. 22a this is true for any DC applied current if β = α, and it is true for currents smaller than a critical current if β > α. In the steady-wall motion regime, the non-adiabatic torque together with the damping constant α dictates the velocity v = βu/α. It is obtained by replacing Hext in (18) by the spin transfer equivalent field βu/(γ Δw ) [29]. Beyond the critical current, the adiabatic term dominates, and the wall slows down (v = u) due to the onset of precession and backward motion accompanied by periodic transformations between Bloch and Néel wall. The critical current thus corresponds to the Walker breakdown of the field-driven dynamics (compare Fig. 21d). So the non-adiabatic spin transfer plays a crucial role in current-induced wall motion: it allows for a non-zero mobility of the wall in a steady-state, viscousflow regime with the mobility proportional to β/α, resulting in the elimination of the intrinsic threshold current even for small values of β. Any threshold current is then only determined by extrinsic pinning. Compared to relatively wide domain walls in in-plane magnetized nanowires (see Fig. 19), which can readily expand to many times their equilibrium size under current torques [38], the domain walls are of Bloch-type in nanowires with an anisotropy perpendicular to the plane of the wire. Well studied are atomically engineered cobalt-nickel superlattices in which interface anisotropies dominate over the volume terms (anisotropy and magnetostatic), thus leading to the perpendicular easy axis. With typical wall widths of some nanometers, the magnetization gradient is large, and consequently a high non-adiabatic effect is to be expected. This makes such materials attractive for current-induced domain wall motion as it pushes the walls like the torque of a magnetic field with an increased STT efficiency compared to in-plane domain walls. Wall velocities of the order of 100 m/s were measured in such films, comparable to the speeds found for field-driven wall motion in bubble material. The spin transfer torques, described so far, are volume effects. Even higher current-induced wall velocities (before Walker breakdown) in perpendicularanisotropy media can be achieved by making use of interface-related spin torque effects that are obtained by depositing the magnetic stripes adjacent to a nonmagnetic conductive layer with strong spin-orbit interaction (SO layer) like heavy metal platinum, palladium, iridium, or tungsten films (see refs. [39, 38] for reviews). The following current-induced SO-torques may contribute: • Spin Hall effect (SHE): If a charge current passes the SO layer, the SHE creates a spin current due to spin-orbit scattering of the conduction electrons. The spin current flows in a direction perpendicular to the charge current toward each surface of the SO layer. It may thus be seen as pure spin current, i.e., without any charge flow in the perpendicular direction. The direction of the spin is perpendicular to both the charge and spin currents as illustrated in Fig. 23a. The interfacial spin accumulation is injected into the adjacent magnetic layer where it exerts a STT [39]

    8 Magnetic Domains No

    429 Heff

    SH

    Heff

    Bloch wall

    SH

    Heff

    Heff

    Néel wall

    tion

    b)

    a)

    Wall mo

    Fig. 23 (a) Illustration of the Spin Hall Effect and its torque τSH . Indicated are the perpendicular domain magnetizations and the central vectors for both rotation senses of a Bloch- (upper-) and Néel wall (lower panel). (b) The Spin-Hall torque on a Néel wall leads to an effective magnetic field that depends on the wall rotation sense, leading to a wall motion along or opposite to the charge current direction. For each sketch, the spin-orbit layer is indicated as bottom layer, being in contact with the ferromagnetic film on top

    τSH ∼ m × σ × m .

    (24)

    Here σ = j × z with j and z being the unit vectors of current and out-of-plane direction, respectively. The spin-Hall torque τSH manipulates the magnetization in the magnetic film as shown for domain walls in Fig. 23a. For Bloch walls of either rotation sense, the torque vanishes, while for Néel walls a transverse torque acts on the wall spins that points in the same direction (along the injected electron spin direction) for both rotation senses of the wall magnetization. The resulting demagnetizing fields finally lead to an effective magnetic field acting on a Néel wall. As illustrated in Fig. 23b, this field is opposite for Néel walls with opposite rotation senses. Consequently, neighboring walls will be pushed in the same direction if they have the same chirality, i.e., opposite directions of their central vectors. Such homochiral Néel walls can be enforced in ultrathin multilayers with interface-induced Dzyaloshinskii-Moriya interaction; see section “Domain Wall Types” and Fig. 19a. • Rashba effect: This spin-orbit-related phenomenon, reviewed in Ref. [40], may be induced by an electric field that is mediated by the symmetry breaking at the interfaces of a typical multilayer composed of heavy-metal/ferromagnetic metal/oxide films and that is oriented perpendicular to the film plane. When an inplane charge current flows in the magnetic layer, the itinerant electrons transform the electrical field into a (Rashba) in-plane magnetic field that is experienced by the flowing electrons and that polarizes them. By exchange interaction they can then exert a field-like SST on the local magnetization of the magnetic film which may result in some pressure on the domain walls [41, 39]. In contrast to the Spin Hall effect, which is a bulk effect, the Rashba effect is an interface property. The direction of domain wall motion and their velocity are determined by the subtle interplay of these phenomena, which can be tuned independently by

    430

    R. Schäfer

    varying the thickness of the SO layers and the composition of the ferromagnetic layer. Current-induced domain wall motion with velocities of several 100 m/s can be achieved by making use of SO effects. Still higher wall velocities of almost 1.000 m/s were measured in nanowires that are formed from synthetic antiferromagnets, i.e., from two antiferromagnetically coupled films via an ultrathin nonmagnetic spacer layer [42, 38]. The high velocity is attributed to the vanishing net magnetic moment in such film systems. Besides the so far discussed nanowires with the current flowing in the plane, STT [34] manifests itself also in sub-micrometer diameter pillars, composed of magnetic multilayers and flooded by a current perpendicular to the plane. Such pillars are the basic units of the SST-MRAM (magnetic random-access memory) concept. Magnetically, they can be well described in the framework of a macrospin model, meaning that the magnetic films of the pillar are assumed to move in a singledomain state rather than by short-wavelength modes. As domains thus do not play a role, we will touch this case just briefly here, leaving details to specialized literature (see refs. [43, 44] for reviews). Assume a ferromagnet/normal-metal/ferromagnet trilayer system as illustrated in Fig. 24a, b. The two ferromagnetic layers of this spin valve-type device are depicted as “fixed” and “free,” indicating that they are less and more susceptible, respectively, to the STT. If unpolarized electrons are entering the fixed layer from the left (Fig. 24a), they will undergo spin-filtering, i.e., by exchange interaction with the magnetic moment of the fixed layer, the flowing electrons have an averaged spin moment parallel to the magnetization of the fixed layer when emerging into the nonmagnetic spacer film. If the spacer is thinner than the spin diffusion length, this current will remain spin-polarized when it enters the free layer, in which the spin components of the incoming electrons transverse to its magnetization are absorbed. As the spin angular momentum is conserved, this lost transverse momentum of the electrical current must be absorbed by the magnetic film, leading to a torque that

    ted smit Tran ctrons ele Fixed layer

    m

    Free layer

    d itte nsmrons a r T lect e

    ted flecrons e R ct ele

    ele

    a)

    I< I>

    ns ctro

    b)

    Heff , m

    Spin transfer torques for:

    0

    0

    m free

    c)

    Fig. 24 (a, b) Illustration of the STT in a spin-valve-type trilayer, leading to the switching of the free layer magnetization either parallel (a) or antiparallel (b) to that of the fixed layer (starting from an assumed perpendicular initial configuration, with the thickness of the nonmagnetic spacer layer exaggerated for visualization purpose). The magnetization of the fixed layer is stabilized by making it thicker than the free layer, by using a material with a larger total moment or by exchange coupling to an antiferromagnetic layer. (c) Schematics of the STT in analogy to Fig. 20

    8 Magnetic Domains

    431

    tends to turn the free layer magnetization toward the orientation of the incident spin polarization. So the parallel alignment of magnetization in the fixed and free layer is stabilized. This STT requires that the moments of the two layers are initially non-collinear, perhaps due to thermal fluctuations (exact parallel or antiparallel alignment would not lead to torques). If the electron flow is opposite (Fig. 24b), the electrons will first undergo spin filtering by the free layer. They will then flow to the fixed layer and apply a torque to that layer. This torque, however, is inefficient as the fixed layer magnetization is held rigidly in place. However, a fraction of the electrons will be reflected back from the interface between normal metal and fixed layer toward the free layer. These reflected electrons have an averaged spin polarization antiparallel to the fixed layer magnetization (the parallel spins are readily transmitted), i.e., opposite to the case discussed before. The free layer now feels a torque that turns the free layer moment away from the fixed layer moment, thus destabilizing the parallel alignment of magnetization and possibly leading to a reversal of the free layer magnetization if the current is sufficiently strong. The fixed layer consequently serves as an electron polarizer by providing the polarized electrons that finally act on the free layer for both current directions. This currentinduced magnetization switching between parallel and antiparallel magnetization of free and fixed layer provides a smart alternative to magnetization switching by magnetic field and is applied in the STT-MRAM for the writing process. The dynamics of the free layer in the spin-valve pillar is phenomenologically described by the LLGS equation (16), in which the τ -term reads τ=

    dmfree I = g(θ ) mfree × (mfree × mfixed ). dt A

    (25)

    The Slonczewski term, g(θ ), represents the material-dependent spin-transfer efficiency with θ being the angle between the magnetization directions of free and fixed layers. The associated torques are illustrated in Fig. 24c. For simplicity, a uniaxial anisotropy along the z-axis is assumed (defining Heff ) with the magnetization of the fixed layer pointing along the same direction. In the absence of STT, the magnetization of the free layer, instantaneously pointing at an angle with respect to z, spirals toward the anisotropy axis, driven by the precessional and damping torques (compare Fig. 20). In the presence of STT, the additional Slonczewski torque (25) can point either in the same direction as the damping torque or opposite to it, depending on the direction of current (compare Figs. 20 and 24c). So the STT can either reinforce the damping torque, making the free layer to relax toward the easy axis faster than without, or it can act against the damping so that free layer magnetization relaxes more slowly. If, in the latter case, the current density exceeds a critical value, the STT can be larger in magnitude than the damping torque, so that the free layer begins to spiral away from the z-axis rather than relaxing toward this direction. If the energy, added to the free layer by the STT, is high enough, the free layer will finally spiral to θ = π , where it can remain in a stable state antiparallel to the fixed layer. Depending on the detailed angular dependence of spin transfer and damping torques, it is also possible that the energies lost to damping and

    432

    R. Schäfer

    gained from the STT are balanced over each precessional cycle. Some dynamical equilibrium may then be reached at an intermediate angle θ , meaning that steadystate precessional oscillations in the microwave frequency range are generated [45]. The precession, which is basically caused by a DC applied current, occurs with large precession angles that are not accessible using magnetic field excitation alone and which are the basis for the so-called spin-transfer oscillators that are envisaged for applications in communication technology.

    References 1. Hubert, A., Schäfer, R.: Magnetic Domains. The Analysis of Magnetic Microstructures. Springer, Berlin/Heidelberg (1998) 2. Kirilyuk, A., Ferré, J., Grolier, V., Jamet, J.P., Renard, D.: Magnetization reversal in ultrathin ferromagnetic films with perpendicular anisotropy. J. Magn. Magn. Mat. 171, 45–63 (1997) 3. Ma, L., Gilbert, D.A., Neu, V., Schäfer, R., Zheng, J.G., Yan, X.Q., Shi, Z., Liu, K., Zhou, S.M.: Magnetization reversal in perpendicularly magnetized L10 FePd/FePt heterostructures. J. Appl. Phys. 116(3), 033922 (2014) 4. He, P., Ma, X., Zhang, J.W., Zhao, H.B., Lüpke, G., Shi, Z., Zhou, S.M.: Quadratic scaling of intrinsic Gilbert damping with spin-orbital coupling in L10 FePdPt films: experiments and Ab Initio Calculations. Phys. Rev. Lett. 110, 077203 (2013) 5. Lopatina, E., Soldatov, I., Budinsky, V., Marsilius, M., Schultz, L., Herzer, G., Schäfer, R.: Surface crystallization and magnetic properties of Fe84.3 Cu0.7 Si4 B8 P3 soft magnetic ribbons. Acta Mat. 96, 10–17 (2015) 6. McCord, J., Schäfer, R.: Domain wall asymmetries in Ni81 Fe19 /NiO: proof of variable anisotropies in exchange bias systems. New J. Phys. 11, 83016 (2009) 7. Flohrer, S., Schäfer, R., McCord, J., Roth, S., Schultz, L., Herzer, G.: Magnetization loss and domain refinement in nanocrystalline tape wound cores. Acta Mat. 54 (12) 3253–3259 (2006) 8. Parkin, S.S.P., Hayashi, M., Thomas, L.: Magnetic domain-wall racetrack memory. Science 320, 190–194 (2008) 9. Landau, L.D., Lifshitz, E.: On the theory of the dispersion of magnetic permeability in ferromagnetic bodies. Phys. Z. Sowjetunion 8, 153–169 (1935) 10. Dzyaloshinskii, I.E.: Theory of helicoidal structures in antiferromagnets. I. Nonmetals. Sov. Phys. JETP 19, 960–971 (1964) 11. Bogdanov, A.N., Yablonsky, D.A.: Thermodynamically stable "vortices" in magnetically ordered crystals. The mixed state of magnets. Zh. Eksp. Teor. Fiz 95(1), 178–182 (1989) 12. Bogdanov, A.N., Hubert, A.: Thermodynamically stable magnetic vortex states in magnetic crystals. J. Magn. Magn. Mater. 138, 255–269 (1994) 13. Yu, X.Z., Onose, Y., Kanazawa, N., Park, J.H., Han, J.H., Matsui, Y., Nagaosa, N., Tokura, Y.: Real-space observation of a two-dimensional skyrmion crystal. Nature (London) 465, 901–904 (2010) 14. Romming, N., Hanneken, C., Menzel, M., Bickel, J.E., Wolter, B., von Bergmann, K., Kubetzka, A., Wiesendanger, R.: Writing and deleting single magnetic skyrmions. Science 341, 636–639 (2013) 15. Bode, M., Heide, M., von Bergmann, K., Ferriani, P., Heinze, S., Bihlmayer, G., Kubetzka, A., Pietzsch, O., Blügel, S., Wiesendanger, R.: Chiral magnetic order at surfaces driven by inversion asymmetry, Nature 447, 190–193 (2007) 16. Leonov, A.O., Monchesky, T.L., Romming, N., Kubetzka, A., Bogdanov, A.N., Wiesendanger, R.: The properties of isolated chiral skyrmions in thin magnetic films. New J. Phys. 18, 065003 (2016) 17. Mühlbauer, S., Binz, B., Jonietz, F., Pfleiderer, C., Rosch, A., Neubauer, A., Georgii, R., Böni, P.: Skyrmion lattice in a chiral magnet. Science 323, 915–919 (2009)

    8 Magnetic Domains

    433

    18. Kiselev, N.S., Bogdanov, A.N., Schäfer, R., Rössler, U.K.: Chiral skyrmions in thin magnetic films: new objects for magnetic storage technologies? J. Phys. D: Appl. Phys. 44, 392001 (2011) 19. Bogdanov, A.: Private communication 20. Hilzinger, R., Rodewald, W.: Magnetic Materials. Publicis Publishing, Erlangen (2013) 21. Herzer, G.: Private communication 22. Soldatov, I., Schäfer, R.: Selective sensitivity in Kerr microscopy. Rev. Sci. Instrum. 88, 073701 (2017) 23. Schäfer, R.: The Magnetic Microstructure of Nanostructured Materials. In: Liu, J., Fullerton, E., Gutfleisch, O., Sellmyer, D. (eds) Nanoscale Magnetic Materials and Applications. Springer, Boston (2009) 24. Malozemoff, A.P., Slonczewski, J.C.: Magnetic Domain Walls in Bubble Materials. Academic Press, New York (1979) 25. Döring, L., Hengst, C., Otto, F., Schäfer, R.: Interacting tails of asymmetric domain walls: theory and experiments. Phys. Rev. B 93, 024414 (2016) 26. Thiaville, A., Rohart, S., Jué, É., Cros, V., Fert, A.: Dynamics of Dzyaloshinskii domain walls in ultrathin magnetic films. Europhys. Lett. 100 (5), 57002 (2012) 27. Thiaville, A., Nakatani, Y.: In: Shinjo, T. (eds.) Nanomagnetism and spintronics, p. 231. Elsevier (2009) 28. Thiaville, A., Nakatani, Y.: In: Hillebrands, B., Thiaville, A. (eds.) Spin dynamics in confined magnetic structures III, p. 161. Springer, Berlin (2006) 29. Mougin, A., Cormier, M., Adam, J.P., Metaxas, P.J., Ferré, J.: Domain wall mobility, stability and Walker breakdown in magnetic nanowires. Europhys. Lett. 78 (5), 57007 (2007) 30. Hubert, A.: Theorie der Domänenwände in geordneten Medien. Springer, Berlin/Heidelberg/New York (1974) 31. Filippov, B.N.: Static properties and nonlinear dynamics of domain walls with a vortexlike internal structure in magnetic films (Review). Low Temp. Phys. 28, 707–738 (2002) 32. Berger, L.: Exchange interaction between ferromagnetic domain wall and electric current in very thin metallic films. J. Appl. Phys. 55, 1954–1956 (1984) 33. Berger, L.: Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev. B 54, 9353 (1996) 34. Slonczewski, J.C.: Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mat. 159, L1–L7 (1996) 35. Thiaville, A., Nakatani, Y., Miltat, J., Suzuki, Y.: Micromagnetic understanding of currentdriven domain wall motion in patterned nanowires. Europhys. Lett. 69, 990–996 (2005) 36. Zhang, S., Li, Z.: Roles of nonequilibrium conduction electrons on the magnetization dynamics of ferromagnets. Phys. Rev. Lett. 93, 127204 (2004) 37. Thanks to T. Moore (Leeds) and J. Augusto (IFW) for straightening the vectors 38. Parkin, S., Yang, S.-H.: Memory on the racetrack. Nat. Nanotechnol. 10, 195 (2015) 39. Khvalkovskiy, A.V., Cros, V., Apalkov, D., Nikitin, V., Krounbi, M., Zvezdin, K.A., Anane, A., Grollier, J., Fert, A.: Matching domain-wall configuration and spin-orbit torques for efficient domain-wall motion. Phys. Rev. B 87, 020402 (2013) 40. Manchon, A., Koo, H.C., Nitta, J., Frolov, S.M., Duine, R.A.: New perspectives for Rashba spin-orbit coupling. Nat. Mater. 14, 871–882 (2015) 41. Miron, I.M., Moore, T., Szambolics, H., Buda-Prejbeanu, L.D., Auffre, S., Rodmacq, B., Pizzini, S., Vogel, J., Bonfim, M., Schuhl, A., Gaudin, G.: Fast current-induced domain-wall motion controlled by the Rashba effect. Nat. Mater. 10, 419–423 (2011) 42. Yang, S.-H., Ryu, K.-S., Parkin, S.: Domain-wall velocities of up to 750 m s−1 driven by exchange-coupling torque in synthetic antiferromagnets. Nat. Nanotechnol. 10, 221–226 (2015) 43. Ralph, D.C., Buhrman, R.: In: Maekawa, S. (ed.) Concepts in spin electronics. Oxford University Press, Oxford (2006) 44. Stiles, M.D., Miltat, J.: In: Hillebrands, B., Thiaville, A. (eds.) Spin dynamics in confined magnetic structures III. Springer-Verlag, Berlin, Heidelberg (2006)

    434

    R. Schäfer

    45. Kiselev, S.I., Sankey, J.C., Krivorotov, I.N., Emley, N.C., Schoelkopf, R.J., Buhrman, R.A., Ralph, D.C.: Microwave oscillations of a nanomagnet driven by a spin-polarized current. Nature 425, 380–383 (2003) Rudolf Schäfer studied Materials Science and received his Ph.D. in Engineering at Erlangen-University in 1990. After Postdoc stays at IBM Research in Yorktown Heights and Forschungszentrum Juelich he moved to IFW Dresden in 1993. His interest areas span magnetic materials with focus on magnetic microstructures.

    9

    Magnetotransport Michael Ziese

    Contents Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Classical Magnetoresistance in Semiconductors and Semimetals . . . . . . . . . . . . . . . . . . . . . . Magnetotransport and Ferromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Anisotropic Magnetoresistance, Planar Hall Effect, and Two-Current Model . . . . . . . . . . . Giant Magnetoresistance (GMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Colossal Magnetoresistance (CMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Tunneling Magnetoresistance (TMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Powder Magnetoresistance (PMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Organic Magnetoresistance (OMAR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantum Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Exotic Tunneling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inverse Spin Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantum Spin Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetoimpedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    436 440 442 442 446 448 451 454 454 454 456 458 460 461 462 462 464 464 467

    Abstract

    In this chapter, magnetotransport phenomena in ferromagnetic materials and heterostructures are discussed. The survey starts with the definition of the electrical resistivity and a discussion of magnetoresistance in normal met-

    M. Ziese () Fakultät für Physik und Geowissenschaften, Universität Leipzig, Leipzig, 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_9

    435

    436

    M. Ziese

    als, semiconductors, and semimetals. This is followed by a description of magnetoresistance processes in ferromagnets: from anisotropic magnetoresistance and critical scattering in bulk ferromagnets to giant magnetoresistance in magnetic multilayers, colossal magnetoresistance in manganites, tunneling magnetoresistance in trilayers with insulating barrier, and to powder and organic magnetoresistance. The two-current model features prominently in the analysis of the majority of these effects. Sections on quantum transport and exotic tunneling phenomena follow. Then transverse electric phenomena are presented: the discussion of the normal and anomalous Hall effect leads to the introduction of spin currents and the spin Hall effects. The chapter concludes with a brief section on magnetoimpedance and an explanation of measurement techniques.

    Basics When a static electric field E is applied to a material, it might respond with a current flow – then it is called an electric conductor – or with a static polarization P , in which case it is called an insulator. Conductors are characterized by Ohm’s law [1]: When all external parameters (temperature, pressure, magnetic field . . . ) are held constant, the current density j flowing through a conductor is proportional to the  applied electric field E: j = σ E .

    (1)

    σ is a tensor of second rank known as the conductivity tensor; it is a material property with units S/m. The inverse tensor ρ = σ−1 is the resistivity tensor with units  m. When a magnetic field H is applied, the conductivity tensor obeys Onsager’s relation [2]: σij (H ) = σj i (−H ) .

    (2)

    The diagonal components are even under magnetic field inversion; the off-diagonal components might have an even as well as an odd contribution. In case of isotropic systems and cubic systems, all three diagonal components are equal and are simply called conductivity σ . In this chapter, the material systems are restricted to solids; according to their conductivity values, conductors might be quantitatively further divided into metals (good conductors) and semiconductors. However, it is more appropriate to use the temperature coefficient of resistivity (TCR) dρ/dT to qualitatively distinguish metals (positive TCR) from semiconductors (negative TCR). More than two thirds of the elements are metals; see Fig. 1. If contributions from ionic motion are neglected, the current flow is carried by electrons that are driven through the material by the applied electric field. In a crystalline material, the electrons are found in Bloch states characterized by a  the wave vector, or h¯ k,  the crystal momentum. The electronic quantum number k,   being periodic in the energy En (k) is a function of band index n and wave vector k,

    9 Magnetotransport

    437

    Fig. 1 Left panel: Room temperature resistivities of the elements as a function of order number. Resistivities diverge toward the r.h.s. of the periodic table (C (graphite), Si, Ge, Te). Certain features recur periodically, such as the resistivity minima of the Cu group elements, the local maxima of the Ti group elements, as well as a maximum at half-filling of the d-shell (Mn, Tc, Re). Resistivity values from [3]. Right panel: Resistivity of Fe1−x Nix alloys at various temperatures [4]. The resistivity increase of the alloy is clearly seen, especially at low temperatures. The situation is further complicated by the martensite-austenite transition that occurs along the alloying route from iron (bcc) to nickel (fcc)

     reciprocal lattice and thus forming energy bands [5]. The electron velocity vn (k) does in general not vanish, i.e., electrons move through periodic lattices without scattering. Electron scattering is due to imperfections in the crystal lattice, either by static imperfections (lattice defects) or by dynamic imperfections such as phonons or magnons. Within the Drude model , the scattering is characterized by a relaxation time τ . The conductivity can then be expressed by the carrier density n, the effective electron mass m∗ , the relaxation time τ , and the electron charge e as [6] σ =

    e2 nτ . m∗

    (3)

    When various scattering mechanisms act in the same conductor, in a first approximation, the resistivities add up: ρ = ρi + ρp + . . . (Matthiessen’s rule, [7]). ρi denotes the scattering by impurities that is mainly temperature independent, and ρp denotes the scattering by phonons that is proportional to T 5 (Bloch T 5 law) at low temperatures, to T 3 for umklapp processes, and to T at high temperatures [5]. The effect of electron-electron scattering is rather small leading to a term proportional to T 2 . As a simple rule, the resistivity of the non-ferromagnetic elements at room temperature is proportional to the temperature, since electron-phonon scattering dominates, whereas alloys have a nearly temperature-independent resistivity due to

    438

    M. Ziese

    the dominance of impurity scattering. This is illustrated in the right panel of Fig. 1 showing the resistivity of Fe1−x Nix alloys for various temperatures. A well-established model describing electron transport in the limit of weak scattering is the semiclassical model of electron dynamics. This builds on the energy  According to Hamilton’s equations, the velocity is related to band structure En (k). the momentum gradient of the Hamiltonian [5]:  = 1 ∂En /∂ k . vn (k) h¯

    (4)

    The time derivative of the crystal momentum is given by the Lorentz force [5]:    × B(  r , t) .  r , t) + vn (k) h¯ k˙ = −e E(

    (5)

    This model was very successful in explaining a wealth of phenomena, especially the distinction between metals, semiconductors, and insulators as systems with or without overlapping energy bands, the values of the carrier density, the occurrence of holes as charge carriers, and magnetoresistance effects. In an applied magnetic  both the electronic energy and the wave-vector component along B induction B, are constant, i.e., the electrons move on constant energy orbits perpendicular to the magnetic field. If an electric field E is applied perpendicular to the magnetic  2 is superimposed on the cyclic induction, a drift motion with velocity E × B/B motion. In the limit of high fields, i.e., when the cyclotron frequency ωc = eB/m∗ is large compared to the scattering rate τ −1 , the E × B drift dominates and leads to a Hall current; see section “Hall Effect ”, if the cyclic orbits are closed, i.e., restricted in space. In this case the magnetoresistance saturates; in the case of open orbits extending between adjacent Brillouin zones, the magnetoresistance increases with increasing magnetic induction without bounds. The electronic motion in a magnetic field leads to a variety of effects (Shubnikov-de Haas effect in case of magnetoresistance) being periodic in 1/B with period Δ

      1 2π e = , B hA(E ¯ F)

    (6)

    where A(EF ) denotes the cross section of an extremal Fermi surface perpendicular to the applied magnetic field. These effects have proven viable tools for Fermi surface studies [5]. If the mean free path = vF τ , where vF denotes the Fermi velocity, is of the order of the lattice spacing a, the Mott-Ioffe-Regel limit = a is reached and the metal approaches insulating behaviour [9]. In the limit of strong disorder, in an alloy or an amorphous material, the electronic wave function may become localized [10], such that a crossover or transition to an insulating state with variable range hoppingdominated conductivity occurs [9]. Analyzing the formation of extended wave functions from the atomic orbitals, Mott [9] introduced the concept of a mobility edge that separates localized states in the tails of the energy bands from extended

    9 Magnetotransport

    439

    states within the energy band centers. In certain compounds, this limit is actually accompanied by a resistivity saturation [11,12], but this is not a universal signature, and the resistivity of other metals just crosses this limit without any features [13]. Electron scattering in the limit of strong disorder is determined by various scattering length scales, and especially the phase coherence length can be much larger than the mean free path. This leads to the possibility that an electron is strongly scattered and retraces its original trajectory coherently, a process known as weak localization [14]. Weak localization leads to corrections to the resistivity being logarithmic in temperature in two-dimensional and proportional to T 1/2 in three-dimensional systems. Furthermore, a magnetic field threading the weakly localized trajectories destroys the phase coherence and leads to a negative magnetoresistance [14]. Spin-orbit coupling has a similar effect but leads to a positive magnetoresistance contribution, see, e.g., the magnetoresistance of Au in Mg [15]. After discussing some of the basics of magnetotransport in metals, one might ask what the specifics of magnetotransport in ferromagnets are. This is nicely highlighted by a set of classic data from Gerlach and Schneiderhan [8]; see Fig. 2. In ferromagnets, electron-magnon scattering is large and contributes another term

    Fig. 2 Resistance and magnetoresistance of a nickel wire; magnetic field and current density were applied along the wire axis. (a) The resistivity of ferromagnetic metals has a slope change at the Curie temperature (615 K in this case, 633 K for pure Ni). (b) The magnetoresistance slope, and therefore also the magnetoresistance in large magnetic fields has a maximum at the Curie temperature due to critical scattering. (c) The magnetoresistance below the Curie temperature shows an anisotropic contribution at low fields and a linear contribution from critical scattering at high fields. (d) Above the Curie temperature, only the linear MR remains. The numbers indicate the measurement temperatures. (After [8])

    440

    M. Ziese

    ρm ∝ T 3/2 to Matthiesen’s rule [16]; see also section “Magnetotransport and Ferromagnetism”. Above the Curie temperature, magnon scattering diminishes, thus often leading to a slope change in the resistivity of ferromagnets; see Fig. 2a. The magnetoresistance measured as a function of magnetic field shows two characteristics: (1) below the Curie temperature TC , see Fig. 2c; the magnetoresistance rises sharply in small fields and decreases linearly in higher field; and (2) above TC , only the linear magnetoresistance remains having a maximum slope at the Curie temperature. The sharp magnetoresistance rise, known as anisotropic magnetoresistance (AMR), is due to a reorientation of the magnetization and is further discussed in section “Magnetotransport and Ferromagnetism”. The linear magnetoresistance is due to magnon scattering and extends to very high fields [17]. Figure 2b shows the slope ∂(ΔR/R)/∂(μ0 H ) as a function of temperature; a clear maximum is observed at the Curie temperature that is due to critical scattering.

    Classical Magnetoresistance in Semiconductors and Semimetals In the simplest approximation, the free electron model, magnetoresistance is absent, but the electrons show a Hall effect ; see section “Hall Effect ”. Electron motion is perpendicular to the magnetic field and can be described by a two-dimensional resistivity or conductivity tensor  ρ=ρ

    1 −ωc τ ωc τ 1

     σ =

    1 ρ(1 + ωc2 τ 2 )

    

    1 ωc τ −ωc τ 1

     .

    (7)

    ρ denotes the longitudinal resistivity; the Hall constant is RH = ρωc τ/B. Note that the diagonal components of the conductivity depend on the magnetic field. If more than one band contributes to the electron transport, the total conductivity tensor is obtained as the sum over the conductivity tensors of all bands. In case of two bands, this yields the following expressions for the longitudinal resistivity ρ and Hall constant R (two-band model, the indices refer to the respective bands): ρ=

    ρ1 ρ2 (ρ1 + ρ2 ) + (ρ1 R22 + ρ2 R12 )B 2 (ρ1 + ρ2 )2 + (R1 + R2 )2 B 2

    (8)

    R=

    R1 ρ22 + R2 ρ12 + R1 R2 (R1 + R2 )B 2 . (ρ1 + ρ2 )2 + (R1 + R2 )2 B 2

    (9)

    These expressions highlight certain features of the classical magnetoresistance. In small magnetic fields the resistivity is proportional to B 2 , whereas it saturates in large fields. The Hall constant is given by a weighted average in small fields and saturates to R → 1/[e(±n1 ± n2 )] in high fields, where the ±-sign indicates either hole or electron conduction. We further see that the strength of the magnetoresistance is determined by ωc τ . Since the relaxation time τ is inversely proportional to the zero field resistivity, a scaling of the magnetoresistance

    9 Magnetotransport

    441

    with B/ρ is expected, which is known as Kohler’s rule [20]. A nonsaturating magnetoresistance is obtained in compensated materials with n1 +n2 = 0. These equations form the basis for a magnetotransport analysis of semiconductors. They were also used to describe the behavior of compensated metals such as Bismuth or zero-gap semiconductors such as graphite. In graphite and Bi, the magnetoresistance is so strong that a magnetic field-induced metal-insulator transition occurs, i.e., the resistivity at low temperatures is larger than at high temperatures, since ωc τ is very large at low temperatures; see Fig. 3. The resistivity and magnetoresistance can be fairly well understood within a three-band model; see Fig. 3a and b [18]. This is somewhat surprising, since the magnetoresistance in general depends on the form of the Fermi surface as well as the dominant scattering mechanism [21]. Further, the semiclassical model is only an approximation and does not provide a full quantum-mechanical description of the magnetotransport. The observation of a magnetoresistance linear in magnetic field, e.g., in Ag2+δ Se [22] and in graphite [23] or multilayer graphene [24], calls for a full quantum-mechanical description [25]. Early quantum theories of galvanomagnetic effects found a scaling of the magnetoresistance in the quantum limit with H p /T q , where the exponents p and q depend on the scattering

    Fig. 3 In-plane resistivity of graphite as a function of (a) temperature at various fields up to 200 mT and (b) magnetic field at various temperatures up to 200 K. The symbols are data and the solid lines fit to a three-band model (After [18]). (c) Kohler plot for the in-plane magnetoresistance of graphite in magnetic fields up to 200 mT. Obviously, Kohler’s rule is violated in graphite. (d) The same magnetoresistance data as in (c) but this time plotted as a function of H /T . Scaling is found for temperatures between 20 K and at least 270 K (After [19])

    442

    M. Ziese

    mechanism [26, 27]. Although the problem of quantum magnetoresistance is far from solved, it is clear that the magnetoresistance of graphite does not obey Kohler’s rule but comes close to a H /T -scaling; see Fig. 3c and d.

    Magnetotransport and Ferromagnetism Anisotropic Magnetoresistance, Planar Hall Effect, and Two-Current Model Spin disorder scattering near the Curie temperature was discussed in section “Basics”. Below the Curie temperature, the magnetoresistance of a ferromagnet depends on the relative direction between the electric current density j and the  [28, 16, 29]. This constitutes the anisotropic magnetoresistance magnetization M (AMR ). Let ρ and ρ⊥ denote the resistivities (extrapolated to zero induction B)  and transverse geometry (j ⊥ M),  respectively; then the in longitudinal (j  M) anisotropic magnetoresistance is defined as Δρ = ρ

    ρ 1 3 ρ

    − ρ⊥ + 23 ρ⊥

    .

    (10)

    In case of a polycrystalline or amorphous ferromagnet, the resistivity depends only on the angle Θ between magnetization and current density, but not on the crystal directions; this gives rise to an angle-dependent longitudinal resistivity ρlong = ρ⊥ + (ρ − ρ⊥ ) cos2 Θ

    (11)

    and a transverse resistivity ρtrans = −(ρ − ρ⊥ ) sin Θ cos Θ .

    (12)

    In case of single crystalline materials, the angle dependence might be much more complicated [29, 30, 31, 32]. As an illustration, Fig. 4 shows the anisotropic magnetoresistance of a single crystalline SrRuO3 film as a function of angle between the applied field of 8 T and the crystallographic axes. The complex angle dependence is obvious. Since the magnetic anisotropy of SrRuO3 is rather strong, the angular dependence is even further complicated, since the magnetization is not always along the magnetic field direction; note especially the sharp magnetoresistance change at about 30◦ from the [110] axis in the (001) plane, which is due to the crossing of a hard magnetic axis [33]. The transverse resistivity is known as planar Hall effect . Due to the absence of any offset in Eq. (12), this is sometimes advocated to be a much more sensitive probe than the AMR. Especially in case of ferromagnets with rather large resistivity,

    9 Magnetotransport

    443

    Fig. 4 Angle dependence of the AMR in a single crystal SrRuO3 film. The current density was along the [001] orthorhombic axis; the magnetic field was rotated in the (a) (110) and (b) (001) plane. Some crystallographic directions are indicated (After [33])

    the planar Hall effect was dubbed giant [34, 35]. The planar Hall effect of La0.84 Sr0.16 MnO3 is shown in Fig. 5. The basis for the understanding of the AMR is Mott’s two-current model [36, 37]. This in turn is based on the s − d model: the band structure of a 3d transition metal might be understood as being composed of a broad free-electronlike s-band superimposed on a narrow tight-binding-like d-band. The comparatively high resistivity of the transition metals might be understood by scattering of mobile s-electrons into d-states, where the latter have a considerable density of states at the Fermi level. Spin-flip scattering events are rare, at least at low temperature, since scattering occurs predominantly at non-magnetic scattering centers. Therefore, electron transport can be viewed as a parallel circuit of spin-up and spin-down conduction channels; see Fig. 6; in each channel, the resistivities ρ ↑ and ρ ↓ consist in general of a small ss- and a large sd-scattering contribution. The spin-flip−1 scattering rate τ↑↓ is negligible at low temperatures. In case of strong ferromagnets, the majority (or spin-up band) is completely filled, thus eliminating the sd-scattering of spin-up electrons and short-circuiting carrier transport along the spin-up channel. Historically, this model was motivated by a resistivity drop below the Curie temperature; see Fig. 6c; note, however, data and discussion in [16]. For a measurement of ρ↓ and ρ↑ , the scattering center has to be defined. This is usually done by alloying a ferromagnetic host metal with a few atomic percent of particular metal impurities. The channel resistivities can then be obtained from deviations of the measured data from Matthiessen’s rule either as a function of temperature or in ternary alloys [16]. Figure 7 shows the channel resistivities as well

    444

    M. Ziese

    Fig. 5 (a) AMR and (b) and (c) planar Hall effect of La0.84 Sr0.16 MnO3 at 120 K. The lines in (a) and (b) are fits of Eqs. (11) and (12) to the data using (a) R − R⊥ = −29 Ω and (b) R − R⊥ = −88 Ω. (After [35])

    Fig. 6 Mott’s two-current model: (a) spin-up and spin-down conduction channels weakly coupled by spin-flip scattering. (b) Short-circuiting by the spin-up channel in case of a strong ferromagnet. (c) Similarity of the temperature dependence of the resistivities of Ni (above TC ), Pd, and Pt; for comparison Ag

    9 Magnetotransport

    445

    Fig. 7 (a) Channel resistivities ρ↓ (solid symbols) and ρ↑ (×ed symbols) of Ni, Co and Fe for the specified 3d transition metal impurities. (b) Channel resistivity ratio α = ρ↓ /ρ↑ as a function of impurity Table 1 Channel resistivities ρ↓ and ρ↑ and α = ρ↓ /ρ↑ for Ni, Co, and Fe hosts. (From [16, 38, 39, 40, 41, 42])

    Impurity/host Ti V Cr Mn Fe Co Ni Rh Re Ir Os

    Ni 105 67 65 75 60 35 21 75 48 64

    ρ↓ (nΩm) Co 110 77 24 100 67

    28 77 38 68

    Fe 40 13 28 17 33 120 64 27 200 43

    Ni 56 130 220 7.5 4 1.8 100 260 280 500

    ρ↑ (nΩm) Co 76 77 73 120 5.4

    28 180 117 235

    α Fe 105 105 125 130 16 24 11 87 22 130

    Ni 1.9 0.5 0.3 10 15 19

    Co 1.4 1.0 0.3 0.8 12

    0.1 0.3 0.2 0.1

    1.0 0.4 0.3 0.3

    Fe 0.4 0.1 0.2 0.1 2 5 6 0.3 9 0.3

    as their ratio α = ρ↓ /ρ↑ for the ferromagnetic hosts Ni, Co, and Fe as a function of the 3d transition metal impurities. Table 1 contains data of the channel resistivities. The data in Fig. 7a conform to the expectation that ρ↑ is small only for transition metal impurities Fe, Co, and Ni. In case of the early 3d elements, the channel resistivities are sizable, and for V and Cr that form virtual bound states at the Fermi level of the ferromagnetic host, α is even smaller than unity. In case of Ti, rather

    446

    M. Ziese

    large channel resistivities are found, but α, at least for the hosts Ni and Co, is larger than unity. The AMR is a spin-orbit interaction effect. It might be described by a model Hamiltonian that takes the exchange interaction, the crystal-field splitting, and the spin-orbit coupling into account [43, 44, 45]:  · S . Hˆ = Hˆ ex + Hˆ CF + AL

    (13)

    Spin-orbit coupling leads to a spin-admixture such that the majority carrier wave functions acquire a small spin-down component and vice versa. The scattering matrix element due to a spherically symmetric scatterer from an s- into a d-state is dependent on the magnetization, i.e., spin direction, thus leading to the AMR. In second-order perturbation theory, one obtains Δρ = γ (α − 1) , ρ

    (14)

    where α = ρ↓ /ρ↑ is the resistivity ratio between up- and down-channels and γ sets the strength of the effect. In second-order perturbation theory γ ∝ (A/Eex )2 or γ ∝ (A/ECF )2 , depending on whether the exchange splitting Eex or crystal field splitting ECF is larger. Figure 8 shows the anisotropic magnetoresistance as a function of channel resistivity ratio α. Up to α 3, the data fall onto a straight line with a slope of 0.03; shown is a linear dependence with a slope γ = 0.01 as suggested in [46].

    Giant Magnetoresistance (GMR ) Giant magnetoresistance is a phenomenon found in metallic multilayers [47, 48, 49] and spin valves [50]. It occurs when the ferromagnetic layers within the heterostructure are antiferromagnetically coupled through the adjacent metallic layers; in a sufficiently large applied magnetic field, the ferromagnetic layers will be ferromagnetically aligned, and the GMR ratio appears as the normalized resistance difference between the antiparallel and parallel configuration. Figure 9 shows extensive magnetization and magnetoresistance data on Co/Ru multilayers that show both the appearance of an antiferromagnetic coupling between the Co layers that oscillates with the Ru layer thickness as well as the accompanying giant magnetoresistance showing the same oscillations. In a first approximation, GMR might be understood within a simple resistor model [52, 53, 54]. One distinguishes two configurations with the current density parallel to the layers (current-in-plane, CIP) and perpendicular to the layers (current-perpendicular-to-plane, CPP); see Fig. 10a; in general CPP GMR is larger than CIP GMR. Within the spirit of the two-current model, the current density is

    9 Magnetotransport

    447

    Fig. 8 Anisotropic magnetoresistance AMR for alloys of Ni as a function of the channel resistivity ratio of the particular alloying element. The red line was calculated from Eq. (14) using γ = 0.01 (After [46] and [16])

    carried by spin-up and spin-down currents within each layer. A multilayer can then be modeled by a parallel circuit of layer resistances in series (CPP) or partially in series (within a mean free path, CIP); see Fig. 10b and [52, 53, 54]. On a more fundamental basis, spin accumulation at the interfaces has to be taken into account. The concept of spin diffusion was studied in early work [55] with measurements of the spin diffusion length λs in Al after spin injection from permalloy. At the interface between a ferromagnet and a normal metal, the conversion of an unpolarized current in the normal metal into spinpolarized currents in the ferromagnet leads to a difference in the electrochemical potentials μ↑ and μ↓ for spin-up and spin-down electrons; see Fig. 10c; this potential difference in turn causes an additional boundary resistance [56]. Using this concept, a quantitative theory of the GMR in CPP configuration was developed [57]. In antiferromagnetic spintronics, the spin valve as a trilayer formed by two ferromagnets separated by a normal metal is replaced by a ferromagnet adjacent to an antiferromagnetic tunnel junction, e.g., in a NiFe/IrMn/MgO/Pt stack [58]. The resistance change is due to tunneling anisotropic magnetoresistance [59]. The coupling to the external field is facilitated by the permalloy layer that rotates the antiferromagnetic magnetic moments by the exchange-spring effect.

    448

    M. Ziese

    Fig. 9 (a)–(f) Magnetization at 300 K (left axis) and transverse magnetoresistance at 4.5 K (right axis) of six Ru/Co superlattices with structure Si(111)/(10 nm) Ru/[(1.6 nm) Co/tRu Ru]20 /(5 nm) Ru and Ru layer thicknesses of 0.8, 1.1, 1.9, 2.7, 3.1, and 3.8 nm. (g) Transverse magnetoresistance and (h) saturation magnetic field as a function of Ru layer thickness. Clear oscillations in the saturation field and the magnetoresistance are observed: when the interlayer coupling is antiferromagnetic with a large saturation field, the magnetoresistance is “giant” (After [51])

    Colossal Magnetoresistance (CMR ) The isotropic magnetoresistance near the Curie temperature of ferromagnetic oxides of the type La0.7 Sr0.3 MnO3 (manganites) is called colossal magnetoresistance [60, 61, 62]. In principle, this magnetoresistance phenomenon is similar to the critical scattering in metallic ferromagnets; compare Fig. 2. The effect, however, is significantly larger, especially for compounds with a rather low Curie temperature [63]. A phenomenological overview of magnetoresistance data is given in Fig. 11a and b; the magnetoresistance was defined by the pessimistic definition Δρ/ρ0 = [ρ0 − ρ(B)] /ρ0 , where ρ0 denotes the resistivity value in zero field. Manganites crystallize in the perovskite structure (RE)1−x Ax MnO3 with rare earth elements RE = La, Nd, Pr, Y, . . . and A = Ca, Sr, Ba, Pb, or alkaline elements [68, 69]. The phase diagram of these materials is rather complex; see Fig. 11c: depending on the doping concentration x, a variety of antiferromagnetic, ferromagnetic, charge, or orbitally ordered phases is observed [67]. Depending on the size of the ions, a ferromagnetic phase is found around a doping level of 30%; maximum Curie temperature is 360 K for Sr-doping.

    Fig. 10 (a) Current-in-plane (CIP) and current-perpendicular-to-plane (CPP) configurations. (b) Resistor model of GMR – upper row: carrier scattering in the ferromagnetic and antiferromagnetic configurations (schematic); lower row, corresponding spin-dependent circuit models. (c) Spin-dependent electrochemical potentials at a ferromagnet-normal metal interface; λsF and λsN denote the spin-diffusion lengths in the ferromagnet and normal metal ((a) and (b) after [54])

    9 Magnetotransport 449

    450

    M. Ziese

    Fig. 11 (a) Zero field resistivity (left axis) and magnetization (right axis) as well as magnetoresistance ratio in a magnetic field of 1 T of La0.7 Ba0.3 MnO3 and La0.7 Ca0.3 MnO3 films. (After [64]). (b) Magnetoresistance ratio of various manganite compounds as a function of the Curie temperature. (After [65] and [63]). (c) Phase diagrams of La1−x Cax MnO3 and Pr1−x Cax MnO3 with the following phases: PMI paramagnetic insulator, FMM ferromagnetic metal, FMI ferromagnetic insulator, AFMI antiferromagnetic insulator, CO charge ordered. TC , TN , TP , and TCO denote the transition temperatures into the magnetically ordered and charge ordered phases, TS denotes the temperature of the structural transition between orthorhombic and rhombohedral phases. (After [66] and [67])

    In a first approximation, the electronic structure might be understood from the octahedral crystal field around the Mn ions. This leads to an energetic splitting of the threefold degenerate t2g states (dxy , dyz , and dzx orbitals) and the twofold degenerate eg states (dx 2 −y 2 and d3z2 −r 2 orbitals); see Fig. 12a. The doping with divalent atoms on the rare earth site introduces a mixed valence of Mn3+ and Mn4+ ions. The strong Hund’s rule coupling leads to a core spin of 3/2 Bohr magnetons of the low-lying t2g levels; the additional electron in the eg levels of a Mn3+ ion can hop to a neighboring Mn4+ ion, when the core spins are aligned; i.e., the hopping integral depends on

    9 Magnetotransport

    451

    Fig. 12 (a) Crystal-field splitting of the Mn 3d-levels in an octahedral environment. The arrows indicate the electron occupation of the levels for Mn3+ and Mn4 ions according to Hund’s rules. The dotted arrows indicate hopping of the eg electron from Mn3+ to Mn4+ via the O 2p orbital. This leads to the two degenerate Mn-O-Mn clusters shown. (c) Calculated band structure and density of states of La0.7 Sr0.3 MnO3 . The band structure is shown near the Γ point in the directions toward the X and K points (After [72])

    the angle Θ between the core spins as cos(Θ/2). This conduction mechanism is called double exchange [70,71]; see Fig. 12a; it explains the strong interdependence between ferromagnetism and conductivity and forms the basis for understanding colossal magnetoresistance. The physics of manganites is really intricate, since not only charge and spin degrees of freedom are coupled, but there is also a coupling to the orbitals and to lattice phonons [73]. Moreover, the inherent disorder created by the doping has to be taken into account. This has led to a variety of models for the colossal magnetoresistance, taking electron-phonon [74], electron-polaron [75], or phase separation [76] into account. At surfaces and interfaces, orbital reconstruction might play a role and might lead to new effects [77]. The strong Hund’s rule splitting leads to a half-metallic [78] band structure; see Fig. 12b [72], and in oxide heterostructures, La0.7 Sr0.3 MnO3 is the established material for ferromagnetic electrodes or spin injectors. The half-metallicity of the manganites in combination with a breakdown of the double-exchange mechanism at extended lattice defects leads to a significant extrinsic magnetoresistance known as grain-boundary magnetoresistance [79, 80, 81].

    Tunneling Magnetoresistance (TMR ) Early work on spin-dependent tunneling focused on tunneling from ferromagnets through a thin insulating barrier into superconductors [82, 83, 84]. The spindependent density of states D↑ (E), D↓ (E) of a ferromagnet can be measured by the asymmetry of the conductance peaks that occurs when electrons tunnel through an

    452

    M. Ziese

    AlOx barrier into a thin Al film in a large parallel field that splits the quasiparticle density of states due to the Zeeman energy; see Fig. 13a. This technique enables measurements of the spin polarization P =

    D ↑ − D↓ . D↑ + D↓

    (15)

    In related experiments with superconducting tips on ferromagnets, the spin polarization was determined by its influence on Andreev reflection [85,86]. Note that the spin polarization determined in different experiments might not be identical, since it might be either related to the density of states (spectroscopic spin polarization, Eq. (15)) or to the current density (P = (j↑ − j↓ )/(j↑ + j↓ ), compare to Mott’s two-current model above) [87]. Tunneling between two ferromagnets through an insulating barrier was discussed for the first time in [88], see data on Co/Ge/Co junctions in Fig. 13b and [104] for a review. In that work, a simple expression for the tunneling magnetoresistance between two ferromagnets with spin polarizations P1 and P2 was derived: T MR =

    R↑↓ − R↑↑ 2P1 P2 = ; R↑↑ 1 − P1 P2

    (16)

    R↑↑ and R↑↓ denote the resistance values for parallel and antiparallel magnetization orientation; note that in the standard pessimistic MR definition used in this chapter T MR = 2P1 P2 /(1 + P1 P2 ). High-quality ferromagnetic tunnel junctions were reported for the first time in [89, 105]; see Fig. 13c. The junction resistance is considerably enhanced in the magnetic field range between the coercive fields of the two layers, when the magnetizations are antiparallel. The tunneling transport mechanism is evidenced by the so-called Rowell-criteria [106, 90]: (i) exponential thickness dependence of the resistance, (ii) quasi-parabolic differential conductance vs. voltage curves, and (iii) insulator-like temperature dependence of the tunnel resistance; see Fig. 13d. It is even possible to measure the spin-resolved density of states using voltage-dependent TMR [107]. As seen from Eq. (16), the TMR ratio diverges in case of half-metallic ferromagnets; see, e.g., data on LSMO [91] in Fig. 13e. However, in case of oxide ferromagnets, the TMR ratio decreases strongly even for temperatures far below the Curie temperature [108]. In certain single crystalline systems, e.g., Fe/MgO/Fe, details of the band structure might enhance the TMR ratio above the value estimated by Eq. (16) [109, 92, 110, 111]; see the data in Fig. 13f. Large TMR ratios were also found for amorphous ferromagnets [112]. A vast number of spin-polarization values can be found in the literature; Table 2 shows a small selection of consolidated values obtained by the transport techniques discussed here in comparison to results from spin-polarized photoemission spectroscopy. The relation of the spin polarization to band structure and magnetic order might be complex, see, e.g., CoGd alloys with compensation point [113].

    9 Magnetotransport

    453

    Fig. 13 (a) Conductance G of a Ni/AlOx /Al junction as a function of applied voltage. The conductance was normalized to the conductance GN in the normal state. T = 0.4 K. The conductance peaks split in magnetic fields applied parallel to the Al film. In a magnetic field, the conductance peak heights are different, since the density of states in Ni is spin dependent. (After [82]). (b) Relative conductance ΔG/G0 = (G↑↑ − G↑↓ )/G0 of a Co/Ge/Co junction normalized to the conductance at 0 V. T = 4.2 K. (After [88]). (c) Tunneling magnetoresistance of a CoFe/Al2 O3 /Co junction at 295 K. (After [89]). (d) Resistance of a IrMn/NiFeCo/CoFe/AlOx /NiFeCo junction as a function of temperature in the parallel and antiparallel magnetization state. (After [90]). (e) Tunneling magnetoresistance of two La2/3 Sr1/3 MnO3 /SrTiO3 /La2/3 Sr1/3 MnO3 junctions showing the strong decrease with temperature. The inset shows a resistance hysteresis loop at 4.2 K. (After [91]). (f) Tunneling magnetoresistance hysteresis loops for a Fe/MgO/Fe junction at 80 K (open symbols) and 293 K (solid symbols). (After [92])

    Table 2 Spin polarization of various ferromagnets determined from transport measurements (second row) and spin-polarized photoemission spectroscopy [93] (third row). LSMO = La0.7 Sr0.3 MnO3 Fe 0.45 [94] 0.4 [98]

    Co +0.42 [94] −0.4 [99]

    Ni +0.31 [94] −0.3 [100]

    LSMO 0.95 [91] 0.9 [101]

    CrO2 0.90 [95] 0.95 [102]

    Fe3 O4 −0.4 [96] −0.55 [103]

    SrRuO3 −0.55 [97] ———–

    Spin-dependent tunneling is used as a surface-sensitive high-resolution magnetic imaging technique [114]. To this end, ferromagnetic or antiferromagnetic tips are used in a STM (scanning tunneling microscope) with atomic resolution, e.g., to measure the sharpness of domain walls [115] and the nanoscale spatial modulations

    454

    M. Ziese

    of the TMR [116] and to investigate the surface spin chirality [117] or single skyrmions [118].

    Powder Magnetoresistance (PMR ) Powder compacts of oxide ferromagnets often show a strong increase in their magnetoresistance compared to the respective bulk materials [119, 120, 121]. Powder magnetoresistance is related to grain boundary or tunneling magnetoresistance: the contact areas between adjacent grains are poorly conducting such that carrier transport occurs by tunneling between regions with different magnetization directions [69, 81].

    Organic Magnetoresistance (OMAR ) Polyfluorene films sandwiched between non-magnetic electrodes (Al, indium-tinoxide (ITO)) showed a sizable negative magnetoresistance in magnetic fields of the order of 100 mT that could be controlled by the applied voltage [122]. Even the sign of the magnetoresistance could be tuned by varying applied voltage and temperature [123, 124] or by adjusting the electron and hole injection in doublelayer OLEDs [125]. The OMAR is believed to be an intrinsic property of the organic semiconductor, not a property of the metal/organic semiconductor/metal sandwich. The magnetic field dependence has a characteristic shape described by B 2 /(B02 + B 2 ), where B0 is a constant of the order of 10 mT. OMAR mechanisms involve the external magnetic field manipulation of paramagnetic entities that are coupled to the local hyperfine field. These entities might be bipolarons, oppositely charged polarons, or excitons formed therefrom [126]. Ferromagnetic ordering in polymers showing OMAR was reported, so there might even be a relation between the negative magnetoresistance and ferromagnetism [127].

    Quantum Transport In previous sections, transport was mainly in the diffusive regime. In case of mean free paths being large compared to sample dimensions, i.e., in the ballistic regime, the sample conductance does not diverge but is found to be quantized in units of G0 = 2e2 / h = 7.74809173 × 10−5 S. The conductance can then be expressed in the form [128]

    G=

    2e2  Ti , h i

    (17)

    9 Magnetotransport

    455

    where Ti denotes the transmittivity of the ith conduction channel. The quantization can be clearly demonstrated in constrictions or nanowires, when only a few conduction channels are available, see, e.g., [129,130]. Equation (17) is valid, when the scattering process does not significantly depend on the spin and when states with different spin degrees of freedom are energetically degenerate. These conditions do not necessarily hold in ferromagnetic nanocontacts. Indeed, break junctions made from Ni wires touching a Ni plate showed conduction with half conductance quanta e2 / h but only in magnetic fields high enough to saturate the Ni wire magnetization [131]; see Fig. 14a and b. This indicates that the spin structure close to the contact is crucial in determining the conductance value. The magnetoresistance of Ni nanocontacts and break junctions showed large values up to 80% [132, 134]; see Fig. 14c. Although the data scatter strongly, a clear trend toward large magnetoresistance values at small conductance values of a few conductance quanta was observed. Similar results were found for break junctions between various half-metallic oxides [133, 134] (see Fig. 14d) but also for nanoconstrictions patterned in half-metallic oxide films [135]. The experimental data on oxides differ from those on metallic ferromagnets in that the conductance is significantly smaller than one conductance unit, such that the transport cannot be ballistic. Domain walls in nanoconstrictions can be very sharp [136]. The

    Fig. 14 Histograms of junction conductance for a variety of Ni wire-Ni plate break junctions in applied magnetic fields of (a) 0 mT and (b) 10 mT. When the Ni wire is in magnetic saturation, the emergence of spin-polarized half conductance quanta e2 / h was observed (After [131]). Magnetoresistance of (c) Ni and (d) magnetite nanocontacts as a function of junction conductance. Note that conductance in case of Ni is ballistic, whereas it is by a tunneling mechanism in case of Fe3 O4 . (c) After [132]. (d) After [133]. The lines are guides to the eye

    456

    M. Ziese

    magnetoresistance of these nanocontacts is modeled by spin scattering at the domain wall [137, 134, 133]: in case of a sharp wall, the situation is similar to that in spindependent tunneling with the tunneling probability being proportional to the product of the density of states left and right to the domain wall; for thicker domain walls, the resulting magnetoresistance is decreased by an adiabatic factor describing the accommodation of the travelling spin to the local spin in the domain wall [137,134]. The domain wall magnetoresistance increases strongly with decreasing domain wall width δ which explains the trends in Figs. 14c and d; in case of oxides, a δ −3 dependence was predicted [81].

    Exotic Tunneling Exotic tunneling refers to spin-dependent tunneling in systems with an active tunneling barrier [138]. In this section, spin blockade, resonant tunneling, spin filters, and ferromagnetic and multiferroic barriers will be briefly discussed. Spin blockade effects occur in systems in which tunneling is via ferromagnetic nanoparticles. Experimentally this might be realized in granular films consisting of ferromagnetic nanoparticles embedded in an insulating matrix [139] or by integrating ferromagnetic nanoparticles in the insulating barrier between ferromagnetic electrodes [144] or paramagnetic electrodes [145,146]. If the electrostatic energy of the charged nanoparticle is sufficiently large to prevent the tunneling of an additional electron onto the nanoparticle, the systems are in Coulomb-blockade regime. The tunneling current then has a steplike dependence on the voltage (Coulomb staircase); this was observed at room temperature for tunneling through Co nanoparticles with diameters 1–4 nm embedded in Al2 O3 [146]. The spin accumulation on the ferromagnetic island was theoretically shown to lead to a further chemical potential shift that in turn yields an additional modulation of the current-voltage characteristics leading to an oscillating tunneling magnetoresistance [147,148,149]. Experimentally, in granular systems, one finds a low temperature upturn in the TMR due to higher order tunneling processes over more than one ferromagnetic island [139, 120]; see Fig. 15a. Further, an oscillating TMR was reported for tunneling between a Co and an Al layer through a Co-Al-O film; see Fig. 15b. Resonant tunneling occurs through some kind of resonant structure, e.g., a quantum well or a double barrier. A classical device is the Esaki diode showing a negative differential resistance in the forward bias regime of heavily doped Ge p-n-junctions [150]. A spin-dependent tunneling device with a quantum well is shown in the inset to Fig. 15c [140]. A Cu quantum well in a Co/AlOx /NiFe structure leads to the oscillation of the TMR with Cu barrier thickness. This is due to the formation of two quantum well states with momenta determined by the Cu Fermi surface that resonate in the well and lead to an efficient injection and extraction into the tunneling barrier [140, 151]. As already indicated in the discussion of spin-blockade effects, spin-dependent tunneling can also be observed between a ferromagnetic and non-magnetic electrode, when these are separated by a thin ferro- or ferrimagnetic material that acts

    9 Magnetotransport

    457

    Fig. 15 (a) TMR of Co-Al-O granular films as a function of temperature. (After [139]). (b) TMR of a thin Co-Al-O layer in CPP geometry as a function of bias voltage. (After [139]). (c) TMR of a Co/AlOx /NiFe spin-valve with a Cu quantum wall as a function of Cu layer thickness. (After [140]). (d) TMR of a La0.7 Sr0.3 MnO3 /NiFe2 O4 /Au spin-filter valve. (After [141]). (e) TMR of a Fe/BaTiO3 /La0.7 Sr0.3 MnO3 tunnel junction for two orientations of the BaTiO3 polarization. (After [142]). (f) TMR of a La0.7 Sr0.3 MnO3 /La0.1 Bi0.9 MnO3 /Au spin-filter valve with multiferroic spin filter. (After [143])

    as a spin filter. Due to the exchange energy an electron experiences when tunneling through the magnetically ordered insulating barrier, the barrier heights for the two spin directions differ by 2Eex [152, 153]. A more detailed analysis shows that not only the barrier height but also the band alignment with the electrodes plays a role [154]. Experimentally, spin filtering was observed for ferromagnetic EuO in Al/EuO/Y/Al junctions by conductance spectroscopy in the superconducting state of Al [155] and ferrimagnetic NiFe2 O4 and multiferroic La0.1 Bi0.9 MnO3 by magnetoresistance measurements in La0.7 Sr0.3 MnO3 /NiFe2 O4 /Au [141] (see Fig. 15d) and La0.7 Sr0.3 MnO3 /SrTiO3 /La0.1 Bi0.9 MnO3 /Au [156] tunneling junctions. Another way of modulating the barrier height is by the use of a ferroelectric barrier with the electric polarization along the tunneling direction. In this case, the barrier height depends on the direction of polarization [157, 158]. This might be due to electrostatic effects, to strain, or to detailed cation shifts at the interface between the ferroelectric and the electrodes [157]. It is often difficult to distinguish between a barrier height effect or extrinsic mechanisms such as resistance switching

    458

    M. Ziese

    that might depend on interfacial effects or the formation of conducting paths within the ferroelectric [159, 160]. The use of ferromagnetic electrodes and a ferroelectric barrier allows for the construction of a four-state memory element. This has been demonstrated, e.g., for Fe/BaTiO3 /La0.7 Sr0.3 MnO3 [142]; see Fig. 15e, for La0.7 Sr0.3 MnO3 /PbZr0.2 Ti0.8 O3 /Co [161] and for La0.7 Sr0.3 MnO3 /PbTiO3 /Co junctions [162]. A four-state logic might also be realized using a multiferroic spin filter. In this case, the spin filter can be modulated by application of a magnetic field but also by application of an electric field via a voltage pulse. Experimentally, such a device was realized for the first time in a La0.7 Sr0.3 MnO3 /La0.1 Bi0.9 MnO3 /Au tunnel junction [143]; see Fig. 15f. For further reading, the review papers [163, 164, 165] are recommended.

    Hall Effect When a magnetic induction B is applied perpendicular to a flat sample, then a transverse electric field Ey compensating the Lorentz force is generated by charge accumulation at the sample edges. This transverse electric field gives rise to a transverse voltage VH , called Hall voltage [166]; see Fig. 16a. In a metal, electrons (−) or holes (+) may be charge carriers, and the normal Hall constant is given by RH =

    Ey ρyx 1 VH d = = =± . Ix B jx B B en

    (18)

    In case two bands are contributing to the carrier transport, the Hall constant was already given in Eq. (9). The Hall angle is defined as tan ΘH = Ey /Ex = ρyx /ρxx . A magnetic field applied to a ferromagnetic metal aligns the magnetization along its direction; see Fig. 16a. The magnetization component M along the magnetic field direction leads to an additional Hall effect contribution, known as extraordinary or anomalous Hall effect. Phenomenologically, the Hall resistivity is then given by [170] ρyx = RH B + μ0 RA M

    (19)

    with the anomalous Hall constant RA . Note that in ferromagnetic films the demagnetizing factor is close to unity and the magnetic field H within the sample is given in terms of the applied field HA by H = HA − NM, such that the magnetic induction is B = μ0 (H + M) = μ0 (HA + (1 − N)M) μ0 HA . Figure 16b shows the Hall resistivity of a La0.7 Sr0.3 MnO3 film as a function of magnetic field. At low fields, the Hall resistivity rises steeply due to the anomalous Hall contribution (see Fig. 16c); at high fields the magnetization M is saturated, and a much smaller slope due to the ordinary Hall effect is observed. The product μ0 RA MS is obtained by extrapolation to B = 0 as indicated in Fig. 16b; RH is obtained from the high field slope. The manganites are hole conductors with a negative anomalous Hall constant.

    9 Magnetotransport

    459

    The anomalous Hall effect is due to the spin-orbit interaction; see section “Spin Currents ”. One distinguishes between extrinsic and intrinsic contributions. Extrinsic contributions are due to scattering: (i) Skew-scattering is the asymmetric scattering of an electron with respect to the plane containing its incident velocity and the ion’s magnetic moment [170, 171, 172]. (ii) Side-jump scattering denotes the side shift of the extrapolated trajectories before and after the scattering event [170, 173]. Both mechanisms predict a scaling RA ∝ ρ n of the anomalous Hall constant with the longitudinal resistivity, with n = 1 for skew scattering and n = 2 for side-jump scattering, compare to Fig. 16d. The intrinsic contribution to the anomalous Hall effect is due to the Berry phase electrons acquire when moving through a spin lattice [174, 175]. An example material for this contribution might be SrRuO3 which has an intricate temperature dependence of the anomalous Hall constant attributed to magnetic monopole-like features at the Weyl points in momentum space [176, 177]; this is illustrated

    Fig. 16 (a) Hall effect in a ferromagnet. (b) Hall resistivity ρyx measured on a 5 nm thick La0.7 Sr0.3 MnO3 film. At higher temperature, the anomalous Hall effect is clearly seen; the extrapolation to B = 0 to determine μ0 RA MS is indicated. (c) Initial Hall effect slopes dρyx /dB for 300 nm thick La0.7 Ca0.3 MnO3 and La0.7 Ba0.3 MnO3 films as well as a Ni sample. (After [64] and [167]). (d) Scaling of the Hall constant with the longitudinal resistivity for the films from (c). Solid and dashed lines indicate the linear and quadratic scaling laws due to skew- and sidejump scattering. (After [64]) (e) Hall resistivity of SrRuO3 illustrating the sign change of the anomalous Hall constant. (After [168]). (f) Hall effect of a SrRuO3 /SrIrO3 bilayer decomposed into the anomalous (AHE) and topological (THE) Hall effect contributions (After [169])

    460

    M. Ziese

    by Fig. 16e showing the sign change of the anomalous Hall constant in SrRuO3 . Further, Berry phase effects lead to the emergence of a topological Hall effect when the charge carriers move through spin structures with complex topology such as skyrmions [178,169]. Although the spin topology does not leave a direct signature in the magnetization component M, an additional Hall resistivity ρT H E , not following the field dependence of the magnetization, is observed. Complex spin structures might be induced by interfacial spin-orbit coupling, as illustrated in Fig. 16f showing the Hall effect of a SrRuO3 /SrIrO3 bilayer [169]. The Hall effect can be used for nanoscale magnetometry, since Hall bars can be fabricated on the micro- and nanometer scale; see, e.g., magnetometry on single iron nanoparticles [179] or the discovery of flux-line lattice melting in Bi2 Sr2 CaCu2 O8 high temperature superconductors [180].

    Spin Currents The charge current density from Eq. (1) is defined as j = (−e)n v with the electronic charge (−e), the carrier density n, and the electronic velocity v. Taking not only drift, but also diffusion currents into account, Eq. (1) is generalized to ji = σ Ei + eD

      ∂n 1 ∂μc = σ Ei + ∂xi e ∂xi

    (20)

    with the electronic diffusion constant D. In terms of a generalized force, the diffusion is not driven by a density gradient, but by the gradient of the chemical potential μc ; the second equation on the right hand side follows from the generalized Einstein equation eD(∂n/∂μc ) = nμ, where μ = σ/(en) is the electronic mobility. In case of a semiconductor, in the non-degenerate case (∂n/∂μc ) = n/(kB T ) which leads back to the Einstein equation. A similar phenomenology holds for spindependent charge currents [181] but is not further followed here. Instead, the following discussion uses a phenomenological approach proposed in [182]. Besides the charge the electron also transports spin. The average spin direction is characterized by the spin polarization vector P , −1 ≤ Pi ≤ 1, such that the spin density is given by nP . The transport of the spin direction is described by a second rank tensor [182, 183] js,ik =

    h¯ nvi Pk . 2

    (21)

    The units of the spin current density js,ik are chosen such that it carries angular momentum h/2. Since the spin transport is coupled to the charge transport, the ¯ spin current density is also driven by electric fields and chemical potential gradients [182, 183]:

    9 Magnetotransport

    

    js,ik

    h¯ =− 2e

    461

          ∂(nPk ) 1 ∂μs,k h¯ =− . σ Ei Pk + eD σ Ei Pk + ∂xi 2e e ∂xi (22)

    The spin-dependent potentials μs,k are defined by generalized Einstein equations eD(∂(nPk )/∂μs,k ) = nμ. In relation with the continuity equation h¯ (nPk ) h¯ ∂(nPk ) ∂js,ik + =− 2 ∂t ∂xi 2 τs

    (23)

    spin drift and diffusion can be predicted. τs denotes the spin relaxation time. New phenomena occur, when spin-orbit coupling is taken into account [182,184, 185]. It was already shown in [182] that the spin-orbit interaction couples the charge and spin current densities in such a way that a term proportional to −(2e/h) ¯ ikl js,kl is added to Eq. (20) and a term proportional to (h/(2e)) ¯ ikl jl is added to Eq. (22). The different signs follow from the fact that the charge current density changes sign under both space and time inversion, whereas the spin current density changes sign only under space inversion. This yields the equations [183, 185]:   j = σ E + eD∇n + αSH σ (E × P ) + αSH eD ∇ × (nP ) (24)    ∂n h¯ ∂(nPk ) . σ Ei Pk + eD − αSH σ ikl El − αSH eDikl js,ik = − 2e ∂xi ∂xl (25) ikl denotes the totally antisymmetric tensor; αSH is the spin Hall angle defined as the ratio of the spin Hall conductivity σs to the diagonal charge conductivity σ : αSH = (2h/e)(σ ¯ s /σ ); for a tabulation of αSH values, see [185]. The cross terms appearing in Eqs. (24), (25) relate to various Hall effect phenomena. The third term on the r.h.s. of Eq. (24) yields a charge current density proportional to both applied electric field and spin polarization, i.e., in the presence of a finite magnetization, an applied electric field induces a Hall effect proportional to the magnetization component perpendicular to the applied electric field. This is exactly the anomalous Hall effect discussed in section “Hall Effect ” and described phenomenologically by Eq. (19).

    Spin Hall Effect The third and fourth term on the r.h.s. of Eq. (25) lead to the spin Hall effect [184], in which a charge current produces a perpendicular spin current that leads to spin accumulation at the boundaries of a finite sample. Consider a narrow flat metal or semiconductor strip of width w (−∞ < x < ∞, −w/2 ≤ y ≤ w/2, 0 ≤ z ≤ d) in an electric field E applied along the x-direction. Solving Eqs. (23), (25) with the boundary condition of vanishing spin current density at the surface

    462

    M. Ziese

    yields a√spin polarization density nPz = (αSH σ λs )/(eD)E sinh(y/λs ), where λs = Dτs denotes the spin diffusion length. At the two sample edges, spin polarization of opposite direction perpendicular to the sample is created within a layer of thickness λs . The spin Hall effect was detected optically [186] in GaAs and electrically in Al [188]. Figure 17a shows results obtained on an n-doped GaAs film: At the sample edges, Kerr rotation signals with an extent of about 10 μm and opposite polarity appear in zero magnetic field. A transverse magnetic field destroys the spin polarization, i.e., the spin Hall effect – contrary to the normal Hall effect – is a zero magnetic field phenomenon.

    Inverse Spin Hall Effect The fourth term on the r.h.s. of Eq. (24) leads to the inverse spin Hall effect, in which a spin current produces a Hall voltage by the accumulation of charges at the boundaries of a finite sample. Consider a narrow metallic or semiconducting half-slab (−∞ < x < ∞, −w/2 ≤ y ≤ w/2, z ≤ 0). At the top surface, spins are injected with a spin polarization along the x-axis. This leads to a nonzero spin current density component js,zx ; solving Eqs. (25), (23) one finds a spin polarization decreasing exponentially into the material. Since there is no charge flow through the surfaces, the charge current generated by ∇ × P must be balanced by an induced electric field along the y-direction: Ey = −αSH (eD)/(σ λs )(nPx0 ) exp(z/λs ); Px0 denotes the spin polarization at the top surface. The inverse spin Hall effect was detected in Pt when spin pumping from an adjacent permalloy layer [189] or by spin diffusion from a Cu-permalloy injector [187]; however, it does not only occur in non-magnetic metals with strong spin-orbit interaction, but also in ferromagnetic permalloy, after spin injection from an adjacent yttrium iron garnet (YIG) [190]. Figure 17d shows the experiment after [187]: spins are injected from a permalloy pad into a Cu junction by driving a current through the ferromagnetic pad along one of the Cu arms. The spin density diffuses further along the Cu arm into the Pt wire, where it creates a Hall voltage by the inverse spin Hall effect.

    Quantum Spin Hall Effect The classical spin current phenomenology was extended by the discovery of the quantum spin Hall effect in two-dimensional systems subjected to a strong magnetic field and the realization that the spin Hall conductance σs = (h/(2e))α ¯ SH σ may be quantized in units of e/(2π ) [191]. In this case, spin-up and spin-down electrons move with opposite chirality in one-dimensional edge states around the sample, e.g., in HgTe quantum wells [192]. A generalization of these observations leads to the concept of the topological insulator with topologically protected conducting surface states around insulating bulk states [193].

    Fig. 17 Spin Hall effect [186]: (a) GaAs film with electrodes; a local Kerr effect probe measures the Kerr rotation across the sample; the Kerr rotation is proportional to the spin polarization. (b) Kerr rotation as a function of position across the sample and magnetic field applied parallel to y. The concentration at the sample edges and the dephasing due to the Hanle effect are clearly seen. (c) Kerr rotation as a function of position across the film. The solid line is a fit to the sinh function derived in the text. Inverse spin Hall effect [187]: (d) Pt film connected via a Cu junction to a permalloy (Py) spin injector. A current I is driven through permalloy and junction by the voltage Vs ; the inverse spin Hall voltage VH across the Pt strip is measured. (e) and (f) inverse spin Hall resistance VH /I at 77 K and 300 K as a function of applied magnetic field

    9 Magnetotransport 463

    464

    M. Ziese

    Magnetoimpedance Magnetoimpedance describes the dependence of the impedance Z of a soft ferromagnetic wire or ribbon on a static magnetic field. In soft ferromagnets such as FeCoSiB amorphous wires, the effect can be substantial for frequencies in the high kHz and low MHz regime and was therefore called giant magnetoimpedance (GMI) [194, 195]. The basis for GMI is the classical skin effect occurring in conductors at high frequencies. Let us consider a soft ferromagnetic ribbon with permeability μ = μr μ0 such that the magnetic induction B is given in terms of the magnetic field by B = μH . If the ribbon is conducting, and Maxwell’s equations can be solved for a current Iac = I0 exp(iωt) flowing along the long ribbon axis (infinite slab of thickness d, −d/2 ≤ x ≤ d/2): Z=R

    R (1 + i)u = . tanh [(1 + i)u] μac

    (26)

    R denotes √ the dc resistance of the ribbon and u = d/(2δ) with the skin depth δ = 2/(ωσ μ0 μr ). The second equation on the r.h.s. is valid for a material with isotropic relative permeability μr . μac denotes the ac-susceptibility of the ribbon, caused by the skin effect, when placed in an alternating magnetic field Hac = H0 exp(iωt) along the ribbon axis. For small skin depths, d/δ  1, μac ∼ (δ/d)(1 − i), and the impedance is much larger than the dc resistance. When a static magnetic field is applied, the relative permeability μr is substantially decreased and the skin depth increased, thus leading to a large decrease of the impedance, i.e., in the zeroth approximation, giant magnetoimpedance can be considered as a magnetically controlled skin effect. Beyond this basic approximation, various aspects have to be taken into account [196]: in eddy current models, the role of different domain structures (circular domains, stripe domains) on the microscopic eddy current loss was studied [197]; in domain models, the dispersion of the circumferential permeability [198] and the different contributions from domain wall motion and magnetization rotation [199] were clarified. Eddy current and domain models successfully explain the relevant features of giant magnetoimpedance up to a frequency of about 100 MHz. In the GHz range, the magnetoimpedance is dominated by ferromagnetic resonance effects [200].

    Measurements The measurement considerations in this section are restricted to homogeneous films of thickness d; for anisotropic films, see [201]. In two-contact measurements, the resistance of the wires, the contacts, and the sample are measured in series; this is

    9 Magnetotransport

    465

    Fig. 18 (a) Four-contact configuration for measurement of the longitudinal voltage V and the Hall voltage VH . (b) van der Pauw geometries. (c) Corbino disk

    acceptable, when the sample resistance dominates, e.g., with metallic samples and soldered or welded contacts. CPP measurements on metallic multilayers are made in two-contact configuration using superconducting contacts. Often, however, the contact resistance is high or is even non-ohmic, when pn or Schottky junctions are formed between contact material and sample. In these cases, two-contact measurements should be avoided; care should always be taken in choosing the appropriate contact material. Consider first the four-contact configuration shown in Fig. 18a. When a current I is injected through a tip into the sample, the current density j flows radially outward from the tip and at distance r is given by j =

    I 1 1 r = E = − ∇Φ 2 ρ ρ 2π r d

    (27)

    where r is a radius vector within the plane of the sample and Φ denotes the electric potential. The latter is found by integration; in the situation shown in Fig. 18a with current I injected at 1 and extracted at 4, the potential difference between 2 and 3 is given by V =

      ρI r2 . ln πd r1

    (28)

    From this, the resistance R = V /I and the resistivity ρ might be obtained. The Hall voltage is given by VH = ρyx I /d, where ρyx denotes the off-diagonal resistivity; see Eq. (7). In this geometry, care has to be taken to ensure a careful alignment of the Hall voltage contacts; any horizontal shift will lead to the addition of a longitudinal voltage contribution to the Hall voltage. In order to minimize the influence of contact size and placement, van der Pauw suggested an alternative method for resistance measurements of thin homogeneous samples with arbitrary shape [202]. Although this method is valid for arbitrarily formed samples, the recommended shapes are shown in Fig. 18b, since these

    466

    M. Ziese

    minimize the influence of the size and exact location of the contacts. Consider a sample with four contacts 1, 2, 3, 4 arranged counterclockwise around its rim, see Fig. 18b, Ikl be a current injected into contact k and extracted from contact l; in case of dc currents eight, in case of ac currents four non-diagonal combinations Ikl , k, l = 1, 2, 3, 4, can be defined. The voltages measured between contacts i and j are labeled Vij . Then eight (four) voltage measurements yield eight (four) positive resistance values: Rij,kl = Vij /Ikl . Measurement consistency requires that Rij,kl = Rj i,lk . The reciprocity theorem [203] requires that R21,34 + R12,43 = R43,12 + R34,21 , R32,41 + R23,14 = R14,23 + R41,32 . The sheet resistance can be calculated from the two averaged resistances (dc case) RA = (R21,34 + R12,43 + R43,12 + R34,21 )/4 and RB = (R32,41 + R23,14 + R14,23 + R41,32 )/4 using the van der Pauw relation     πd πd exp − (29) RA + exp − RB = 1 . ρ ρ If the sample has a line of symmetry, RA = RB , and ρ=

    πd RA . ln(2)

    (30)

    In case of Hall measurements, the procedure is similar; with use of ac currents, the diagonal resistances R24,13 and R13,24 are measured. The Hall resistivity loop as a function of applied magnetic field is then obtained as ρyx = (R24,13 − R13,24 )d .

    (31)

    If one is only interested in the Hall constant, resistivity values for positive and negative magnetic field should be averaged. Hall effect measurements are only possible, when the sample has edges placed in such a way that charges accumulate. This was the case in the configurations discussed so far. These correspond to a vanishing current density component perpendicular to the longitudinal current density component. Measurements in the absence of a transverse electric field, i.e., a Hall field, can be performed in the so-called Corbino disk configuration; see Fig. 18c. Here the conducting annulus has contacts around the outer and inner rims. In the absence of a magnetic field, the current density flows in radial direction; in a perpendicular magnetic field, the current density flows in spirals. In this configuration, the conductance related to σxx is measured. In case of measurements of the spin accumulation, nonlocal resistivity measurements are performed [55]. Usually four contacts are defined on the sample in order to perform local four-point measurements. Further contacts are defined outside the current path for the measurement of voltages generated by the diffusion of carriers. Care has to be taken to define a unique ground potential in these measurements [204].

    9 Magnetotransport

    467

    References 1. Ohm, G.S.: Die galvanische Kette. T. H. Riemann, Berlin (1827) 2. Onsager, L.: Reciprocal relations in irreversible processes. Phys. Rev. 37, 405 (1931) 3. Lide, D.R. (ed.): CRC Handbook of Chemistry and Physics, 84th edn. CRC Press, Boca Raton (2003) 4. Ho, C.Y., Ackerman, M.W., Wu, K.Y., Havill, T.N., Bogaard, R.H., Matula, R.A., Oh, S.G., James, H.M.: Electrical resistivity of ten selected binary alloy systems. J. Phys. Chem. Ref. Data 12, 183 (1983) 5. Ashcroft, N.W., Mermin, N.D.: Solid State Physics. Holt-Saunders Japan, Tokyo (1981) 6. Drude, P.: Zur Elektronentheorie der Metalle. Ann. Phys. 306, 566 (1900) 7. Matthiessen, A., von Bose, M.: On the influence of temperature on the electric conducting power of metals. Phil. Trans. R. Soc. Lond. 152, 1 (1862) 8. Gerlach, W., Schneiderhan, K.: Ferromagnetismus und elektrische Eigenschaften. Widerstand, magnetische Widerstandsänderung und wahre Magnetisierung beim Curie-Punkt. Ann. Phys. 398, 772 (1930) 9. Mott, N.F.: Electrons in disordered structures. Adv. Phys. 16, 49 (1967) 10. Anderson, P.W.: Absence of diffusion in certain random lattices. Phys. Rev. 109, 1492 (1958) 11. Mooij, J.H.: Electrical conduction in concentrated disordered transition metal alloys. Phys. Stat. Sol. (A) 17, 521 (1973) 12. Fisk, Z., Webb, G.W.: Saturation of the high-temperature normal-state electrical resistivity of superconductors. Phys. Rev. Lett. 36, 1084 (1976) 13. Millis, A.J., Hu, J., Das Sarma, S.: Resistivity saturation revisited: results from a dynamical mean field theory. Phys. Rev. Lett. 82, 2354 (1999) 14. Lee, P.A., Ramakrishnan, T.V.: Disordered electronic systems. Rev. Mod. Phys. 57, 287 (1985) 15. Bergmann, G.: Consistent temperature and field dependence in weak localization. Phys. Rev. B 28, 515 (1983) 16. Campbell, I.A., Fert, A.: Transport properties of ferromagnets. In: Wohlfarth, E.P. (ed.) Ferromagnetic Materials, vol. 3, pp. 751–800. North-Holland Publishing Company, Amsterdam (1982) 17. Raquet, B., Viret, M., Sondergard, E., Céspedes, O., Mamy, R.: Electron-magnon scattering and magnetic resistivity in 3d ferromagnets. Phys. Rev. B 66, 024433 (2002) 18. Du, X., Tsai, S.-W., Maslov, D.L., Hebard, A.F.: Metal-insulator-like behavior in semimetallic bismuth and graphite. Phys. Rev. Lett. 94, 166601 (2005) 19. Kempa, H., Kopelevich, Y., Mrowka, F., Setzer, A., Torres, J.H.S, Höhne, R., Esquinazi, P.: Magnetic-field-driven superconductor-insulator-type transition in graphite. Solid State Commun. 115, 539 (2000) 20. Kohler, M.: Zur magnetischen Widerstandsänderung reiner Metalle. Ann. Phys. 424, 211 (1938) 21. Beer, A.C.: Galvanomagnetic Effects in Semiconductors. Academic Press, New York (1963) 22. Xu, R., Husmann, A., Rosenbaum, T.F., Saboungi, M.-L., Enderbya, J.E., Littlewood, P.B.: Large magnetoresistance in non-magnetic silver chalcogenides. Nature 390, 57 (1997) 23. McClure, J.W., Spry, W.J.: Linear magnetoresistance in the quantum limit in graphite. Phys. Rev. 165, 809 (1968) 24. Friedman, A.L., Tedesco, J.L., Campbell, P.M., Culbertson, J.C., Aifer, E., Perkins, F.K., Myers-Ward, R.L., Hite, J.K., Eddy, C.R. Jr., Jernigan, G.G., Gaskill, D.K.: Quantum linear magnetoresistance in multilayer epitaxial graphene. Nano Lett. 10, 3962 (2010) 25. Abrikosov, A.A.: Quantum magnetoresistance. Phys. Rev. B 58, 2788 (1998) 26. Argyres, P.N.: Quantum theory of longitudinal magneto-resistance. J. Phys. Chem. Solids 4, 19 (1958) 27. Adams, E.N., Holstein, T.D.: Quantum theory of transverse galvano-magnetic phenomena. J. Phys. Chem. Solids 10, 254 (1959)

    468

    M. Ziese

    28. Thomson, W.: On the electro-dynamic qualities of metals: effects of magnetization on the electric conductivity of nickel and of iron. Proc. R. Soc. (Lond.) 8, 546 (1857) 29. McGuire, T.R., Potter, R.I.: Anisotropic magnetoresistance in ferromagnetic 3d alloys. IEEE Trans. Mag. 11, 1018 (1975) 30. Döring, W.: Die Abhängigkeit des Widerstandes von Nickelkristallen von der Richtung der spontanen Magnetisierung. Ann. Phys. 424, 259 (1938) 31. Döring, W., Simon, G.: Die Richtungsabhängigkeit der Magnetostriktion. Ann. Phys. 460, 373 (1960) 32. Birss, R.R.: Symmetry and Magnetism. North-Holland, Amsterdam (1964) 33. Ziese, M., Vrejoiu, I., Hesse, D.: Structural symmetry and magnetocrystalline anisotropy of SrRuO3 films on SrTiO3 . Phys. Rev. B 81, 184418 (2010) 34. Tang, H.X., Kawakami, R.K., Awschalom, D.D., Roukes, M.L.: Giant planar Hall effect in epitaxial (Ga,Mn)As devices. Phys. Rev. Lett. 90, 107201 (2003) 35. Bason, Y., Klein, L., Yau, J.-B., Hong, X., Ahn, C.H.: Giant planar Hall effect in colossal magnetoresistive La0.84 Sr0.16 MnO3 thin films. Appl. Phys. Lett. 84, 2593 (2004) 36. Mott, N.F.: The electrical conductivity of transition metals. Proc. R. Soc. 153, 699 (1936) 37. Mott, N.F.: The resistance and thermoelectric properties of the transition metals. Proc. R. Soc. 156, 368 (1936) 38. Dorleijn, J.W.F., Miedema, A.R.: A quantitative investigation of the two current conduction in nickel alloys. J. Phys. F 5, 487 (1975) 39. Dorleijn, J.W.F.: Electrical conduction in ferromagnetic metals. Philips Res. Rep. 31, 287 (1976) 40. Fert, A., Campbell, I.A.: Electrical resistivity of ferromagnetic nickel and iron based alloys. J. Phys. F 6, 849 (1976) 41. Ross, R.N., Price, D.C., Williams, G.: Resistivity and magnetoresistance anisotropy associated with s, p impurities in ferromagnetic Ni and Co. J. Phys. F 8, 2367 (1978) 42. Ross, R.N., Price, D.C., Williams, G.: Spontaneous magnetoresistance anisotropy and deviations from Matthiessen’s rule induced by 5s, p impurities in Fe. J. Magn. Magn. Mater. 10, 59 (1979) 43. Smit, J.: Magnetoresistance of ferromagnetic metals and alloys at low temperatures. Physica 17, 612 (1951) 44. Campbell, I.A., Fert, A., Jaoul, O.: The spontaneous resistivity anisotropy in Ni-based alloys. J. Phys. C 3, S95 (1970) 45. Malozemoff, A.P.: Anisotropic magnetoresistance with cubic anisotropy and weak ferromagnetism: a new paradigm. Phys. Rev. B 34, 1853 (1986) 46. Jaoul, O., Campbell, I.A., Fert, A.: Spontaneous resistivity anisotropy in Ni alloys. J. Magn. Magn. Mater. 5, 23 (1977) 47. 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) 48. Baibich, M.N., Broto, J.M., Fert, A., Van Dau, F.N., Petroff, F., Eitenne, P., Creuzet, G., Friederich, A., Chazelas, J.: Giant magnetoresistance of (001)Fe/(001)Cr magnetic superlattices. Phys. Rev. Lett. 61, 2472 (1988) 49. Binasch, G., Grünberg, P., Saurenbach, F., Zinn, W.: Enhanced magnetoresistance in layered magnetic structures with antiferromagnetic interlayer exchange. Phys. Rev. B 39, 4828 (1989) 50. Dieny, B.: Giant magnetoresistance in spin-valve multilayers. J. Magn. Magn. Mater. 136, 335 (1994) 51. Parkin, S.S.P., More, N., Roche, K.P.: Oscillations in exchange coupling and magnetoresistance in metallic superlattice structures: Co/Ru, Co/Cr, and Fe/Cr. Phys. Rev. Lett. 64, 2304 (1990) 52. Edwards, D.M., Mathon, J., Muniz, R.B.: A resistor network theory of the giant magnetoresistance in magnetic superlattices. IEEE Trans. Mag. 21, 3548 (1991) 53. Mathon, J.: Exchange interactions and giant magnetoresistance in magnetic multilayers. Contemp. Phys. 32, 143 (1991)

    9 Magnetotransport

    469

    54. Mathon, J.: Phenomenological theory of giant magnetoresistance. In: Ziese, M., Thornton, M.J. (eds.) Spin Electronics, pp. 71–88. Springer, Berlin/Heidelberg (2001) 55. Johnson, M., Silsbee, R.H.: Spin-injection experiment. Phys. Rev. B 37, 5326 (1988) 56. van Son, P.C., van Kempen, H., Wyder, P.: Boundary resistance of the ferromagneticnonferromagnetic metal interface. Phys. Rev. Lett. 58, 2271 (1987) 57. Valet, T., Fert, A.: Theory of the perpendicular magnetoresistance in magnetic multilayers. Phys. Rev. B 48, 7099 (1993) 58. Park, B.G., Wunderlich, J., Martí, X., Holý, V., Kurosaki, Y., Yamada, M., Yamamoto, H., Nishide, A., Hayakawa, J., Takahashi, H., Shick, A.B., Jungwirth, T.: A spin-valve-like magnetoresistance of an antiferromagnet-based tunnel junction. Nat. Mater. 10, 347 (2011) 59. Gould, C., Rüster, C., Jungwirth, T., Girgis, E., Schott, G.M., Giraud, R., Brunner, K., Schmidt, G., Molenkamp, L.W.: Tunneling anisotropic magnetoresistance: a spin-valve-like tunnel magnetoresistance using a single magnetic layer. Phys. Rev. Lett. 93, 117203 (2004) 60. von Helmolt, R., Wecker, J., Holzapfel, B., Schultz, L., Samwer, K.: Giant negative magnetoresistance in perovskitelike La2/3 Ba1/3 MnOx ferromagnetic films. Phys. Rev. Lett. 71, 2331 (1993) 61. Jin, S., Tiefel, T.H., McCormack, M., Fastnacht, R.A., Ramesh, R., Chen, L.H.: Thousandfold change in resistivity in magnetoresistive La-Ca-Mn-O films. Science 264, 413 (1994) 62. Searle, C.W., Wang, S.T.: Studies of the ionic ferromagnet (LaPb)MnO3 . Electric transport and ferromagnetic properties. Can. J. Phys. 48, 2023 (1970) 63. Fontcuberta, J., Martínez, B., Seffar, A., Piñol, S., García-Muñoz, J.L., Obradors, X.: Colossal magnetoresistance of ferromagnetic manganites: structural tuning and mechanisms. Phys. Rev. Lett. 76, 1122 (1996) 64. Ziese, M., Srinitiwarawong, C.: Extraordinary Hall effect in La0.7 Ca0.3 MnO3 and La0.7 Ba0.3 MnO3 thin films. Europhys. Lett. 45, 256 (1999) 65. Ramirez, A.P.: Colossal magnetoresistance. J. Phys.: Condens. Matter 9, 8171 (1997) 66. Fujishiro, H., Fukase, T., Ikebe, M.: Charge ordering and sound velocity anomaly in La1−x Srx MnO3 (x ≥ 0.5). J. Phys. Soc. Jpn. 67, 2582 (1998) 67. Shenoy, V.B., Rao, C.N.R.: Electronic phase separation and other novel phenomena and properties exhibited by mixed-valent rare-earth manganites and related materials. Phil. Trans. R. Soc. A 366, 63 (2008) 68. Jonker, G., van Santen, J.: Ferromagnetic compounds of manganese with perovskite structure. Physica 16, 337 (1950) 69. Coey, J.M.D., Viret, M., von Molnár, S.: Mixed-valence manganites. Adv. Phys. 48, 167 (1999) 70. Zener, C.: Interaction between the d-shells in the transition metals. Ferromagnetic compounds of manganese with perovskite structure. Phys. Rev. 82, 403 (1951) 71. de Gennes, P.G.: Effects of double exchange in magnetic crystals. Phys. Rev. 118, 141 (1960) 72. Böttcher, D., Henk, J.: Magnetic properties of strained La2/3 Sr1/3 MnO3 perovskites from first principles. J. Phys.: Condens. Matter 25, 136005 (2013) 73. Salomon, M.B., Jaime, M.: The physics of manganites: structure and transport. Rev. Mod. Phys. 73, 583 (2001) 74. Millis, A.J., Littlewood, P.B., Shraiman, B.I.: Double exchange alone does not explain the resistivity of La1−x Srx MnO3 . Phys. Rev. Lett. 74, 5144 (1995) 75. J. M. D Coey, Viret, M., Ranno, L., and Ounadjela, K.: Electron localization in mixed-valence manganites. Phys. Rev. Lett. 75, 3910 (1995) 76. Dagotto, E., Hotta, T., Moreo, A.: Colossal magnetoresistant materials: the key role of phase separation. Phys. Rep. 344, 1 (2001) 77. Ziese, M. and Vrejoiu, I.: Properties of manganite/ruthenate superlattices with ultrathin layers. Phys. Status Solidi RRL 7, 243 (2013) 78. de Groot, R.A., Mueller, F.M., van Engen, P.G., Buschow, K.H.J.: New class of materials: half-metallic ferromagnets. Phys. Rev. Lett. 50, 2024 (1983) 79. Hwang, H.Y., Cheong, S.-W., Ong, N.P., Batlogg, B.: Spin-polarized intergrain tunneling in La2/3 Sr1/3 MnO3 . Phys. Rev. Lett. 77, 2041 (1996)

    470

    M. Ziese

    80. Gupta, A., Gong, G.Q., Xiao, G., Duncombe, P.R., Lecoeur, P., Trouilloud, P., Wang, Y.Y., Dravid, V.P., Sun, J.Z.: Grain-boundary effects on the magnetoresistance properties of perovskite manganite films. Phys. Rev. B 54, R15629 (1996) 81. Ziese, M.: Extrinsic magnetotransport phenomena in ferromagnetic oxides. Rep. Prog. Phys. 65, 143 (2002) 82. Tedrow, P.M., Meservey, R.: Spin-dependent tunneling into ferromagnetic nickel. Phys. Rev. Lett. 26, 192 (1971) 83. Tedrow, P.M., Meservey, R.: Spin polarization of electrons tunneling from films of Fe, Co, Ni, and Gd. Phys. Rev. B 7, 318 (1973) 84. Meservey, R., Tedrow, P.M.: Spin-polarized electron tunneling. Phys. Rep. 238, 173 (1994) 85. Soulen, R.J. Jr., Byers, J.M., Osofsky, M.S., Nadgorny, B., Ambrose, T., Cheng, S.F., Broussard, P.R., Tanaka, C.T., Nowak, J., Moodera, J.S., Barry, A., Coey, J.M.D.: Measuring the spin polarization of a metal with a superconducting point contact. Science 282, 85 (1998) 86. Bugoslavsky, Y., Miyoshi, Y., Clowes, S.K., Branford, W.R., Lake, M., Brown, I., Caplin, A.D., Cohen, L.F.: Possibilities and limitations of point-contact spectroscopy for measurements of spin polarization. Phys. Rev. B 71, 104523 (2005) 87. Mazin, I.I.: How to define and calculate the degree of spin polarization in ferromagnets. Phys. Rev. Lett. 83, 1427 (1999) 88. Julliere, M.: Tunneling between ferromagnetic films. Phys. Lett. 54A, 225 (1975) 89. Moodera, J.S., Kinder, L.R., Wong, T.M., Meservey, R.: Large magnetoresistance at room temperature in ferromagnetic thin film tunnel junctions. Phys. Rev. Lett. 74, 3273 (1995) 90. Åkerman, J.J., Slaughter, J.M., Dave, R.W., Schuller, I.K.: Tunneling criteria for magneticinsulator-magnetic structures. Appl. Phys. Lett. 79, 3104 (2001) 91. Bowen, M., Bibes, M., Barthélémy, A., Contour, J.-P., Anane, A., Lemaître, Y., Fert, A.: Nearly total spin polarization in La2/3 Sr1/3 MnO3 from tunneling experiments. Appl. Phys. Lett. 82, 233 (2003) 92. Faure-Vincent, J., Tiusan, C., Jouguelet, E., Canet, F., Sajieddine, M., Bellouard, C., Popova, E., Hehn, M., Montaigne, F., Schuhl, A.: High tunnel magnetoresistance in epitaxial Fe/MgO/Fe tunnel junctions. Appl. Phys. Lett. 82, 4507 (2003) 93. Johnson, P.D., Güntherodt, G.: Spin-polarized photoelectron spectroscopy as a probe of magnetic systems. In: Kronmüller, H., Parkin, S.S.P. (eds.) Handbook of Magnetism and Advanced Magnetic Materials, pp. 1635–1657. Wiley, Chichester (2007) 94. Monsma, D.J., Parkin, S.S.P.: Spin polarization of tunneling current from ferromagnet/Al2 O3 interfaces using copper-doped aluminum superconducting films. Appl. Phys. Lett. 77, 720 (2000) 95. Yates, K.A., Branford, W.R., Magnus, F., Miyoshi, Y., Morris, B., Cohen, L.F., Sousa, P.M., Conde, O., Silvestre, A.J.: The spin polarization of CrO2 revisited. Appl. Phys. Lett. 91, 172504 (2007) 96. Hu, G., Suzuki, Y.: Negative spin polarization of Fe3 O4 in magnetite/manganite-based junctions. Phys. Rev. Lett. 89, 276601 (2002) 97. Nadgorny, B., Osofsky, M.S., Singh, D.J., Woods, G.T., Soulen, R.J. Jr., Lee, M.K., Bu, S.D., Eom, C.B.: Measurements of spin polarization of epitaxial SrRuO3 thin films. Appl. Phys. Lett. 82, 427 (2003) 98. Sinkovi´c, B., Shekel, E., Hulbert, S.L.: Spin-resolved iron surface density of states. Phys. Rev. B 52, R8696 (1995) 99. Rampe, A., Hartmann, D., Weber, W., Popovic, S., Reese, M., Güntherodt, G.: Induced spin polarization and interlayer exchange coupling of the systems Rh/Co(0001) and Ru/Co(0001). Phys. Rev. B 51, 3230 (1995) 100. Eib, W., Alvarado, S.F.: Spin-polarized photoelectrons from nickel single crystals. Phys. Rev. Lett. 37, 444 (1976) 101. Park, J.-H., Vescovo, E., Kim, H.-J., Kwon, C., Ramesh, R., Venkatesan, T.: Magnetic properties at surface boundary of a half-metallic ferromagnet La0.7 Sr0.3 MnO3 . Phys. Rev. Lett. 81, 1953 (1998)

    9 Magnetotransport

    471

    102. Dedkov, Y.S., Fonin, M., König, C., Rüdiger, U., Güntherodt, G., Senz, S., Hesse, D.: Roomtemperature observation of high-spin polarization of epitaxial CrO2 (100) island films at the fermi energy. Appl. Phys. Lett. 80, 4181 (2002) 103. Fonin, M., Pentcheva, R., Dedkov, Y.S., Sperlich, M., Vyalikh, D.V., Scheffler, M., Rüdiger, U., Güntherodt, G.: Surface electronic structure of the Fe3 O4 (100): evidence of a half-metal to metal transition. Phys. Rev. B 72, 104436 (2005) 104. Tsymbal, E.Y., Mryasov, O.N., LeClair, P.R.: Spin-dependent tunnelling in magnetic tunnel junction. J. Phys.: Condens. Matter 15, R109 (2003) 105. Moodera, J.S., Nowak, J., van de Veerdonk, R.J.M.: Interface magnetism and spin wave scattering in ferromagnet-insulator-ferromagnet tunnel junctions. Phys. Rev. Lett. 80, 2941 (1998) 106. Brinkman, W.F., Dynes, R.C., Rowell, J.M.: Tunneling conductance of asymmetrical barriers. J. Appl. Phys. 41, 1915 (1970) 107. De Teresa, J.M., Barthélémy, A., Fert, A., Contour, J.P., Lyonnet, R., Montaigne, F., Seneor, P., Vaurès, A.: Inverse tunnel magnetoresistance in Co/SrTiO3 /La0.7 Sr0.3 MnO3 : new ideas on spin-polarized tunneling. Phys. Rev. Lett. 82, 4288 (1999) 108. Jo, M.-H., Mathur, N.D., Todd, N.K., Blamire, M.G.: Very large magnetoresistance and coherent switching in half-metallic manganite tunnel junctions. Phys. Rev. B 61, R14905 (2000) 109. Mathon, J., Umerski, A.: Theory of tunneling magnetoresistance of an epitaxial Fe/MgO/Fe(001) junction. Phys. Rev. B 63, 220403(R) (2001) 110. 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 (2004) 111. Parkin, S.P.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., Samant, M., Yang, S.-H.: Giant tunnelling magnetoresistance at room temperature with MgO (100) tunnel barriers. Nat. Mater. 3, 862 (2004) 112. Djayaprawira, D.D., Tsunekawa, K., Nagai, M., Maehara, H., Yamagata, S., Watanabe, N., Yuasa, S., Suzuki, Y., Ando, K.: 230% room-temperature magnetoresistance in CoFeB/MgO/CoFeB magnetic tunnel junctions. Appl. Phys. Lett. 86, 092502 (2005) 113. Kaiser, C., Panchula, A.F., Parkin, S.S.P.: Finite tunneling spin polarization at the compensation point of rare-earth-metal-transition-metal alloys. Phys. Rev. Lett. 95, 047202 (2005) 114. Wiesendanger, R.: Single-atom magnetometry. Curr. Opin. Solid State Mater. Sci. 15, 1 (2011) 115. Ding, H.F., Wulfhekel, W., Kirschner, J.: Ultra sharp domain walls in the closure domain pattern of Co(0001). Europhys. Lett. 57, 100 (2002) 116. Oka, H., Tao, K., Wedekind, S., Rodary, G., Stepanyuk, V.S., Sander, D., Kirschner, J.: Spatially modulated tunnel magnetoresistance on the nanoscale. Phys. Rev. Lett. 107, 187201 (2011) 117. Ferriani, P., von Bergmann, K., Vedmedenko, E.Y., Heinze, S., Bode, M., Heide, M., Bihlmayer, G., Blügel, S., Wiesendanger, R.: Atomic-scale spin spiral with a unique rotational sense: Mn monolayer on W(001). Phys. Rev. Lett. 101, 027201 (2008) 118. 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) 119. Coey, J.M.D., Berkowitz, A.E., Balcells, L., Putris, F.F., Parker, F.T.: Magnetoresistance of magnetite. Appl. Phys. Lett. 72, 734 (1998) 120. Coey, J.M.D., Berkowitz, A.E., Balcells, L., Putris, F.F., Barry, A.: Magnetoresistance of chromium dioxide powder compacts. Phys. Rev. Lett. 80, 3815 (1998) 121. Coey, J.M.D.: Powder magnetoresistance. J. Appl. Phys. 85, 5576 (1999) 122. Francis, T.L., Mermer, Ö., Veeraraghavan, G., Wohlgenannt, M.: Large magnetoresistance at room temperature in semiconducting polymer sandwich devices. New J. Phys. 6, 185 (2004) 123. Bloom, F.L., Wagemans, W., Kemerink, M., Koopmans, B.: Separating positive and negative magnetoresistance in organic semiconductor devices. Phys. Rev. Lett. 99, 257201 (2007)

    472

    M. Ziese

    124. Bergeson, J.D., Prigodin, V.N., Lincoln, D.M., Epstein, A.J.: Inversion of magnetoresistance in organic semiconductors. Phys. Rev. Lett. 100, 067201 (2008) 125. Hu, B., Wu, Y.: Tuning magnetoresistance between positive and negative values in organic semiconductors. Nat. Mater. 6, 985 (2007) 126. Wohlgenannt, M., Bobbert, P.A., Koopmans, B.: Intrinsic magnetic field effects in organic semiconductors. MRS Bull. 39, 590 (2014) 127. Majumdar, S., Lill, J.-O., Rajander, J., Majumdar, H.: Observation of ferromagnetic ordering in conjugated polymers exhibiting OMAR effect. Org. Electron. 21, 66 (2015) 128. Landauer, R.: Conductance determined by transmission: probes and quantised constriction resistance. J. Phys.: Condens. Matter 1, 8099 (1989) 129. Wharam, D.A., Thornton, T.J., Newbury, R., Pepper, M., Ahmed, H., Frost, J.E.F., Hasko, D.G., Peacock, D.C., Ritchie, D.A., Jones, G.A.C.: One-dimensional transport and the quantisation of the ballistic resistance. J. Phys. C: Solid State Phys. 21, L209 (1988) 130. Ohnishi, H., Kondo, Y., Takayanagi, K.: Quantized conductance through individual rows of suspended gold atoms. Nature 395, 780 (1998) 131. Ono, T., Ooka, Y., Miyajima, H., Otani, Y.: 2e2 /h to e2 /h switching of quantum conductance associated with a change in nanoscale ferromagnetic domain structure. Appl. Phys. Lett. 75, 1622 (1999) 132. García, N., Muñoz, M., Zhao, Y.-W.: Magnetoresistance in excess of 200% in ballistic ni nanocontacts at room temperature and 100 oe. Phys. Rev. Lett. 82, 2923 (1999) 133. Versluijs, J.J., Bari, M.A., Coey, J.M.D.: Magnetoresistance of half-metallic oxide nanocontacts. Phys. Rev. Lett. 87, 026601 (2001) 134. Chung, S.H., Muñoz, M., García, N., Egelhoff, W.F., Gomez, R.D.: Universal scaling of ballistic magnetoresistance in magnetic nanocontacts. Phys. Rev. Lett. 89, 287203 (2002) 135. Céspedes, O., Watts, S.M., Coey, J.M.D., Dörr, K., Ziese, M.: Magnetoresistance and electrical hysteresis in stable half-metallic La0.7 Sr0.3 MnO3 and Fe3 O4 nanoconstrictions. Appl. Phys. Lett. 87, 083102 (2005) 136. Bruno, P.: Geometrically constrained magnetic wall. Phys. Rev. Lett. 83, 2425 (1999) 137. Tatara, G., Zhao, Y.-W., Muñoz, M., García, N.: Domain wall scattering explains 300% ballistic magnetoconductance of nanocontacts. Phys. Rev. Lett. 83, 2030 (1999) 138. Gregg, J.F., Petej, I., Jouguelet, E., Dennis, C.: Spin electronics – a review. J. Phys. D: Appl. Phys. 35, R121 (2002) 139. Yakushiji, K., Mitani, S., Ernult, F., Takanashi, K., Fujimori, H.: Spin-dependent tunneling and coulomb blockade in ferromagnetic nanoparticles. Phys. Rep. 451, 1 (2007) 140. Yuasa, S., Nagahama, T., Suzuki, Y.: Spin-polarized resonant tunneling in magnetic tunnel junctions. Science 297, 234 (2002) 141. Lüders, U., Bibes, M., Bouzehouane, K., Jacquet, E., Contour, J.-P., Fusil, S., Bobo, J.-F., Fontcuberta, J., Barthélémy, A., Fert, A.: Spin filtering through ferrimagnetic NiFe2 O4 tunnel barriers. Appl. Phys. Lett. 88, 082505 (2006) 142. Garcia, V., Bibes, M., Bocher, L., Valencia, S., Kronast, F., Crassous, A., Moya, X., EnouzVedrenne, S., Gloter, A., Imhoff, D., Deranlot, C., Mathur, N.D., Fusil, S., Bouzehouane, K., Barthélémy, A.: Ferroelectric control of spin polarization. Science 327, 1106 (2010) 143. Gajek, M., Bibes, M., Fusil, S., Bouzehouane, K., Fontcuberta, J., Barthélémy, A., Fert, A.: Tunnel junctions with multiferroic barriers. Nat. Mater. 6, 296 (2007) 144. Schelp, L.F., Fert, A., Fettar, F., Holody, P., Lee, S.F., Maurice, J.L., Petroff, F., Vaurès, A.: Spin-dependent tunneling with coulomb blockade. Phys. Rev. B 56, R5747 (1997) 145. Guéron, S., Deshmukh, M.M., Myers, E.B., Ralph, D.C.: Tunneling via individual electronic states in ferromagnetic nanoparticle. Phys. Rev. Lett. 83, 4148 (1999) 146. Graf, H., Vancea, J., Hoffmann, H.: Single-electron tunneling at room temperature in cobalt nanoparticles. Appl. Phys. Lett. 80, 1264 (2002) 147. Barna´s, J., Fert, A.: Magnetoresistance oscillations due to charging effects in double ferromagnetic tunnel junctions. Phys. Rev. Lett. 80, 1058 (1998) 148. Barna´s, J., Fert, A.: Effects of spin accumulation on single-electron tunneling in a double ferromagnetic microjunction. Europhys. Lett. 44, 85 (1998)

    9 Magnetotransport

    473

    149. Barna´s, J., Fert, A.: Interplay of spin accumulation and coulomb blockade in double ferromagnetic junctions. J. Magn. Magn. Mater. 192, L391 (1999) 150. Esaki, L.: New phenomenon in narrow Germanium p − n junctions. Phys. Rev. 109, 603 (1958) 151. Itoh, H., Inoue, J., Umerski, A., Mathon, J.: Quantum oscillation of magnetoresistance in tunneling junctions with a nonmagnetic spacer. Phys. Rev. B 68, 174421 (2003) 152. Moodera, J.S., Santos, T.S., Nagahama, T.: The phenomena of spin-filter tunnelling. J. Phys.: Condens. Matter 19, 165202 (2007) 153. Kok, M., Beukers, J.N., Brinkman, A.: Spin-polarized tunneling through a ferromagnetic insulator. J. Appl. Phys. 105, 07C919 (2009) 154. Caffrey, N.M., Fritsch, D., Archer, T., Sanvito, S., Ederer, C.: Spin-filtering efficiency of ferrimagnetic spinels CoFe2 O4 and NiFe2 O4 . Phys. Rev. B 87, 024419 (2013) 155. Santos, T.S., Moodera, J.S.: Observation of spin filtering with a ferromagnetic EuO tunnel barrier. Phys. Rev. B 69, 241203(R) (2004) 156. Gajek, M., Bibes, M., Varela, M., Fontcuberta, J., Herranz, G., Fusil, S., Bouzehouane, K., Barthélémy, A., Fert, A.: La2/3 Sr1/3 MnO3 -La0.1 Bi0.9 MnO3 heterostructures for spin filtering. J. Appl. Phys. 99, 08E504 (2006) 157. Tsymbal, E.Y., Kohlstedt, H.: Tunneling across a ferroelectric. Science 313, 181 (2006) 158. Garcia, V., Bibes, M.: Ferroelectric tunnel junctions for information storage and processing. Nat. Commun. 5, 4289 (2014) 159. Kohlstedt, H., Petraru, A., Szot, K., Rüdiger, A., Meuffels, P., Haselier, H., Waser, R., Nagarajan, V.: Method to distinguish ferroelectric from nonferroelectric origin in case of resistive switching in ferroelectric capacitors. Appl. Phys. Lett. 92, 062907 (2008) 160. Pantel, D., Haidong, H., Goetze, S., Werner, P., Kim, D.J., Gruverman, A., Hesse, D., Alexe, M.: Tunnel electroresistance in junctions with ultrathin ferroelectric Pb(Zr0.2 Ti0.8 )O3 barriers. Appl. Phys. Lett. 100, 232902 (2012) 161. Pantel, D., Goetze, S., Hesse, D., Alexe, M.: Reversible electrical switching of spin polarization in multiferroic tunnel junctions. Nat. Mater. 11, 289 (2012) 162. Quindeau, A., Fina, I., Marti, X., Apachitei, G., Ferrer, P., Nicklin, C., Pippel, E., Hesse, D., Alexe, M.: Four-state ferroelectric spin-valve. Sci. Rep. 5, 9749 (2015) 163. Béa, H., Gajek, M., Bibes, M., Barthélémy, A.: Spintronics with multiferroics. J. Phys.: Condens. Matter 20, 434221 (2008) 164. Bibes, M., Villegas, J.E., Barthélémy, A.: Ultrathin oxide films and interfaces for electronics and spintronics. Adv. Phys. 60, 5 (2011) 165. Fusil, S., Garcia, V., Barthélémy, A., Bibes, M.: Magnetoelectric devices for spintronics. Annu. Rev. Mater. Res. 44, 91 (2014) 166. Hall, E.H.: On a new action of the magnet on electric currents. Am. J. Math. 2, 287 (1879) 167. Pugh, E.M., Rostoker, N.: Hall effect in ferromagnetic materials. Rev. Mod. Phys. 25, 151 (1953) 168. Ziese, M., Vrejoiu, I.: Anomalous and planar Hall effect of orthorhombic and tetragonal SrRuO3 layers. Phys. Rev. B 84, 104413 (2011) 169. Matsuno, J., Ogawa, N., Yasuda, K., Kagawa, F., Koshibae, W., Nagaosa, N., Tokura, Y., Kawasaki, M.: Interface-driven topological Hall effect in SrRuO3 -SrIrO3 bilayer. Sci. Adv. 2, e1600304 (2016) 170. Hurd, C.M.: Pressing electricity. In: Chien, C.L., Westgate, C.R. (eds.) The Hall Effect and Its Applications, pp. 1–54. Plenum Press, New York (1980) 171. Smit, J.: The spontaneous Hall effect in ferromagnetics I. Physica 21, 877 (1955) 172. Smit, J.: The spontaneous Hall effect in ferromagnetics II. Physica 24, 39 (1958) 173. Berger, L.: Side-jump mechanism for the Hall effect of ferromagnets. Phys. Rev. B 2, 4559 (1970) 174. Lyanda-Geller, Y., Chun, S.H., Salamon, M.B., Goldbart, P.M., Han, P.D., Tomioka, Y., Asamitsu, A., Tokura, Y.: Charge transport in manganites: hopping conduction, the anomalous hall effect, and universal scaling. Phys. Rev. B 63, 184426 (2001)

    474

    M. Ziese

    175. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A.H., Ong, N.P.: Anomalous Hall effect. Rev. Mod. Phys. 82, 1539 (2010) 176. Fang, Z., Nagaosa, N., Takahashi, K.S., Asamitsu, A., Mathieu, R., Ogasawara, T., Yamada, H., Kawasaki, M., Tokura, Y., Terakura, K.: The anomalous Hall effect and magnetic monopoles in momentum space. Science 302, 92 (2003) 177. Chen, Y., Bergman, D.L., Burkov, A.A.: Weyl fermions and the anomalous Hall effect in metallic ferromagnets. Phys. Rev. B 88, 125110 (2013) 178. Neubauer, A., Pfleiderer, C., Binz, B., Rosch, A., Ritz, R., Niklowitz, P.G., Böni, P.: Topological Hall effect in the A phase of MnSi. Phys. Rev. Lett. 102, 186602 (2009) 179. Li, Y., Xiong, P., von Molnár, S., Wirth, S., Ohno, Y., Ohno, H.: Hall magnetometry on a single iron nanoparticle. Appl. Phys. Lett. 80, 4644 (2002) 180. Zeldov, E., Majer, D., Konczykowski, M., Geshkenbein, V.B., Vinokur, V.M., Shtrikman, H.: Thermodynamic observation of first-order vortex-lattice melting transition in Bi2 Sr2 CaCu2 O8 . Nature 375, 373 (1995) 181. Zhang, S.: Spin Hall effect in the presence of spin diffusion. Phys. Rev. Lett. 85, 393 (2000) 182. Dyakonov, M.I., Perel, V.I.: Current-induced spin orientation of electrons in semiconductors. Phys. Lett. 35A, 459 (1971) 183. Dyakonov, M.I., Khaetskii, A.V.: Spin Hall effect. In: Dyakonov, M.I. (ed.) Spin Physics in Semiconductors, pp. 211–243. Springer, Berlin/Heidelberg (2008) 184. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834 (1999) 185. Sinova, J., Valenzuela, S.O., Wunderlich, J., Back, C.H., Jungwirth, T.: Spin Hall effects. Rev. Mod. Phys. 87, 1213 (2015) 186. Kato, Y.K., Myers, R.C., Gossard, A.C., Awschalom, D.D.: Observation of the spin Hall effect in semiconductors. Science 306, 1910 (2004) 187. Kimura, T., Otani, Y., Sato, T., Takahashi, S., Maekawa, S.: Room-temperature reversible spin Hall effect. Phys. Rev. Lett. 98, 156601 (2007) 188. Valenzuela, S.O., Tinkham, M.: Direct electronic measurement of the spin Hall effect. Nature 442, 176 (2006) 189. 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) 190. Miao, B.F., Huang, S.Y., Qu, D., Chien, C.L.: Inverse spin Hall effect in a ferromagnetic metal. Phys. Rev. Lett. 111, 066602 (2013) 191. Bernevig, B.A., Zhang, S.-C.: Quantum spin Hall effect. Phys. Rev. Lett. 96, 106802 (2006) 192. König, M., Wiedmann, S., Brüne, C., Roth, A., Buhmann, H., Molenkamp, L.W., Qi, X.-L., Zhang, S.-C.: Quantum spin Hall insulator state in HgTe quantum wells. Science 318, 766 (2007) 193. Hasan, M.Z., Kane, C.L.: Topological insulators. Rev. Mod. Phys. 82, 3045 (2010) 194. Beach, R.S., Berkowitz, A.E.: Giant magnetic field dependent impedance of amorphous FeCoSiB wire. Appl. Phys. Lett. 64, 3652 (1994) 195. Panina, L.V., Mohri, K.: Magneto-impedance effect in amorphous wires. Appl. Phys. Lett. 65, 1189 (1994) 196. Phan, M.-H., Peng, H.-X.: Giant magnetoimpedance materials: fundamentals and applications. Progress Mater. Sci. 53, 323 (2008) 197. Panina, L.V., Mohn, K., Uchyama, T., Noda, M.: Giant magneto-impedance in Co-rich amorphous wires and films. IEEE Trans. Magn. 31, 1249 (1995) 198. Yoon, S.S., Kim, C.G.: Separation of reversible domain-wall motion and magnetization rotation components in susceptibility spectra of amorphous magnetic materials. Appl. Phys. Lett. 78, 3280 (2001) 199. Chen, D.-X., Muñoz, J.L.: Ac impedance and circular permeability of slab and cylinder. IEEE Trans. Magn. 35, 1906 (1999) 200. Yelon, A., Ménard, D., Britel, M., iureanu, P.: Calculations of giant magnetoimpedance and of ferromagnetic resonance response are rigorously equivalent. Appl. Phys. Lett. 69, 3084 (1996)

    9 Magnetotransport

    475

    201. Montgomery, H.C.: Method for measuring electrical resistivity of anisotropic materials. J. Appl. Phys. 42, 2971 (1971) 202. van der Pauw, L.J.: A method of measuring the resistivity and Hall coefficient on lamellae of arbitrary shape. Philips Tech. Rev. 20, 220 (1959) 203. Büttiker, M.: Symmetry of electrical conduction. IBM J. Res. Develop. 32, 317 (1988) 204. Johnson, M.: Spin accumulation in gold films. Phys. Rev. Lett. 70, 2142 (1993)

    Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films Mark L. M. Lalieu

    10

    and Bert Koopmans

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magneto-Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Basics of Magneto-Optics from a Macroscopic Perspective . . . . . . . . . . . . . . . . . . . . . . . . . Micropic Understanding of Magneto-Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Measuring Magnetism Using MOKE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Measuring Ultrafast Magnetization Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism . . . . . . . . . . . . . . . . Conceptual Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ultrafast Laser-Induced Loss of Magnetic Order . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . All-Optical Switching of Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Laser-Pulse-Excited Spin Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    478 480 480 485 488 497 502 502 505 521 527 534 537

    Abstract

    Optical tools have played an important role in the field of thin film magnetism, and the spintronics that has emerged from it. Initially, the role of optics was limited to providing sensitive and versatile magneto-optical probes of the magnetic behavior. In the 1990s, it became clear that ultrashort laser pulses can also be successfully used in a reverse approach to manipulate magnetic order at down to femtosecond time scales, an approach we refer to as opto-magnetism. This chapter provides a basic introduction and a review of developments in both magneto-optics and opto-magnetism, as mostly applied to metallic thin film ferri- and ferromagnetic systems. In the second part of the chapter, after a

    M. L. M. Lalieu · B. Koopmans () Department of Applied Physics, Eindhoven University of Technology, Eindhoven, The Netherlands 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_10

    477

    478

    M. L. M. Lalieu and B. Koopmans

    conceptual introduction of ultrafast laser-induced magnetization dynamics, three processes are treated in more detail: laser-induced loss of magnetic order, alloptical switching of magnetization, and pulsed laser-induced spin currents.

    Introduction This chapter discusses how light can be used to unveil magnetic ordering phenomena in thin film systems but also how it has become a unique tool to modify magnetism at unprecedentedly short time scales. Key to probing magnetism is the notion that polarized light changes its polarization state when interacting with a magnetically ordered material. This was discovered by the English scientist Michael Faraday in 1845. He found that the plane of polarization is rotated when light propagates through a magnetized material [1]. The effect is nonreciprocal, i.e., it changes sign upon reversal of the magnetization, it is cumulative in the sense that the polarization rotation is proportional to the optical path length through the material under investigation, and it has its origin in the different refractive index experienced by left-handed and right-handed circularly polarized light. A similar effect occurs when light is reflected from a magnetized material. However, in reflection it lacks the cumulative behavior, and therefore the rotation is generally much smaller. For this reason, it was only discovered 30 years later in 1877 by the Scottish scientist John Kerr [2]. The effects in transmission and reflection are referred to as the Faraday effect and the magneto-optical (MO) Kerr effect (MOKE), respectively. Nowadays, MOKE has become one of the most versatile and widely used characterization techniques in research on (thin film) ferromagnetic metals. Its popularity got a strong impulse in the 1980s with the advance of thin film magnetism that lead to the birth of spintronics. The complementary approach, using light not to probe magnetism, but to manipulate magnetic order, has a much shorter history. Among the simplest applications is heating a ferromagnetic thin film by a focused laser spot and thereby reducing the magnetic moment according to equilibrium thermodynamics. Since the magnetic anisotropy has an even stronger temperature dependence than the magnetization itself, this approach can be used for thermomagnetic writing. It was applied e.g. in the MO recording strategies investigated in the 1990s [3]. Apart from this technological drive, researchers were challenged by the question as to how rapidly the magnetization could be quenched upon sudden laser heating. The pioneering experiments on this matter were conducted by Beaurepaire and co-workers in 1996 [4]. The authors investigated laser-induced magnetization dynamics in the strongly nonequilibrium regime using femtosecond laser pulses. They found that the demagnetization in ferromagnetic nickel thin films was completed well within a picosecond, much faster than expected; see Fig. 1a. This observation has become a benchmark in the field. Unfortunately, both the first author of the seminal paper, Eric Beaurepaire, and his co-author Jean-Yves Bigot passed away in 2018.

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    479

    Fig. 1 Pioneering examples of ultrafast opto-magnetism – the reverse of magneto-optics. (a) Iconic measurement by Beaurepaire et al. showing the measured MO contrast as a function of pump-probe delay time for a Ni thin film. It provided the first evidence for sub-ps loss of magnetization after laser heating with a fs laser pulse. Figure taken from Ref. [4]. (b) Artistic impression of the first ever reported all-optical switching of magnetization, showing alternatively up (white) and down (gray) magnetic domains written by laser pulses with opposite helicities, based on a real data set. (Figure taken from Ref. [5] with kind permission from APS)

    Fueled by the novel insight, a new field in magnetism emerged. Much effort is still being devoted to understanding the ultrafast laser-induced demagnetization, but this phenomenon turned out just to be a tip of an iceberg. Many new mechanisms to manipulate ferromagnetic matter by short pulses of light have been discovered since – a field that we will broadly refer to as opto-magnetism. Among the highlights was the demonstration in 2007 that a single fs laser pulse can be used to “write” the magnetic moment of ferrimagnetic GdFeCo deterministically, where the final direction is merely determined by the helicity of the circularly polarized light [5], Fig. 1b. Soon thereafter, an alternative form of all-optical switching (AOS) of the magnetization was found, which is independent of the laser polarization and leads to a toggle-type behavior that reverses the magnetization at the impact of each laser pulse [6, 7]. Furthermore, while AOS was originally restricted to ferrimagnetic alloys (materials with spin subsystems with oppositely oriented magnetizations), very similar results have been found for synthetic ferrimagnets [8], i.e., multilayered thin films composed of layers with opposite magnetization and even simple ferromagnetic films with perpendicular magnetic anisotropy, such as Pt/Co/Pt [9]. In all these cases, however, the AOS turned out not be the result of a single fs pulse but a cumulative effect needing many pulses. In parallel to these discoveries, it was found that spin currents are produced when ultrashort laser pulses are absorbed in ferromagnetic thin films [10, 11]. The spin angular momentum carried by these currents can affect the local magnetization dynamics [10] but can also exert a spin-transfer torque on a nearby, noncollinearly

    480

    M. L. M. Lalieu and B. Koopmans

    aligned magnetic layer [12, 13]. Altogether, the field of opto-magnetism is in a highly excited phase, where new discoveries are being made, while experiment and theory are collaborating to unravel the underlying microscopic mechanisms. This chapter provides an introduction to both magneto-optics and optomagnetism, particularly focusing on ferro- and ferrimagnetic metallic thin films. In section “Magneto-Optics”, we first introduce the basics of magneto-optical phenomena, both from a phenomenological and microscopic perspective. Then, we will treat a variety of magneto-optical schemes to study static magnetic order in (multilayered) thin film systems, followed by time-resolved extensions thereof employing pulsed pump-probe laser strategies. Specific implementations to obtain layer- or element-specific information and MO microscopy will be discussed. In section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism”, we switch to the use of laser pulses to excite magnetization dynamics. After an introduction of relevant concepts, we focus on three specific contemporary topics: ultrafast demagnetization, all-optical switching, and laser-induced spin currents. We conclude with some final remarks and a brief outlook in section “Conclusions and Outlook”.

    Magneto-Optics In this section, the principles of the magneto-optical Kerr effect are introduced, both from a phenomenological and a microscopic perspective. Different schemes for MOKE measurements on metallic thin films are explained, both for static and ultrafast time-resolved studies. Ways to separate magnetic signal from different layers or elements, and Kerr microscopy are addressed.

    Basics of Magneto-Optics from a Macroscopic Perspective Magneto-optics describes the interaction of light with magnetized matter, leading to a change in its intensity or polarization state. In this chapter, we mostly focus on the application to ferro- and ferrimagnetic metals, which have an optical skin depth of typically 10–20 nm only. Experiments are thus limited to the reflection geometry or using thin enough films on top of an optically transparent substrate. For ultrathin films, the rotation of the polarization is typically of the order of tens of millidegrees only. Yet, they have been proven to provide a versatile and very sensitive measure of the magnetization using a variety of experimental schemes. As an intriguing example, minute laser-induced dynamic effects of less than a percent of change in the magnetization in sub-nm ferromagnetic films can be routinely measured with high signal to noise, as will be discussed throughout this chapter. In such experiments, the sensitivity to the optical polarization is typically 10−7 − 10−8 rad, corresponding to rotating the holder of a 1 cm polarizer by no more than a nm at its circumference. In this chapter, we will focus on reflection studies, unless otherwise specified.

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    481

    Complementary to the refractive effects introduced before, magnetic circular dichroism (MCD) is an important manifestation of the interaction of light with magnetic matter. As an example, performing a Faraday experiment on a semiconductor magnetized along the propagation direction of the light, and choosing light with a photon energy above the band gap of the material, left- and right-handed circularly polarized light will be absorbed in a different way. Similarly, for ferromagnetic metals such circular dichroism provides a tool for measuring magnetic properties, provided the optical thickness of the specimen under consideration allows for transmission experiments. The fundamental interaction by light and matter is described most generically by the nonlocal and nonlinear optical susceptibility χ . It relates the induced dielectric polarization P to the electric field E in a rather informal notation via  P = 0

    χ

    (1)

    (r, r  ) · E(r  )dr  + 0

     χ

    (2)

    (r, r  , r  ) : E(r  )E(r  )dr  dr  + · · · ,

    (1) where χ is the n-th order optical susceptibility tensor of rank n + 1 and 0 is the permittivity of free space. Here, P (t) = P e−iωt denotes the induced polarization at optical frequency ω. In the right-hand side, only (combinations of) electrical fields E(t) = Ee−iωi t oscillating at (a set of) frequencies (ω1 , ..., ωn ) that contribute to a polarization at ω = ω1 + ... + ωn are considered. Although in this chapter we mainly consider the linear response (n = 1), we will briefly address (magnetizationinduced) optical second-harmonic generation (SHG) as a relevant nonlinear optical technique (n = 2). For completeness, Eq. (1) includes nonlocal optical response beyond the electric dipole approximation, such as electric quadrupole and magnetic dipole processes. However, such nonlocal extensions are considered to be beyond the scope of this chapter. Only considering the local and linear optical response, the dielectric tensor  of rank 2 is related to the (linear) optical susceptibility by (n)

      (1) ij = 0 δij + χij ,

    (2)

    where i and j ∈ {x, y, z} and δij is the Kronecker-delta function. For isotropic materials, ij = 0 δij . Here,  is the relative permittivity or dielectric constant. Note that in this convention  is dimensionless, whereas elements of the dielectric tensor ij have units C2 · N−1 · m−2 . The dielectric constant is related to the complex refractive index n˜ = n + iκ via  = n˜ 2 , in which n is the refractive index and κ is the extinction coefficient. In order to understand the basics of magneto-optics, we consider an optically isotropic material with uniform magnetization M. Then, the dielectric tensor reads: ⎛

    ⎞ xx xy xz  = ⎝ −xy xx yz ⎠ . −xz −zy xx

    (3)

    482

    M. L. M. Lalieu and B. Koopmans

    All these tensor elements are complex parameters, with a real part (representing an induced dielectric polarization in-phase with the driving field) and an imaginary part (a polarization 90 degrees out of phase). Furthermore, note that whereas the diagonal elements ii are even in M, the nonreciprocal, off-diagonal elements ij transform antisymmetrically under reversal of the magnetization, M ↔ −M. The latter property provides the “magnetic contrast,” being either a change of the polarization of the light after reflection or transmission or a change of dichroic signal, upon reversal of M. The off-diagonal element ij is related to the component ˆ jˆ. Thus, for spatially isotropic materials carrying a magnetization of M parallel to i× in the z-direction, the dielectric tensor reduces to ⎛

    ⎞ ⎛ ⎞ xx xy 0 1 iQ 0  = ⎝ −xy xx 0 ⎠ = xx ⎝ −iQ 1 0 ⎠ , 0 0 1 0 0 xx

    (4)

    where in the second step, we adopted the convention of introducing a magnetooptical parameter Q related to the dielectric tensor via xy = ixx Q. For any (multi)layered material consisting of an arbitrary number of layers, the MO Kerr and Faraday rotation at a specific angle of incidence are fully specified by the thickness and dielectric tensor elements of all the layers. The MO response is most conveniently calculated using a transfer matrix approach [14]. Here, we restrict ourselves to a few very simple cases. For convenience, we adapt a Jones matrix formalism [15], in which the light’s electrical polarization is described by a vector with two components, representing two orthogonal components of E. Unless otherwise specified, the two components are assumed to represent the s- and ppolarization, respectively, i.e., field components perpendicular (s, from the German senkrecht) and parallel (p) to the plan of incidence, as sketched in Fig. 2a. Within the Jones formalism, each optical operation (reflection, transmission, or propagation) acting on a plane wave can be represented by a 2 × 2 matrix. The resulting electrical field after the operation is then just calculated by multiplying the relevant matrix with the incident field vector. As a consequence of the complex nature of the dielectric tensor, the magnetization-induced change of the light’s polarization, referred to as the complex Kerr rotation θ˜ , is a complex quantity consisting of a real and imaginary part. Let us consider the simplest case of a polar configuration with light incident at arbitrary angle to an isotropic thin film sample having an out-of-plane magnetization. The Jones matrix of this sample can be written as

    rss rsp rps rpp

    ,

    (5)

    where the diagonal elements represent the ordinary reflection coefficients for sand p-polarized light. The off-diagonal elements give rise to the magneto-optical effects. They are (generally) orders of magnitude smaller than the diagonal ones

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    483

    polar

    longitudinal

    transverse

    (c)

    (d)

    (e)

    p s

    (a)

    (b)

    Fig. 2 (a) Definition of s- and p-polarization. (b) Illustration of the MO Kerr effect, in which linearly polarized light gains a Kerr rotation θ and ellipticity ε upon reflection of a magnetic thin film, with ε defined as the ratio of the minor to the major axis of the elliptically polarized light. (c-e) Three configurations of MOKE; (c) polar configuration, (d) longitudinal configuration, and (e) transversal configuration

    and change sign when reversing M. Considering the incident light to be s-polarized and normalizing all field amplitudes to the incident one, the reflected field can easily be derived to be

    rss rsp rps rpp





    1 1 . = rss 0 rps /rss

    (6)

    Thus, we found that the polarization has rotated by an angle ±θ˜ = ±θ + iε = rps /rss , where ± refers to a magnetization in the ±ˆz-direction and θ and ε are the Kerr rotation and ellipticity, respectively; see Fig. 2b. Readers familiar with ellipsometry will immediately realize the importance of having two complementary optical responses for extracting physical information from more complex (multi)layered systems. In section “Layer-Specific MOKE”, it will become clear how one can profit from this. It is of relevance to stress that upon p-polarized incidence, the complex Kerr rotation is given by ±θ˜ = rsp /rpp = rps /rss . This means that, in general, the MO signal will depend on the chosen polarization of the incident light. Aiming for a calculation of the (complex) MO Kerr and Faraday rotation, a more transparent description is obtained in terms of the eigen modes of Eq. (4), which are circularly polarized electro-magnetic waves. Considering propagation along z, we can define dielectric elements + and − , experienced by the right- and left circularly polarized eigen modes, respectively. They are related to xy via + − − = ixy .

    (7)

    In this notation, modes with opposite circular polarization experience a different complex refractive index, n˜ ± , according to Δn˜ ≡ n˜ + − n˜ − =

    ixy = −nQ. ˜ 0 n˜

    (8)

    484

    M. L. M. Lalieu and B. Koopmans

    Due to the difference between n˜ + and n˜ − and restricting ourselves to real values, right and left circularly polarized light will experience a relative phase difference while propagating in a direction collinear with M. Considering linearly polarized light, which is a linear superposition of circularly polarized light with opposite handedness,



    1 1 1 1 1 1 =√ √ +√ , (9) 0 i −i 2 2 2 propagating through a lossless medium (n˜ = n, κ = 0) over a distance d, one readily obtains for the transmitted electrical field, neglecting an overall time-dependent part e−iωt



    ein+ ωd/c 1 ein− ωd/c 1 ind cos(nQωd/2c) . (10) + =e sin(nQωd/2c) i −i 2 2 This result describes a rotation of the initial polarization that is proportional with d, which is just the Faraday effect. Similarly, it is straightforward to verify that dissipative media with magnetic circular dichroism, i.e., Δκ = 0, generally show both a polarization rotation and a change in ellipticity. Of more relevance in the context of thin film magnetism is the difference in the reflection coefficient experienced by right and left circularly polarized light, which for normal incidence on a semi-infinitely thick material magnetized perpendicular to the surface reads r± =

    1 − n˜ ± . 1 + n˜ ±

    (11)

    It can be easily verified that reflected linearly polarized light will experience a (complex) rotation of its polarization, given by θ˜ =

    inQ ˜ , −1

    (12)

    which is just the MO Kerr effect at normal incidence. As stated before, at nonnormal incidence, the complex Kerr rotation depends on the polarization of the incident light. Similar expressions for (multi)layered thin film systems and arbitrary configurations can readily be derived using the earlier mentioned transfer matrix method [14]. In the analysis so far, magneto-optics was described in terms of a dielectric tensor that incorporates the magnetization implicitly. Alternatively, one may linearize the M-dependence and introduce a material-dependent – but magnetizationindependent – tensor α accordingly. This description is particularly illustrative when, in section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism”, we will treat the so-called inverse-Faraday effect (IFE). For a spatially isotropic material with a finite magnetization, the induced dielectric polarization P can be written as

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    P = αE × M,

    485

    (13)

    where α is the diagonal element of α, and E is again the amplitude of the oscillating electric field, and M is the static magnetization vector. Equation (13) provides an intuitive picture of both the Faraday effect and MOKE. Assuming Mˆz and incident light with wave vector qM and polarization Ex, ˆ an induced oscillatory dielectric polarization P yˆ is generated, emitting electro-magnetic radiation coherent with the incident light but with a polarization orthogonal to it. Superposition of the transmitted or reflected incident wave with the orthogonally polarized one is observed as a rotation of the plane of polarization. To complete this phenomenological introduction of MO effects, three different configurations in which MOKE can be applied are introduced. They are distinguished by the orientation of the magnetization vector with respect to the sample’s surface normal and the plane of incidence of the light; see Fig. 2c-e. In the polar geometry, the magnetization vector is perpendicular to the magnetic thin film (or sample surface). In the longitudinal configuration, the magnetization is parallel to the film and lies within the plane of incidence. In this case, a non-normal incidence of the light needs to be chosen, since a normal incidence would result in qP , for which the MO effects vanish; see Eq. (13). This is in contrast to the polar Kerr effect, in which case perpendicular incidence can be considered the most simple and efficient configuration. Finally, the transverse configuration refers to the case where the magnetization vector is parallel to the magnetic film and perpendicular to the plane of incidence. This configuration requires p-polarized light, because the signal vanishes for pure s-polarized light (Eq. 13). In the transverse configuration, no change of polarization can be observed, but the Kerr effect manifests itself as a change in intensity upon reversal of the magnetization. As a rule of thumb, the polar effect is typically an order of magnitude stronger than the longitudinal effect.

    Micropic Understanding of Magneto-Optics The phenomenological analysis so far was purely based on symmetry. Aiming for a microscopic understanding, the elements of the dielectric tensor – both the diagonal and the off-diagonal ones – can be related to the electronic states of the material. Within the electric dipole approximation, the interaction between light and matter can be treated equivalently [16] using either an interaction Hamiltonian eE · r (sometimes referred to as velocity gauge) or −(e/m)A · p (length gauge), where e is the electron charge, m its mass, A is the vector potential, and r and p = −ih∇ ¯ are the position and momentum operator, respectively. For isolated atoms this leads to the well-known dipole selection rule ΔLz = ±1, where Lz is the z-component of the angular momentum operator, which allows for transitions between s- and p-states, and between p- and d-states, but not between s- and d-states. Within a single-particle description, the dielectric tensor elements can be expanded as a sum over contributions due to “vertical” optical transitions (k n = k n ), introducing matrix elements

    486

    M. L. M. Lalieu and B. Koopmans i Πnn  = Ψn |pi |Ψn ,

    (14)

    where |Ψn are single-particle electronic states with wave vector k n . Following Ref. [17], converting equations from c.g.s. to S.I. units and expressing elements of the dielectric tensor ij (ω) = 0 δij + iσij (ω)/ω rather than the optical conductivity σij (ω), one obtains ij (ω) = 0 δij +

    f (εn ) − f (εn ) Π i  Π j  e2 n n nn . 2V ω  ω ω − ω hm ¯ nn nn + i/τ 

    (15)

    n,n

    Here, the summation is over all states n (n ) with energy εn (εn ) and hω ¯ nn = εn − εn . Furthermore, f (εn ) refers to the Fermi-Dirac distribution function, and V is the volume of the crystal. Finite lifetime effects have been taken into account in a phenomenological way, by introducing a life time τ for all transitions. In order to capture the magneto-optics in a transparent way, the matrix elements can be written in an alternative form that reflects transitions by circularly polarized light, i.e., photons carrying a positive or negative quantum of angular momentum:   ± Πnn  = Ψn |pi − ipj |Ψn ,

    (16)

    for light propagating in the direction iˆ × jˆ. In this notation, the off-diagonal elements of the dielectric tensor read [17, 18]

    ij (ω) =

    ie2 2hm ¯ 2V ω

     2  j 2 occ. unocc.

    Πni  n  − Πn n  n

    n

    ωn2 n − ω2

    , with i = j,

    (17)

    where now the summation over n ranges over all occupied states and the summation over n over all empty states. The full response including the corresponding real part can be obtained by Kramers-Kronig transformation [18]. It can be shown that the quantity |Πn+ n |2 − |Πn− n |2 , relevant for the nonreciprocal off-diagonal elements ij , vanishes once summed over any full set of degenerate states if spin-orbit coupling is neglected.1 This reflects the fact that magneto-optics probes (a weighted average of) orbital angular momentum rather than spin ordering. Thus, although usually it is assumed that magneto-optics is sensitive to the spin ordering, this is only true in the case of finite spin-orbit coupling and assuming a strict correlation between spin and orbital momenta. Exact summation rules can be derived that show that the integrated spectral weight of a certain nonreciprocal tensor element corresponds to a component of the total orbital momentum vector [20]. More loosely speaking, one can say that magneto-optics measures the orbital moment, whereas the spin moment can only be derived by assuming a fixed relation between the two. This should be kept in mind particularly when performing

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    487

    experiments on strongly nonequilibrium systems, as will be discussed later in section “Ultrafast Laser-Induced Loss of Magnetic Order”. The inability to directly probe the spin moments is reflected in the MO response of elementary (3d transition-metal) ferromagnets and their alloys. In these materials, the orbital moments are strongly quenched by crystal-field effects, so their magnetization is strongly dominated by the spin ordering, while orbital moments provide only a small contribution. As a consequence, MO effects are relatively small. Examples of experimental MO spectra for the elementary ferromagnets nickel, cobalt, and iron, displaying xy as a function of photon energy, are depicted in Fig. 3. Features such as peaks in the spectrum can be assigned to interband transitions. Together with the zero crossings, they lead to the notion that the magnitude of the Kerr effect for a specific material and even its sign depend strongly on the photon energy. A simple intuitive picture explaining peaks in xy for a transition-metal ferromagnet is sketched in Fig. 4. Panel (a) shows a simplified density of states for the d-bands of spin-up and spin-down electrons, split by an exchange splitting of several tenths of eV. The bands further split up by a weaker spin-orbit coupling, described by

    a

    b

    0.8 Im

    3

    (eV 2)

    0.4

    xy

    /

    0.0 Re

    -0.2

    (h )2

    / xy

    (h )2

    Im

    2

    0

    0.2

    0

    (eV 2)

    0.6

    4

    -0.4

    0 Re

    -1

    -0.6 -0.8

    1

    0

    1

    2

    3

    4

    5

    -2

    6

    0

    1

    2

    c

    4

    5

    6

    4 Im

    (eV 2)

    2

    0

    3

    1

    (h )2

    xy

    /

    3

    h (eV)

    h (eV)

    0 -1 Re -2

    0

    1

    2

    3

    4

    5

    6

    h (eV)

    Fig. 3 Experimental MO spectra showing the square of the real (black) and imaginary (red) value of (h¯ ω)2 xy /0 for (a) nickel, (b) cobalt, and (c) iron, as a function of the photon energy h¯ ω. The data points are adopted from Ref. [21], and the solid lines are guides for the eye

    488

    M. L. M. Lalieu and B. Koopmans

    a

    b

    c

    Fig. 4 (a) Simplified spin-resolved density of states (up and down electrons), showing exchange and spin-orbit split d-bands and one of the p-states out of the broad sp-band. (b) Peaks in the imaginary part of the diagonal dielectric tensor element xx for spin-conserving transitions between p and d states for up and down electrons, as derived from (a). (c) Similar construction of the imaginary part of xy . The real parts, both in (b) and (c), can be obtained by Kramers-Kronig transformations. (Figure adapted from Refs. [22, 23])

    Hso = ξ L · S,

    (18)

    where ξ is the material-dependent spin-orbit coupling parameter. The splitting into Lz = −2 · · · + 2 sublevels is opposite for the up and down bands, as apparent from Eq. (18). Considering optical transitions between a single (degenerate) p-state out of the broad sp-band and the band of nondegenerate d-states and taking into account the dipole selection rule ΔLz = ±1 for right and left circularly polarized light, respectively, one readily derives the schematic spectra for the absorptive (imaginary) parts of the diagonal (xx ) and non-diagonal (+ − − = ixy ) dielectric tensor elements; see Fig. 4b, c, respectively. When interfacing the transition metals with heavy elements like Pt or Pd, spinorbit effects can be enhanced. This may lead to larger orbital moments with enhancement of the MO Kerr effect and other consequences such as perpendicular magnetic anisotropy. As an example, experimental values for the Kerr rotation of the polarization axis as observed in Co/Pt multilayers range typically up to 0.5◦ [24].

    Measuring Magnetism Using MOKE Different configurations for performing MOKE experiments will be briefly introduced, starting with a simple crossed-polarizer approach and followed by more sensitive schemes using a balanced photodiode and polarization modulation. Subsequent sections capture modes of operation that further enhance the applicability, addressing layer-specific studies, spectroscopic MOKE, Kerr microscopy, and other advanced approaches.

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    489

    Different Configurations We consider a reflection experiment with linearly polarized light, incident on a flat surface at a certain angle of incidence. The polarization state of the light will be expressed as an angle θin , according to tan(θin ) = Es /Ep , where s and p refer to the s- and p-polarized components (Fig. 2a). Upon reflection, the polarization state may be changed, and we denote the change of rotation by θ ≡ θout − θin , where “in” and “out” refer to the incoming and reflected light. It should be realized that a change of polarization state can already be induced without any MO interaction and even from an isotropic material. This generally occurs when the incident light is neither s- nor p-polarized. The resulting effects thereof are widely used in ellipsometry. For the analysis of MO measurements, it is convenient to distinguish between contributions to the total (complex) polarization rotation that transfer symmetrically (“S”) and antisymmetrically (“A”) under reversal of the magnetization M, θ˜T = θ˜S + θ˜A . Then, the complex Kerr rotation θ˜ = θ + i as introduced in section “Basics of Magneto-Optics from a Macroscopic Perspective” is obtained from θ˜ =

    θ˜T (M) − θ˜T (−M) 2

    (19)

    – a common experimental procedure to extract magnetic information in polarization-sensitive studies. In this manner, as will be demonstrated later (section “Measuring Ultrafast Magnetization Dynamics”), any nonmagnetic contribution to the optical signal in a MOKE measurement can be eliminated. In the generic case of a sample with a depth profile of the magnetization, and possibly having different magnetic sublattices that are labeled with an index i, the magnetization is described by M i (z) (with z ≥ 0 and as measured with respect to its surface at z = 0). For such a system, we can relate the measured (complex) Kerr rotation to the magnetization profile via θ˜ =

     i



    F˜ i (z) · M i (z)dz,

    (20)

    0

    where F˜ i (z) is a vector of generalized Fresnel coefficients, depending on all details of the experimental configuration, the wavelength of light being used, and the sample layout. F˜ i (z) determines the sensitivity to the magnetization of sublattice i at depth z and will generally decay at a length scale corresponding to the optical skin depth in the material. The simplest case occurs when (1) one measures a single magnetic film with uniform magnetization M along a known direction or (2) having chosen the configuration or sample layout such that all components of F˜ (z) are approximately zero, except for those of a single layer (and single sublattice) with uniform M i along a known direction. In that case, M i can be taken out of the integral, and the generic relation reduces to

    490

    M. L. M. Lalieu and B. Koopmans

    θ˜ = F˜ M.

    (21)

    This provides two complementary channels to measure the same magnetic parameter M, i.e., the real and imaginary part of θ˜ = θ +iε. Clearly, in order to optimize the signal-to-noise ratio, the channel with the largest component of F should be chosen. Moreover, optical components, such as wave plates, can be used to effectively measure an optimal linear combination of θ and ε. Having two complementary channels can be exploited in many ways. First of all, in the simple case described by Eq. (21), the complementary channels can be used to verify the assumptions that are made. For example, if reversing the inplane magnetization shows up differently in rotation and ellipticity in a longitudinal MOKE experiment, it could be a signature of a perpendicular component being mixed in due to a small misalignment of the applied magnetic field. More importantly, however, the complementary channels can be used to separately measure different magnetic components, such as the switching of two different magnetic layers (see section “Layer-Specific MOKE”), to differentiate between an out-ofplane and in-plane component of the magnetization or to distinguish two different magnetic sublattices in an alloy. The simplest realization of a MOKE experiment exploits a crossed polarizer configuration, as sketched in Fig. 5a. Usually, a collimated light source (often a CW or pulsed laser) is polarized (P1) before being focused on the sample, the latter to establish spatial resolution. The reflected light is collimated and sent through a polarizer (P2), dubbed the analyzer, set at an angle α nearly – but not exactly – crossed to the polarization axis of P1, i.e., α = |αP1 − αP2 | = π/2 + γ , for small γ . For simplicity, we assume P1 to be aligned along the s- or p-polarization axis. In that case, the intensity after the analyzer is given by I = I0 R(γ 2 + 2γ θ + θ 2 + ε2 ),

    (22)

    where I0 is the intensity of the incident light, R = rr ∗ is the reflection coefficient for the intensity, and further reflection and absorption losses of optical components

    a

    b

    c

    Fig. 5 Illustration of three different schemes for performing MOKE. (a) A crossed-polarizer configuration, in which an analyzer (P2) is almost crossed with a polarizer (P1). (b) A balanced photodiode configuration, in which a polarizing beam splitter (PBS) is adjusted so as to balance signals of two orthogonal polarization components. (c) A polarization modulation configuration, in which a PEM is used to establish high-frequency polarization modulation, and a lock-in amplifier detects signals at the fundamental and second-harmonic frequency of the modulation. A quarterwave plate (QWP) can be inserted to select a linear combination of rotation and ellipticity

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    491

    are neglected. The terms quadratic in θ and ε can usually be neglected because of the smallness of the Kerr effect and, moreover, are symmetric under reversal of M and thereby are difficult to distinguish from the first term. The 2γ θ term is the Kerr rotation signal of interest. While it looks as if the sensitivity increases upon opening the analyzer (i.e., increasing γ ), this comes at the cost of an increased background signal, described by the first, γ 2 , term. Thus, a compromise should generally be sought to optimize signal to noise. Finally, note that although according to Eq. (22) the crossed-polarizer configuration seems to be insensitive to the Kerr ellipticity, adding a quarter-wave plate (at an angle of 45◦ ) between the sample and the analyzer reverses the role of rotation and ellipticity. In an alternative approach, a polarizing beam splitter (PBS) is adjusted to balance the signal from two photodiodes, measuring orthogonal polarization components, as sketched in Fig. 5b. In this balanced photodiode scheme, with the beam splitter set such that it splits components with polarization angles of +45◦ and −45◦ , the measured intensity simply reads (neglecting terms quadratic in θ and ε) I = I0 Rθ.

    (23)

    In comparison with the crossed-polarizer configuration, where noise in the intensity of the light source enters via the γ 2 term, the balanced photodetector approach is – in lowest order – just limited by electronic noise and noise related to the light detection. Another way of enhancing signal-to-noise is by using polarization modulation, as depicted in Fig. 5c. A popular version uses a high-frequency photo-elastic modulator (PEM), typically operating at f = 50 kHz. The PEM is installed before the sample, and the incident light has its linear polarization axis making an angle of 45◦ with the main axis of the modulator. The Jones matrix of the PEM is given by

    1 0 , (24) 0 eiA0 cos(2πf t) where A0 determines the amplitude of the oscillatory retardation between the two orthogonal polarization components of the light. Setting A0 = π/2, the light after the PEM oscillates between left- and right-handed circular polarization. Two simple configurations can be distinguished. For the configuration where the main axis of the modulator is set to 45◦ with respect to the plane of incidence of the light on the sample, we refer to Refs. [25, 26]. Here, we will briefly discuss the simpler case where the main axis of the modulator is set to 0◦ . The signal from the photodetector, positioned after the analyzer that is aligned with the PEM at the p-polarization axis, is sent to a lock-in amplifier. It can be shown [26] that dc and harmonic signals are generated. Up to second order in the polarization modulation frequency f , they read Idc = [1/2 + J0 (A0 )θ ] RI0 ≈ RI0 /2,

    (25)

    I1f = J1 (A0 )εRI0 ,

    (26)

    I2f = J2 (A0 )θ RI0 ,

    (27)

    492

    M. L. M. Lalieu and B. Koopmans

    where Jn (A0 ) is the n-th order Bessel function. Thus, it is seen that ellipticity ε (via the 1f -signal) and rotation θ (via the 2f -signal) – or, alternatively, complementary linear combinations thereof when using an additional wave plate – can be measured simultaneously.

    Layer-Specific MOKE In case of a multilayer structure including multiple magnetic layers, the measured value of ε or θ is proportional to a weighted average of the magnetization in the different layers, where the weighting factors depend on the details of the structure (materials, layer thicknesses) and measurement configuration (wavelength, polarization, and angle of incidence), c.f., Eq. (20). In this section, an extension to the previously discussed MOKE setup will be introduced, which makes it possible to perform layer-specific measurements in a magnetic bilayer system, i.e., individually measuring the magnetization of one of the two magnetic layers [27, 28]. The layerspecific technique discussed here is based on the PEM configuration of the MOKE setup, but it is noted that it works for other MOKE configurations as well. A phenomenological explanation of the layer-specific MOKE measurement technique starts by representing the MO response of a single magnetic layer by a Kerr vector θ˜ = θ + iε = Ωeiξ in the complex Kerr plane spanned by ε and θ . The Kerr vector is described by a Kerr amplitude Ω and Kerr angle ξ , as illustrated in Fig. 6a. Within this representation, the I1f (I2f ) signal in the ordinary MOKE setup measures the projection of the Kerr vector on the imaginary (real) axis. In the case of a magnetic bilayer, the Kerr vector describes the MO signal of the complete structure and is equal to the sum of the Kerr vectors of the individual layers, θ˜tot = θ˜1 + θ˜2 = F˜1 M1 + F˜2 M2 , as illustrated in Fig. 6b, and where we again simplified Eq. (20). Measuring, for instance, I2f (i.e., θ ), yields the projection a

    b

    c

    Fig. 6 (a) The Kerr vector in the complex plane spanned by the Kerr ellipticity ε and the Kerr rotation θ. It is described by a Kerr amplitude Ω and Kerr angle ξ , and represents the MO response of a single magnetic layer. The measured values of ε or θ in the MOKE setup are the projections of the Kerr vector on the imaginary or real axes. (b) Total Kerr vector representing the MO response of a magnetic bilayer. The total Kerr vector θ˜tot is equal to the sum of the Kerr vectors of the individual layers, θ˜1 and θ˜2 . Rotating the projection axis P to be perpendicular to θ˜1 causes the MO signal of this layer to be excluded in the projection. (c) Kerr vector for different depths di of a magnetic layer within a structure, with d1 < d2 < d3 . Burying the magnetic layer deeper into the structure causes a rotation of the Kerr vector, as well as a decrease in its amplitude. (Figures adapted from Ref. [27])

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    493

    of θ˜tot on the real axis, also called the projection axis P , and thus includes the MO signal of both layers. Measurement of an individual layer would be possible when the projection axis P is rotated to be perpendicular to one of the two Kerr vectors, e.g., θ˜1 , as illustrated in Fig. 6b. In that case, the projection of θ˜1 is zero, and the measured signal only includes the projection of θ˜2 . This shows that by rotating the projection axis, it is possible to individually measure the magnetization in one of the two layers in a magnetic bilayer by removing the contribution of the other layer to the measured MO signal. One possible method to rotate the projection axis is by adding a quarter-wave plate (QWP) to the MOKE setup [28]. As indicated in Fig. 5c, the QWP is placed in between the first polarizer and the PEM. Depending on the angle of the QWP, I1f and I2f measure linear combinations of ε and θ , corresponding to a rotation of the projection axis. Using Jones matrix calculations [15], the first- and secondharmonic contributions to the measured intensity of the light can be calculated to be   I1f = 2I0 R θ sin(2α) + ε cos2 (2α) J1 (A0 ) (28)   I2f = 2I0 R θ cos2 (2α) − ε sin(2α) J2 (A0 ), in which α is the angle of the QWP. It can be seen that for both I1f and I2f , there are certain angles of the QWP α = α0 at which the MO signal disappears (i.e., I1f/2f (α0 ) = 0). When the Kerr vectors of the two magnetic layers (i = 1, 2) in a magnetic bilayer have different Kerr angles ξi , as is the case in Fig. 6b, the MO signal of the individual layers will disappear at different angles α0,i . When the QWP angle is set to α0,i of layer i, only the magnetization of the other layer is measured, resulting in a layer-specific measurement. One way to assure different Kerr angles is by using two magnetic layers made of different materials. This was used in Ref. [28], in which the magnetization dynamics of the individual Ni and Fe layers in a Fe/Ru/Ni multilayer were measured separately at different angles of the QWP (α0,Ni = α0,Fe ). An additional advantage of this technique over other element-specific measurement techniques is the ability to perform layer-specific measurements on magnetic bilayers containing two identical magnetic layers. This is made possible by a depth dependence of the MO signal [27, 29], as illustrated in Fig. 6c. The figure illustrates the Kerr vector of a (single) magnetic layer for different depths di within a structure, with d1 < d2 < d3 . Burying the magnetic layer deeper into the structure causes a rotation of the Kerr vector, as well as a decrease in its amplitude. The rotation of the Kerr vector with depth causes the Kerr angle of two identical magnetic layers in a magnetic bilayer to be (slightly) different, resulting in a different α0 for the two layers and making it possible to perform a layer-specific measurement. An experimental demonstration of the depth dependence of α0 for a single outof-plane magnetized Co layer buried in Pt is presented in Fig. 7a. In this figure, the I2f signal coming from the magnetization in the Co layer is measured as a function of the QWP angle, for different thicknesses of the Pt capping layer. The I2f

    494

    M. L. M. Lalieu and B. Koopmans

    a

    MOKE Signal (norm.)

    b

    1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0 -0.5 -1.0 -250 -200 -150 -100

    -50

    0

    50

    100

    150

    200

    250

    Field (mT) Fig. 7 (a) Measurement of the I2f signal coming from the magnetization in the Co layer as a function of the QWP angle, for different thicknesses of the Pt capping layer. The inset highlights the shift in α0 when the Co layer is buried under an increasing amount of Pt. The solid lines are fits to the data using Eq. (28). (b) Layer-specific measurement performed on a Ta/Pt/Co/Pt/Ru/Pt/Co/AlOx (bottom to top) multilayer, showing three hysteresis loops measured at three different angles α of the QWP. The top panel shows the hysteresis curve in which both Co layers are visible, in which the outer (inner) switches correspond to the magnetization in the bottom (top) Co layer. The middle and bottom curves show layer-specific measurements, in which the signal of one of the two layers is excluded from the measurement

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    495

    signal is determined from the height of the hysteresis curve measured using polar MOKE. The solid lines are fits using Eq. (28), showing a very good agreement. It can be seen that for each curve there are two angles of the QWP at which the I2f signal is zero. These angles correspond to the α0 angles of the Co layer. More importantly, highlighted in the inset, a clear shift in α0 of about 0.4 ◦ nm−1 is seen when the Co layer is buried under an increasing amount of Pt. This shift correlates to the rotation of the Kerr vector, as illustrated in Fig. 6c, whereas the decrease in the signal amplitude with increasing Pt thickness corresponds to the decrease of the Kerr amplitude. The prediction that the small shift in α0 allows for a layer-specific measurement in a magnetic bilayer containing two magnetic layers made from the same material is demonstrated in Fig. 7b. The presented measurement is performed on a Ta(2)/Pt(10)/Co(1)/Pt(0.4)/Ru(1)/Pt(0.4)/Co(1.3)/AlOx (2.0) (bottom to top and thickness in nanometer) multilayer, containing two out-of-plane magnetized Co layers that are antiferromagnetically coupled by a thin Pt/Ru/Pt spacer. At α = 3.57◦ (middle panel), a sensitivity to only the top layer is achieved, resulting in a singlecomponent hysteresis curve. Similarly, at α = 5.01◦ (bottom panel), only the bottom layer is seen, while at arbitrary angle (top panel), both layers show up. This measurement clearly demonstrates that although the difference in α0 for the two Co layers is quite small (1.44◦ ), a layer-specific measurement is possible, even when the two magnetic layers are made of the same material, which is not possible with element-specific measurement techniques such as X-ray magnetic circular dichroism, discussed in the next section. For completeness, it is noted that the difference in α0 for the two layers may not only be due to their different depth but may also be partially ascribed to subtle differences in their electronic and optical properties due to different neighboring layers.

    MO Spectroscopy Another approach to disentangle magnetic information in more complex systems is provided by selecting appropriate wavelengths. While most MOKE setups operate with a single laser or LED operating in the visible or near-infrared, a complete MO spectrum can be measured using either a broadband light source in combination with a spectrometer or a wavelength-tunable laser system. Clearly, as with ordinary ellipsometry, a more complete deconvolution of the MO response in multilayer systems is possible from such a spectrum, provided enough information about the wavelength-dependent diagonal and off-diagonal elements of the dielectric tensor of the different layers is known. More specifically, in some alloys it is known that different elements contribute to different parts of the MO spectrum. As an example, in rare-earth transition-metal ferrimagnetic alloys, it is known that the high-wavelength part (600 nm) is most sensitive to the transition-metal atoms (and their magnetic moments) and the (ultra)violet part to the rare-earth atoms. This has been used e.g. to disentangle the different response of the sublattices after femtosecond laser heating [30]. Extending our scope to sources well outside the visible range, X-ray techniques are known to be extremely powerful. In fact, X-ray magnetic circular dichroism

    496

    M. L. M. Lalieu and B. Koopmans

    (XMCD) is basically just a special type of MO measurement, though with the added value of being element-specific by probing the core levels. Moreover, in special cases, XMCD provides a quantitative separation between spin and orbital contributions to the magnetic moments. This separation is based on specific sum rules, as derived, for instance, for the transition-metal L-edges [31, 32]. Interestingly, table-top versions of core level spectroscopy based on high-harmonic generation (HHG) of fs laser pulses to produce XUV pulses with photon energies up to ∼100 eV have become available [33]. They can be thought to bridge the gap between simple MOKE setups and the much more demanding user-facility synchrotron-based techniques. Nowadays, several groups are performing pumpprobe MO measurements using the HHG pulses as a probe. In contrast to the fs-XMCD measurements, the approach has been limited to the M-edge of 3d elements rather than the L-edge and the use of magnetic linear dichroism (MLD). This development is of particular interest for studies of ultrafast magnetic processes, e.g., see Ref. [34]. Some examples of full element-specific and spin-orbit-resolved XMCD as well as table-top core level spectroscopy applied to the exploration of fs magnetization dynamics will be discussed in section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism”. Finally, in the context of MO spectroscopy, we briefly address options provided by higher-order nonlinear terms in Eq. (1), which offer a complementary optical view of magnetism. Most relevant in the present context is the process of magnetization-induced second-harmonic generation (MSHG) [35], described (2) by the second-order optical susceptibility tensor χ . In this process, occurring at high laser fluences, light is generated at twice the frequency of the incident light. Based on symmetry arguments, it can be shown that all tensor elements vanish in the bulk of centrosymmetric media. It is from this that SHG derives its intrinsic surface/interface sensitivity. In the late 1980s, it was realized that the magnetization in a centrosymmetric ferromagnetic metal does break time but not spatial inversion symmetry. As a consequence, MSHG provides an optical measurement scheme that is – within the electric dipole approximation – inherently sensitive to the magnetization at the interface. Absolute MSHG signals are many orders of magnitude smaller than their linear counterparts. However, this can be compensated for by the fact that the relative MO contrast can be huge, and in specific configurations, it can be tuned at will up to 100 % [36]. Some examples of MSHG as applied to investigations of ultrafast magnetization dynamics will be discussed in section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism”.

    MOKE Microscopy One of the advantages of using MOKE as a magnetometry tool is its spatial resolution. Even with long-focal-length lenses, a resolution of a couple of tens of micrometer resolution is easily achieved in a very simple setup. Such resolution allows local magnetization measurements on thin film samples where one (or more) of the layers has been given a position-dependent thickness via a “wedge growth”

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    497

    technique, in which a mechanical knife-edge shutter is withdrawn during deposition of the respective layer. Additionally, local measurements of MOKE can be easily implemented in a UHV environment. In the context of ultrathin films and magnetic surfaces, it has been dubbed surface magneto-optic Kerr effect (SMOKE) [37]. Aiming for a higher resolution, toward the diffraction limit and beyond, a number of approaches have been developed. In scanning Kerr microscopy a tightly focused laser spot is used for imaging. The spatial resolution is limited by the Rayleigh criterion: λ s= , (29) 2NA where s is the smallest distance between features that can be distinguished, λ is the used wavelength, and NA is the numerical aperture (defined as n sin(θ ) with θ the collection angle of the microscope objective and n the refractive index of the medium). However, this approach is not used in practice for separating nearby magnetic features but for measuring the magnetization of nano-objects significantly smaller than the diffraction limit. This approach is made possible by the superior sensitivity of MOKE [38] and was later dubbed “nano-MOKE.” In principle, it can be implemented using any of the MOKE detection schemes discussed previously. A complementary, and widely used, approach is wide-field Kerr microscopy, see Fig. 8a. In fact, this approach is similar to polarization microscopy. It generally makes use of an intense white light source for illumination, a high-NA microscope objective, and a crossed-polarizer detection scheme (Fig. 5a). The resolution can be further enhanced by using an immersion lens (n > 1). In the simplest, polar, geometry, the full entrance aperture is used. In that case, the incident light propagates on average perpendicular to the sample’s surface, which leads to a sensitivity to the perpendicular component of M. In order to gain sensitivity to in-plane components, light from an asymmetrical part of the entrance aperture should be selected. When doing this with a four-quadrant detector and measuring different combinations of signals from the four quadrants, full vectorial resolution can be achieved; see Fig. 8b. Wide-field Kerr microscopy has been used for decades to measure micromagnetic domains and their dynamics. Moreover, it has been extended to the regime of GHz dynamics, often based on stroboscopic approaches with a variable delay between trains of electronically generated magnetic field pulses to excite the dynamics and laser pulses for MO probing [41, 42]. Extending the domain of optics further, toward the use of X-rays or photoelectron spectroscopies, fascinating developments have been reported in the past two decades. For more details on those novel microscopy techniques, see  Chap. 24, “Magnetic Imaging and Microscopy” of this handbook.

    Measuring Ultrafast Magnetization Dynamics MOKE provides a very powerful way to measure magnetization dynamics at ultrafast time scales using fs laser pulses. This is most easily done in a pump-

    498

    M. L. M. Lalieu and B. Koopmans

    a ~16 fps CCD camera

    Polar

    Longitudinal Longitudinal with transverse sensitivity

    Transverse

    Transverse Tube lens Analyzer

    Collector

    Xe arc lamp

    Aperture diaphragm (centered)

    Field Diaphragm

    Compensator

    Polarizer

    Back focal plane Objective lens

    sample

    sample

    b

    50 mm

    Fig. 8 (a) Schematic drawing of a wide-field Kerr microscope. Figure adapted from Ref. [39]. (b) Example of Kerr microscopy, showing a domain pattern in a permalloy patterned film element of 240 nm thickness, imaged in longitudinal sensitivities with light incidence from different directions. (Figure adapted from Ref. [40] Reproduced from [40], with the permission of AIP Publishing)

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    499

    probe scheme using two laser pulses, i.e., the first pulse is used to excite (pump) the magnetization dynamics and the second (mechanically time-delayed) pulse is used to probe it. The ultimate time resolution is achieved by using two pulses derived from the same laser source to avoid electronic jitter. To obtain a high sensitivity, a stroboscopic scheme can be employed, in which a train of high-repetition-rate laser pulses is focused on the sample. In such cases, the approach is commonly referred to as time-resolved MOKE (TR-MOKE). Linearizing Eq. (21) for a simple system in which the MO signal depends solely on a single magnetization component, it is simple to see that the pump-pulseinduced complex Kerr rotation can be written as ˜ = F0 ΔM(t) + M0 ΔF (t), Δθ(t)

    (30)

    where F0 and M0 are the generalized Fresnel coefficient and the magnetization in the absence of a pump pulse. This equation carries an important warning. The first term on the right-hand side is proportional to the laser-induced change in magnetization, ΔM(t), and is the targeted signal reflecting the genuine magnetization dynamics. The second term, however, does not. It is due to a pump-induced modification of the proportionality factor, ΔF (t). We will return to this issue in section “Ultrafast Laser-Induced Loss of Magnetic Order”, but the message here is that one should always be aware of the occurrence of possible “artifacts.” An implementation of TR-MOKE based on the polarization modulation scheme is sketched in Fig. 9 and will be explained later. A simpler realization exploits the crossed-polarizer configuration, which was introduced in Fig. 5a. Both the pump and probe pulses can be derived from the same laser source using a beam splitter (BS) or frequency converter by a nonlinear crystal. Collimated probe pulses are polarized before being passed along a mechanical delay line to adjust the time delay between the two pulses, whereafter they are directed to the sample. Pump and probe pulses are focused to (at least partially) overlapping spots on the sample. It can be of advantage to use a probe spot that is smaller than the pump spot, in order to probe a region where the change in magnetization is more or less homogeneous. The influence of the pump beam on the polarization state of the reflected probe pulse is measured using an analyzer and any type of photodetector. To measure ˜ the pump-induced changes to the magnetization, measurements of θ(t) with and without pump-pulse excitation are compared. Often, to enhance the sensitivity, a mechanical chopper is placed in the pump beam that periodically blocks the pump pulses (see Fig. 9), and a lock-in amplifier is used to directly measure Δθ˜ (t). From Eq. (22), it is easy to derive that the pump-induced change in output signal is described in lowest order of θ˜ by [26]: ΔI (t)/I0 = 2γ Δθ (t)R0 + 2γ θ0 ΔR(t) + γ 2 ΔR(t),

    (31)

    where R0 and ΔR(t) are the reflectivity and pump-induced transient thereof. Care has to be taken to rule out artificial signals due to a ΔR(t) of nonmagnetic origin. The term γ 2 ΔR(t) transforms symmetrically under reversal of M, and can as such

    500

    M. L. M. Lalieu and B. Koopmans

    Fig. 9 Schematic overview of the TR-MOKE setup in a polarization modulation scheme. Using a delay line to alter the path length of the probe pulse, the magnetization can be measured as a function of time after the pump-pulse excitation by adjusting the pump-probe time delay (Δt)

    be easily removed using a procedure corresponding to Eq. (19). In contrast, the term 2γ θ0 ΔR(t) transforms antisymmetrically and cannot easily be distinguished from magnetization dynamics. Bigot et al. argued that part of this drawback of the crossed-polarizer approach is avoided by performing measurements at a multitude of analyzer angles [43]. Also the second configuration to measure MOKE, as introduced in Fig. 5b, can be easily adapted for time-resolved studies [44]. When working exactly at the balanced configuration, a dependency on ΔR(t) can be avoided [26], and Eq. (23) yields ΔI (t)/I0 = 2Δθ (t)R0 .

    (32)

    Note that when required, a sensitivity to the complementary ellipticity channel is obtained by using a quarter-wave plate, an option also available for the crossedpolarizer configuration. Finally, the polarization modulation scheme can be applied, as is illustrated in Fig. 9 (see also Fig. 5c). Using the configuration that was used to derive Eqs. (25)– (27) and choosing A0 = 2.405, for which J0 (A0 ) = 0, one finds [26] ΔI1f (t) = 2J1 (A0 )Δε(t), Idc

    (33)

    ΔI2f (t) = 2J2 (A0 )Δθ (t). Idc

    (34)

    Similar to the balanced photodiode approach, no artifacts proportional to ΔR(t) appear in lowest order. However, small misalignments can easily give rise to nonmagnetic contributions. For that reason, it is always recommended to measure the dynamic response at two opposite orientations of the magnetization whenever possible. Then, Eq. (19) can be used to deduce the magnetization dynamics. An example of such a measurement is shown in Fig. 10. In this experiment, the magnetization in a 10 nm thick Ni film was measured as a function of the pump-

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    501

    Fig. 10 Typical TR-MOKE measurement performed on an in-plane magnetized 20 nm thick Ni film, in the presence of an external field applied almost perpendicular to the film surface. Top panel: time-resolved Kerr rotation measured with the two opposite magnetization directions. Bottom panel: magnetic (red) and nonmagnetic (blue) contributions to the measured signal, calculated using the difference and sum of the data measured for the opposite magnetization directions, respectively. The cartoons illustrate four stages during the observed dynamics; (I) equilibrium state before laser-pulse excitation with the magnetization aligned to the effective field (red dotted line); (II) laser-pulse excitation, causing demagnetization and a change in the anisotropy and thereby the effective field; (III) recovery of the magnetization; (IV) magnetization precession triggered by the sudden change in effective field direction.( Figure taken from Ref. [45])

    probe time delay. The figure displays both the fast demagnetization dynamics in the first few picoseconds and a laser-induced magnetization precession at a longer time scale, which is discussed in the following section. The top panel shows the dynamics measured for the two opposite magnetization directions (black), set by the external field applied almost perpendicular to the sample surface (H = ±160 kA/m). The pure magnetic response to the laser-pulse excitation is calculated using the difference of the two curves (Eq. 19) and is shown by the open dots in the bottom panel (red). The solid symbols (blue) represent the laser-pulse-induced changes in the optical signal due to “nonmagnetic” contributions, which is calculated using the sum of the measured data for the two opposite magnetization directions and arises e.g. due to state-filling that can occur during pump-probe overlap (see section “Experimental Demonstration of Laser-Induced Demagnetization”). In this specific measurement, it only shows some small features around zero pump-probe delay, but in other cases these nonmagnetic contributions to the signal can become more prominent, or even dominant. The setup as displayed in Fig. 9 can be considered as a basic or standard TRMOKE arrangement. A simple extension can be made in order to reduce the nonmagnetic contributions by using a two-color scheme, in which the pump and

    502

    M. L. M. Lalieu and B. Koopmans

    probe pulses have different wavelengths. This can be accomplished, for instance, by sending the probe beam through a frequency doubler. Also, the previously discussed layer-specific MOKE technique can be used by adding a QWP to the setup, as was shown for the static MOKE in Fig. 5c and which allows to individually measure the magnetization dynamics in one of the two layers in a magnetic bilayer, e.g., see Refs. [28, 46]. Analogous to the TR-MOKE setup, time-resolved MCD measurements can be performed by using fs X-ray or XUV pulses for the probe beam, allowing element-specific and spin-orbit-resolved MCD measurements of the laser-induced magnetization dynamics. An example of such time-resolved elementspecific XMCD measurement, performed to investigate the laser-pulse-induced magnetization reversal in a GdFeCo alloy, is discussed in section “All-Optical Switching of Magnetization”. Lastly, time-resolved MSHG measurements can be achieved by recording the probe-pulse-induced second-harmonic generation as a function of the pump-probe delay time [47].

    Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism Having discussed the use of magneto-optics to probe magnetic phenomena, we now address the possibility of manipulating the magnetic state by optics – referred to as opto-magnetism. After a conceptual introduction to the field, three contemporary topics will be addressed in more detail: (i) the ultrafast loss of magnetic order, socalled femtosecond (fs) laser-induced demagnetization, (ii) the all-optical switching of the magnetization with fs laser pulses, and (iii) the ability to optically generate and exploit fs spin currents in nanostructured magnetic systems.

    Conceptual Introduction The 1990s saw an entirely new development, facilitated by the commercial availability of fs pulsed laser sources, which began to be used to explore the ultimate limits of magnetization dynamics. A major outcome of these studies is that we no longer see optics as merely a powerful tool to read the magnetic state but also as a very efficient tool to manipulate it. As such, a new field of opto-magnetism has emerged, in which magnetic ordering is affected by the light. Opto-magnetism provides a method to control the magnetic state beyond the traditional precessional dynamics as described most conveniently by the LandauLifshitz-Gilbert equation of motion [48] dM α = −γ μ0 (M × Heff ) + dt M



    dM M× . dt

    (35)

    Here, an effective field H eff is introduced that applies a precessional torque on the magnetization M as described by the first term on the right-hand side, where γ is

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    503

    the (positive) gyromagnetic ratio and μ0 the vacuum permeability. The second term describes a phenomenological damping via the Gilbert damping parameter α, which allows the system to dissipate energy and thereby converge to the lowest energy state, in which M is aligned with H eff . The precessional dynamics in ferromagnets proceed at frequencies of up to 10s of GHz for applied magnetic fields not exceeding 1 MA/m (μ0 H ∼ 1 Tesla) and conserves the magnitude of the local magnetization. Opto-magnetism provides a route to change not only the orientation but also the magnitude of M, and it does so on a time scale that is typically up to three orders of magnitude faster. A first (basic) mechanism to exploit pulsed lasers to manipulate ferromagnetic order is laser heating. Pulsed lasers have enough energy per pulse to heat up a ferromagnetic thin film of any material and with a thickness comparable to the skin depth of the light, to well above its Curie temperature. Increasing the temperature will lower the equilibrium magnetization, and the system will strive to re-establish thermal equilibrium between the spin- and other degrees of freedom, lowering its magnetization. Finding the relevant time scale associated with losing magnetic order was one of the fundamental quests that inspired researchers in the 1990s. Beaurepaire and co-workers were the first to find that the relevant time scale was well below a picosecond [4]. This rapid loss of magnetization was referred to as ultrafast demagnetization. Effects of laser heating can be even more spectacular if they provoke a transition between contrasting magnetic phases. A trivial example is laser heating through the Curie temperature, driving a system from the ferromagnetic to the paramagnetic state. Then, the magnetization can regrow in the opposite direction upon cooling down when a static reversed magnetic field is applied [49], as is shown in Fig. 11a. A more interesting situation occurs for materials like FeRh, which can be grown

    Fig. 11 (a) Recovery of the magnetization of GdFeCo alloy after a full laser-induced demagnetization in the presence of a applied magnetic field in the original direction (triangles up), zero field (squares), and reverse field (triangles down). (Figure taken from Ref. [49]. with kind permission of APS) (b) Laser-induced growth of magnetization in FeRh as measured by TR-MOKE (filled symbols), compared to the transient reflectivity ΔR. (Figure taken from Ref. [50] with kind permission from APS)

    504

    M. L. M. Lalieu and B. Koopmans

    such that it is antiferromagnetic around room temperature but ferromagnetic at higher temperatures. In this case, fs laser heating could be expected to cause a laserinduced growth of ferromagnetic order, as is indeed observed [50, 51]; see Fig. 11b. Consensus has been reached that the final magnetic phase transition is completed within several tens of ps [52, 53]. Apart from affecting the equilibrium magnetization, fs laser heating can be used to optically generate magnetization precession [54,55,45], as was shown in Fig. 10, or to launch standing spin waves [56]. In these approaches, the laser-pulse excitation causes a rapid sub-ps change of the magnetic anisotropy, after which the dynamics is governed by the LLG equation, Eq. (35). A similar approach has been used to optically modify the interface exchange bias between a ferromagnetic film and an adjacent antiferromagnetic film [57]. Besides the thermal effects, the electromagnetic photon field associated with the laser pulse itself can induce changes in the magnetic state, as mediated by spinorbit coupling [58, 59]. Although such effects certainly exist, their relevance to explaining the experimentally observed sub-ps dynamics in metallic systems is not always evident, as will be discussed in section “Theories for Femtosecond Demagnetization”. However, in special materials spectacular effects have unambiguously been identified. For instance, in 2017 it was shown that linearly polarized fs laser pulses break the degeneracy between meta-stable magnetic states in transparent nonmetallic iron garnets, leading to optically induced switching with exceptionally low energy absorption [60]. All-optical switching of magnetization (AOS) in metallic systems was discovered by Stanciu et al. in 2007 [5]. It was shown that magnetization of GdFeCo alloys can be “written” by a single fs laser pulse in a helicity-dependent (HD) manner, i.e., the final state (magnetization up or down) is determined entirely by the helicity of the light (left- and right-handed circular polarization), as was illustrated in Fig. 1b. Later, it was found that helicity-independent single-pulse toggle switching played a decisive role in these seminal experiments, governed by a dynamic exchange mechanism. In this scenario, strong exchange forces (corresponding to a field up to > 1000 T) drive an ultrafast reversal of the oppositely oriented magnetic moments at the Gd and Fe/Co atomic sites [6] (see section “All-Optical Switching of Magnetization”). From a conceptual point of view, it is of relevance to introduce the inverse Faraday effect (IFE), which is another effect of the photon field. It elegantly reflects the symmetry between magneto-optics and opto-magnetism. The IFE is reciprocal to the ordinary Faraday effect and was proposed in the 1960s. It was first experimentally demonstrated for a nonabsorbing material by van der Ziel et al. [61]. In the language of Eq. (13), the IFE describes the generation of a temporary (quasi-static) magnetization M in the material according to M = αE × E ∗ ,

    (36)

    where E is the amplitude of the oscillating (optical) electrical field as defined before, E ∗ its complex conjugate, and α a material constant characterizing the size of the IFE in the material. It should be emphasized that the cross product E × E ∗

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    505

    vanishes for linearly polarized light. For circularly polarized light, we see that the induced magnetization reverses upon reversal of the helicity. Furthermore, it should be stressed that the induced M is coherent in the sense that it is setup by the laserfield and decays within a decoherence time. It has been suggested that the IFE may be of relevance for HD-AOS of metallic ferrimagnets, which is discussed in section “All-Optical Switching of Magnetization”. Laser excitation has been demonstrated to be a source of laser-induced spin currents – whether by spin-polarized ballistic electrons, of “superdiffusive” nature, or driven by the demagnetization or thermal gradients. These spin currents trigger a variety of magnetization dynamics, further elaborated on in section “Laser-Pulse-Excited Spin Currents”.

    Ultrafast Laser-Induced Loss of Magnetic Order We now focus on the process of fs laser-induced demagnetization. A historical review of early experiments is followed by a collection of key-experimental results that should guide the explanation. Then, the issue of conservation of angular momentum is revisited, after which theories proposed to explain the ultrafast demagnetization are analyzed. Finally, a selection of more recent advances in the field is discussed.

    Experimental Demonstration of Laser-Induced Demagnetization Experiments addressing the time scale at which magnetic order can be quenched after sudden (laser) heating started in the 1980s [62]. The first real time-resolved experiments were performed in the early 1990s by Vaterlaus and co-workers [63, 64]. Ferromagnetic gadolinium, a rare-earth ferromagnet with a Curie temperature just below room temperature, was heated by 10 ns pump pulses. The magnetic order was measured by analyzing the spin-polarization of the electrons photoemitted by 60 ps probe pulses. The relaxation time of the magnetization was found to be 100 ± 80 ps. A similar spin-lattice relaxation time was calculated by Hübner and Bennemann [65]. This was the background for the pioneering experiments by Eric Beaurepaire, Jean-Yves Bigot, and co-workers in 1996. They were the first to study demagnetization by using +bk,σ |k, ↓>,

    (39)

    for σ =↑, ↓ and where labeling is chosen such that for all states b < 0.5. Then, a spin-mixing parameter < b2 > can be calculated after weighing it over all possible scattering processes. This spin-mixing parameter scales with Z 4 , where Z is the ionic charge in units of e, and typically b2 in Eq. (39), highlighting the importance of spin-orbit interaction. One readily sees that no magnetization dynamics occurs, i.e., dm/dt = 0, if asf = 0, as expected. Two other approaches, although rather contrasting at first sight, reproduce magnetization dynamics very similar to the M3TM. Firstly, the Landau-Lifshitz Bloch (LLB) equations provide an extension of the LLG equation, by adding

    516

    M. L. M. Lalieu and B. Koopmans

    an additional differential equation for the longitudinal component of M. Atxitia et al. treated the longitudinal magnetization dynamics after pulsed laser heating for arbitrary S = n/2 (where n is an integer). They derived an analytic differential equation for the longitudinal magnetization relaxation for S = 1/2 and showed that it can be exactly mapped on Eq. (43) [108]. As such, the M3TM can be considered a microscopic quantum-mechanical derivation of the longitudinal part of the LLB. Consequently, identical dynamics is reproduced. Secondly, atomistic LLG (ALLG) models produce very similar M(t) traces as well [109]. In ALLG calculations, a large grid of atomic magnetic moments of fixed magnitude is considered, in which neighbors couple via interatomic exchange, and stochastic thermal noise is introduced to account for (electronic) temperature effects. Angular momentum dissipation to the lattice is described by the phenomenological Gilbert damping parameter α, Eq. (35). Assuming that α at the atomic scale has the same magnitude as the one describing mesoscopic magnetization precession, very similar magnetization dynamics upon laser heating is obtained compared to LLB and M3TM [109]. The ALLG approach lacks both the numerical simplicity and the quantum mechanical footing of the M3TM but, on the other hand, allows for treating spin fluctuations, stochasticity, and the inclusion of realistic spin wave modes [110]. The assumption that the microscopic and mesoscopic α are identical may seem a bold statement, but it has interesting consequences. In ALLG, the demagnetization time scales as α −1 , and as such α plays a role similar to asf in the M3TM. Actually, also in the M3TM, a simple relation between the mesoscopic α and τM has been derived, by a simplified quantum mechanical calculation of damping in precessional dynamics based on the same Hamiltonian used in the M3TM. The result thereof is [55] τM ≈ c0

    h¯ 1 , k B TC α

    (45)

    where c0 is a constant of the order of unity, typically ∼1/8, but depending on details. Taking a Ni thin film as an example, using α equal to a few hundredths and TC = 630 K, one readily obtains τM ∼ 100 fs in close agreement with experimental observations. An intuitive interpretation of Eq. (45) is that atomic fluctuations in the local magnetization damp out (or thermalize) at the same rate as a macroscopic magnetization would do in a field set by the atomic exchange field (governed by TC and typically 1000s of Tesla) and governed by the macroscopic α. Theoretical predictions for Elliott-Yafet spin-flip probabilities of transition-metal ferromagnets have been obtained by ab initio calculations of the Eliashberg function by Carva et al., while order of magnitude estimates were obtained by calculating the < b2 > parameter by Fähnle et al., [111]. Resulting values are typically 0.05– 0.10, in close agreement with the interpretation of experiments using the M3TM [85] (section “Towards Quantitative Understanding”). Nevertheless, the relevance of this agreement has been disputed. In a rigid band-structure calculation adapting a Stoner picture of magnetism (in which magnetization is changed by variation of the atomic magnetic moment), Carva et al. find a negligible demagnetization

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    517

    using the calculated ab initio value, whereas Schellekens et al. argue that close agreement with experimental results is obtained by implementing them in the Weiss or Heisenberg picture (in which the magnetization dynamics is due to thermally disordered orientation of fixed atomic moments) [112]. The basic M3TM makes use of severe approximations that can be lifted when desired. One can implement a more realistic electronic density of states, phonon dispersion, or extensions from S = 1/2 to arbitrary spin multiplets. Also, the description of itinerant magnetism in a pure Weiss picture may be considered inadequate. The s-d model applied to ultrafast magnetization dynamics in Ref. [113] provides an interesting step beyond. We treated phonon-mediated Elliott-Yafet spin-flip scattering in quite some detail. Spin-flip e-e scattering is another possibly relevant scattering mechanism, in which one of the two mutually scattering electrons flips its spin. Although it could be argued that this process does not conserve angular moment, we highlighted before that it will be facilitated by the non-spin-conserving character of the Hamiltonian and will lead to a dump of angular momentum in the rotational degrees of motion of the crystal. Steil et al. included this type of Coulomb scattering, without the need of referring to a phononic spin bath [97, 114]. They implemented this in a Stonerlike description in which electrons are scattering between different spin bands. Reasonable agreement with experiments was found assuming spin-flip probabilities of 0.15 and 0.30, for Co and Ni. From this, a μT -model emerged, in which dynamics is described in terms of equilibration of temperatures and chemical potentials of the electron subsystems simultaneously [115]. This approach bares some similarities with the s-d model of Ref. [113]. Finally, the increase of computational power and refinement of codes have fueled activities on many-body theory and ab initio calculations. Ultimately, this would enable to trace the angular momentum transfer from a microscopic perspective. Töws and Pastor performed small-cluster calculations distinguishing local and itinerant electrons and exploring the interplay between spin-orbit interactions and interatomic hopping, as well as the role of the photon field [96]. Realistic time scales of demagnetization of typically 100 fs were obtained. Although the lattice degree of freedom was not included in this work, the authors stated that the leaking angular momentum would manifest itself as a global rotation of the rigid lattice if it were explicitly included in the calculation. Fully ab initio calculations of laser-induced magnetization dynamics have been reported by Sharma and co-workers [116, 117]. They employed time-dependent density functional theory in an all-electron solid-state code for the case of unconstrained noncollinear spins, as driven by short pulses of electro-magnetic radiation. Despite the fact that the amount of demagnetization observed was at least an order of magnitude smaller than realistic experiments would show and the phonons were not included yet, this field is rapidly making progress.

    Towards Quantitative Understanding We conclude our treatment of laser-induced fs demagnetization by a selection of studies that aimed at a more quantitative understanding.

    518

    M. L. M. Lalieu and B. Koopmans

    The pulse-length dependence of the laser-induced demagnetization has been investigated systematically for pump pulses from 40 fs up to 7 ps [118]. It was demonstrated that all M(t) curves could be reconstructed by simply convoluting the shortest-pulse measurement with the respective pulse profiles. This behavior speaks again in favor of a simple thermodynamic picture in which little role is played by highly excited electrons and the laser field itself. Roth et al. have presented a detailed study of laser-induced demagnetization of a Ni thin film as a function of temperature (from 80 K up to 480 K), presented in Fig. 16a, and laser fluences ranging up to almost complete demagnetization; see Fig. 14b [85]. Upon increasing temperature, they observed a transition from a fast demagnetization followed by a remagnetization to a two-step demagnetization process, where after a fast decay, the magnetization continues to decrease at a slower rate. The behavior at low and high temperatures has been dubbed type-I and typeII, resp., as will be discussed in more detail below. Interestingly, the authors found that all their M(t) traces could be described by a single set of parameters – most specifically a single value of the spin-flip probability asf independent of temperature and fluence. This way, they established a value asf in excellent agreement with the ab initio estimates calculated by Carva et al., which range from 0.04 to 0.10, depending on calculational details [119]. Despite this agreement, more work is still needed to assess the relative contribution of e-p and e-e scattering to the demagnetization. In the experiments by Roth et al., a characteristic slowing down is observed in the τM versus ambient temperature and τM versus fluence dependence when approaching full quenching. Such a slowing down at higher temperature and fluence is a very characteristic behavior and was also reported for Ni and Co in Ref. [103]. The same characteristic behavior was reported in a comparative study on laser-induced demagnetization in Co films compared to Co/Pt multilayers [120]. Measuring at a whole range of fluences, it was observed that τM for Co/Pt (40–110 fs) is consequently below that for Co (50–300 fs); see Fig. 16b. Fitting the trends with the M3TM revealed a fourfold increase of asf for the Co/Pt multilayers, which was assigned to increased spin-orbit scattering with the heavy Pt atoms. Following up the pioneering work by Vaterlaus [63], magnetization dynamics of gadolinium has received intense interest, and important contributions have come in particular by Bovensiepen and co-workers, using a variety of refined experimental techniques, including MSHG [121], TR-MOKE [122], fs-XMCD [75], and time-resolved angle-resolved photoemission spectroscopy [123]. Gadolinium is an interesting material with a large local magnetic moment carried by the 4f electrons and a much smaller moment due to an induced splitting of the itinerant 5d bands, adding up to a total magnetic moment of 7.55 μB . Femtosecond excitation could well cause strongly nonequilibrium dynamics, in which 4f and 5d magnetic moments do not decay simultaneously. However, comparison of TR-MOKE and fsXMCD data, predominantly probing 5d and 4f , respectively, revealed that already after a few hundred fs the two subsystems display almost identical dynamics indicative for their efficient mutual exchange scattering [122]. One of the most striking findings is that the demagnetization in Gd proceeds in two steps. Apart from a slow component at ∼50 ps, similar to the claim in the

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    (a)

    519

    (b)

    (c) 1.0

    M/M0

    0.9 0.8 0.7 0.6 0.5 -50

    0

    50

    100

    150

    200

    Delay (ps)

    Fig. 16 (a) Ambient temperature dependence of the laser-induced sub-ps demagnetization of a 15 nm Ni film at constant laser fluence. A transition from a type-I to type-II behavior is observed when increasing the ambient temperature from 300 to 480 K. Figure taken from Ref. [85]. (b) Demagnetization time τM as a function of the demagnetization amplitude, measured in a Co/Pt multilayer and a Co film. Figure taken from Ref. [120]. (c) Time-resolved XMCD signals measured on a 10 nm Gd film using 50 fs laser pump pulses and ps (open symbols) or fs (solid symbols) X-ray probe pulses. The solid red curve is a fit to the data using the M3TM. (Data points taken from Ref. [75] with kind permission of APS)

    original work [63], fs-XMCD studies revealed an initial much more rapid decay within the first ps that accounts for only a part of the total demagnetization [75]; see Fig. 16c. As mentioned before, this behavior has been denoted as type-II, in contrast to ordinary type-I behavior with a single continuous decay as observed for, e.g., Ni and Co [103]. A similar two-step demagnetization was also observed in the fs-XMCD for ferromagnetic Tb films, with an identical 1 ps initial decay, but the slower decay being a factor of 5 faster than for Gd. The initial faster component was assigned to the presence of hot electrons that speed up spin-lattice interaction during the first ps, in line with the 1 ps that is needed in both materials to equilibrate the electron and phonon system. The differences between Gd and Tb were assigned to the different orbital momenta in their respective ground state, being L = 0 and L = 3, respectively. The peculiar type-II behavior observed for Gd has also been addressed in the context of the M3TM in Ref. [103]. Surprisingly, in that work it was found that type-II dynamics is a natural outcome whenever τM becomes larger than τE . In such a case, a single scattering mechanism naturally provides a two-step decay with two different time constants. The first rapid decay occurs as long as the electron temperature is still (far) above the lattice temperature. Once the electron and phonon

    520

    M. L. M. Lalieu and B. Koopmans

    temperatures have converged, the demagnetization process continues at a slower rate, albeit driven by exactly the same microscopic process as during the first phase. Independent support for the feasibility of type-II dynamics has come from a T -dependent study on Ni [85], presented in Fig. 16a. There it was shown that for Ni, the magnetization dynamics slows down so significantly at higher starting temperatures that τM becomes larger than τE . A transition from type-I to type-II behavior in the very same material is then expected, as was experimentally observed indeed. In the nickel experiment, it is clear that no second mechanism is active. This argument could be used as a smoking gun for the Gd and Tb case, although final conclusions will require further, more detailed studies. A study trying to intuitively understand the wide variety in demagnetization time as observed for different materials was reported by Müller et al. [76]. They compared demagnetization of materials with a relatively low spin polarization P (Ni and the Heusler alloy Co2 MnSi) with that of materials with almost a 100% spin polarization at EF (Fe3 O4 , LSMO, and CrO2 ). For the low-P materials, a very rapid demagnetization τM ∼ 1 ps was measured, whereas the high-P materials all showed a demagnetization at a 100 ps time scale, in some cases displaying a type-II behavior. Also other materials (several oxides, selenides and sulfides) as measured in Ref. [77] were included in the analysis. All these materials are assumed to be close to half-metallic, i.e., having P = 1, and all of them have indeed showed long demagnetization times ranging from a few ps to 1000 ps. Thus, a relation between P and τM was concluded on. This conclusion can be considered to be consistent with a spin-flip scattering model, since a significant reduction of asf could be expected if one of the two spin bands has a gap in its band structure around EF . Another prediction from LLB, ALLG, and M3TM is provided by the inverse proportionality between τM and the Gilbert damping constant α, cf. Eq. (45). In Ref. [55], where measurements of τM and α on the very same Ni thin film were reported, a very close quantitative agreement with Eq. (45) was found. However, several studies trying to verify the trend predicted by the equation for series of samples did not find a consistent picture. As an example, Radu et al. measured τM and α for permalloy as a function of impurity doping for several dopants, Ho, Dy, Tb, and Gd [124]. A consistent reduction of τM with increasing doping concentration and α −1 was not observed. The authors concluded that the M3TM (as well as the LLB and ALLG) seems to be oversimplified for treating the case of 4f impurities. Although within the two decades after the discovery of sub-ps demagnetization by Beaurepaire our understanding has increased enormously, full consensus about the detailed mechanisms is still not being reached. Development of new experimental approaches to investigate laser-induced magnetization dynamics will be essential to make further progress. Among the outstanding questions, it is still intensively debated whether the Stoner or Heisenberg picture is the most appropriate for describing the ultrafast magnetic processes. As to this question, time- and spinresolved photoelectron spectroscopy provides additional insight by showing how spin-polarized bands are evolving after laser excitation [125, 126]. Moreover, in most descriptions of the magnetization dynamics, the spin system is approximated to be internally in thermodynamic equilibrium. Just like with the electron and lattice

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    521

    system, we know that this may be a poor approximation at a 100 fs time scale. Indeed, experiments using time- and energy-resolved photoelectron spectroscopy have resolved that at ultimate short time scales, one cannot treat the spin polarization of the conduction band as a uniformly varying rigid system any longer [127]. However, whether such nonequilibrium effects significantly affect the main channel of angular momentum transfer is yet unknown.

    All-Optical Switching of Magnetization The field of femtomagnetism gained a new boost by the discovery of all-optical switching of magnetization in 2007 [5]. All-optical switching (AOS) describes the reversal of the magnetization in a magnetic material by laser-pulse excitation, without the use of any additional stimulus (e.g., an external magnetic field). The discovery of AOS came entirely unexpected and gained much attention because the optical switch is both fast (picosecond time scale) and energy-efficient and thus shows high potential to be used in future magnetic memory applications. With such applications in prospect, as well as a drive for fundamental knowledge, the observation of AOS initiated a field of research that rapidly developed, pursuing full comprehension of the mechanism underlying the AOS. In the following, a selected overview of the research performed in this pursuit, and the resulting status quo is given.

    All-Optical Switching in Ferrimagnetic Alloys The first demonstration of AOS was done using a magnetic thin film made of a GdFeCo alloy [5]. This is a rare-earth transition-metal (RE-TM) alloy, which has two distinct magnetic sublattices that are occupied by the Gd (RE) and FeCo (TM) atoms. The alloy is a ferrimagnetic material, in which the two (ferro)magnetic sublattices are coupled antiferromagnetically, and a (net) spontaneous magnetization is maintained due to unequal sublattice magnetizations (typically at room temperature). Figure 17a presents the pioneering measurement performed by Stanciu et al., demonstrating AOS in a 20 nm thick GdFeCo film using single circularly polarized femtosecond (fs) laser pulses [5]. The figure shows the out-of-plane magnetization in the GdFeCo layer, initially containing both an up (white) and down (black) domain. After sweeping a pulsing laser beam across the sample, once with left (σ − ) and once with right (σ + ) circularly polarized laser pulses, it can be seen that domains of reversed magnetization are “written” for specific helicity and magnetization direction combinations. In other words, the final state of the magnetization in these domains is determined by the helicity of the laser pulse. Using these observations, the AOS was explained as a two-step process. First the spin system is heated close to the Curie temperature, whereafter a lightinduced effective field, attributed to the IFE [61, 128, 129] (section “Conceptual Introduction”), was assumed to switch the magnetization. The direction of the effective field is set by the helicity of the light, making the AOS a helicity-dependent process.

    522

    (a)

    M. L. M. Lalieu and B. Koopmans

    (b) Normalized XMCD (%)

    100 Gd 50 0 –50

    Fe

    –100 –1

    0

    1

    2

    3

    Pump–probe delay (ps)

    Fig. 17 (a) Pioneering measurement demonstrating single-pulse all-optical switching in a GdFeCo alloy, showing the out-of-plane magnetization in the layer, initially containing both an up (white) and down (black) domain. Using both left (σ − ) and right (σ + ) circular polarized laser pulses, it was demonstrated that the magnetization in the exposed area could be reversed and that the final direction of the magnetization was determined by the helicity of the light. (Figure taken from Ref. [5] with kind permission from APS). (b) Time-resolved element-specific x-ray magnetic circular dichroism measurements on a GdFeCo thin film, showing the magnetization of the Gd and Fe sublattices in time after fs laser-pulse excitation. (Figure taken from Ref. [6] with kind permission from Springer Nature)

    In the years that followed, the importance of both the laser-induced heating and helicity effects for the AOS was under debate [130, 131, 132, 133, 134, 135]. Macrospin Landau-Lifshitz-Bloch [132] and atomistic Landau-Lifshitz-Gilbert [134] calculations showed that AOS could indeed be achieved via a laser-induced effective-field pulse (in combination with the demagnetization). However, apart from the high amplitude of the magnetic field pulse needed for the switch (≈20 T), the induced magnetic field needed to persist in the material for a much longer time than the duration of the laser pulse itself. This resulted in the notion of some kind of helicity memory mechanism in the RE-TM system [133, 135]. More clarity came when it was observed that the magnetization in GdFeCo can be switched using single linearly polarized laser pulses [7], proving the (single-pulse) switching mechanism to be a purely thermal process, which was simultaneously established theoretically as well [136]. Moreover, it was demonstrated that the helicity dependence presented in Fig. 17a was the result of magnetic circular dichroism (MCD) [137], which describes the difference in absorption for left and right circularly polarized light in magnetic materials (section “Basics of Magneto-Optics from a Macroscopic Perspective”). In the case of AOS, there is a minimum amount of energy that needs to be absorbed for the switch. Therefore, the MCD opens up a fluence window where the AOS is helicity-dependent. Insight into the thermal single-pulse switching mechanism was actually already observed earlier using time-resolved XMCD measurements [6] (section “MO Spectroscopy”). In these measurements, the magnetizations of the individual Gd and Fe sublattices in a GdFeCo thin film were measured in time after fs laserpulse excitation and in the presence of an external magnetic field. The result of

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    523

    the measurement is presented in Fig. 17b, where the normalized XMCD signal of both magnetic sublattices is plotted as a function of pump-probe delay. The observed behavior is typically divided into three subsequent phases: (i) the magnetization of both sublattices rapidly demagnetizes after the laser-pulse excitation. The magnetizations on the two sublattices quench at different rates, with the Fe magnetization demagnetizing much faster and reaching zero while the Gd is still demagnetizing. (ii) As a result of exchange scattering, angular momentum is transferred from the demagnetizing Gd to the Fe sublattice, resulting in a buildup of Fe magnetization along the Gd magnetization direction. This results in a transient ferromagnetic-like state in the ferrimagnet. (iii) Finally, the system cools down on a longer time scale. With the relatively large magnetization buildup in the Fe, the antiferromagnetic exchange coupling drives the switch of the Gd magnetization. Eventually, the system recovers to the ferrimagnetic state, having its magnetization reversed with respect to the initial state and demonstrating the AOS. Soon it was concluded that – in agreement with the previous scenario – there are three key ingredients necessary to facilitate the AOS: (i) two magnetic sublattices; (ii) an antiferromagnetic exchange interaction between the two sublattices, resulting in an antiparallel alignment of the sublattice magnetizations; and (iii) a difference in demagnetization rate for the two sublattices. This was verified theoretically by introducing these ingredients in various models, including a general theoretical framework [136], atomistic spin models [6,7,138,139], and the M3TM [140], which all demonstrated the capability of thermal single-pulse switching. Experimentally, the range of materials for which AOS was found rapidly expanded, including different RE-TM alloys, multilayers, heterostructures, and RE free syntheticferrimagnetic heterostructures [141, 8, 142]. The common features in all these material systems are the previously mentioned needed ingredients, supporting their necessity. The importance of the helicity of the light, however, was not elucidated for some of the systems.

    All-Optical Switching in Ferromagnetic Systems With the AOS mechanism and its requirements believed to be well understood, the observation of AOS in ferromagnetic thin films, multilayers, and granular media came as a surprise [9]. These ferromagnetic systems have a single magnetic layer, thus lacking all of the ingredients previously considered necessary. Therefore, the observation questioned the validity of the established understanding of the AOS mechanism. Consequently, the inverse Faraday effect as a driving mechanism was reconsidered. For instance, by adding an effective field pulse (as well as dipolar interactions) to the M3TM discussed earlier (section “Theories for Femtosecond Demagnetization”), the AOS images obtained in experiments on ferromagnets could be reproduced [9, 143]. Additionally, ab initio calculations on the inverse Faraday effect predicted the high field amplitudes required to model the AOS [144]. Apart from the need of a deeper insight in the switching mechanism on a fundamental level, the discovery gave the research field an additional boost since the ferromagnetic materials showing AOS were the same systems already heavily used in the field of spintronics. This observation could thus ease the integration of AOS in future spintronic devices.

    524

    M. L. M. Lalieu and B. Koopmans

    The inconsistency between the AOS mechanism resolved for the GdFeCo alloys and the observation of AOS in ferromagnetic systems was reconciled by the discovery that there are two different AOS mechanisms at play [145]. For the GdFeCo system, it was confirmed that the AOS is indeed a single-pulse helicityindependent switch, corresponding to the earlier discussed thermal single-pulse switching mechanism. For both a ferromagnetic Pt/Co/Pt multilayer and a ferrimagnetic TbCo alloy, however, no single-pulse AOS was found. For these systems, a helicity-dependent multiple-pulse mechanism was revealed. A measurement of this switching mechanism in the Pt/Co/Pt multilayer is shown in Fig. 18a. The figure

    (a)

    -

    (b)

    M

    +

    1 Hz

    M

    15 µm No pulse 100 pulses 600 pulses 3,600 pulses

    Fig. 18 (a) Measurement of the multiple-pulse AOS switching mechanism in a Pt/Co/Pt sample, showing the normalized magnetization (measured as the anomalous Hall voltage VHall ) as a function of time, while simultaneously the structure is exposed to femtosecond laser pulses which are either left circular (σ − ), right circular (σ + ), or linearly (π ) polarized. (Figure taken from Ref. [145] with kind permission from APS) (b) Measurement demonstrating the helicity-dependent laser-pulse-induced domain wall motion. The direction of the domain wall motion is set by the helicity of the laser pulses, causing either the up (light) or down (dark) magnetized domain to grow in size. (Figure taken from Ref. [146])

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    525

    shows the normalized magnetization (measured by the anomalous Hall voltage VHall ) as a function of time, while simultaneously the structure is exposed to fs laser pulses which are either left circularly (σ − ), right circularly (σ + ), or linearly (π ) polarized. As can be seen in the left panel, the magnetization in the Co layer is lost during the first few laser pulses, independent of the polarization of the light. In the case of the circularly polarized light, the magnetization recovers during subsequent laser pulses (right panel), where the direction of the recovered magnetization is set by the helicity of the light. No magnetization recovery is observed for the linearly polarized laser pulses. This measurement clearly showed the helicity dependence in the multiple-pulse AOS mechanism. For the ferromagnetic thin films, the helicity dependence was proposed to arise from laser-pulse-induced domain wall (DW) motion, which on itself was experimentally demonstrated in Co/Pt multilayers [146, 147]. As presented in Fig. 18b, the latter experiments showed that an up-down DW (i.e., a DW separating an up (light) from a down (dark) magnetized domain) moves in opposite direction when exposed by a train of left (σ − ) or right (σ + ) circularly polarized laser pulses. In the case of the multidomain state generated by the first few laser pulses (left panel Fig. 18a), the helicity-dependent DW motion causes the domains with a certain magnetization direction to expand, causing the magnetization in the exposed area to “grow” into the direction set by the helicity of the laser pulses. Both the earlier discussed inverse Faraday effect and the magnetic circular dichroism were suggested as a possible origin of the helicity-dependent DW motion [145,146,147]. Around the same time of the discovery of the two different switching mechanisms also came the observation of a domain-size criterion for the helicitydependent multiple-pulse AOS [148]. Using magnetic-layer-thickness-dependent studies, it was revealed that the helicity-dependent AOS is observed when the magnetic domain size during cooldown after the laser-pulse excitation is larger than the laser spot size. This criterion can be seen as an extension to an earlier found low-remanence criterion for ferrimagnetic alloys [149]. The latter showed that the helicity-dependent multiple-pulse AOS in ferrimagnetic alloys is only found when the remanent magnetization is below a specific value, resolving the earlier observation that helicity-dependent AOS was only found when the magnetization compensation temperature was close to room temperature [8]. Both criteria demonstrate the need of a low dipolar energy to prevent (small) domain formation during the multiple-pulse AOS. The AOS in the (high-anisotropic) granular media was also found to be a helicity-dependent multiple-pulse process [150]. However, a different (stochastic) cumulative switching mechanism was proposed, based on a helicity-dependence of the magnetization switching probability of the single-domain grains [151, 152, 150, 153]. This helicity-dependence was again linked to both the inverse Faraday effect and the magnetic circular dichroism.

    New Directions in All-Optical Switching With the integration of AOS in future memory applications in mind, the research on AOS shifted back toward the GdFeCo alloys because it was the only material

    526

    M. L. M. Lalieu and B. Koopmans

    system for which the single-pulse AOS mechanism was found, which is needed for fast future spintronics. Realizing that the AOS is a purely thermal process, it was found that the switch could also be obtained using ultrafast hot-electron pulses [154, 155] (analogous to the hot-electron pulse-induced demagnetization discussed in section “Key Observations in Laser-Induced fs Demagnetization”). These hot-electron pulses were generated by laser-pulse excitation of the Pt layer in a Pt/NM/GdFeCo multilayer, whereafter the hot-electron pulse flows through the thick nonmagnetic conductive spacer (NM) toward the GdFeCo layer. There, the hot-electron pulse will deposit its energy and induce the magnetization reversal. Later, using a photoconductive switch, it was demonstrated that the fast switch can also be induced by a picosecond charge-current pulse [156]. Alongside the investigation on the different modes of stimuli for the thermal AOS, research was done trying to achieve the single-pulse AOS in other material systems that are more relevant for spintronic applications. For instance, it was demonstrated that a ferromagnetic layer can be switched using single laser pulses when it is deposited on top of a GdFeCo layer, in which case the ferromagnet switches along with the GdFeCo alloy due to the exchange coupling between the two layers [157]. A similar mutual AOS can be achieved when the GdFeCo layer and the ferromagnet (FM) are separated by a nonmagnetic (NM) spacer layer to create a GdFeCo/NM/FM spin-valve structure. In this case, the single-pulse AOS in the GdFeCo alloy can be transferred to the ferromagnet by the spin current generated by the Gd sublattice upon laser-pulse excitation [158, 159]. A different approach was taken, both theoretically [162] and experimentally [160], by using a synthetic-ferrimagnetic multilayer which mimics the earlier discussed properties of the GdFeCo alloy that are needed for the thermal singlepulse AOS. Figure 19a shows the magnetization of a Pt/Co/Gd stack, toggling up (light) and down (dark) upon excitation with subsequent fs laser pulses [160]. This toggling behavior is characteristic for the thermal single-pulse AOS. The use of such synthetic-ferrimagnetic multilayer allows for easy fabrication and (interface) engineering. Moreover, the specific Pt/Co/Gd stack possesses a perpendicular magnetic anisotropy (PMA) [160], large spin-orbit torques (spin Hall effect), and a sizable interfacial Dzyaloshinskii-Moriya interaction (iDMI) resulting in chiral Neél domain walls [161], making the structure an ideal candidate to facilitate the integration of AOS with spintronic devices such as the racetrack memory [163]. The potential of the stack for spintronic integration is demonstrated in Fig. 19c, showing field-free “on-the-fly” AOS in a Pt/Co/Gd (micron-sized) racetrack [161]. The figure shows the normalized anomalous Hall signal (∝ Mz ) measured in the Hall cross located at one end of the racetrack, as illustrated in Fig. 19b. The measured signal registers magnetic domains passing the cross shortly after they are written at the other side of the racetrack using thermal single-pulse AOS (red dotted lines). The domains are transported through the wire as soon as they are written by an electrical current that is sent continuously through the wire, combining the spin Hall effect in the heavy-metal Pt seed layer, the PMA in the magnetic layer, and the chiral Neél walls for coherent and efficient domain wall motion. Additionally, recent experiments have demonstrated that the DW velocity in synthetic-antiferromagnetic [164] and (synthetic) ferrimagnetic [165, 166]

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films Fig. 19 (a) Kerr microscope image of the magnetization in a Pt/Co/Gd stack after laser-pulse excitation. The labels indicate the number of subsequent linearly polarized laser pulses the spot is exposed to. Light and dark regions represent the up and down (out-of-plane) magnetization directions, respectively. Figure taken from Ref. [160]. (b,c) Measurement demonstrating “on-the-fly” AOS in a (micron-sized) Pt/Co/Gd racetrack. (Figure taken from Ref. [161] with kind permission from APS)

    527

    200 um

    (a)

    1

    2

    3

    4

    5

    6

    7

    8

    9

    10

    LI 2

    LI 1

    Ga+

    J

    J

    (b) LI 2 (norm.)

    1.0 0.5 0.0 -0.5 -1.0 0

    5

    (c)

    10

    15

    20

    25

    30

    35

    40

    45

    50

    55

    60

    Time (s)

    systems dramatically increases when having a compensated magnetization or angular momentum, respectively. Combined with the easy fabrication and engineering of the synthetic-ferrimagnetic multilayers used for the single-pulse AOS, this shows great prospect for future integrated photonic memory applications.

    Laser-Pulse-Excited Spin Currents Around the same time of the first observation of all-optical switching, it was also discovered that the fs laser-pulse excitation of ferromagnetic layers can induce a nonlocal transfer of angular momentum [10]. In other words, it was demonstrated that a current of spin polarized electrons, i.e., a spin current, can be generated upon fs laser-pulse excitation of a ferromagnetic thin film. As was the case for all-optical switching, part of the excitement over the observation derived from its potential use in future spintronic applications. Within the field of spintronics, spin currents are heavily used to manipulate magnetic information in (future) data storage devices, e.g., to write data in magnetic random access memory [167] or transport data in the earlier introduced magnetic racetrack memory [163]. Conventionally, these spin currents are generated electronically on a nanosecond time scale. The manipulation of the magnetization can be pushed to the ultrafast time scale by using the optically

    528

    M. L. M. Lalieu and B. Koopmans

    generated spin currents, which operate at the sub-picosecond time scale. On a more fundamental level, the discovery of the fs laser-pulse-excited spin current sparked a debate about the underlying mechanism of ultrafast demagnetization, discussing the importance of both local and nonlocal dissipation of angular momentum. In the following, a selected overview of the research performed on the optically generated spin currents and their ability to control the magnetization in magnetic thin films is given.

    Optically-Induced Spin Transfer The presence of laser-pulse-excited spin currents was first demonstrated in a collinear magnetic bilayer, consisting of two identical out-of-plane magnetized FM layers that were separated by a nonmagnetic conductive spacer [10]. As illustrated in Fig. 20a, the magnetizations in the FM layers can either be in a parallel (P) or an antiparallel (AP) alignment. Upon laser-pulse excitation, both FM layers demagnetize, and angular momentum is exchanged between the two layers by the spin currents that are generated in each layer and are injected into the other layer. The spin current leaving each layer is polarized along the magnetization in the concerning layer, as shown in Fig. 20a. As a result, in the case of a parallel (antiparallel) alignment, majority (minority) spins are injected in each layer after laser-pulse excitation, which increases (decreases) the magnetic moment in the layer and thereby hinders (assists) the demagnetization. The measured effect of the transfer of angular momentum on the demagnetization in either layer is

    (c)

    (b)

    H

    probe Fe

    MgO(001)

    pump

    M

    Au

    t Ru = 0.4 nm

    Normalized Δθ/θ

    (a)

    AP P

    0 mSHG

    0.5 Delay (ps)

    1.0

    Fig. 20 (a) Illustration of the magnetic collinear bilayer used to demonstrate the presence of laserpulse-excited spin currents. (b) Demagnetization traces measured with the magnetic bilayer in the parallel (P) and antiparallel (AP) configuration, showing an increase in demagnetization speed and amplitude for the AP alignment with respect to the P alignment. ((a,b) Figure taken from Ref. [10] with kind permission of Springer Nature) (c) Illustration of the second-harmonic generation pump-probe measurement on a Fe/Au bilayer, directly measuring the laser-induced spin current injected from the Fe layer into the nonmagnetic Au layer. (Figure taken from Ref. [168] with kind permission from APS)

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    529

    shown in Fig. 20b, which indeed shows that both the speed and amplitude of the demagnetization is increased for the antiparallel alignment with respect to the parallel alignment. Performing similar measurements on a bilayer with an insulating spacer showed no difference in the demagnetization for the parallel and antiparallel configurations, confirming the presence of angular momentum exchange by the laser-pulse-excited spin currents in the case of the conducting spacer. Soon after the discovery, a first model on the generation mechanism of the optically excited spin current was developed [11, 169]. In this model, a spin current is generated as a result of spin-dependent lifetimes and velocities of the excited hot electrons. As a specific example, the hot majority electrons in Ni have a longer inelastic lifetime as compared to minority electrons [τ0,↑ > τ0,↓ , Eq. (40)], and thereby will dominate the spin current. The spin current is considered to propagate in the “superdiffusive” regime, starting in the ballistic regime just after excitation and relaxing to normal diffusion on a longer time scale. Moreover, it was claimed that the demagnetization measured in thin nickel films could be explained purely based on this nonlocal transfer of angular momentum away from the probed area, starting the discussion on the local versus nonlocal dissipation of angular momentum as the driving force of fs laser-induced demagnetization. Even though the occurrence of optically induced spin currents was demonstrated in experiments such as shown in Fig. 20a, b, determining their relative contribution to sub-ps demagnetization as compared to local dissipation of angular momentum turned out to be a challenge. While some experiments were explained using the optically excited spin current as the dominant source of demagnetization [170, 171, 172, 173], other demagnetization studies showed no sign of its contribution [93]. An increasing amount of experiments, however, demonstrated that both the local and nonlocal angular momentum dissipation mechanisms are of importance and that the relative contribution of either is dependent on the precise material system [34,29,174,46]. Using element-specific XMCD measurements, even a laser-induced enhancement of the Fe magnetization in a Fe/Ru/Ni multilayer was claimed [170, 33, 34]. The enhancement to values up to 10% above the saturation magnetization was attributed to the spin current generated in the (parallel) Ni layer being absorbed in the Fe layer. This observation, however, could not be reproduced using different measurement techniques, i.e., by XMCD measurements at a different absorption edge of the Fe and Ni [174], or using the earlier discussed layer-specific MOKE technique [28] (see section “Layer-Specific MOKE”). The first direct detection of the optically excited spin current was done using a Fe/Au bilayer [168]. An illustration of the measurement is shown in Fig. 20c. A spin current was generated in the Fe layer by laser-pulse excitation at the Fe side, which was injected into the (thick) nonmagnetic Au layer and propagated toward the outer Au surface. Using a probe pulse incident from the Au side, the spin current was measured at the Au surface via magnetization-induced second-harmonic generation (MSHG) (see section “MO Spectroscopy”). Later, the spin accumulation resulting from this laser-pulse-excited spin-current injection into the nonmagnetic layer in a FM/NM bilayer was also measured using time-resolved MOKE [13, 175] and using the layer-specific MOKE technique [46].

    530

    M. L. M. Lalieu and B. Koopmans

    With the increasing amount of research on the optically excited spin currents, more mechanisms for spin current generation were proposed. At least four different mechanisms have been identified. (i) As conjectured in the original work by Malinowski [10], an optically generated spin-polarized distribution of ballistic (hot) electrons can act as a source of spin current. (ii) Adding inelastic scattering events, one arrives at the earlier discussed superdiffusive spin current model [11, 169]. (iii) Beyond the superdiffusive mechanism, assuming all electrons to be thermalized, a laser-induced temperature gradient across a ferromagnetic layer may still result in a longer-lived spin current generation at the FM/NM interface due to the spindependent Seebeck effect [176]. This concept was extended to the nonthermal regime in Ref. [177]. (iv) Also the rapid change of M itself has been proposed as a source. More specifically, the angular momentum conservation of the electronmagnon coupling causes the demagnetization to generate spin-polarized electrons at a rate of −dM/dt, which propagate into the neighboring layers. This process was modeled in the diffusive regime by Choi et al. [13]. Strongly related, but within the perspective of an s-d model, sudden laser-driven demagnetization will induce a temporal splitting of the s-electron chemical potential, which also act as a source of spin current into neighboring layers [113].

    Optical Spin-Transfer Torque Alongside the investigation on the underlying mechanism, research focused on the optically generated spin current itself and its ability to manipulate the magnetization in a second FM layer on an ultrafast time scale [12, 13, 176, 178, 179]. As an example, experiments were performed using noncollinear bilayers, as illustrated in Fig. 21a. In these experiments, the optically excited spin current generated in an outof-plane magnetized layer (FM1), called the emitter, travels through a conductive spacer layer and is injected into an in-plane (transversally) magnetized layer (FM2), called the collector. The injected transverse spins are absorbed in the collector, which results in a spin-transfer torque (STT) on its magnetization [180], canting the magnetization toward the direction of the injected spins. One way to measure the magnetization canting is by performing such an experiment in the presence of an external magnetic field that is aligned parallel to the (initial) direction of the magnetization in the collector. After the initial canting of the magnetization due to the absorbed spin-current pulse, the magnetization is no longer aligned with the applied magnetic field, resulting in a damped precessional motion of the magnetization around the applied field direction [56] (see Eq. (35)). An example of such a precession measurement for a [Co/Ni]4 /Co/Cu/Co system is presented in Fig. 21b. The figure shows the measured Kerr rotation as a function of time after the fs laser-pulse excitation (demagnetization signal is subtracted), measured with a polar TR-MOKE setup (section “Measuring Ultrafast Magnetization Dynamics”), and for two different configurations of the noncollinear magnetizations. The measured dynamics is a superposition of two precessions, which are individually shown by the red and blue solid curves for the top measurement. The blue curve corresponds to the precession of the out-of-plane magnetization in the [Co/Ni]4 emitter (blue arrow) and originates from a laser-induced change in the

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    (a)

    Pt

    Heat current

    (b)

    531

    FM1

    Cu

    FM2

    M

    spin current

    M

    0.2

    Δ

    Kerr

    (mdeg)

    0.1 0.0 –0.1 0.1 0.0 –0.1 0

    (c)

    250

    500 Delay (ps)

    750

    1,000

    8

    Efficiency (%)

    7 6 5 4 3 2 1 0

    0

    1

    2

    3

    4

    5

    6

    Co thickness (nm) Fig. 21 (a) Illustration of a noncollinear bilayers used to measure the laser-induced spin-transfer torque. (Figure taken from Ref. [13] with kind permission from Springer Nature) (b) Time-resolved measurements of the out-of-plane component of the magnetization for two different configurations of the noncollinear bilayer. (Figure adapted from Ref. [12] with kind permission from Springer Nature) (c) Efficiency as a function of the collector layer thickness, with the efficiency defined as the ratio of spins absorbed by the collector to the spins lost during demagnetization in the emitter. (Figure taken from Ref. [179] with kind permission from APS)

    532

    M. L. M. Lalieu and B. Koopmans

    anisotropy [12, 56] (as discussed in section “Conceptual Introduction”). The red curve, however, corresponds to the precession of the in-plane magnetization in the Co collector (red arrow) and is driven by the optical STT mechanism demonstrated in Fig. 21a. Comparing the dynamics for the two different configurations shows that the phase of the magnetization precession in the collector is set by the orientation of the magnetization in the emitter (i.e., the polarization of the generated spin current), as is expected for the laser-induced STT mechanism. This laser-induced STT in the noncollinear bilayer can be used as a tool to probe the optically excited spin current and has been used to investigate the transport of the laser-pulse-excited spin current through different spacer layers [12], the optical excitation of spin currents in ferrimagnetic alloys [159], and the dependence of the optical spin current generation on the emitter thickness [179]. The latter showed that the full thickness of the emitter (an out-of-plane magnetized [Co/Ni]N multilayer) contributes to the generated spin current (at least up to 3.4 nm), which is consistent with recently performed theoretical calculations using an extended superdiffusive spin current model [181] but contradicts an earlier observed optical spin-emission region in Fe of ≈1 nm [182]. Additionally, the absorption of the optically excited spin current in the collector was investigated using a wedge-shaped collector layer. The result of this measurement is presented in Fig. 21c. The figure shows the efficiency as a function of the Co collector thickness, with the efficiency defined as the ratio of spins absorbed by the collector to the spins lost during demagnetization in the [Co/Ni]4 emitter. It can be seen that the spin current is absorbed very locally near the injection interface, i.e., 90 % of the transverse spins are absorbed within the first ≈2 nm of the collector layer.

    Optical THz Spin Wave Excitation As it turns out, the very local absorption of the short and intense optically generated spin current pulse near the injection interface provides ideal conditions for the excitation of higher-order (sub)THz standing spin waves in the collector [178, 179]. This is demonstrated for a noncollinear bilayer with a 5.5 nm thick Co collector layer in Fig. 22a. The measured polar TR-MOKE signal shows both the first-order standing spin wave (0.55 THz) as well as the uniform precession (≈10 GHz). As illustrated in the inset of the figure, the higher-order THz spin waves are excited via the creation of a strong gradient in the canting angle of the magnetization, created by the very local absorption of the optically generated spin current near the interface. This highly nonequilibrium magnetization state leads to (damped) standing spin waves, which are driven by the exchange interaction and can be excited without the presence of an externally applied field. The frequency of the standing spin waves is determined by their wavelength, which in turn is set by the thickness of the collector and reaches above 1 THz for a 3 nm thick Co layer [179]. In the measurement presented in Fig. 22a, only the first-order standing spin wave is observed. Higher-order modes were observed in a different measurement performed on a noncollinear bilayer with two (perpendicularly aligned) in-plane magnetized layers and with a thicker collector (a 14 nm thick Fe layer) [178]. The result of that measurement is presented in Fig. 22b. In this figure, the left

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    533

    4.0

    (a) MOKE signal (arb. unit.)

    3.5 3.0 2.5

    0.55 THz

    2.0 1.5 1.0 0.5 0.0 -0.5

    10.23 GHz 0

    5

    20

    40

    60

    80

    100

    120

    140

    160

    180

    200

    Delay (ps)

    (b)

    1.2

    (c) ×20

    n =4

    Frequency (THz)

    n =3 n =2

    0.8

    n =1 n =0

    0.4 DFe = 280 meV·Å2

    0

    0 1 2 3 4 5 Eigenmode number n = k d /π

    6

    d = 14 nm

    Fig. 22 (a) Measurement of the (sub)THz standing-spin-wave excitation using a noncollinear bilayer with a 5.5 nm thick Co collector layer. The figure shows the polar TR-MOKE signal measured as a function of time after the laser-pulse excitation. Both the first-order standing spin wave (0.55 THz) as well as the uniform precession (≈10 GHz) are visible. The inset shows the gradient in the magnetization within the collector layer after the optical excitation and the resulting uniform and first-order standing spin wave. Figure taken from [179]. (b) Left panel: Fourier spectrum of the magnetization dynamics measured in a 14 nm thick Fe collector layer after laserpulse excitation. Right panel: calculated spin-wave dispersion, in which the dots represent the frequencies corresponding to the different standing-spin-wave orders. There is a match between the peaks in the Fourier spectrum and the calculated standing-spin-wave frequencies up to n = 4 (solid dots). (c) Illustration of the different standing-spin-wave orders. ( (b,c) Figure taken from Ref. [178])

    panel shows the Fourier spectrum of the magnetization dynamics measured in the collector after laser-pulse excitation. The right panel displays the calculated spin wave dispersion, in which the dots represent the frequencies corresponding to the

    534

    M. L. M. Lalieu and B. Koopmans

    different standing-spin-wave orders. As can be seen, there is a match between the peaks in the Fourier spectrum and the calculated standing-spin-wave frequencies up to n = 4 (solid dots), demonstrating that standing spin waves up to the fourthorder mode were excited in the noncollinear bilayer by fs laser-pulse excitation; see also Fig. 22c. Recently, the higher-order standing spin wave excitation via the ultrafast spin-transfer torque from an injected fs transverse spin current pulse was reproduced using micromagnetic modeling [183]. Moreover, an experimental design that allows for the lateral propagation of these laser-pulse-excited spin waves was suggested. These findings show that in addition to its general importance in the field of spintronics, the optically excited spin currents could also be of high potential for future THz magnonics.

    Conclusions and Outlook At the end of this chapter, we briefly highlight some of the exciting trends in the field and sketch promising routes for future research. Magneto-optics is known for almost two centuries, and experimental approaches including Kerr microscopy have well matured. Nevertheless, applications to new systems and phenomena keep introducing new demands and opportunities. The rise of thin film magnetism and the birth of spintronics in the last two decades of the previous century have pushed applications of the magneto-optic Kerr effect enormously. A similar boost has come from recent trends in spintronics, such as novel device architectures based on magnetic domain walls that are driven by spinorbit torques, which has caused a revival of Kerr microscopy. Another trigger pushing methodological developments has come from the new field of opto-magnetism, requiring magneto-optical techniques with femtosecond resolution. In many cases this can be easily established by pump-probe configurations and measuring in a stroboscopic manner. But such an approach is of limited value if stochasticity plays a role. This faces new challenges, in particular when aiming at microscopy. Developments of single-shot wide-field magneto-optical microscopy and other experimental approaches with combined spatial resolution down to the nanometer scale and sub-ps time resolution are emerging and will most probably grow in importance. An interesting case is provided by establishing depth resolution, which could certainly be pushed beyond todays achievements. As briefly discussed in this chapter, together with experimental progress in using visible and near-infrared (laser) light for magneto-optical and opto-magnetic experiments, rapid developments using other parts of the electromagnetic spectrum have been witnessed. In this sense X-ray techniques are particularly noteworthy because of their element specificity, as well as distinguishing spin and orbital magnetism. Table-top implementations using high-harmonic generation will enable much more widespread use and possibly new spectroscopic approaches. Finally, a rapid growth of THz studies is being seen. They provide access to ac conductivity of electrons and a complementary view on sub-ps magnetization dynamics and spin currents.

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    535

    In the field of opto-magnetism, a trend toward further refinement of the optical manipulation will certainly continue. Prominent issues on the agenda for the forthcoming decade will be improving the energy efficiency as well as spatial resolution of all-optical switching. As to the latter, first attempts of using plasmonics for further downscaling beyond the diffraction limit have been reported, but it will take intensive research efforts to push this direction toward more robust control. Another issue relates to exploring the ultimate data rates possible. Note that the fact that we can switch magnetization by sub-ps laser pulses is scientifically of interest, but for applications the data rate – determined by the minimum time duration between two succeeding pulses to establish two independent switching processes – will be the decisive factor. Pushing all-optical manipulation (switching) of magnetic matter to a next level of control will certainly rely on a better understanding of the underlying processes. The basics of the single-pulse exchange-driven toggle mechanism are well understood and can be qualitatively reproduced by different models. However, a number of outstanding questions are indicating that our knowledge is far from complete. It seems that Gd is an essential element to establish single-pulse AOS, but it is not known yet why. Also, it has been claimed that spin transport may be of crucial relevance for successful AOS, but its importance has yet to be confirmed. More general, insight is lacking as to the relative role of electron-mediated spin currents, exchange scattering, and magnon spin currents in the AOS process. Beyond understanding the toggle-switching scenario, many open questions exist with respect to the origin of helicity-dependent switching, whether it can be pushed toward the single-pulse regime and about the role of the laser pulse length. Beyond using rare-earth-transition ferromagnetic metal alloys and layered synthetic ferrimagnets, it is of significant interest to explore other materials that could display the same or similar phenomena, maybe – at least with respect to some aspects – in a superior way. In this context, entirely different means of nonthermal switching in magnetic dielectric materials has been successfully started to be explored. These dielectrics provide a complementary route toward highly efficient, low power switching. For sure the switching mechanism is by far superior when it comes to dissipated laser power, but due to the low absorption, it needs extremely high laser power to excite. Moreover, the specific materials may be less easily integrated in devices, and future research should clarify its true potential. Another highly interesting route is linked to the present explosion in the use of antiferromagnets for spintronics. Ideas are emerging as to how antiferromagnets can be switched using ultrashort laser pulses. A particularly noteworthy trend that we signaled in this chapter is the merge of magneto-optics, opto-magnetism, and spintronics. We foresee that this development will further evolve. On the one hand, it provides and interesting route toward novel applications in a More than Moore context (NB: More than Moore refers to a trend in modern chip technology to embed new and complementary functionalities onchip, rather than further pushing miniaturization according to Moore’s law). On the other hand, it provides a beautiful playground for fundamental research. The demonstrated ability to generate spin currents by fs laser pulses at the nanometer

    536

    M. L. M. Lalieu and B. Koopmans

    scale allows for a complementary view on spintronic concepts and in particular to explore their ultimate limits. We emphasize that the recent finding that AOS can be established in exactly the type of easy-engineerable ultrathin-layered systems used for spintronics will further fuel the successful merge of opto-magnetism and spintronics. In this context it is also worthwhile mentioning recent THz studies on ultrafast magnetization dynamics, as another means of driving ac free-electron (spin) currents. Toward further understanding of the ultrafast laser-induced magnetization dynamics, including AOS, an important role will inevitably be played by refinement of theory and increase of numerical abilities. As to magneto-optics, ab initio density functional schemes for calculating magneto-optical response have been successfully introduced already by the end of the previous century, but the problem describing femto-second opto-magnetism is orders of magnitude more complex. In the past decade, a number of semiempirical theoretical frameworks have been developed that can successfully reproduce – and sometimes even predict – a wide range of experiments. However, it is only since very recently that fully ab initio approaches, such as time-dependent density functional theory, have been able to make contact with laser-excited femtomagnetism. While before 2010 it was still unimaginable to include all involved degrees of freedom (electronic, lattice, and spin) with enough accuracy in a time-dependent ab initio code, due to more efficient codes and increase of computational power, promising breakthroughs are now being reported. It is expected that such ab initio approaches will become of increasing importance to guide the field further in the forthcoming decade. Some final words are reserved for sketching the application potential of the exciting phenomena discussed in the present chapter. Commercial use of magnetooptical recording dates back to the mini-disk system in the 1990s. Despite high expectations in the decade that followed, next generations of optical storage technology moved away from magneto-optics. However, a very exciting combination of magnetic hard disk recording with optics is presently becoming more mature. In heat-assisted magnetic recording (HAMR), an optical waveguide is integrated with the magnetic write head. Laser light is focused to a 50 nm spot using plasmonic tricks, to locally heat up the storage medium and lower its coercivity. This process enables using harder magnetic media and thus pushes data storage densities. Very excitingly, after 20 years of industrial development, this approach is now at the brink of being launched commercially. This success of HAMR might fuel expectations for other application using combinations of light and magnetism. Actually, ever since the first reports on ultrafast laser-induced magnetization dynamics, and in particular AOS, it has been (maybe overoptimistically) positioned as baring enormous potential for applications. Its potential for recording yet has to be proven, but a particular appealing direction is to consider applications beyond (disk-based) storage technology. Both magnetooptics and opto-magnetism could be envisioned to be implemented in photonic integrated circuits, to enrich their functionality. Magneto-optics has been considered for realizing a compact on-chip optical isolator already for decades, albeit with little success so far. However, first research initiatives have been established that aim at

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    537

    a broader implementation of spintronic and magnetic functionalities – in particular include AOS – in integrated photonics. It is a fascinating thought that one day the merge of magneto-optics, opto-magnetism, and spintronics might lead to spin off in such an applied photonic context.

    Notes 1 An

    exception is provided by topological MO effects for crystals with a specific magnetic point group symmetry with a chiral spin ordering [19]. 2 Note that in the presence of spin-orbit coupling S and L are no good quantum numbers either, but we might have expected the total J to be still conserved, but this is apparently not the case.

    References 1. Faraday, M.: On the magnetization of light and the illumination of magnetic lines of force. Trans. Roy. Soc. 5, 592 (1864) 2. Kerr, J.: On the rotation of the plane of polarization by reflection from the pole of a magnet. Phil. Mag. 3, 321–343 (1877) 3. Mansuripur, M.: The Physical Principles of Magneto-Optical Recording. Cambridge University Press (1995) 4. Beaurepaire, E., Merle, J.-C., Daunois, A., Bigot, J.-Y.: Ultrafast spin dynamics in ferromagnetic nickel. Phys. Rev. Lett. 76, 4250–4253 (1996) 5. Stanciu, C.D., Hansteen, F., Kimel, A.V., Kirilyuk, A., Tsukamoto, A., Itoh, A., Rasing, T.: All-optical magnetic recording with circularly polarized light. Phys. Rev. Lett. 99, 047601 (2007) 6. Radu, I., Vahaplar, K., Stamm, C., Kachel, T., Pontius, N., Durr, H.A., Ostler, T.A., Barker, J., Evans, R.F.L., Chantrell, R.W., Tsukamoto, A., Itoh, A., Kirilyuk, A., Rasing, T., Kimel, A.V.: Transient ferromagnetic-like state mediating ultrafast reversal of antiferromagnetically coupled spins. Nature 472, 205–208 (2011) 7. Ostler, T.A., Barker, J., Evans, R.F.L., Chantrell, R.W., Atxitia, U., Chubykalo-Fesenko, O., El Moussaoui, S., Le Guyader, L., Mengotti, E., Heyderman, L.J., Nolting, F., Tsukamoto, A., Itoh, A., Afanasiev, D., Ivanov, B.A., Kalashnikova, A.M., Vahaplar, K., Mentink, J., Kirilyuk, A., Rasing, T., Kimel, A.V.: Ultrafast heating as a sufficient stimulus for magnetization reversal in a ferrimagnet. Nat. Commun. 3, 666 (2012) 8. Mangin, S., Gottwald, M., Lambert, C-H., Steil, D., Uhlír, V., Pang, L., Hehn, M., Alebrand, S., Cinchetti, M., Malinowski, G., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: Engineered materials for all-optical helicity-dependent magnetic switching. Nat. Mater. 13, 286–292 (2014) 9. Lambert, C-H., Mangin, S., Varaprasad, B.S.D.C.S., Takahashi, Y.K., Hehn, M., Cinchetti, M., Malinowski, G., Hono, K., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: All-optical control of ferromagnetic thin films and nanostructures. Science 345, 1337–1340 (2014) 10. Malinowski, G., Della Longa, F., Rietjens, J.H.H., Paluskar, P.V., Huijink, R., Swagten, H.J.M., Koopmans, B.: Control of speed and efficiency of ultrafast demagnetization by direct transfer of spin angular momentum. Nat. Phys. 4, 855–858 (2008) 11. Battiato, M., Carva, K., Oppeneer, P.M.: Superdiffusive spin transport as a mechanism of ultrafast demagnetization. Phys. Rev. Lett. 105, 027203 (2010) 12. Schellekens, A.J., Kuiper, K.C., de Wit, R.R.J.C., Koopmans, B.: Ultrafast spin-transfer torque driven by femtosecond pulsed-laser excitation. Nat. Commun. 5, 4333 (2014) 13. Choi, G-M., Min, B-C., Lee, K-J., Cahill, D.G.: Spin current generated by thermally driven ultrafast demagnetization. Nat. Commun. 5, 4334 (2014)

    538

    M. L. M. Lalieu and B. Koopmans

    14. Zak, J., Moog, E.R., Liu, C., Bader, S.D.: Universal approach to magneto-optics. J. Magn. Magn. Mat. 89, 107–123 (1990) 15. Jones, R.C.: A new calculus for the treatment of optical systems. J. Opt. Soc. Am. 31, 488–493 (1941) 16. Rzazewski, K., Boyd, R.W.: Equivalence of interaction Hamiltonians in the electric dipole approximation. J. Mod. Opt. 51, 1137–1147 (2004) 17. Oppeneer, P.M.: Magneto-optical kerr spectra. In: Buschow, K.H.J. (ed.) Handbook of Magnetic Materials: Volume 13, pp. 229–422. Elsevier Science (2001) 18. Bennett, H.S., Stern, E.A.: Faraday effects in solids. Phys. Rev. 137, A448–A470 (1965) 19. Feng, W., Zhou, X., Hanke, J-P., Guo, G-Y., Blügel, S., Mokrousov, Y., Yao, Y.: Topological magneto-optical effect and its quantization in noncoplanar antiferromagnets (2018) arXiv:1811.05803. 20. Oppeneer, P.M.: Magneto-optical spectroscopy in the valence-band energy regime: relationship to the magnetocrystalline anisotropy. J. Magn. Magn. Mat. 188, 275–285 (1998) 21. Krinchik, G.S., Artem’ev, V.A.: Magneto-optical properties of Ni, Co, and Fe in the ultraviolet visible, and infrared parts of the spectrum. Zh. Eksp. Teor. Fiz. 53, 1901–1912 (1968) 22. Schellekens, A.J.: Manipulating spins. Ph.d. thesis, Eindhoven University of Technology (2014) 23. Ebert, H.: Magneto-optical effects in transition metal systems. Rep. Prog. Phys. 59, 1665– 1735 (1996) 24. Hashimoto, S., Ochiai, Y., Aso, K.: Film thickness dependence of magneto-optical and magnetic properties in Co/Pt and Co/Pd multilayers. J. Appl. Phys. 67, 4429–4431 (1990) 25. Koopmans, B., van Kampen, M., Kohlhepp, J.T., de Jonge, W.J.M.: Femtosecond spin dynamics of epitaxial Cu(111)/Ni/Cu wedges. J. Appl. Phys. 87, 5070 (2000) 26. Koopmans, B.: Laser-induced magnetization dynamics. In: Hillebrands, B., Ounadjela, K. (eds.) Spindynamics in Confined Magnetic Structures II, vol. 87, pp. 253–320. Springer (2003) 27. Hamrle, J., Ferré, J., Nývlt, M., Višˇnovský, Š.: In-depth resolution of the magneto-optical kerr effect in ferromagnetic multilayers. Phys. Rev. B 66, 224423 (2002) 28. Schellekens, A.J., de Vries, N., Lucassen, J., Koopmans, B.: Exploring laser-induced interlayer spin transfer by an all-optical method. Phys. Rev. B 90, 104429 (2014) 29. Wieczorek, J., Eschenlohr, A., Weidtmann, B., Rösner, M., Bergeard, N., Tarasevitch, A., Wehling, T.O., Bovensiepen, U.: Separation of ultrafast spin currents and spin-flip scattering in Co/Cu(001) driven by femtosecond laser excitation employing the complex magnetooptical Kerr effect. Phys. Rev. B 92, 174410 (2015) 30. Khorsand, A.R., Savoini, M., Kirilyuk, A., Kimel, A.V., Tsukamoto, A., Itoh, A., Rasing, T.: Element-specific probing of ultrafast spin dynamics in multisublattice magnets with visible light. Phys. Rev. Lett. 110, 107205 (2013) 31. Thole, B.T., Carra, P., Sette, F., van der Laan, G.: X-ray circular dichroism as a probe of orbital magnetization. Phys. Rev. Lett. 68, 1943–1946 (1992) 32. Carra, P., Thole, B.T., Altarelli, M., Wang, X.: X-ray circular dichroism and local magnetic fields. Phys. Rev. Lett. 70, 694–697 (1993) 33. Mathias, S., La o vorakiat, C., Shaw, J.M., Turgut, E., Grychtol, P., Adam, R., Rudolf, D., Nembach, H.T., Silva, T.J., Aeschlimann, M., Schneider, C.M., Kapteyn, H.C., Murnane, M.M.: Ultrafast element-specific magnetization dynamics of complex magnetic materials on a table-top. J. Electron. Spectrosc. Relat. Phenom. 189, 164–170 (2013) 34. Turgut, E., La o vorakiat, C., Shaw, J.M., Grychtol, P., Nembach, H.T., Rudolf, D., Adam, R., Aeschlimann, M., Schneider, C.M., Silva, T.J., Murnane, M.M., Kapteyn, H.C., Mathias, S.: Controlling the competition between optically induced ultrafast spin-flip scattering and spin transport in magnetic multilayers. Phys. Rev. Lett. 110, 197201 (2013) 35. Pan, R-P., Wei, H.D., Shen, Y.R.: Optical second-harmonic generation from magnetized surfaces. Phys. Rev. B 39, 1229–1234 (1989) 36. Koopmans, B., Janner, A-M., Jonkman, H.T., Sawatzky, G.A., van der Woude, F.: Strong bulk magnetic dipole induced second-harmonic generation from C60 . Phys. Rev. Lett. 71, 3569– 3572 (1993)

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    539

    37. Qiu, Z.Q., Bader, S.D.: Surface magneto-optic kerr effect. Rev. Sci. Instrum. 71, 1243–1255 (2000) 38. Allwood, D.A., Xiong, G., Cooke, M.D., Cowburn, R.P.: Magneto-optical Kerr effect analysis of magnetic nanostructures. J. Phys. D Appl. Phys. 36, 2175–2182 (2003) 39. Schäfer, R.: Investigation of domains and dynamics of domain walls by the magneto-optical Kerr effect. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials. Wiley (2007) 40. Soldatov, I.V., Schäfer, R.: Selective sensitivity in kerr microscopy. Rev. Sci. Instrum. 88, 073701 (2017) 41. Buess, M., Höllinger, R., Haug, T., Perzlmaier, K., Krey, U., Pescia, D., Scheinfein, M.R., Weiss, D., Back, C.H.: Fourier transform imaging of spin vortex eigenmodes. Phys. Rev. Lett. 93, 077207 (2004) 42. Neudert, A., McCord, J., Chumakov, D., Schäfer, R., Schultz, L.: Small-amplitude magnetization dynamics in permalloy elements investigated by time-resolved wide-field Kerr microscopy. Phys. Rev. B 71, 134405 (2005) 43. Bigot, J.-Y., Guidoni, L., Beaurepaire, E., Saeta, P.N.: Femtosecond spectrotemporal magneto-optics. Phys. Rev. Lett. 93, 077401 (2004) 44. Ju, G., Vertikov, A., Nurmikko, A.V., Canady, C., Xiao, G., Farrow, R.F.C., Cebollada, A.: Ultrafast nonequilibrium spin dynamics in a ferromagnetic thin film. Phys. Rev. B 57, R700– R703 (1998) 45. Dalla Longa, F.: Laser-induced magnetization dynamics. Ph.d thesis, Eindhoven University of Technology (2008) 46. Hofherr, M., Maldonado, P., Schmitt, O., Berritta, M., Bierbrauer, U., Sadashivaiah, S., Schellekens, A.J., Koopmans, B., Steil, D., Cinchetti, M., Stadtmüller, B., Oppeneer, P.M., Mathias, S., Aeschlimann, M.: Speed and efficiency of femtosecond spin current injection into a nonmagnetic material. Phys. Rev. B 96, 100403(R) (2017) 47. Hohlfeld, J., Matthias, E., Knorren, R., Bennemann, K.H.: Nonequilibrium magnetization dynamics of nickel. Phys. Rev. Lett. 78, 4861–4864 (1997) 48. Miltat, J., Albuquerque, G., Thiaville, A.: An introduction to micromagnetics in the dynamic regime. In: Hillebrands, B., Ounadjela, K. (eds.) Spindynamics in Confined Magnetic Structures I, vol. 83, pp. 1–33. Springer (2002) 49. Hohlfeld, J., Gerrits, T., Bilderbeek, M., Rasing, T., Awano, H., Ohta, N.: Fast magnetization reversal of GdFeCo induced by femtosecond laser pulses. Phys. Rev. B 65, 012413 (2001) 50. Ju, G., Hohlfeld, J., Bergman, B., van de Veerdonk, R.J.M., Mryasov, O.N., Kim, J.-Y., Wu, X., Weller, D., Koopmans, B.: Ultrafast generation of ferromagnetic order via laser-induced phase transformation in FeRh thin films. Phys. Rev. Lett. 93, 197403 (2004) 51. Thiele, J.-U., Buess, M., Back, C.H.: Spin dynamics of the ferromagnetic-toantiferromagnetic phase transition in FeRh on a sub-picosecond time scale. Appl. Phys. Lett. 85, 2857–2859 (2004) 52. Radu, I., Stamm, C., Pontius, N., Kachel, T., Ramm, P., Thiele, J.-U., Dürr, H.A., Back, C.H.: Laser-induced generation and quenching of magnetization on FeRh studied with timeresolved X-ray magnetic circular dichroism. Phys. Rev. B 81, 104415 (2010) 53. Pressacco, F., Uhlí˜r, V., Gatti, M., Nicolaou, A., Bendounan, A., Arregi, J.A., Patel, S.K.K., Fullerton, E.E., Krizmancic, D., Sirotti, F.: Laser induced phase transition in epitaxial FeRh layers studied by pump-probe valence band photoemission (2018). arXiv:1803.00780 54. van Kampen, M., Koopmans, B., Kohlhepp, J.T., de Jonge, W.J.M.: Laser-induced precession in canted-spin ferromagnets. J. Magn. Magn. Mat. 240, 291–293 (2002) 55. Koopmans, B., Ruigrok, J.J.M., Longa, F.D., de Jonge, W.J.M.: Unifying ultrafast magnetization dynamics. Phys. Rev. Lett. 95, 267207 (2005) 56. van Kampen, M., Jozsa, C., Kohlhepp, J.T., LeClair, P., Lagae, L., de Jonge, W.J.M., Koopmans, B.: All-optical probe of coherent spin waves. Phys. Rev. Lett. 88, 227201 (2002) 57. Ju, G., Nurmikko, A.V., Farrow, R.F.C., Marks, R.F., Carey, M.J., Gurney, B.A.: Ultrafast optical modulation of an exchange biased ferromagnetic/antiferromagnetic bilayer. Phys. Rev. B 58, R11857–R11860 (1998)

    540

    M. L. M. Lalieu and B. Koopmans

    58. Hübner, W., Zhang, G.P.: Ultrafast spin dynamics in nickel. Phys. Rev. B 58, R5920–R5920 (1998) 59. Bigot, J.-Y., Vomir, M., Beaurepaire, E.: Coherent ultrafast magnetism induced by femtosecond laser pulses. Nat. Phys. 5, 515–520 (2009) 60. Stupakiewicz, A., Szerenos, K., Afanasiev, D., Kirilyuk, A., Kimel, A.V.: Ultrafast nonthermal photo-magnetic recording in a transparent medium. Nature 542, 71–74 (2017) 61. van der Ziel, J.P., Pershan, P.S., Malmstrom, L.D.: Optically-induced magnetization resulting from the inverse Faraday effect. Phys. Rev. Lett. 15, 190–193 (1965) 62. Agranat, M.B., Ashikov, S.I., Granovskii, A.B., Rukman, G.I.: Interaction of picosecond laser pulses with the electron, spin and phonon subsystem of nickel. Zh. Eksp. Teor. Fiz. 86, 1376– 1379 (1984) 63. Vaterlaus, A., Beutler, T., Meier, F.: Spin-lattice relaxation time of ferromagnetic gadolinium determined with time-resolved spin-polarized photoemission. Phys. Rev. Lett. 67, 3314–3317 (1991) 64. Vaterlaus, A., Beutler, T., Guarisco, D., Lutz, M., Meier, F.: Spin-lattice relaxation in ferromagnets studied by time-resolved spin-polarized photoemission. Phys. Rev. B 46, 5280– 5286 (1992) 65. Hübner, W., Bennemann, K.H.: Simple theory for spin-lattice relaxation in metallic rare-earth ferromagnets. Phys. Rev. B 53, 3422–3427 (1996) 66. Scholl, A., Baumgarten, L., Jacquemin, R., Eberhardt, W.: Ultrafast spin dynamics of ferromagnetic thin films observed by fs spin-resolved two-photon photemission. Phys. Rev. Lett. 79, 5146–5149 (1997) 67. Koopmans, B., van Kampen, M., Kohlhepp, J.T., de Jonge, W.J.M.: Ultrafast magneto-optics in nickel: magnetism or optics? Phys. Rev. Lett. 85, 844–847 (2000) 68. Koopmans, B., van Kampen, M., de Jonge, W.J.M.: Experimental access to femtosecond spin dynamics. J. Phys. Cond. Mat. 15, S723–S736 (2003) 69. Regensburger, H., Vollmer, R., Kirschner, J.: Time-resolved magnetization-induced second-harmonic generation from the Ni(110) surface. Phys. Rev. B 61, 14716–14719 (2000) 70. Guidoni, L., Beaurepaire, E., Bigot, J.-Y.: Magneto-optics in the ultarfast regime: thermalization of spin populations in ferromagnetic films. Phys. Rev. Lett. 89, 017401 (2002) 71. Oppeneer, P.M., Liebsch, A.: Ultrafast demagnetization in nickel: theory of magneto-optics for non-equilibrium electron distributions. J. Phys. Cond. Mat. 16, 5519–5530 (2004) 72. Dewhurst, J.K., Shallcross, S., Gross, E.K.U., Sharma, S.: Substrate-controlled ultrafast spin injection and demagnetization. Phys. Rev. A 10, 044065 (2018) 73. Dewhurst, J.K., Elliott, P., Shallcross, S., Gross, E.K.U., Sharma, S.: Laser-induced intersite spin transfer. Nano Lett. 18, 1842–1848 (2018) 74. Beaurepaire, E., Maret, M., Halté, V., Merle, J.-C., Daunois, A., Bigot, J.-Y.: Spin dynamics in CoPt3 alloy films: a magnetic phase transition in the femtosecond time scale. Phys. Rev. B 58, 12134–12137 (1998) 75. Wietstruk, M., Melnikov, A., Stamm, C., Kachel, T., Pontius, N., Sultan, M., Gahl, C., Weinelt, M., Dürr, H.A., Bovensiepen, U.: Hot-electron-driven enhancement of spin-lattice coupling in Gd and Tb 4f ferromagnets observed by femtosecond X-ray magnetic circular dichroism. Phys. Rev. Lett. 106, 127401 (2011) 76. Müller, G.M., Walowski, J., Djordjevic, M., Miao, G-X., Gupta, A., Ramos, A.V., Gehrke, K., Moshnyaga, V., Samwer, K., Schmalhorst, J., Thomas, A., Hütten, A., Reiss, G., Moodera, J.S., Münzenberg, M.: Spin polarization in half-metals probed by femtosecond spin excitation. Nat. Mater. 8, 56–61 (2009) 77. Ogasawara, T., Ohgushi, K., Tomioka, Y., Takahashi, K.S., Okamoto, H., Kawasaki, M., Tokura, Y.: General features of photoinduced spin dynamics in ferromagnetic and ferrimagnetic compounds. Phys. Rev. Lett. 94, 087202 (2005) 78. Steil, D., Alebrand, S., Roth, T., Krauß, M., Kubota, T., Oogane, M., Ando, Y., Schneider, H.C., Aeschlimann, M., Cinchetti, M.: Band-structure-dependent demagnetization in the Heusler alloy Co2 Mn1−x Fex Si. Phys. Rev. Lett. 105, 217202 (2010)

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    541

    79. Wanh, J., Khodaparast, G.A., Kono, J., Slupinski, T., Oiwa, A., Munekata, H.: Ultrafast optical manipulation of ferromagentic order in InMnAs/GaSb. J. Supercond. 16, 373–377 (2003) 80. Wang, J., Sun, C., Kono, J., Oiwa, A., Munekata, H., Cywinski, L., Sham, J.J.: Ultrafast quenching of ferromagnetism in InMnAs induced by intense laser irradiation. Phys. Rev. Lett. 95, 167401 (2005) 81. Kise, T., Ogasawara, T., Ashida, M., Tomioka, Y., Tokura, Y., Kuwata-Gonokami, M.: Ultrafast spin dynamics and critical behavior in half-metallic ferromagnet: Sr2 FeMoO6 . Phys. Rev. Lett. 85, 1986–1989 (2000) 82. Rhie, H.S., Dürr, H.A., Eberhardt, W.: Femtosecond electron and spin dynamics in Ni/W(110) films. Phys. Rev. Lett. 90, 247201 (2003) 83. Dürr, H.A., Stamm, C., Kachel, T., Pontius, N., Mitzner, R., Quast, T., Holldack, K., Khan, S., Lupulescu, C., Aziz, E.F., Wietstruk, M., Eberhardt, W.: Ultrafast electron and spin dynamics in nickel probed with femtosecond X-ray pulses. IEEE Trans. Magn. 44, 1957–1961 (2008) 84. Boeglin, C., Beaurepaire, E., Halté, V., López-Flores, V., Stamm, C., Pontius, N., Dürr, H.A., Bigot, J.-Y.: Distinguishing the ultrafast dynamics of spin and orbital moments in solids. Nature 465, 458–461 (2010) 85. Roth, T., Schellekens, A.J., Alebrand, S., Schmitt, O., Steil, D., Koopmans, B., Cinchetti, M., Aeschlimann, M.: Temperature dependence of laser-induced demagnetization in Ni: a key for identifying the underlying mechanism. Phys. Rev. X 2, 021006 (2012) 86. Bonetti, S., Hoffmann, M.C., Sher, M.-J., Chen, Z., Yang, S.-H., Samant, M.G., Parkin, S.S.P., Dürr, H.A.: THz-driven ultrafast spin-lattice scattering in amorphous metallic ferromagnets. Phys. Rev. Lett. 117, 087205 (2016) 87. Beaurepaire, E., Turner, G.M., Harrel, S.M., Beard, M.C., Bigot, J.-Y., Schmuttenmaer, C.A.: Coherent terahertz emission from ferromagnetic films excited by femtosecond laser pulses. Appl. Phys. Lett. 84, 3465–3467 (2004) 88. Hilton, D.J., Averitt, R.D., Meserole, C.A., Fisher, G.L., Funk, D.J., Thompson, J.D., Taylor, A.J.: Terahertz emission via ultrashort-pulse excitation of magnetic metal films. Opt. Lett. 29, 1805–1807 (2004) 89. Huisman, T.J., Mikhaylovskiy, R.V., Tsukamoto, A., Rasing, T., Kimel, A.V.: Simultaneous measurements of terahertz emission and magneto-optical kerr effect for resolving ultrafast laser-induced demagnetization dynamics. Phys. Rev. B 92, 104419 (2015) 90. Bergeard, N., Hehn, M., Mangin, S., Lengaigne, G., Montaigne, F., Lalieu, M.L.M., Koopmans, B., Malinowski, G.: Hot-electron-induced ultrafast demagnetization in Co/Pt multilayers. Phys. Rev. Lett. 117, 147203 (2016) 91. Ferté, T., Bergeard, N., Malinowski, G., Abrudan, R., Kachel, T., Holldack, K., Hehn, M., Boeglin, C.: Ultrafast hot-electron induced quenching of Tb 4f magnetic order. Phys. Rev. B 96, 144427 (2017) 92. Wilson, R.B., Yang, Y., Gorchon, J., Lambert, C-H., Salahuddin, S., Bokor, J.: Electric current induced ultrafast demagnetization. Phys. Rev. B 96, 045105 (2017) 93. Schellekens, A.J., Verhoeven, W., Vader, T.N., Koopmans, B.: Investigating the contribution of superdiffusive transport to ultrafast demagnetization of ferromagnetic thin films. Appl. Phys. Lett. 102, 252408 (2013) 94. Einstein, A., de Haas, W.J.: Experimental proof of the existence of Ampére’s molecular currents. KNAW Proc. 18 I, 696–711 (1915) 95. Frenkel, V.Y.: On the history of the Einstein’de Haas effect. Usp. Fiz. Nauk 128, 545–557 (1979) 96. Töws, W., Pastor, G.M.: Many-body theory of ultrafast demagnetization and angular momentum transfer in ferromagnetic transition metals. Phys. Rev. Lett. 115, 217204 (2015) 97. Krauß, M., Roth, T., Alebrand, S., Steil, D., Cinchetti, M., Aeschlimann, M., Schneider, H.C.: Ultrafast demagnetization of ferromagnetic transition metals: the role of the coulomb interaction. Phys. Rev. B 80, 180407 (2009) 98. Fähnle, M., Haag, M., Illg, C.: Is the angular momentum of a ferromagnetic sample after exposure to a fs laser pulse conserved? J. Magn. Magn. Mat. 347, 45–46 (2013)

    542

    M. L. M. Lalieu and B. Koopmans

    99. Elliott, R.J.: Theory of the effect of spin-orbit coupling on magnetic resonance in some semiconductors. Phys. Rev. 96, 266–279 (1954) 100. Rhie, H.S., Dürr, H.A., Eberhardt, W.: Femtosecond evolution of electronic interactions at the Ni(111) surface. Appl. Phys. A 82, 9–14 (2005) 101. Aeschlimann, M., Bauer, M., Pawlik, S., Weber, W., Burgermeister, R., Oberli, D., Siegmann, H.C.: Ultrafast spin-dependent electron dynamics in fcc Co. Phys. Rev. Lett. 79, 5158–5161 (1997) 102. Dalla Longa, F., Kohlhepp, J.T., de Jonge, W.J.M., Koopmans, B.: Influence of photon angular momentum on ultrafast demagnetization in nickel. Phys. Rev. B 75, 224431 (2007) 103. Koopmans, B., Malinowski, G., Dalla Longa, F., Steiauf, D., Fahnle, M., Roth, T., Cinchetti, M., Aeschlimann, M.: Explaining the paradoxical diversity of ultrafast laser-induced demagnetization. Nat. Mater. 9, 259–265 (2010) 104. Zhang, G.P., Hübner, W.: Femtosecond spin dynamics in the time domain. J. Appl. Phys. 85, 5657–5659 (1999) 105. Zhang, G.P., Hübner, W.: Laser-induced ultrafast demagnetization in ferromagnetic metals. Phys. Rev. Lett. 85, 3025–3028 (2000) 106. Zhang, G.P., Hübner, W., Lefkidis, G., Bai, Y., George, T.F.: Paradigm of the time-resolved magneto-optical Kerr effect for femtosecond magnetism. Nat. Phys. 5, 499–502 (2009) 107. Carva, K., Battiato, M., Oppeneer, P.M.: Is the controversy over femtosecond magneto-optics really solved? Nat. Phys. 7, 665 (2011) 108. Atxitia, U., Chubykalo-Fesenko, O.: Ultrafast magnetization dynamics rates within the Landau-Lifshitz-Bloch model. Phys. Rev. B 84, 144414 (2011) 109. Kazantseva, N., Nowak, U., Chantrell, R.W., Hohlfeld, J., Rebei, A.: Slow recovery of the magnetisation after a sub-picosecond heat pulse. Europhys. Lett. 81, 27004 (2008) 110. Evans, R.F.L., Fan, W.J., Chureemart, P., Ostler, T.A., Ellis, M.O.A., and Chantrell, R.W.: Atomistic spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter. 26, 103202 (2014) 111. Steiauf, D., Fähnle, M.: Elliott-Yafet mechanism and the discussion of femtosecond magnetization dynamics. Phys. Rev. B 79, 140401 (2009) 112. Schellekens, A.J., Koopmans, B.: Comparing ultrafast demagnetization rates between competing models for finite temperature magnetism. Phys. Rev. Lett. 110, 217204 (2013) 113. Tveten, E.G., Brataas, A., Tserkovnyak, Y.: Electron-magnon scattering in magnetic heterostructures far out of equilibrium. Phys. Rev. B 92, 180412(R) (2015) 114. Mueller, B.Y., Baral, A., Vollmar, S., Cinchetti, M., Aeschlimann, M., Schneider, H.C., Rethfeld, B.: Feedback effect during ultrafast demagnetization dynamics in ferromagnets. Phys. Rev. Lett. 111, 167204 (2013) 115. Mueller, B.Y., Rethfeld, B.: Thermodynamic μT model of ultrafast magnetization dynamics. Phys. Rev. B 90, 144420 (2014) 116. Krieger, K., Dewhurst, J.K., Elliott, P., Sharma, S., Gross, E.K.U.: Laser-induced demagnetization at ultrashort time scales: predictions of TDDFT. J. Chem. Theory Comput. 11, 4870–4874 (2015) 117. Elliott, P., Müller, T., Dewhurst, J.K., Sharma, S., Gross, E.K.U.: Ultrafast laser induced local magnetization dynamics in Heusler compounds. Scient. Rep. 6, 38911 (2016) 118. Fognini, A., Salvatella, G., Gort, R., Michlmayr, T., Vaterlaus, A., Acremann, Y.: The influence of the excitation pulse length on ultrafast magnetization dynamics in nickel. Struct. Dyn. 2, 024501 (2015) 119. Carva, K., Battiato, M., Legut, D., Oppeneer, P.M.: Ab initio theory of electron-phonon mediated ultrafast spin relaxation of laser-excited hot electrons in transition-metal ferromagnets. Phys. Rev. B 87, 184425 (2013) 120. Kuiper, K.C., Roth, T., Schellekens, A.J., Schmitt, O., Koopmans, B., Cinchetti, M., Aeschlimann, M.: Spin-orbit enhanced demagnetization rate in Co/Pt-multilayers. Appl. Phys. Lett. 105, 202402 (2014) 121. Melnikov, A., Radu, I., Bovensiepen, U., Krupin, O., Starke, K., Matthias, E., Wolf, M.: Coherent optical phonons and parametrically coupled magnons induced by femtosecond laser excitation of the Gd(0001) surface. Phys. Rev. Lett. 91, 227403 (2003)

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    543

    122. Sultan, M., Atxitia, U., Melnikov, A., Chubykalo-Fesenko, O., Bovensiepen, U.: Electronand phonon-mediated ultrafast magnetization dynamics of Gd(0001). Phys. Rev. B 85, 184407 (2012) 123. Frietsch, B., Carley, R., Gleich, M., Teichmann, M., Bowlan, J., Weinelt, M.: Fluencedependent dynamics of the 5d6s exchange splitting in Gd metal after femtosecond laser excitation. Jpn. J. Appl. Phys. 55, 07MD02 (2016) 124. Radu, I., Woltersdorf, G., Kiessling, M., Melnikov, A., Bovensiepen, U., Thiele, J.-U., Back, C.H.: Laser-induced magnetization dynamics of lanthanide-doped permalloy thin films. Phys. Rev. Lett. 102, 117201 (2009) 125. Turgut, E., Zusin, D., Legut, D., Carva, K., Knut, R., Shaw, J.M., Chen, C., Tao, Z., Nembach, H.T., Silva, T.J., Mathias, S., Aeschlimann, M., Oppeneer, P.M., Kapteyn, H.C., Murnane, M.M., Grychtol, P.: Stoner versus Heisenberg: ultrafast exchange reduction and magnon generation during laser-induced demagnetization. Phys. Rev. B 94, 220408(R) (2016) 126. Eich, S., Plötzing, M., Rollinger, M., Emmerich, S., Adam, R., Chen, C., Kapteyn, H.C., Murnane, M.M., Plucinski, L., Steil, D., Stadtmüller, B., Cinchetti, M., Aeschlimann, M., Schneider, C.M., Mathias, S.: Band structure evolution during the ultrafast ferromagneticparamagnetic phase transition in cobalt. Sci. Adv. 3, e1602094 (2017) 127. Gort, R., Bühlmann, K., Däster, S., Salvatella, G., Hartmann, N., Zemp, Y., Holenstein, S., Stieger, C., Fognini, A., Michlmayr, T.U., Bähler, T., Vaterlaus, A., Acremann, Y.: Early stages of ultrafast spin dynamics in a 3d ferromagnet. Phys. Rev. Lett. 121, 087206 (2018) 128. Pershan, P.S., van der Ziel, J.P., Malmstrom, L.D.: Theoretical discussion of the inverse Faraday effect, Raman scattering, and related phenomena. Phys. Rev. 143, 574–583 (1966) 129. Kimel, A.V., Kirilyuk, A., Usachev, P.A., Pisarev, R.V., Balbashov, A.M., Rasing, T.: Ultrafast non-thermal control of magnetization by instantaneous photomagnetic pulses. Nature 435, 655–657 (2005) 130. Rebei, A., Hohfeld, J.: The magneto-optical barnett effect: Circularly polarized light induced femtosecond magnetization reversal. Phys. Lett. A 372, 1915–1918 (2008) 131. Hohlfeld, J., Stanciu, C.D., Rebei, A.: Athermal all-optical femtosecond magnetization reversal in GdFeCo. Appl. Phys. Lett. 94, 152504 (2009) 132. Vahaplar, K., Kalashnikova, A.M., Kimel, A.V., Hinzke, D., Nowak, U., Chantrell, R., Tsukamoto, A., Itoh, A., Kirilyuk, A., Rasing, T.: Ultrafast path for optical magnetization reversal via a strongly nonequilibrium state. Phys. Rev. Lett. 103, 117201 (2009) 133. Steil, D., Alebrand, S., Hassdenteufel, A., Cinchetti, M., Aeschlimann, M.: All-optical magnetization recording by tailoring optical excitation parameters. Phys. Rev. B 84, 224408 (2011) 134. Vahaplar, K., Kalashnikova, A.M., Kimel, A.V., Gerlach, S., Hinzke, D., Nowak, U., Chantrell, R., Tsukamoto, A., Itoh, A., Kirilyuk, A, Rasing, T.: All-optical magnetization reversal by circularly polarized laser pulses: Experiment and multiscale modeling. Phys. Rev. B 85, 104402 (2012) 135. Alebrand, S., Hassdenteufel, A., Steil, D., Cinchetti, M., Aeschlimann, M.: Interplay of heating and helicity in all-optical magnetization switching. Phys. Rev. B 85, 092401 (2012) 136. Mentink, J.H., Hellsvik, J., Afanasiev, D.V., Ivanov, B.A., Kirilyuk, A., Kimel, A.V., Eriksson, O., Katsnelson, M.I., Rasing, T.: Ultrafast spin dynamics in multisublattice magnets. Phys. Rev. Lett. 108, 057202 (2012) 137. Khorsand, A.R., Savoini, M., Kirilyuk, A., Kimel, A.V., Tsukamoto, A., Itoh, A., Rasing, T.: Role of magnetic circular dichroism in all-optical magnetic recording. Phys. Rev. Lett. 108, 127205 (2012) 138. Wienholdt, S., Hinzke, D., Carva, K., Oppeneer, P.M., Nowak, U.: Orbital-resolved spin model for thermal magnetization switching in rare-earth-based ferrimagnets. Phys. Rev. B 88, 020406 (2013) 139. Evans, R.F.L., Ostler, T.A., Chantrell, R.W., Radu, I., Rasing, T.: Ultrafast thermally induced magnetic switching in synthetic ferrimagnets. Appl. Phys. Lett. 104, 082410 (2014) 140. Schellekens, A.J., Koopmans, B.: Microscopic model for ultrafast magnetization dynamics of multisublattice magnets. Phys. Rev. B 87, 020407 (2013)

    544

    M. L. M. Lalieu and B. Koopmans

    141. Alebrand, S., Gottwald, M., Hehn, M., Steil, D., Cinchetti, M., Lacour, D., Fullerton, E.E., Aeschlimann, M., Mangin, S.: Light-induced magnetization reversal of high-anisotropy TbCo alloy films. Appl. Phys. Lett. 101, 162408 (2012) 142. Hassdenteufel, A., Schubert, C., Schmidt, J., Richter, P., Zahn, D.R.T., Salvan, G., Helm, M., Bratschitsch, R., Albrecht, M.: Dependence of all-optical magnetic switching on the sublattice magnetization orientation in Tb-Fe thin films. Appl. Phys. Lett. 105, 112403 (2014) 143. Cornelissen, T.D., Córdoba, R., Koopmans, B.: Microscopic model for all optical switching in ferromagnets. Appl. Phys. Lett. 108, 142405 (2016) 144. Berritta, M., Mondal, R., Carva, K., Oppeneer, P.M.: Ab initio theory of coherent laserinduced magnetization in metals. Phys. Rev. Lett. 117, 137203 (2016) 145. El Hadri, M.S., Pirro, P., Lambert, C.-H., Petit-Watelot, S., Quessab, Y., Hehn, M., Montaigne, F., Malinowski, G., Mangin, S.: Two types of all-optical magnetization switching mechanisms using femtosecond laser pulses. Phys. Rev. B 94, 064412 (2016) 146. Medapalli, R., Afanasiev, D., Kim, D.K., Quessab, Y., Manna, S., Montoya, S.A., Kirilyuk, A., Rasing, T., Kimel, A.V., Fullerton, E.E.: Multiscale dynamics of helicity-dependent alloptical magnetization reversal in ferromagnetic co/pt multilayers. Phys. Rev. B 96, 224421 (2017) 147. Quessab, Y., Medapalli, R., El Hadri, M.S., Hehn, M., Malinowski, G., Fullerton, E.E., Mangin, S.: Helicity-dependent all-optical domain wall motion in ferromagnetic thin films. Phys. Rev. B 97, 054419 (2018) 148. El Hadri, M.S., Hehn, M., Pirro, P., Lambert, C-H., Malinowski, G., Fullerton, E.E., Mangin, S.: Domain size criterion for the observation of all-optical helicity-dependent switching in magnetic thin films. Phys. Rev. B 94, 064419 (2016) 149. Hassdenteufel, A., Schmidt, J., Schubert, C., Hebler, B., Helm, M., Albrecht, M., Bratschitsch, R.: Low-remanence criterion for helicity-dependent all-optical magnetic switching in ferrimagnets. Phys. Rev. B 91, 104431 (2015) 150. Takahashi, Y.K., Medapalli, R., Kasai, S., Wang, J., Ishioka, K., Wee, S.H., Hellwig, O., Hono, K., Fullerton, E.E.: Accumulative magnetic switching of ultrahigh-density recording media by circularly polarized light. Phys. Rev. A 6, 054004 (2016) 151. Gorchon, J., Yang, Y., Bokor, J.: Model for multishot all-thermal all-optical switching in ferromagnets. Phys. Rev. B 94, 020409 (2016) 152. Ellis, M.O.A., Fullerton, E.E., Chantrell, R.W.: All-optical switching in granular ferromagnets caused by magnetic circular dichroism. Sci. Rep. 6, 30522 (2016) 153. John, R., Berritta, M., Hinzke, D., Müller, C., Santos, T., Ulrichs, H., Nieves, P., Walowski, J., Mondal, R., Chubykalo-Fesenko, O., McCord, J., Oppeneer, P.M., Nowak, U., Münzenberg, M.: Magnetisation switching of FePt nanoparticlerecording medium by femtosecond laser pulses. Scient. Rep. 7, 4114 (2017) 154. Wilson, R.B., Gorchon, J., Yang, Y., Lambert, C-H., Salahuddin, S., Bokor, J.: Ultrafast magnetic switching of gdfeco with electronic heat currents. Phys. Rev. B 95, 180409 (2017) 155. Xu, Y., Deb, M., Malinowski, G., Hehn, M., Zhao, W., Mangin, S.: Ultrafast magnetization manipulation using single femtosecond light and hot-electron pulses. Adv. Mater. 29, 1703474 (2017) 156. Yang, Y., Wilson, R.B., Gorchon, J., Lambert, C-H., Salahuddin, S., Bokor, J.: Ultrafast magnetization reversal by picosecond electrical pulses. Sci. Adv. 3, e1603117 (2017) 157. Gorchon, J., Lambert, C-H., Yang, Y., Pattabi, A., Wilson, R.B., Salahuddin, S., Bokor, J.: Single shot ultrafast all optical magnetization switching of ferromagnetic Co/Pt multilayers. Appl. Phys. Lett. 111, 042401 (2017) 158. Iihama, S., Xu, Y., Deb, M., Malinowski, G., Hehn, M., Gorchon, J., Fullerton, E.E., Mangin, S.: Single-shot multi-level all-optical magnetization switching mediated by spin transport. Adv. Mater. 30, 1804004 (2018) 159. Choi, G-M., Min, B-C.: Laser-driven spin generation in the conduction bands of ferrimagnetic metals. Phys. Rev. B 97, 014410 (2018) 160. Lalieu, M.L.M., Peeters, M.J.G., Haenen, S.R.R., Lavrijsen, R., Koopmans, B.: Deterministic all-optical switching of synthetic ferrimagnets using single femtosecond laser pulses. Phys. Rev. B 96, 220411(R) (2017)

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    545

    161. Lalieu, M.L.M., Lavrijsen, R., Koopmans, B.: Integrating all-optical switching with spintronics. Nat. Commun. 10, 110 (2019) 162. Gerlach, S., Oroszlany, L., Hinzke, D., Sievering, S., Wienholdt, S., Szunyogh, L., Nowak, U.: Modeling ultrafast all-optical switching in synthetic ferrimagnets. Phys. Rev. B 95, 224435 (2017) 163. Parkin, S., Yang, S-H.: Memory on the racetrack. Nat. Nanotechnol. 10, 195–198 (2015) 164. Yang, S-H., Ryu, K-S., Parkin, S.: Domain-wall velocities of up to 750 m s−1 driven by exchange-coupling torque in synthetic antiferromagnets. Nat. Nanotechnol. 10, 221–226 (2015) 165. Kim, K-J., Kim, S.K., Hirata, Y., Oh, S-H., Tono, T., Kim, D-H., Okuno, T., Ham, W.S., Kim, S., Go, G., Tserkovnyak, Y., Tsukamoto, A., Moriyama, T., Lee, K-J., Ono, T.: Fast domain wall motion in the vicinity of the angular momentum compensation temperature of ferrimagnets. Nat. Mater. 16, 1187–1192 (2017) 166. Caretta, L., Mann, M., Büttner, F., Ueda, K., Pfau, B., C.M. Günther, Hessing, P., Churikova, A., Klose, C., Schneider, M., Engel, D., Marcus, C., Bono, D., Bagschik, K., Eisebitt, S., Beach, G.S.D.: Fast current-driven domain walls and small skyrmions in a compensated ferrimagnet. Nat. Nanotechnol. 13, 1154–1160 (2018) 167. Kent, A.D., Worledge, D.C.: A new spin on magnetic memories. Nat. Nanotechnol. 10, 187– 191 (2015) 168. Melnikov, A., Razdolski, I., Wehling, T.O., Papaioannou, E.T., Roddatis, V., Fumagalli, P., Aktsipetrov, O., Lichtenstein, A.I., Bovensiepen, U.: Ultrafast transport of laser-excited spinpolarized carriers in Au/Fe/MgO(001). Phys. Rev. Lett. 107, 076601 (2011) 169. Battiato, M., Carva, K., Oppeneer, P.M.: Theory of laser-induced ultrafast superdiffusive spin transport in layered heterostructures. Phys. Rev. B 86, 024404 (2012) 170. Rudolf, D., La-O-Vorakiat, C., Battiato, M., Adam, R., Shaw, J.M., Turgut, E., Maldonado, P., Mathias, S., Grychtol, P., Nembach, H.T., Silva, T.J., Aeschlimann, M., Kapteyn, H.C., Murnane, M.M., Schneider, C.M., Oppeneer, P.M.: Ultrafast magnetization enhancement in metallic multilayers driven by superdiffusive spin current. Nat. Commun. 3, 1037 (2012) 171. Eschenlohr, A., Battiato, M., Maldonado, P., Pontius, N., Kachel, T., Holldack, K., Mitzner, R., Föhlisch, A., Oppeneer, P.M., Stamm, C.: Ultrafast spin transport as key to femtosecond demagnetization. Nat. Mater. 12, 332–336 (2013) 172. Khorsand, A.R., Savoini, M., Kirilyuk, A., Rasing, T.: Optical excitation of thin magnetic layers in multilayer structures. Nat. Mater. 13, 101–102 (2014) 173. Eschenlohr, A., Battiato, M., Maldonado, P., Pontius, N., Kachel, T., Holldack, K., Mitzner, R., Föhlisch, A., Oppeneer, P.M., Stamm, C.: Optical excitation of thin magnetic layers in multilayer structures Reply. Nat. Mater. 13, 102–103 (2014) 174. Eschenlohr, A., Persichetti, L., Kachel, T., Gabureac, M., Gambardella, P., Stamm, C.: Spin currents during ultrafast demagnetization of ferromagnetic bilayers. J. Phys. Condens. Matter. 29, 384002 (2017) 175. Choi, G-M., Cahill, D.G.: Kerr rotation in Cu, Ag, and Au driven by spin accumulation and spin-orbit coupling. Phys. Rev. B 90, 214432 (2014) 176. Choi, G-M., Moon, C-H., Min, B-C., Lee, K-J., Cahill, D.G.: Thermal spin-transfer torque driven by the spin-dependent Seebeck effect in metallic spin valves. Nat. Phys. 11, 576–582 (2015) 177. Alekhin, A., Razdolski, I., Ilin, N., Meyburg, J.P., Diesing, D., Roddatis, V., Rungger, I., Stamenova, M., Sanvito, S., Bovensiepen, U., Melnikov, A.: Femtosecond spin current pulses generated by the nonthermal spin-dependent Seebeck effect and interacting with ferromagnets in spin valves. Phys. Rev. Lett. 119, 017202 (2017) 178. Razdolski, I., Alekhin, A., Ilin, N., Meyburg, J.P., Roddatis, V., Diesing, D., Bovensiepen, U., Melnikov, A.: Nanoscale interface confinement of ultrafast spin transfer torque driving non-uniform spin dynamics. Nat. Commun. 8, 15007 (2017) 179. Lalieu, M.L.M., Helgers, P.L.J., Koopmans, B.: Absorption and generation of femtosecond laser-pulse excited spin currents in noncollinear magnetic bilayers. Phys. Rev. B 96, 014417 (2017) 180. Stiles, M.D., Zangwill, A.: Anatomy of spin-transfer torque. Phys. Rev. B 66, 014407 (2002)

    546

    M. L. M. Lalieu and B. Koopmans

    181. Baláž, P., Žonda, M., Carva, K., Maldonado, P., Oppeneer, P.M.: Transport theory for femtosecond laser-induced spin-transfer torques. J. Phys. Condens. Matter 30, 115801 (2018) 182. Alekhin, A., Bürstel, D., Melnikov, A., Diesing, D., Bovensiepen, U.: Ultrafast laser-excited spin transport in Au/Fe/MgO(001), relevance of the Fe layer thickness. In: Bigot, J.Y., Hubner, W., Rasing, T., Chantrell, R. (eds.) Ultrafast Magnetism I, vol. 159, pp. 241–243. Springer Proceedings in Physics (2014) 183. Ulrichs, H., Razdolski, I.: Micromagnetic view on ultrafast magnon generation by femtosecond spin current pulses. Phys. Rev. B 98, 054429 (2018) 184. Hellman, F., Hoffmann, A., Tserkovnyak, Y., Beach, G.S.D., Fullerton, E.E., Leighton, C., MacDonald, A.H., Ralph, D.C., Arena, D.A., Dürr, H.A., Fischer, P., Grollier, J., Heremans, J.P., Jungwirth, T., Kimel, A.V., Koopmans, B., Krivorotov, I.N., May, S.J., Petford-Long, A.K., Rondinelli, J.M., Samarth, N., Schuller, I.K., Slavin, A.N., Stiles, M.D., Tchernyshyov, O., Thiaville, A., Zink, B.L.: Interface-induced phenomena in magnetism. Rev. Mod. Phys. 89, 025006 (2017) 185. Kirilyuk, A., Kimel, A.V., Rasing, T.: Ultrafast optical manipulation of magnetic order. Rev. Mod. Phys. 82, 2731 (2010) 186. Koopmans, B.: Time-resolved Kerr-effect and spin dynamics in itinerant ferromagnets. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials: Volume 3 Novel Techniques for Characterizing and Preparing Samples, pp. 1589–1613. Wiley (2007)

    Further Reading 1. 2. 3. 4. 5. 6. 7.

    Qiu, Z.Q., Bader, S.D.: Surface magneto-optic kerr effect. [37] Soldatov, I.V., Schäfer, R.: Advances in quantitative Kerr microscopy. Phys. Rev. B 95, 014426 (2017) Oppeneer, P.M.: Magneto-optical Kerr spectra. [17] Hellman, F., et al.: Chapter IV in “Interface-induced phenomena in magnetism”. [184] Kirilyuk, A., et al.: Ultrafast optical manipulation of magnetic order. [185] Koopmans, B.: Laser-induced magnetization dynamics. [26] Koopmans, B.: Time-resolved Kerr-effect and Spin Dynamics in Itinerant Ferromagnets. [186]

    Mark L. M. Lalieu received his PhD from the Eindhoven University of Technology in 2019. Within the Physics of Nanostructures group, he worked in the field of laser-induced ultrafast magnetization dynamics and its integration with the fields of spintronics and magnonics.

    10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films

    547

    Bert Koopmans received his PhD from the University of Groningen. After a short stay at the Radboud University Nijmegen, and three years as a Humboldt Fellow at the Max-Planck Institute in Stuttgart, he joined the Eindhoven University of Technology, where since 2003 he chairs the group Physics of Nanostructures. His research interests encompass spintronics, nanomagnetism and ultrafast magnetization dynamics.

    Magnetostriction and Magnetoelasticity

    11

    Dirk Sander

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Classification of Magnetoelastic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Anisotropic Magnetostriction (Joule Magnetostriction) . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetovolume Effects, Spontaneous Magnetostriction, Forced Magnetostriction, and Invar Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Villari Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . E Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetomechanical Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wiedemann Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Matteucci Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Field-Induced Strain Phenomena, Which Differ from Joule Magnetostriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetoelasticity and Joule Magnetostriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Derivation of the Magnetostrictive Strain Tensor: Cubic Case . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction of Polycrystalline Cubic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Derivation of the Magnetostrictive Strain Tensor: Hexagonal Case . . . . . . . . . . . . . . . . . . . Magnetostriction of Polycrystalline Hexagonal Materials . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction and Stress-Induced Magnetic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetoelastic Effects in Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Experimental Determination of Magnetostriction and Magnetoelastic Coupling . . . . . . . . . . Magnetoelastic Coupling in Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction and Magnetoelasticity: Physical Origin and Insights from Theory . . . . . . . Compilation of Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetoelastic and Elasticity Data for Bulk Transition Metals . . . . . . . . . . . . . . . . . . . . . . Theoretical and Experimental Values of Magnetoelastic Coupling Coefficients and Their Strain Dependence (Table 4) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction Data of Amorphous Fe Alloys (Table 5) . . . . . . . . . . . . . . . . . . . . . . . . . .

    550 551 551 552 553 553 555 556 557 557 560 560 564 565 566 567 569 570 571 574 577 577 578 579

    D. Sander () Max Planck Institute of Microstructure Physics, Halle, 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_11

    549

    550

    D. Sander

    Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si, Fe-Ga-Al, and Fe-Ga-Ge Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Zero Magnetostriction Alloys and Soft Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction Data for Paramagnetic Metals and Alloys . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction Data for Bi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction Data for Tb, Dy, and Ho . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction Data of TbFe2 (Terfenol) and Tb27 Dy73 Fe2 (Terfenol-D) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetostriction Data of Oxide Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Notations for Lattice Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Relation Between λ and B for the Hexagonal System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    579 580 582 582 582 584 585 585 586 586

    Abstract

    The physical concepts magnetostriction and magnetoelasticity are presented. Spontaneous volume magnetostriction and saturation linear magnetostriction are distinguished. Various magnetoelastic phenomena are introduced, but the emphasis is on magnetostriction in bulk samples and thin films. The equations for magnetostrictive and magnetoelastic coefficients are derived for cubic, hexagonal, and isotropic systems. Experiments on the measurement of the linear magnetostriction λi and magnetoelastic coupling coefficients Bj are discussed. Ab initio-based theory elucidates the physical origin of magnetostrictive effects in metals at the electronic level, although accurate calculations are often elusive. The magnetoelastic properties of nm thin films may deviate in magnitude and sign from the bulk values. Both experiment and theory identify substrate-induced lattice strain as a driving force for this deviation. Data on magnetostriction and magnetoelasticity are compiled, including those of highly magnetostrictive systems, such as (Tb,Dy)Fe2 (Terfenol) and (Fe,Ga) (Galfenol).

    Introduction Magnetostriction is the relative change of length of a sample when it is magnetized [1,2,3,4,5,6,7,8,9,10,11,12,13]. It is customary to distinguish between spontaneous volume magnetostriction, which is a change of volume in the magnetically ordered state that is almost independent of applied magnetic field, and anisotropic linear magnetostriction, also known as Joule magnetostriction, which saturates with the magnetization. The latter phenomenon was discovered by James Prescott Joule in 1841 [14] and described in 1847 [15, 16]. Joule used an ingenious mechanical lever mechanism to magnify 3000 times the tiny elongation upon magnetization of a 2-foot-long Fe bar to make it measurable. He found a relative change of length upon magnetization of order 1.4 × 10−6 . He also found that the net sample volume remained unchanged by immersing the bar in water in a sealed container and looking for a displacement of water in a capillary tube when the bar was magnetized. These historic observations are consistent with today’s understanding of the phenomenon now known as Joule magnetostriction.

    11 Magnetostriction and Magnetoelasticity

    551

    Magnetostriction is not necessarily tiny. Although a typical order of magnitude is 10−5 for bulk 3d metals [17], values as large as 10−3 have been reported for highly magnetostrictive alloys, such as FeTb (Terfenol) and GaFe (Galfenol), and for some rare earth elements [18, 19, 20, 21, 22]. Anisotropic magnetostriction, or Joule magnetostriction, is the main focus of this chapter. There are numerous other effects related to magnetoelasticity. These effects describe the coupling between magnetization and elasticity, thermal expansion, sample volume, strain, and torsion. They relate to the observed deformation of a sample upon magnetization. The inverse effect, the change of the magnetic configuration by imposing stress or torque on a sample, is also exploited to study magnetostriction. These effects are described in section “Classification of Magnetoelastic Effects”. The derivation of the respective magnetostrictive coefficients from a measurement of change of length upon magnetization is outlined in section “Magnetoelasticity and Joule Magnetostriction”. It is shown that the underlying coupling between lattice strain and magnetization direction is well described in the framework of magnetoelasticity. The mathematical description of magnetoelasticity in view of the tensor properties of crystal elasticity is elucidated, and corresponding expressions for magnetostrictive coefficients and magnetoelastic coupling coefficients are presented. Sections “Magnetostriction and Stress-Induced Magnetic Anisotropy” and “Magnetoelastic Effects in Films” discuss stress-induced magnetic anisotropy and magnetoelastic effects in thin films. Section “Experimental Determination of Magnetostriction and Magnetoelastic Coupling” describes experiments to measure magnetostriction and magnetoelastic coupling in bulk and in films. Insights from theory on the physical origin of magnetoelasticity on the electronic level of metals and 3d oxides are presented in section “Magnetostriction and Magnetoelasticity: Physical Origin and Insights from Theory”. A compilation of magnetostrictive and magnetoelastic coupling data concludes this chapter in section “Compilation of Data”.

    Classification of Magnetoelastic Effects Magnetostriction describes the strain of a sample upon magnetization. A further, closely related manifestation of magnetoelasticity is the inverse effect: the impact of an externally applied strain, e.g., via stress or torsion, on the magnetism of the sample. This gives rise to numerous magnetoelastic effects, which are categorized in this section.

    Anisotropic Magnetostriction (Joule Magnetostriction) Experiments show that an unmagnetized ferromagnet strains when it is magnetized to saturation. It is observed that the strain along the magnetization direction differs in sign and magnitude from the strain measured perpendicular to the magnetization

    552

    D. Sander

    λ >0 M = Ms

    H

    λ TC

    T < TC ω>0 Mavg = 0 H=0

    Fig. 1 Schematic view of magnetostriction in a ferromagnet. Below the Curie temperature TC , a spontaneous lattice strain, the spontaneous volume magnetostriction ω, is observed, even in the absence of a magnetic field. The application of a field H induces an anisotropic lattice strain. Positive (negative) magnetostriction, λ > 0 ( λ < 0 ), gives an elongation (compression) along the field direction and an opposite strain perpendicular to it. The sketch highly exaggerates the volume and linear strains of order 10−6 to 10−2

    direction so that the net volume is conserved. This situation is schematically illustrated in Fig. 1. Formally, this magnetization-induced lattice strain is ascribed to magnetoelastic contributions to the energy density. Qualitatively, these contributions are expressed by terms such as fme = B. Here, B is the magnetoelastic coupling coefficient (units J/m3 ), and the dimensionless lattice strain is . The derivative of the magnetoelastic energy density with respect to lattice strain gives a stress ∂fme /∂ = τme = B, which strains the lattice to the point where elastic stresses compensate the magnetization-induced stress [2, 17]. In general, a three-dimensional magnetostrictive strain state results. This qualitative description already points at the importance of the interplay between magnetoelasticity and crystal elasticity for the magnitude of magnetostriction. The link between the two becomes obvious in the quantitative derivation of magnetostrictive strains below in section “Magnetoelasticity and Joule Magnetostriction”.

    Magnetovolume Effects, Spontaneous Magnetostriction, Forced Magnetostriction, and Invar Effects The spontaneous volume magnetostriction in Fig. 1 is defined as the difference in atomic volume between the magnetically ordered state and the paramagnetic state. Thus, the temperature-driven transition at the Curie temperature brings with it a change of lattice volume, which deviates from the volume extrapolated using the thermal expansion α(T ) observed above the Curie temperature. Consequently, the

    11 Magnetostriction and Magnetoelasticity

    553

    temperature dependence α(T ) shows distinct variations near the magnetic ordering temperature [4, 23]. This is also true for antiferromagnets and other types of magnetic order. The spontaneous volume magnetostriction of a metal can be slightly altered by applying a large external magnetic field, and this gives rise to forced volume magnetostriction [24]. The spontaneous volume magnetostriction, if positive, can produce a lattice expansion to compensate or cancel the usual contraction of lattice spacing with deceasing temperature, leading to vanishingly small or even negative thermal expansion – lattice expansion upon cooling. The former effect is found in invar alloys such as Fe64 Ni36 [25] and the latter in Mn3 (Cu(1−x) Gex )N [26] alloys. Invar generally refers to a material with a vanishingly small thermal expansion near room temperature due to its large positive spontaneous volume magnetostriction [27, 28]. Elinvar refers to a material with a vanishingly small temperature variation of its elastic properties with temperature. For further discussion of magnetovolume and invar effects, see [23, 25, 29, 28, 30].

    Villari Effect The Villari effect describes a change of magnetic susceptibility due to straining the sample [31] by an external stress. It is easier to magnetize a material with positive magnetostriction in the direction of a tensile stress. The stress-induced lattice strain induces a change of the magnetic configuration such that those domains grow in size which have their local magnetization oriented parallel with the lattice strain. This change of domain configuration affects the susceptibility and the magnetization curve below saturation. The impact of stress on the magnetic anisotropy of a sample is discussed in section “Magnetostriction and Stress-Induced Magnetic Anisotropy”. The stress-induced change of domain configuration is schematically illustrated in Fig. 8.

    E Effect A material which is not magnetically saturated is composed of magnetic domains. The magnetic domains are characterized by a saturated magnetization with a constant magnetization direction within each domain. This magnetization direction changes from domain to domain. Different domains in a cubic ferromagnet with [100] anisotropy are separated by 90◦ and 180◦ domain walls. This description is a valid approximation in many cases, but it should not disguise the fact that more complex situations such as vortex structures sometimes exist [9, 33]. The application of a tensile stress to a material with positive magnetostriction rotates the magnetization parallel to the stress direction, and corresponding domains grow in size. The magnetostrictive strain λ then adds to the elastic strain, and the material appears softer. The application of a tensile stress to a material with negative magnetostriction aligns the magnetization perpendicular to the stress direction. A

    554

    D. Sander

    positive magnetostrictive strain λ⊥ still appears along the stress direction, and the material appears again softer. Thus, irrespective of the sign of magnetostriction, a magnetic material with domains that are non-collinear with the applied stress in zero field always appears elastically softer during a magnetization process as compared to a sample which is magnetically saturated. This variation of elasticity is termed E effect. Figure 2 presents an example of the E effect for Ni [32]. The Young modulus E is affected by the magnetization state of Ni below the Curie temperature of TC = 360 ◦ C. The demagnetized sample shows a considerably reduced Young modulus of 178 GPa below TC , as compared to the magnetically saturated sample with 211 GPa. Here, a relative reduction of the Young modulus of (Esat − Emagn=0 /Emagn=0 = 0.185) is observed. The magnitude of the E effect depends on numerous factors. The crystallographic orientation of the sample, its magnetic domain structure, magnitude and orientation of the applied strain (stress), and the direction and magnitude of an applied magnetic field need to be considered for a quantitative description of the effect [32, 4, 10]. The resulting complexity is alleviated by considering approximations such as a magnetic domain orientation orthogonal to the stress direction, isotropic elastic and magnetic properties, and minute strains in the realm of acoustic vibrations. Under these assumptions, the reduction of the Young modulus due to magnetostrictive contributions to strain is approximated by [10] 9λ2S H 2 Eel − Eλ = Eel . Eλ μMS HS3

    (1)

    Here, λS , H , HS , and MS are the saturation magnetostriction, the applied magnetic field, the magnetic anisotropy field, and the saturation magnetization, respectively. The √ E effect manifests as a change of the resonance frequency, which scales as E, and can be tuned by an applied magnetic field. The effect is strong for

    220 M=MS

    210 E (GPa)

    Fig. 2 E effect. Young modulus E of Ni for a magnetized sample (M = MS , upper curve (blue)) and for a demagnetized sample (M = 0, lower curve (red)) [32]. The Young modulus of the demagnetized sample is smaller than that of the fully magnetized sample below the Curie temperature TC of Ni of 360 ◦ C

    TC

    200 190 M=0

    180 100

    200 300 temperature (oC)

    400

    11 Magnetostriction and Magnetoelasticity

    0.8 transversal

    0.6 (Eel-Eλ)/Eλ

    Fig. 3 Data of the E effect obtained from the flexural vibration of a reed [34]. The reed has been annealed in fields transversal and longitudinal to its length. The applied field is oriented along the length of the reed. The strain induced by the flexural vibration leads to a stronger reduction of elasticity for transversal domain orientation

    555

    0.4

    0.2

    longitudinal

    0 0

    1

    2

    applied field (mT)

    amorphous ferromagnetic metals with vanishing magnetic anisotropy and sizeable magnetostriction. Here, a relative reduction of E by 45% has been reported [34], as indicated in Fig. 3. Pronounced magnetic anisotropy opposes the effect, as the stress-induced reconfiguration of magnetic domains is hampered. Figure 3 presents an example for the relative variation (Eel − Eλ )/Eλ as a function of an applied field along the length of a reed. The reduction of the Young modulus is more pronounced for a sample with magnetic domain orientation along the transversal direction (along the width of the reed) in zero field as compared to a sample with longitudinal magnetic domain orientation (along the length of the reed) in zero field. The transversal domains are rotated in the magnetizing field toward the longitudinal direction, and this makes their further alignment susceptible to the vibration-induced lattice strain. The effect vanishes with increasing field, where Eλ approaches Eel . The E effect affects elasticity only in the regime of minute strains of the order of magnetostriction. Thus, typically lattice vibrations and small amplitude flexural vibrations are influenced, but not elasticity in the range of larger strains of order 1%.

    Magnetomechanical Damping There is a loss of energy associated with any elastic vibration of a sample. Experiments show that this loss is larger for ferromagnetic than for non-ferromagnetic samples [4, 6]. The reason is that vibrations lead to flexural strains which interact with the magnetic configuration via magnetoelastic coupling. This link to magnetization opens a further loss channel, and a stronger damping results. This is advantageous for applications where a strong intrinsic damping is required to avoid catastrophic vibrational amplitudes near resonance, such as in turbines, but it is

    556

    D. Sander

    disadvantageous for applications in ultrasonic transducers, where it waists energy [36]. The magnetomechanical loss is larger for samples with large magnetoelastic coupling, and it increases with increasing vibrational amplitude. The impact of the vibration-induced strain on the magnetic domain configuration is related to the Villari effect. As a result, the local magnetization of an unsaturated sample changes during vibration, which gives rise to local eddy current and magnetomechanical losses. The former increases with increasing frequency of vibration, whereas the latter is independent of frequency, and it is the prevailing loss channel at low frequencies in the range of Hz. The flexural vibration-induced change of magnetization is an hysteric process, which involves irreversible displacements of domain walls. This leads to losses in the vibrational energy, and the sample vibration is damped. The magnitude of damping depends in a subtle manner on the magnetization state of the sample. A sample with zero net magnetization in zero field shows a larger damping as compared to a magnetically saturated sample. This situation is illustrated in Fig. 4. The curves show the amplitude of a torsional vibration of a 1 mm Fe wire in an axial magnetic field (top curve, blue) and without an axial magnetic field (bottom curve, red) [35]. The torsion-induced magnetization change is suppressed by the applied field, and a lower damping results as compared to the field-free case.

    Wiedemann Effect

    Fig. 4 Decay of the torsion amplitude of a 1-mm-diameter Fe wire with (top, blue curve) and without (bottom, red curve) an applied magnetic field along the wire axis [35]. The magnetic field suppresses magnetomechanical damping by reducing torsion-induced magnetization changes of the wire

    torsion amplitude (arb. units)

    The Wiedemann effect [37] is the torsional twist observed in a sample upon the application of an helical magnetic field. The helical field results from the superposition of an axial magnetic field produced by a surrounding coil and a circular field produced by a current passing through a wire-shaped or thin-walled tube sample. The situation is schematically illustrated in Fig. 5, and the resulting sample torsion γ is indicated. The effect is the basis of a most sensitive measurement of magnetostriction, as the method can be applied in resonance with a torsional

    axial field 10 mT 0

    no axial field 0

    0

    4

    8

    12 16 time (s)

    20

    24

    11 Magnetostriction and Magnetoelasticity

    Iwire

    557

    γ

    Icoil Fig. 5 Schematic of the Wiedemann effect. A coil surrounding the sample gives an axial magnetic field along the sample axis. The current through the wire (sample) produces a radial field, which in conjunction with the axial field results in an helical magnetization of the sample. Magnetostriction in the sample causes a sample twist, which induces a torsion γ , as depicted in the schematic of the cross section

    vibration, whereby the resulting sample twist is largely amplified. A sensitivity of 10−13 for the measurement of magnetostriction has been reported [38].

    Matteucci Effect The converse to the Wiedemann effect is the Matteucci effect, the appearance of a helical magnetization component at the end of a twisted sample upon changing the axial sample magnetization [39]. The effect has been exploited in micro-fluxgate sensors, where a helical magnetization is imposed on an amorphous Fe77.5 Ni7.5 B15 wire of 0.125 mm diameter. A field sensitivity of order 100 nT has been reported [40].

    Magnetic Field-Induced Strain Phenomena, Which Differ from Joule Magnetostriction Joule magnetostriction results from the strain dependence of the magnetic anisotropy. It is observed as a field-driven change of the linear sample dimensions. However, there are other phenomena which describe a change of linear dimensions of a sample for an applied magnetic field [41]. Also there, the term magnetostriction is used, irrespective of the different physical origin. Here we mention three such phenomena. We treat shape-memory alloys, where a magnetic field drives a large strain in the martensitic phase of a material with two twin variants. Huge strains of order several percent are observed. We also present results on field-driven spin-state transition in Co3+ and on the interactions between a magnetic field and pinned Abrikosov flux lattices in type-II superconductors. These effects give strains of the order of 2 × 10−3 and 2 × 10−4 , respectively.

    Ferromagnetic Shape-Memory Alloys (FSMA) The ferromagnetic shape-memory effect is observed for materials, which undergo a diffusionless martensitic phase transition between a high-temperature cubic phase

    558

    D. Sander

    and a low-temperature phase of lower symmetry. The dominant lattice distortion is a shear strain. Sample strains in these phase transitions can be of order several percent. Such a large strain in the martensitic phase is accommodated by the formation of twin variants [42]. This means that structural domains with different c-axis orientations, the twin variants, are characteristic of these materials. The easy magnetization direction is along the c-axis. Field-induced strain may be triggered by a field-driven phase transition or by a field-driven strain developing exclusively in the martensitic phase, where the materials, in their martensitic phase, show a pronounced dependence of lattice strain on magnetic field, and field-driven strains of order several percent are possible. The huge magnetic field-induced strain arises as the Zeeman energy favors the growth of one variant at the expense of the other. This strain occurs fully within in the martensitic phase. This field-induced strain is schematically depicted in Fig. 6. The magnetic field strains in the martensitic phase are due to field-induced twinboundary motion [42, 43, 10, 11]. As a result of the formation of crystallographic twins in the martensitic phase, the magnetization direction varies from one twin variant to the other. Provided that the magnetic anisotropy of the system is sufficiently strong, an applied magnetic field will not simply rotate the magnetization. Instead, the growth of the twin variant with favorable magnetization direction with respect to the applied field is favored, and sizable strain results. The driving force for the elimination of one twin variant and the growth of another is the minimization of the Zeeman energy in the applied magnetic field.

    cubic cooling T < Tm

    H=0

    T > Tm

    applying a field H

    heating T > Tm

    H εH

    Fig. 6 Illustration of the ferromagnetic shape-memory effect. The parent material is cubic above the martensitic temperature Tm . A martensitic phase transition occurs upon cooling below Tm . A lattice strain of order 1% is observed, which is accompanied by the creation of a twin structure with two variants. The application of a magnetic field H drives the growth of one variant, and a field-induced strain H is observed. Heating above Tm restores the material to its cubic phase. The white arrows indicate the magnetization direction within the variants

    11 Magnetostriction and Magnetoelasticity

    559

    The effect is observed in the Heusler alloy Ni2 MnGa. This material has a martensitic transition between a high-temperature cubic phase and low-temperature tetragonal phase. The transition temperature is near Tmart = 276 K. The dominant mechanism of this first-order phase transition is the occurrence of shear strains in the low-temperature phase. The tetragonal phase is characterized by a huge lattice strain along the c-axis of order −7%, whereas the lateral strain along a and b is only of order +1.3% with respect to the parent cubic phase. It is possible to move the phase transition temperature Tmart to higher values by modifying the electron-per-atom ratio by alloying [11]. Strains of up to 10% are achieved for the transition into a orthorhombic phase [44]. It has been found that the temperature dependence of the Ni 3d states plays a crucial role in the electronic origin of the ferromagnetic shape-memory effect [45]. This supports the view that the martensitic instability can be described as a JahnTeller effect [45]. The ferromagnetic shape-memory effect differs from Joule magnetostriction due to its different physical origin; no strain dependence of the magnetic anisotropy is necessary. Also, the effect shows a threshold field, below which no strain is observed. In addition, the magnetic field-induced strain shows a large hysteresis, whereas magnetostriction in hard axis magnetization is an anhysteretic process. The field-induced strain in FSMAs is usually observed only in a narrow temperature range below Tmart and above a lower cutoff temperature [11].

    Magnetic Field-Driven Spin-State Transition in La(1−x) Srx CoO3−δ (x ≥ 0.3) Large magnetic field-induced strain of order 1 × 10−3 has been observed Co3+ compounds such as LaSrCoO3 [46] and LaCoO3 [47], at low temperature (25 K) and in high fields (14 T). This field-driven strain increases monotonically with field, without saturating. It reaches 2.2×10−3 at 14 T. It has been ascribed to a field-driven spin transition between a low-spin state and a Jahn-Teller distorted intermediatespin state [46]. This effect does not rely on the ferromagnetic state of the sample, and this contrasts with Joule magnetostriction. Magnetostriction in Superconductors The application of a magnetic field along the c-axis of the high Tc superconductor Bi2 Sr2 CaCu2 O8 in its superconducting state at 5 K leads to a transverse magnetostriction in the a, b plane of −2 × 10−4 at 6 T [48]. The negative sample strain in the plane perpendicular to the field direction is ascribed to a compressive stress, exerted by pinned Abrikosov vortices which are pushed together by an increasing field. After this overview of effects due to magnetoelastic coupling and magnetic fieldinduced strain, we now focus on Joule magnetostriction. Relations useful for the description and measurement of magnetostriction and magnetoelastic coupling in cubic and hexagonal bulk samples and films are derived in the following sections.

    560

    D. Sander

    Magnetoelasticity and Joule Magnetostriction We start from the expressions for the elastic fel and the magnetoelastic energy density fme for a sample with a cubic lattice. These expressions can be found in many textbooks [7, 8, 10, 12] and reviews [2, 17]. We use the tensor notation for the lattice strain ij [49]. Confusion is possible because there are different conventions for representing strain. The contracted Voigt notation and engineering strains are alternative descriptions, but they give different expressions for the off-diagonal strain. The different conventions are discussed in the Appendix.

    Derivation of the Magnetostrictive Strain Tensor: Cubic Case The magnetoelastic strain tensor me gives the three-dimensional strain state of a sample upon saturation magnetization along the direction (α1 , α2 , α3 ), where the αi are the direction cosines of the magnetization with respect to the cubic axes. The magnetization-induced strain always refers to a reference state of zero net magnetization, as given by a demagnetized sample. This demagnetized reference state is formally implemented by assuming a random orientation of the magnetization. By convention, this reference state is characterized by zero magnetostrictive strain. This convention gives rise to terms of −1/3 in the expression of the αi2 , reflecting the average of cos2 (ϑ) over the unit sphere (see Eq. (17)). Note that such a demagnetized reference state may be hard to realize experimentally. Irrespective of this, the resulting expressions can be used also for calculating the change of strain between two well-defined magnetization directions as outlined below. The role of the reference state for magnetostriction measurements is illustrated in Fig. 7. The elastic energy density 1 2 2 2 2 2 2 +22 +33 )+c12 (11 22 +11 33 +22 33 )+2c44 (23 +13 +12 )... c11 (11 2 (2) and the magneto elastic energy density

    fel =

    1 1 1 fme = B1 (11 (α12 − ) + 22 (α22 − ) + 33 (α32 − ) 3 3 3 +2B2 (23 α2 α3 + 13 α1 α3 + 12 α1 α3 ) . . .

    (3)

    are given in second and first order in strain , respectively [17]. The coefficients of the rank two strain tensor are ij . The elastic stiffness constants are given by cij (units N/m2 ), and the magnetoelastic coupling coefficients are given by Bi (units J/m3 ). Note that this equation is given in a short-hand notation for the cij which disguises that the elastic stiffness is described by a rank four tensor. Appropriate tensor transformations need to be applied for the study of directional dependencies [17]. The dots indicate that higher-order terms are omitted, but may be important

    11 Magnetostriction and Magnetoelasticity

    561

    εII ε

    (a)

    εII

    ε

    T

    M=MS H=HS

    εII

    (b)

    1

    3λ 2 S

    0

    Δε

    ε

    T

    Mavg=0 H=0

    0 2

    HS

    H II εII

    ε

    T

    M=MS H=HS

    Mavg=0 H=0

    Fig. 7 (a) Starting from a demagnetized state with a random magnetization configuration, saturation along the field direction gives the lattice strain  along that direction and ⊥ perpendicular to it. (b) Note that a demagnetized state with a collinear domain orientation along the field direction does not show any magnetostrictive strain change upon saturation. To avoid the impact of the often undefined reference state at zero field, a strain measurement during the reorientation of the magnetization direction between two saturated states 1 (vertical) and 2 (horizontal) gives the change of strain as  = 32 λS

    [17, 50]. This will be discussed below for fme , where the Bi are found to be strain dependent. We also note that an alternative description exists within the symmetry-invariant notation [10]. Here, the free energy expression is expanded in a set of orthogonal harmonic functions. For a thorough discussion, see [51, 52, 8]. The components of the magnetoelastic strain tensor are obtained by calculating the strain derivatives of the total energy density ∂f/∂ij with f = fel + fme , setting the expressions to zero, and solving the resulting six equations for ij . Some simplification is possible by exploiting α12 + α22 + α32 = 1, and the following magnetoelastic strain tensor results: ⎛ me

      B1 α12 − 13

    ⎜ − c11 −c12 ⎜ ⎜ = ⎜ − B2 α1 α2 ⎜ 2c44 ⎝ B2 α1 α3 − 2c44

    ⎞ 1 α2 − B22cα44   B1 α22 − 13



    c11 −c12

    2 α3 − B22cα44

    1 α3 − B22cα44

    ⎟ ⎟ ⎟ B2 α2 α3 ⎟ . − 2c44 ⎟   ⎠ B1 α32 − 13 − c11 −c12

    (4)

    562

    D. Sander

    The magnetostrictive strain δ , which is observed in the direction β = (β1 , β2 , β3 ), where the βi are direction cosines of the strain measurement direction with respect to the cubic axes, is calculated from δ

    = βi βj meij 3

    3

    i=1 j =1

    =− −

    B1 c11 − c12

    (5)

    1 1 1 α12 − β12 + α22 − β22 + α32 − β32 3 3 3

    B2 (α1 α2 β1 β2 + α1 α3 β1 β3 + α2 α3 β2 β3 ) . c44

    (6)

    This equation gives zero strain for a state with random magnetization direction. We calculate the magnetostrictive strains δ for different magnetization and strain measurement directions. This defines recipes for extracting the tabulated magnetostrictive coefficients λ100 and λ111 . We refrain from calling the λ constants, as their value may be changed due to lattice strain, or in the monolayer thickness range [53,54,55,56,17,57,58,50]. The first remark refers to a potential strain dependence of the magnetoelastic coupling coefficients Bi and the second to the contributions of interface effects to Bi in atomically thin samples. Magnetization Along [100]; Strain Measurement Along [100] This defines a procedure to measure λ100 , starting from the demagnetized state with random domain orientation. With α = (1, 0, 0) and β = (1, 0, 0), we obtain δ 2B1 = λ100 = − . 3(c11 − c12 )

    (7)

    Magnetization Along [111]; Strain Measurement Along [111] This defines a procedure to measure λ111 , starting state with √ √ from√the demagnetized random domain orientation. With α = β = (1/ 3, 1/ 3, 1/ 3), we obtain δ B2 = λ111 = − . 3c44

    (8)

    We see that a negative magnetoelastic coupling coefficient Bi induces a positive lattice strain. This is plausible by realizing that the energy density f of the system is lowered by the magnetostrictive strain. Thus, a reduction of f calls for opposite signs for B and , as found above.

    11 Magnetostriction and Magnetoelasticity

    563

    Finally, we obtain the magnetostrictive strain tensor λ by replacing the expressions of Bi in Eq. (4) with the definitions of λ in Eqs. (7) and (8). ⎛ ⎜ ⎜ λ = ⎜ ⎝

    

    3 1 2 2 λ100 α1 − 3 3 2 λ111 α1 α2 3 2 λ111 α1 α3

    



    3 3 α1 α2 λ111 α1 α3 2 λ111 ⎟   2 ⎟ 1 3 3 2 α λ − λ α α ⎟ 2 3 2 2 100 3 2 111  ⎠ 1 3 3 2 2 λ111 α2 α3 2 λ100 α3 − 3

    We can now express the magnetostrictive strain coefficients λ.

    δ

    (9)

    in terms of the magnetostriction

    3 3

    δ = βi βj λij

    (10)

    i=1 j =1

    =

      1 3 λ100 α12 β12 + α22 β22 + α32 β32 − 2 3

    +3λ111 (α1 β1 α2 β2 + α1 β1 α3 β3 + α2 β2 α3 β3 )

    (11)

    We apply this expression first to derive the difference between longitudinal and transversal magnetostrictive strain, namely, strain along the magnetization direction 100 −  100 and perpendicular to it. To this end, we calculate the difference of strain 100 010 for a fixed magnetization direction along [100] and strain measurements along [100] and along the perpendicular direction [010]. Magnetization Along [100]; Strain Measurement Along [100] and [010] We introduce superscripts and subscripts to identify the magnetization direction and the strain measurement direction, respectively. 1 3 100 100 100 100 100 = λ100 ; 010 = − λ100 ; 100 − 010 = λ100 2 2

    (12)

    It follows that λ100 can be derived from two strain measurements along and perpendicular to the saturation magnetization direction, irrespective of a reference state with unknown magnetization distribution. We also see that the transversal magnetostrictive strain is opposite in sign and half in magnitude of the longitudinal magnetostrictive strain. The same result is obtained for the strain along [001]. Thus, the sample volume, given by the trace of the strain tensor, remains constant. The values for λ111 are accessible from measurement along other directions. Magnetization Along [110]; Strain Measurement Along [110] and [1–10]

    110 110 =

    1 3 1 3 3 110 110 110 = λ100 − λ111 ; 110 − 1−10 = λ111 λ100 + λ111 ; 1−10 4 4 4 4 2

    (13)

    564

    D. Sander

    Magnetostriction of Polycrystalline Cubic Materials We may regard an isotropic polycrystalline sample as a polycrystal with random orientation of crystallites. In this case, the distinction between λ100 and λ111 is obsolete, and we introduce λs instead. Such a sample should also have vanishing elastic anisotropy, and this is reflected by c44 = (c11 −c12 )/2; see Eqs. (7), (8). Thus, also a single magnetoelastic coupling coefficient B is sufficient for the description. A compact expression of the magnetostrictive strain of the isotropic sample follows from Eq. (11) as 3 δ 1 = λs cos2 (ϑ) − . 2 3

    (14)

    The angle ϑ is measured between the magnetization direction and the strain measurement direction. The reference to the crystal axes is lost. Determination of the Isotropic Magnetostriction A strain measurement along the direction of magnetization, λ , and perpendicular to it, λ⊥ , gives λ − λ⊥ =

    3 λs , 2

    (15)

    as is depicted in Fig. 7a. We can relate the isotropic magnetostriction λs to λ100 and λ111 , if the directional dependencies of the latter two are considered. To this end, we revert to Eq. (11) with βi = αi and obtain   3 δ 1 4 4 4 = λ100 α1 + α2 + α3 − + 3λ111 α12 α22 + α12 α32 + α22 α32 . 2 3

    (16)

    The directional dependence is given by terms cos2 (ϑ), cos4 (ϑ), and cos2 (ϑ) sin2 (ϑ). A random orientation of crystallites, which is the basis of the isotropic sample described here, is considered by averaging the directional dependence over the unit sphere. We calculate the respective averages

    cos (ϑ)avg = 2







    0

    cos4 (ϑ)avg =



    0

    cos2 (ϑ) sin(ϑ)dϑdϕ/(4π ) =

    1 3

    cos4 (ϑ) sin(ϑ)dϑdϕ/(4π ) =

    1 5

    0

    0

    cos2 (ϑ) sin2 (ϑ)avg =

    π

    π

    0

    0

    π

    cos2 (ϑ) sin3 (ϑ)dϑdϕ/(4π ) =

    2 15

    (17)

    11 Magnetostriction and Magnetoelasticity

    565

    and finally obtain the isotropic magnetostriction λs λs =

    2 3 λ100 + λ111 . 5 5

    (18)

    This expression was first presented by Akulov [59] and was later criticized for neglecting elastic anisotropy [60], which leads to a correction, which can be negligible or sizable, depending on the material [60].

    Derivation of the Magnetostrictive Strain Tensor: Hexagonal Case We follow the same procedure as for the cubic case above. We start from the following expressions for the elastic and the magnetoelastic energy density [17]. We choose a coordinate system where the z-axis coincides with the c-axis and the x-axis with one of the unit vectors of the basal plane. The reference state is ascribed to magnetization parallel to the c-axis ( α3 = 1, α1 = α2 = 0). fel =

    1 1 2 2 2 c11 (11 + 22 ) + c33 33 + c12 11 22 + c13 (11 33 + 22 33 ) 2 2 2 2 2 + 2c44 (23 + 13 + (c11 − c12 )12 + ...

    (19)

    fme = B1 (11 α12 + 212 α1 α2 + 22 α22 ) + B2 (1 − α32 )33 + B3 (1 − α32 )(11 + 22 ) + 2B4 (23 α2 α3 + 13 α1 α3 ) + . . .

    (20)

    We calculate the strain derivative of the total energy density, set it to zero, and solve for the ij . We then calculate the magnetostrictive strain for four different magnetization states [61, 62] and obtain the following definitions for the Bi : B1 = −(c11 − c12 )(λA − λB ); B2 = −c13 (λA + λB ) − c33 λC B3 = −c12 λA − c11 λB − c13 λC ; B4 = c44 (λA + λC − 4λD )

    (21)

    For convenience, we present the expressions for λi in terms of the Bj in the Appendix. We obtain the following magnetostriction tensor: ⎛

    λhex

    ⎞ λA α12 + λB α22 (λA − λB )α1 α2 λACD α1 α3 = ⎝ (λA − λB )α1 α2 λA α22 + λB α12 λACD α2 α3 ⎠ λACD α1 α3 λACD α2 α3 λC (1 − α32 )

    (22)

    with the abbreviationλACD = − 12 (λA + λC − 4λD ). This gives the magnetostrictive strain δ in the hexagonal system

    566

    D. Sander

    δ

    hex = βi βj λij 3

    3

    (23)

    i=1 j =1

    = λA (α1 β1 + α2 β2 )(α1 β1 + α2 β2 − α3 β3 ) + λB (α2 β1 − α1 β2 )2 +λC β3 (α1 α3 β1 + α2 α3 β2 + (α32 − 1)β3 ) +λD 4α3 β3 (α1 β1 + α2 β2 ).

    (24)

    The individual magnetostriction coefficients λi are defined by the following orientations of magnetization α = (α1 , α2 , α3 ) and strain measurement β = (β1 , β2 , β3 ): λA : α = (1, 0, 0), β = (1, 0, 0) β = (0, 1, 0) λB : α = (1, 0, 0), β = (0, 0, 1) λC : α = (1, 0, 0), 1 1 √ √ λD : α = ( , 0, ), β = ( √1 , 0, √1 ) 2

    2

    2

    (25)

    2

    Magnetostriction of Polycrystalline Hexagonal Materials We derive the expression for the effective magnetostriction in a polycrystal with random orientation of grains with hexagonal symmetry. We distinguish between λ and λ⊥ for strain measurements along the magnetization direction and perpendicular to it, respectively. Starting point is Eq. 24, where we set βi = αi to derive λ . λ = λA ((1 − α32 )2 − α32 (1 − α32 )) + 4λD α32 (1 − α32 ) = λA (sin4 (ϑ) − cos2 (ϑ) sin2 (ϑ)) + 4λD cos2 (ϑ) sin2 (ϑ)

    (26)

    Here, ϑ is the angle between the magnetization direction and the c-axis. With the calculated averages of the trigonometric functions from Eq. (17) and with 8 (sin4 (ϑ))avg = 15 , we get λ =

    2 8 λA + λD . 5 15

    (27)

    The relative volume change of magnetostriction ω is given by the trace of the magnetostrictive strain tensor of the hexagonal system in Eq. (22). It is linked to both λ and λ⊥ via ω = (λA + λB + λC ) sin2 (ϑ) = λ + 2λ⊥ .

    (28) (29)

    11 Magnetostriction and Magnetoelasticity

    567

    The volume magnetostriction of a sample with hexagonal symmetry depends on the orientation of the magnetization with respect to the c-axis. Replacing sin2 (ϑ) with its average 23 , inserting λ , and solving for λ⊥ give λ⊥ =

    1 (2λA + 5λB + 5λC − 4λD ). 15

    (30)

    Thus, the magnetostrictive strain λ , λ⊥ of a sample with random orientation of crystallites with hexagonal symmetry depends on all magnetostrictive coefficients λ A . . . λD .

    Magnetostriction and Stress-Induced Magnetic Anisotropy Inspection of Eqs. (3) and (20) reveals that magnetoelasticity is intimately linked to magnetic anisotropy insofar as it involves terms which depend on the direction of magnetization. Consequently, the application of stress τ induces a lattice strain , and this strain contributes via the magnetoelastic coupling coefficients Bi to the energy density. It is important to note that it is the stress-induced strain which drives the corresponding change of magnetic anisotropy. This impact of an imposed lattice strain on magnetic anisotropy is often referred to as inverse magnetostrictive effect. Theory exploits the link between lattice strain and magnetoelastic effects by identifying magnetoelasticity as the derivative of magnetic anisotropy with respect to strain [63, 64]. This aspect is elucidated in section “Magnetostriction and Magnetoelasticity: Physical Origin and Insights from Theory”. Here, we briefly note that for 3d metals the corresponding physics is due to the spin-orbit interaction. Its dependence on lattice strain is a daunting topic. This is also due to the small energy scales, which typically are in the μeV per atom range. Reliable calculations with the required numerical accuracy remain challenging. Thus, phenomenological pictures are still used to describe stress-induced magnetic anisotropy. Here, we focus on illustrative examples to provide a basic description of the matter for a cubic system. We calculate the stress-induced lattice strain and insert the obtained strain in the equation of the magnetoelastic energy density fel of Eq. (3). Then, the orientational dependence of the expression is analyzed. The stress-strain relation is derived from the expression of the elastic energy density of Eq. 2 by exploiting τij = ∂fme /∂ij . In the following, we neglect shear stress by setting τij = 0 for i = j and focus on the cubic system. We obtain τ11 = c11 11 + c12 22 + c12 33 τ22 = c12 11 + c11 22 + c12 33 τ33 = c12 11 + c12 22 + c11 33 and solve these equations for ii to derive the stress-induced lattice strain.

    (31)

    568

    D. Sander

    Isotropic Stress An isotropic stress τ = τii gives rise to an isotropic strain of magnitude  = ii = τ/(c11 + 2c12 ). An isotropic contribution to fel of magnitude B1 τ/(c11 + 2c12 ) results. Thus, the magnetic anisotropy, defined as energy difference between two magnetization directions, remains unaffected. Uniaxial Stress We assume an uniaxial stress τ11 = τ, τ22 = τ33 = 0. A three-dimensional strain state results, where the lattice strain differs along the direction of applied stress from its value in the plane perpendicular to the stress direction. We obtain c11 + c12 (c11 − c12 )(c11 + 2c12 ) c12 . = −τ (c11 − c12 )(c11 + 2c12 )

    11 = τ 22 = 33

    (32)

    The ratio −22 /11 = −33 /11 defines the Poisson ratio ν = c12 /(c11 + c12 ). Insertion into Eq. 3 and omitting isotropic terms give the stress contribution to magnetic anisotropy as τ fme =τ

    B1 α2 c11 − c12 1

    3 = − τ λ100 cos2 (ϑ), 2

    (33)

    where ϑ gives the angle between the direction of stress and magnetization. Thus, an uniaxial magnetic anisotropy results. How large are stress-induced magnetic anisotropies? Can they potentially change the easy magnetization direction as given by the magnetocrystalline anisotropy? Stress-induced magnetic anisotropies scale as Bi . Typical values for B are of order several MJ/m3 . The strain in bulk samples is certainly always below 0.01, and thus the stress contribution in a bulk sample can reach 0.1 MJ/m3 . This is a sizable contribution to magnetic anisotropy in comparison to K1 = 0.05 MJ/m3 of bulk Fe [17, 65, 58]. For systems with smaller magnetocrystalline anisotropy, such as Ni or amorphous metals, stress-induced magnetic anisotropy is often the dominant source. In epitaxial ultrathin films, much larger strains of up to several percent may prevail, and stress-induced anisotropy should always be considered [17, 65]. In view of the impact of magnetoelasticity on anisotropy, it is important to realize that terms B, Bτ, τ λ contribute. Thus, the application of a tensile stress τ > 0 to a sample with positive magnetostriction (λ > 0, B < 0) has the same impact on anisotropy as the application of a compressive stress (τ < 0) to a sample with negative magnetostriction. If λ is positive, then it is easier to magnetize the material in the direction of a tensile stress. This link between stress and magnetic anisotropy plays a role for the domain distribution in stressed materials and for the approach to

    11 Magnetostriction and Magnetoelasticity

    λ>0

    (a)

    τII

    τII

    V(MII)>V(M ) T

    (b) Mavg=0 τ=0

    569

    MII MS

    λ0 τ=0

    τII, λV(MII)

    0

    H II τ

    T

    Fig. 8 The impact of stress on the magnetization curve of a magnetostrictive material. (a) Tensile stress favors a longitudinal domain configuration in a material with positive magnetostriction. (b) A transversal domain configuration is favored for negative magnetostriction. Magnetization along the tensile stress direction gives a reduced effective anisotropy in (a) (red curve) and a larger anisotropy in (b) (blue curve)

    saturation magnetization. The interplay between stress and domain configuration of a sample is schematically illustrated in Fig. 8. The interplay between stress, strain, and domain configuration is also relevant for the E effect discussed above.

    Magnetoelastic Effects in Films Magnetoelastic effects are of significant importance for understanding magnetic anisotropy in thin films [17, 58, 65, 66, 50]. The reason is that magnetoelasticity couples the lattice strain to the magnetization direction via the coefficients Bi , as expressed in Eqs. (2) and (20). Films are, as result of the deposition on a substrate, almost always strained. The situation is often transparent for epitaxial pseudomorphic films on single crystal substrates, where the misfit between film and substrate gives the lattice strain [67, 68, 69]. The resulting film strain is anisotropic. It differs in sign and magnitude for the in-plane and the out-of-plane direction [17]. This three-dimensional strain state of the film modifies the magnetic anisotropy. Easy magnetization directions, which differ from those of the respective bulk materials, are commonly observed. Magnetoelasticity is often the key driving force [17, 65]. Numerous experiments have established that magnetoelasticity in films may deviate from the bulk [53, 54, 57, 55, 56, 41, 17, 58, 50]. The sign and magnitude of Bi are found to change with lattice strain and film thickness. A theoretical understanding of this phenomenon is progressing [70, 71, 72, 73, 74, 75, 76, 77]. Theory provides physical insights in support of a strain-dependent correction B1eff = B1 + D1 ij ,

    (34)

    570

    D. Sander

    leading to an effective magnetoelastic coupling coefficient B1eff . The magnitude of the correction term D1 is sizable, of order 100–1000 MJ/m3 . Table 4 provides an overview of corresponding experimental and theoretical data. Thus, strains in the percent range may already change the effective magnetoelastic coupling considerably.

    Experimental Determination of Magnetostriction and Magnetoelastic Coupling A direct measurement of magnetostriction in bulk samples is possible by measuring the change of length of a sample upon magnetization as described by Eqs. (11) and (24). The resulting sample distortion is schematically indicated in Fig. 9. Strain measurements by strain gauges, by capacitive and optical measurements, have been described to measure the usually small relative change of length of order 10−6 with high precision [78]. Also, the variation of the tunnel current between tip and magnetostrictive sample, as induced by a magnetostrictive change of length of the tunnel barrier, has been analyzed to derive magnetostriction data [79, 80]. It appears that among all techniques the capacitance method, where the change of length of the sample induces a change of distance of the plate separation in a capacitor, has the highest sensitivity and reliability. It is a bulk, rather than a surface measurement where the sample is part of the electrode of a parallel-plate capacitor, and its change of capacity is measured with a three-terminal capacitance measurement [81, 82, 83, 84, 85, 86, 87, 88, 89]. Thus, even minute magnetostriction of order 10−10 in paramagnetic samples has been detected [90]. The accurate measurement of magnetostrictive strain requires the consideration of the form effect. The magnetization of a sample leads to a demagnetizing field within the sample volume which is related to the demagnetizing factor of the

    (a)

    (b)

    Fig. 9 (a) Magnetostriction in bulk samples is accessible from the magnetization-induced change of the lattice strain λ , λ⊥ . (b) Magnetization of a film induces a biaxial magnetoelastic stress, which is related to the magnetoelastic coupling coefficients Bi . A thin substrate undergoes a anticlastic curvature with radii of curvature of opposite sign along and perpendicular to the magnetization direction. Curvature measurements give Bi ; see Eq. (36)

    11 Magnetostriction and Magnetoelasticity

    571

    sample shape. Thus, for magnetostatic reasons alone, a magnetized sample tends to elongate along the magnetization direction, thereby lowering its demagnetizing energy. This leads to the superposition of the form effect-driven lattice strain with the magnetostrictive lattice strain. When λ is small, both can be of comparable magnitude, and corresponding corrections are required [91]. The Wiedemann Effect: Magnetization-Induced Sample Twist Alternatively, the Wiedemann effect can be applied to deduce magnetostriction from a magnetization-induced sample twist (Fig. 5). The helical twist is straightforwardly measured with high sensitivity by reflecting a laser beam from the sample onto a position-sensitive detector. This method is extremely useful for samples in the form of thin ribbons and wires [92,38], where strain measurements are otherwise difficult to implement. Quantitative analysis reveals [92] a non-monotonic dependence of the torsion per unit sample length ξ on the magnetic field produced by coil and current. The maximum torsion per length is related to the isotropic magnetostriction constant λs via ξmax =

    2.2λs , a

    (35)

    where the sample thickness is 2a. Indirect Methods to Determine Magnetoelastic Properties Magnetoelasticity also reveals itself by its impact on the magnetization process. Thus, indirect access to magnetoelasticity is possible from an analysis of magnetization curves [53, 54]. Straining a magnetostrictive sample modifies its magnetic properties. Magnetoelasticity couples applied stress with magnetic anisotropy and magnetic susceptibility. The contribution of an externally applied stress to the magnetic anisotropy of the system is given by Eq. (33). The resulting change of the effective magnetic anisotropy is accessible from the change of magnetization curves and from the modified initial susceptibility [54]. Straining an amorphous sample with negative magnetostriction increases the effective magnetic anisotropy and lowers the initial susceptibility. Other techniques which have the potential to detect changes of the magnetic anisotropy upon the application of stress, such as small angle magnetization rotation (SAMR) [93, 94], ferromagnetic resonance (FMR) [95, 96], and Brillouin light scattering (BLS) [97], give access to magnetoelastic coupling. The analysis of the data requires modeling of the contributions to magnetic anisotropy and its strain dependence to extract the magnetoelastic coupling coefficients reliably.

    Magnetoelastic Coupling in Films The indirect methods described above may also be used to study magnetoelastic effects in thin films. However, they lack the direct link between the experimental

    572

    D. Sander

    observable and the magnetoelastic coupling coefficients that exists in magnetoelastic stress measurements, which are described next. Magnetostriction is a property of bulk samples which are free to strain upon changing the magnetization state. Ferromagnetic films are usually bonded to a substrate, which inhibits film strain upon changing the film magnetization. Instead, a magnetoelastic stress arises, and the resulting strain depends on the stiffness of the film-substrate composite. This magnetoelastic stress is of magnitude Bi , and it amounts to several MPa. Although it is three orders of magnitude less than typical epitaxial misfit stress, which amounts to GPa for a misfit in the 1% range [17]. The magnetoelastic stress can still be detected by the stress-induced curvature of the film-substrate composite. Figure 10 gives a schematic of a corresponding experimental arrangement which is used to prepare ferromagnetic monolayers and analyze their stress and magnetic properties under ultrahigh vacuum conditions in situ [50, 98]. Figure 10 provides a sketch of an optical two-beam curvature measurement in conjunction with a magnet system for magnetoelastic studies under ultrahigh vacuum conditions. The optical deflection measurement is mounted to a window

    Fig. 10 (a) Sketch of an ultrahigh vacuum chamber for in situ film growth with magnets for magnetoelasticity studies. (b) Optical curvature measurement. The optics is mounted outside of the vacuum chamber, and the sample is in the vacuum chamber. 1: sample manipulator, 2: substrate with film (3) on front surface, 4: laser diode with optics, 5: beam splitter, 6: mirrors, 7: piezo transducer, 8: split photodiode. (c) Schematic curvature change upon an in-plane magnetization reorientation of a film-substrate composite with positive Beff

    11 Magnetostriction and Magnetoelasticity

    573

    flange of the vacuum chamber for in situ stress measurements during film growth and during magnetization processes. The experiment involves a measurement of the change of substrate curvature upon an in-plane reorientation of the magnetization direction [17, 65, 99]. The curvature along the sample length is measured for magnetization along its length and magnetization along its width. The magnetoelastic stress produces an anticlastic curvature of the sample. Let us assume a cubic (100)-film with negative B1 (λ100 > 0). The sample length is along [100]. Magnetization along [100] induces a compressive stress along [100], as the film has the tendency to expand along its magnetization direction. This induces a negative curvature along the sample length. A tensile stress results along the perpendicular in-plane direction. Magnetization along the film width [010] rotates the compressive stress to the film width, and a tensile stress along the film length results. An anticlastic curvature results in both magnetization states. The sample changes its curvature from curved downward to curved upward upon magnetization reorientation. This situation is schematically shown in Fig. 10c. This curvature change is directly linked to the effective magnetoelastic coupling coefficient Beff [17] Beff

    Es ts2 = 6(1 + νs )tf

    

     1 length 1 width − . Rx Rx

    (36)

    Note that the elastic properties of the substrate enter via the Young modulus Es and the Poisson ratio νs . Substrate and film thickness are given by ts and tf , respectively. The superscripts give the magnetization direction. Different magnetoelastic coupling coefficients Bi can be measured depending on the crystallographic orientation of the film. Table 1 gives an overview of the required experiments to measure Bi [99]. The quantitative analysis of magnetoelastic curvature measurements needs to take crystalline anisotropy and substrate clamping into account [100, 99]. The use of substrates with a length-to-width ration larger than three minimizes the

    Table 1 Effective magnetoelastic coupling coefficient Beff for films with the given symmetry and orientation for an in-plane reorientation of the magnetization. The Bi are defined in Eq. (20). x1 and x2 denote the direction along film length and film width, respectively Film structure Cubic, film axes along x1 , x2 Cubic, film axes rotated by 45◦ with respect to x1 Hexagonal, c-axis in-plane, 2 domains parallel to x1 , x2 Hexagonal, c-axis in-plane, 2 domains rotated by 45◦ with respect to x1 Hexagonal, c-axis in-plane, 3 domains rotated by 120◦ with respect to x1

    Beff B1 B2 1 2 (B1

    − B2 + B3 )

    B4 1 4 (B1

    − B2 + B3 ) + 12 B4

    574

    D. Sander

    detrimental effect of substrate clamping on the two-dimensional curvature state. The radius of curvature is in the range 100 m–100 km for magnetoelastic stress measurements of nm thin films. Magnetization-induced curvature measurements have been successfully applied to measure magnetoelastic coupling coefficients in films as thin as a few atomic layers [101, 102, 103, 104, 56, 17].

    Magnetostriction and Magnetoelasticity: Physical Origin and Insights from Theory Magnetostriction and magnetoelasticity are due to the coupling between magnetism and atomic structure [63,105,106]. This is brought about by the spin-orbit coupling. It can be understood as the coupling between the electron spin moment with the magnetic field created by its own orbital motion. This orbital motion is coupled to the lattice by the Coulomb potential of the atom cores. As a result, the energy of the system changes depending on the orientation of the magnetization with respect to the lattice. This defines the magnetocrystalline anisotropy, and it leads to the occurrence of specific easy magnetization directions [107, 108, 109, 110, 111, 112, 105, 113, 106, 114]. The magnetocrystalline anisotropy itself depends on lattice strain [106]. Thus, the total energy of the system with a specific magnetization orientation can be lowered by a concurrent lattice strain [64]. The strain dependence of the magnetocrystalline anisotropy is the basis for current calculations of magnetoelastic effects [63, 115, 76, 113, 106, 116]. Despite the considerable advance of electronic structure calculations, the theoretical study of magnetic anisotropy and magnetoelasticity in the framework of density functional theory remains a challenging task [73, 106]. Highly accurate numerical calculation of the total energy density ftot of a sample upon a small lattice strain identifies the magnitude of magnetostriction as the strain which minimizes ftot . Corresponding calculations for Ni presented in Fig. 11 [64] reveal a parabolic dependence of ftot on lattice strain. Note the small magnitude of the energy scale of 0.1 μeV per atom. An amazingly high numerical accuracy is required to tackle magnetoelastic effects reliably. The calculated magnetostriction values, as given by the minima of the plot, have the correct sign, but are roughly a exp. exp. factor four larger than the experimental values (λ100 = −64.5 × 10−6 and λ111 = −28.3 × 10−6 ) [17]. The small magnitude of magnetic anisotropy of 3d transition metals of order several μeV per atom requires high numerical accuracy for the theoretical treatment of this topic. The magnetocrystalline anisotropy EMCA has been calculated for bulk metals, 3d films, and nanostructures [117, 113, 118]. Theory also derived the variation of EMCA with lattice strain, and this property is related to magnetoelasticity. The slope of the plot EMCA as a function of lattice strain is related to the magnetoelastic coupling coefficients [119, 105].

    11 Magnetostriction and Magnetoelasticity

    -0.05

    Ni

    trigonal

    λ111 energy (μeV/atom)

    Fig. 11 Calculated total energy density in dependence of a tetragonal and a trigonal strain of Ni [64]. The minima identify the magnetostriction constants λ100 : −245 × 10−6 and λ111 : −110 × 10−6 . Note the small energy scale of order 100 neV or 1 mK per atom

    575

    -0.15

    -0.25

    tetragonal

    λ100

    -200 -100 distortion (10-6)

    Fig. 12 Calculated change of density of states upon uniaxial compression of bulk Fe from a tight-binding calculation. The plot indicates a shift to lower (higher) energy of bands with dz2 (red) (dx 2 −y 2 ) (blue) symmetry. This changes the spin orbit coupling interaction, and consequently the magnetic anisotropy of the system is modified [115]

    0.04

    d z2

    d

    2

    2

    0.02 0 -0.02 -0.04 -3

    -2

    -1

    0

    1

    Ab initio calculations offer some insights into the physical origin at the electronic level of the strain dependence of the magnetocrystalline anisotropy [115]. Figure 12 shows the calculated change of the density of states upon uniaxial compression of bulk Fe. This tetragonal strain shifts and broadens electronic bands. As a consequence, the spin-orbit coupling between bands of different symmetry near the Fermi energy is changed, and this drives a strain-induced change of magnetic anisotropy [115]. Recent measurements of magnetoelastic coupling in atomic layers have revealed that film strain may modify the magnetoelastic coupling coefficients in a way that can be described by higher-order strain contributions to the magnetoelastic energy density [50]. Calculations [71, 70, 76, 74, 75, 77] support the phenomenological introduction of strain-modified magnetoelasticity, as discussed in section “Magnetoelastic Effects in Films”.

    576

    D. Sander

    A comparison between different calculational schemes within density functional theory and experimental data is presented in Table 4. Inspection of the data reveals significant quantitative differences between experiment and theory which illustrate the challenge presented by accurate calculation of magnetostrictive properties. Nonetheless, recent theoretical work has succeeded in shining fresh light on the electronic origin of the intriguingly large magnetostriction of some Fe alloys [120,121,122,116] (see section “Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si, Fe-Ga-Al, and Fe-Ga-Ge Alloys”) and of the large positive volume magnetostriction of invar alloys [28, 30]. The electrons responsible for magnetism in transition metals are largely delocalized, and one refers to this class of materials as itinerant magnets. However, ferromagnetism also exists in 4f metals, albeit with significantly lower ordering temperatures, and the electronic picture of magnetism is quite different. The origin of rare-earth magnetism is ascribed to localized 4f electrons. The 4f ferromagnetic elements show a pronounced magnetic anisotropy, which is ascribed to a strong intra-atomic spin-orbit coupling of the 4f electrons [123]. This intra-atomic coupling may exceed the anisotropic interaction with all other charges of the metal [124]. In the limiting case of infinite anisotropy, one may envision magnetostriction in 4f metals as being due to the magnetic field-induced rotation of the highly anisotropic charge distribution of the 4f electrons. Prototypical examples are Tb and Er. The charge distribution of the former is oblate and that of the latter prolate. One may expect that the magnetostrictive strain into the direction of a strong magnetic field differs in sign for both elements. This is indeed calculated and observed experimentally [123]. Tb expands along the c-axis for magnetization along the c-axis, whereas Er contracts for the same measurement. In view of the paramount importance of the 4f charge distribution for magnetism, one refers to single-ion anisotropy in these elements to distinguish it from the itinerant magnetism of 3d metals. Antiferromagnetism and ferrimagnetism are found in oxides of 3d metals. Superexchange between 3d cations, mediated by O2− anions, drives the magnetic order. Examples are the antiferromagnetic monoxides MnO, FeO, CoO, NiO, and ferrites of the general composition MFe2 O4 (M = Cr, Mn, Fe, Co, Ni, Cu, Zn) [10], respectively. The antiferromagnetic oxides show a lattice distortion below the Néel temperature TN , which can be described as a magnetostrictive phenomenon in a wider sense [125]. The discussion of magnetostriction in these oxides and ferrites relies on the single-ion anisotropy model [125, 126, 10]. A striking example is CoO, which is cubic (a = 4.261 Å) above TN = 288 K. It contracts at lower temperature along the spin axis (c-axis) and expands in the perpendicular plane (ab-plane) [127]. Extrapolation to 0 K gives a strain of magnetic origin along c by −1.5% and an expansion of +0.9% along a. This large lattice strain has been ascribed to a repulsive Coulomb interaction between the 3d electrons of Co2+ cations, which are responsible for the magnetism and the charge distribution of the O2− anions [126]. The regular orthogonal environment seen by the Co-cation, as produced by the nearest neighbor O-anions, is decisive for an understanding on the electronic level. A crystal field results, which splits the electronic levels

    11 Magnetostriction and Magnetoelasticity

    577

    of Co2+ into double-degenerate eg states at higher-energy and lower-lying tripledegenerate t2g states. The orbital moment of Co2+ in CoO is not quenched, and it is free to align along the spin axis. The corresponding charge distribution is of pancake shape, with its wider extension in the ab-plane, giving rise to the repulsive Coulomb interaction in that plane. The Co2+ cation in CoFe2 O4 ferrite of spinel structure experiences a crystal field of different symmetry. The 3d electronic states of the cations interact also with next-nearest neighbor Co-cations, in addition to interactions with nearest neighbor O-anions in the orthogonal arrangement. The former gives rise to an interaction potential of trigonal symmetry with respect to the < 111 >-axis. This splits the triple-degenerate t2g state into a lower-lying singlet and a higher doublet state. The filling of this electronic-level scheme leads also to an unquenched orbital moment, and a remarkably large magnetostriction at room temperature of λ100 = −590 × 10−6 results [128, 126]. It is ascribed to the strain dependence of the crystal field energy. If the theoretical description of magnetostriction is demanding for bulk samples, this is even more true for the theory of magnetoelasticity of films [50]. In view of the electronic origin of magnetic anisotropy and its strain dependence, it is to be expected that electronic hybridization between film and substrate, or film and adsorbate layer, modifies not only the atomic structure near the interface [129, 130] but also spin polarization [131, 132, 129], magnetic anisotropy [65], and magnetoelasticity [17, 58, 50]. Thus, with decreasing film thickness down to single atomic layers, the discussion of magnetoelasticity in terms of lattice strain alone misses important aspects.

    Compilation of Data Magnetoelastic and Elasticity Data for Bulk Transition Metals Tables 2 and 3 present magnetoelastic coupling and magnetostrictive coefficients, and elasticity data, respectively, for selected 3d bulk metals. A recent study on the magnetostriction of binary and ternary Fe alloys [91] gives values for pure Fe at room temperature and at 77 K (in parentheses) λ100 = 23.3(22.6) × 10−6 and B1 = −3.4(−3.5) MJ/m3 . These values are in close agreement with the values presented in Table 2. Table 2 Room temperature values of λ and B [17]. Note that Eq. (3.4) in [17] should read B4 = c44 (λA + λC − 4λD ). The Bi are calculated from the λi . The required elasticity data are in Table 3. The values for fcc Co are extrapolated from data for PdCo alloys [17] Element zbcc Fe fcc Co fcc Ni hcp Co

    B1 −3.43 −9.2 9.38 −8.1

    B2 7.83 7.7 10 −29

    λ100 24.1 75 −64.5

    λ111 −22.7 −20 −28.3

    B3

    B4

    λA

    λB

    λC

    λD

    28.2

    37.4

    −50

    −107

    126

    −105

    578

    D. Sander

    Table 3 Elastic stiffness constants cij , Young modulus E in GPa. The latter is calculated for the (0001) and (100) orientation for hcp and cubic elements, respectively. The Poisson ratio is given by ν. (Data from [17]) Element hcp Co fcc Co bcc Fe fcc Ni

    c11 307 242 229 249

    c12 165 160 134 152

    c44 75.5 128 115 118

    c13 103

    c33 358

    E 211 114 131 133

    ν 0.49 0.40 0.37 0.38

    Theoretical and Experimental Values of Magnetoelastic Coupling Coefficients and Their Strain Dependence (Table 4) Table 4 Compilation of theoretical values and experimental results of magnetoelastic coupling coefficients Bi and the nonlinear strain-dependent coefficient Di with B eff = Bi + Di  [58]. B1 was derived from the published λ100 with Eq. 7 and the elastic constants from Table 3. Experimental data for fcc Co are extrapolated from data for PdCo alloys [17]. All calculations are for bulk under specific strain states. (All values are given in MJ/m3 .) LSDA local-spin-density approximation, GGA generalized-gradient approximation fcc Co Exp. bulk [17] LSDA [133] LSDA [73, 74] LSDA [113] GGA [133] GGA [73, 74] GGA [113] Exp. film Co/Ir(100) [58] fcc Ni Exp. bulk [17] LSDA [133] LSDA [73, 74] LSDA [113] GGA [133] GGA [73, 74] GGA [113] Exp. film Ni/Cu(100) [134] Exp. film Ni/Ir(100) [58] bcc Fe Exp. bulk [17] LSDA [133] LSDA [73, 74] LSDA [113]

    B1

    D1

    B2

    2D2

    −9.2 −11.3 −15.9 −13.8 −6.9 −9.8 −5.9 3.5

    – – 212 – – 186 – −842

    7.7 – 3 10.6 – 4.5 – 1.8

    – – 58 – – −71 930

    9.38 9.2 12.6 13.9 8.1 10.2 13.7 9.4 1.3

    – – −103 – – −53 – −234 273

    10 – 16.9 38.8 – 11.1 – 10 6.6

    – – −132 – – −47 – – −408

    −3.43 −7.4 −10.1 −8.3

    – – 337 –

    7.83 – −7.0 −8.9

    – – −40 –

    B3

    B4

    D4

    (continued)

    11 Magnetostriction and Magnetoelasticity

    579

    Table 4 (continued) GGA [133] GGA [73, 74] GGA [113] Exp. film Fe/MgO(100) and Fe/Cr/MgO(100) [57] Fe/Ir(100) [58] Fe/W(100) [17] hcp Co Exp. bulk [17] Exp. film Co/W(001) [135]

    B1 −4.1 −2.4 −4.8 −3.4

    D1 – 383 – 1100

    B2 – −3.9 – 7.8

    2D2 – 18 – −365

    −3.6 −1.2

    155 200

    – –

    – –

    −8.1



    −29



    B3

    B4

    D4

    28.2

    37.4 3.4

    1346

    Magnetostriction Data of Amorphous Fe Alloys (Table 5) Table 5 Low temperature and room temperature values of λS in 10−6 for amorphous alloys [11]

    Alloy a-Fe80 B20 a-Fe40 Ni40 B20 a-Co80 B20

    λS , 4.2 K 48 20 −4

    λS , 300 K 32 14 −4

    Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si, Fe-Ga-Al, and Fe-Ga-Ge Alloys Magnetostrictive and elastic data of these binary and ternary alloys [136] have been measured and compiled for different alloy compositions with high resolution on the composition scale [91]. Table 6 presents specific examples of those alloy compositions where magnetostriction is much larger or smaller than that of pure Fe. Note that a tenfold increase of magnetostriction of Fe71.2 Ga28.8 as compared to Fe is observed, although the magnetoelastic coupling coefficient is only twice as large as that of Fe. This indicates a contribution of elastic softening to the remarkably large magnetostrictive strain in this alloy. Equation 7 shows that a small value of (c11 − c12 ) leads to a large magnetostriction λ100 for a given magnetoelastic coupling B1 . Ultrasound spectroscopy revealed a decrease of the tetragonal shear modulus 0.5(c11 − c12 ) from the Fe bulk value of 48 to 6.8 GPa for a Ga content of 27.2% at room temperature [137]. In addition to lattice softening, density functional theory has revealed the atomic arrangement of Ga in Fe-Ga alloys as a further important aspect. The Ga-induced change of electronic structure is of key importance for the understanding of large magnetostriction in these alloys [138, 139]. A structural origin of large magnetostriction, due to nanoscale precipitates of different crystallographic order, has been proposed [140, 141, 142, 143], but this could not be supported by theory [139].

    580

    D. Sander

    Table 6 Room temperature values of λ100 and B1 for cubic Fe alloys with different composition [91]. All λ in 10−6 , B1 in MJ/m3

    Alloy Fe86.6 Ga13.4 Fe81.8 Ga18.2 Fe71.2 Ga28.8 Fe89.9 Ge10.1 Fe86.6 Ge13.4 Fe81.5 Ge18.5 Fe95 Si5 Fe87.8 Si12.2 Fe80.2 Si19.8 Fe93.3 Al6.7 Fe86.6 Al13.4 Fe79.8 Al20.2

    Fig. 13 Contour plots of λ100 for Fe-Al-Ga (left) and Fe-Ge-Ga (right) alloys. The grid spacing is 5 atomic %. Note the rapid change of sign of λ100 with slight additions of Ge in Fe-Ge-Ga. (Adapted from [91])

    Fe-Al-Ga Al

    λ100 +270

    λ100 143 185 242 55 11 −96 33 −10 −21 38 75 105

    B1 −12.6 −10.6 −6.49 −5.6 −1.0 −5.25 −4.3 1.1 2.3 −4.7 −8.11 −8.31

    Fe-Ge-Ga Ge

    +200 +120 +10 −27 −100

    Fe

    Ga

    Fe

    Ga

    The decrease of λ100 with increasing Ge content, leading to negative λ100 above 14% Ge content [91], has been tackled by theory [120], which finds that at small Ge content a Ge-induced softening of the tetragonal shear modulus contributes to a large magnetostriction. The calculated softening of the tetragonal shear modulus is in qualitative agreement with experimental results [91]. At higher Ge content, the availability of eg holes above the Fermi energy induces a negative magnetostriction [120]. It has been predicted that the addition of Cu (Fe79.7 Ga18.7 Cu1.6 ) should lead to an enhanced magnetostriction of λ100 = 550 × 10−6 [116], but this is not observed. However, doping melt-spun Fe83 Ga17 with 0.23 atomic % Tb has been reported to lead to a four times increased magnetostriction as compared to the binary alloy [144]. This increase has been ascribed to a localized increased spinorbit interaction in proximity to Tb [144] (Fig. 13).

    Zero Magnetostriction Alloys and Soft Magnetic Materials Specific applications require ultra-soft magnetic materials. They are characterized by small coercivity (low magnetization reversal fields) and high permeability. Examples are permalloy (FeNi alloy with more than 35% Ni); mumetal (NiFe

    11 Magnetostriction and Magnetoelasticity

    581

    alloy with roughly 80% Ni and small additions of other metals such as Cu, Cr, Mo, balance Fe); sendust (Fe alloy with 10 weight % Si and 5% Al); Fe-Si alloys, MnZn and NiZn ferrites, and amorphous (FeCo)B alloys; and nanocrystalline Fe-based alloys, which find widespread applications in magnetic sensors, transformers, electrical motors, and magnetic shielding [145, 146, 10]. Achieving these material properties is innately linked with the demand for vanishingly small magnetic anisotropy. An important contribution to magnetic anisotropy is magnetoelastic coupling, and consequently small magnetostriction is required. This reduces the strain-induced magnetic anisotropy, which arises inevitably in manufacturing and shaping (bending) the material for use in applications. Thermal annealing after manufacturing, often in an H2 atmosphere, may relax strain and is often performed to obtain ultra-soft magnets. Amorphous and nanocrystalline materials can be of high interest in corresponding applications, as their effective magnetocrystalline anisotropy may average out to zero, provided that the crystallographic grain size is smaller than the magnetic domain size. One example is the alloy Fe73.5 Cu1 Nb3 Si13.5 B9 with a suppressed magnetocrystalline anisotropy and zero magnetostriction [147]. In the quest for low magnetic anisotropy, one can exploit that different elements contribute to magnetic anisotropy with opposing signs, leading to zero net magnetic anisotropy. This venue is followed for permalloy, where the composition Ni78 Fe22 shows almost zero magnetocrystalline anisotropy and almost zero magnetostriction. Bringing both properties to zero at exactly the same composition can be achieved by adding further elements such as Mo and Cu. This is the material of choice when highest initial permeability of up to 106 μ0 is required [10]. A critical aspect for application is that permalloy films show a magnetostrictive behavior which depends on film thickness [56]. Figure 14 reveals that zero magnetostriction is only observed for films thicker than 7 nm, whereas thinner films

    0.5 0 λS (10-6)

    Fig. 14 Room temperature saturation magnetostriction of sputtered polycrystalline Ni81 Fe19 films, deposited on amorphous alumina on a Si(001) substrate. Films thinner than 7 nm show nonzero magnetostriction. (Adapted from [56])

    -0.5 -1.0 -1.5 -2.0 -2.5 -3.0

    0

    5 10 15 20 25 30 35 thickness (nm)

    582

    D. Sander

    exhibit magnetostriction of −2.5 × 10−6 at 3 nm. This suggests that other effects, such as sample roughness, film strain, and electronic hybridization between film and substrate, may lead to a qualitatively different magnetoelastic behavior in atomically thin films as compared to bulk [17, 65, 50].

    Magnetostriction Data for Paramagnetic Metals and Alloys Experiments reveal for paramagnetic samples a magnetic field-induced strain λ which is proportional to H 2 [90, 148]. The magnetostriction is ascribed to the volume dependence of the magnetic susceptibility [149, 90, 150]. Table 7 gives the slope of experimental data when λ is plotted as a function of H 2 .

    Magnetostriction Data for Bi Magnetostrictive strain measurements in Bi show pronounced oscillations, which are ascribed to the de Haas-van Alphen effect [151]. At a field of 10 T, a longitudinal magnetostriction of −10 × 10−6 is measured along the binary axis [151].

    Magnetostriction Data for Tb, Dy, and Ho Magnetostriction of the rare earth metals Tb, Dy, and Ho is large at cryogenic temperature with strains of order 10−3 [152, 153]. These elements show different magnetic order with increasing temperature, but none at room temperature.

    Table 7 Experimental data of the magnetostriction in 10−10 /T2 along the field direction obtained at liquid helium and liquid hydrogen temperatures in fields of up to 10 T [90, 148]

    Metal Ti Zr V Nb Ta Mo W Ru Rh Pd Ir Pt Rh50 Ir50 Rh50 Pd50 Pd67 Pt33 Pd33 Pt67

    λ −0.69 −3.79 4.18 3.6 1.25 7.12 1.83 −0.9 7.0 −25 2.4 −20.0 6.0 17.0 −11.0 −50

    11 Magnetostriction and Magnetoelasticity

    583

    Magnetostriction measurements of Tb [152], Dy, and Ho [153] as a function of field and at different temperatures reveal a considerable complexity, as indicated in Fig. 15. This is ascribed to different magnetic phases, which are probed by temperature and field variations [153]. Note the large magnitude of magnetostriction λ of order 10−3 at low temperature, two–three orders of magnitude larger than 3d transition metals (Table 2). Comparably large magnetostrictive strains have been reported below 150 K and in large fields of order several T for amorphous random anisotropy alloys containing 4f elements [154, 155].

    (a)

    x10-3 Tb 3.2

    78.7 K, λb

    2.4 1.6

    224.3 K, λc

    λ

    0.8 224.5 K, λb

    0

    79.7 K, λc

    -0.8 -1.6

    225.4 K, λa

    c

    a Hb

    -2.4

    79.6 K, λa

    0 0.4 0.8 1.2 1.6 2 2.4 2.8 H(T) (b) x10-3 Dy

    λ

    3

    144 K, λc

    2

    90 K, λc

    1

    44 K, λc

    0 -1 -2 -3 -4

    a H

    c 20 K, λc b

    144 K, λb 85 K, λb 20 K, λb

    0 0.5 1 1.5 2 2.5 H(T)

    (c) x10-3 Ho

    λ

    22 K, λa 85 K, λa

    4

    3

    45 K, λc 77 K, λc

    2

    4.2 K, λc

    1

    4.2 K, λb

    0 -1 -2

    c

    a Hb

    0 0.5 1

    45 K, λb 77 K, λb 77 K, λa

    4.2 K, λa 45 K, λa

    1.5 2 2.5 H(T)

    Fig. 15 Magnetostrictive strain for (a) Tb, (b) Dy, and (c) Ho measured along different directions with the field along the a and b direction for Dy and Tb and Ho, respectively. (Adapted from [152, 153])

    584

    D. Sander

    Magnetostriction Data of TbFe2 (Terfenol) and Tb27 Dy73 Fe2 (Terfenol-D) The Laves phase alloys TbFe2 (Terfenol) and DyFe2 show a large magnetostriction also at room temperature; see Fig. 16. Large magnetic fields are required to reach saturation magnetostriction close to 1.3 × 10−3 of the isotropic polycrystal. The necessity of high fields for saturation has been ascribed to the pronounced magnetic anisotropy of the material. SmFe2 shows a comparable magnitude of magnetostriction, although with opposite sign [21]. Curiously, the different magnetostrictive coefficients differ vastly, λ111  λ100 , and crystals with predominant (111) orientation give the largest strain [21]. Tb and Dy lead to fourth-order cubic magnetic anisotropy of opposite sign in the respective rare earth-Fe2 compound. Thus, the alloy composition of Tb-DyFe can be tuned such as to minimize the magnetic anisotropy, still preserving a large room temperature magnetostriction. The material has been named Terfenol-D (Tb27 Dy73 Fe2 ). It reaches a saturation magnetostriction of order 2.2 × 10−3 at room temperature in moderate fields of 0.3 T [21]. Alloying of Terfenol-D with other elements has not lead to higher magnetostriction, but specific material properties may be tuned by alloying [156].

    x 10-3

    1.0

    λll

    295 K

    0.8

    3.5

    TbFe2

    λll 3/2 λS

    0 λ

    TbFe2 λ 1.0

    2.0 H(T)

    1.5

    0.5

    -0.8 0

    2.0

    1.0

    T

    -0.6

    DyFe2

    T

    λ

    0.2

    -0.4

    TbFe2 (2.5 T)

    2.5

    DyFe2

    -0.2

    )

    3.0

    0.6 0.4

    TbFe2 (H=

    8

    x 10-3

    DyFe2 (2.5 T)

    0 3.0

    0

    50 100 150 200 250 300 T(K)

    Fig. 16 Magnetostrictive strain at 295 K (left) and temperature dependence of magnetostriction of the isotropic polycrystal 32 λS = λ − λ⊥ (right) for TbFe2 and DyFe2 . Note that a large magnetostriction prevails at room temperature. (Adapted from [18])

    11 Magnetostriction and Magnetoelasticity Table 8 Magnetostriction λ100 , λ111 , and λS (for polycrystals) in 10−6 at room temperature, unless noted otherwise [128, 18, 11, 13]

    585 Material Fe3 O4 Fe3 O4 (124 K) γ Fe2 O3 CoFe2 O4 Co1.1 Fe1.9 O4 Co0.9 Fe2.2 O4 Co0.3 Zn0.2 Fe2.2 O4 Co0.3 Mn0.4 Fe2.0 O4 Ni0.8 Fe2.2 O4 MgFe2 O4 Li0.5 Fe2.5 O4 Mn0.98 Fe1.86 O4 Mn0.6 Zn0.1 Fe2.1 O4 NiFe2 O4 Y3 Fe5 O12 (YIG)

    λ100 −19 −23

    λ111 81 55

    −670 −250 −590 −210 −200 −36

    120 120 110 65 −4

    −35 −14

    −1 14

    −1.4

    −1.6

    λS 41 24 −5 −110 −210 −18 −40 −17 −6 −8 −15 3 −26 −2

    Magnetostriction Data of Oxide Materials Exchange interactions in oxides are usually antiferromagnetic, but they can lead to ferromagnetic order in structures such as spinel or garnet. Some values are listed in Table 8. Values for cobalt-containing are large on account of the large spin-orbit interaction for the Co2+ cation.

    Appendix Notations for Lattice Strain There are different conventions for the definition of strain, and it is mandatory essential to distinguish between tensor strain ij and engineering strain γij [49,17]. The two conventions differ by a factor of two for the off-diagonal elements such that 2ij = γij for i = j . Note that the use of the tensor strain is required if it is transformed according to the tensor transformation rules to account for rotated coordinate systems [49]. Throughout this contribution, all strain expressions are given by the tensor strain. For completeness, we note that also the contracted Voigt notation is often used. Starting from the tensor strain, the replacement rule given in Table 9 applies to combine two subscripts into one. Note that factors of two are inserted for the components with subscripts 4, 5, and6 to get 4 = 223 , 5 = 213 , and6 = 212 .

    586

    D. Sander

    Table 9 Subscript replacement rule for the contracted Voigt notation. Note that factors of two are inserted for the components with subscripts 4, 5, and6 to get 4 = 223 , 5 = 213 , and6 = 212 Convention Tensor notation: Contracted Voigt notation:

    Subscripts 11 1

    22 2

    33 3

    23, 32 4

    31, 13 5

    12, 21 6

    Relation Between λ and B for the Hexagonal System We present here the relation between λi and Bj from Eq. (21). λA =

    2 −c c ) B2 (c11 − c12 )c13 + B3 (−c11 + c12 )c33 + B1 (c13 11 33 a

    (37)

    2 +c c ) B2 (c11 − c12 )c13 + B3 (−c11 + c12 )c33 + B1 (−c13 11 33 (38) a −B2 (c11 + c12 ) + (B1 + 2B3 )c13 λC = (39) a −B4 a + (−B2 (c11 − c12 )(c11 + c12 − c13 ) B3 (c11 − c12 )(2c13 − c33 ) λD = + 4ac44 4ac44

    λB =

    +

    2 − c c ))c B1 (c11 c13 − c12 c13 + c13 11 33 44 4ac44

    (40)

    2 + (c11 + c12 )c13 ) a = (c11 − c12 )(−2c13

    References 1. Becker, R., Döring, W.: Ferromagnetismus. Springer, Berlin (1939) 2. Kittel, C.: Physical theory of ferromagnetic domains. Rev. Mod. Phys. 21, 541–583 (1949) 3. Lee, E.W.: Magnetostriction and magnetomechanical effects. Rep. Prog. Phys. 18(1), 184 (1955) 4. Kneller, E.: Ferromagnetismus. Springer, Berlin (1962) 5. Kittel, C.: Introduction to Solid State Physics. Wiley, New York (1963) 6. Bozorth, R.M.: Ferromagnetism. IEEE-Press, Piscataway (1993) 7. Chikazumi, S.: Physics of Ferromagnetism. Oxford University Press, Oxford (2010) 8. du Trémolet de Lacheisserie, E.: Magnetostriction: Theory and Applications of Magnetoelasticity. CRC-Press, Boca Raton (1993) 9. Hubert, A., Schäfer, R.: Magnetic Domains. Springer, Berlin (1998) 10. O’Handley, R.C.: Modern Magnetic Materials: Principles and Applications. Wiley, New York (2000) 11. O’Handley, R.C.: Magnetostrictive materials and magnetic shape memory materials. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials, vol. 4, pp. 2401–2427. Wiley, New York (2007) 12. Moral, A.: Handbook of Magnetostriction and Magnetostrictive Materials. 2 volumes. Del Moral Publisher, University of Zaragoza, Zaragoza (2008)

    11 Magnetostriction and Magnetoelasticity

    587

    13. Coey, J.M.D.: Magnetism and Magnetic Materials. Cambridge University Press, Cambridge (2010) 14. Joule, J.P.: On a new class of magnetic forces. Ann. Electricity Magn. Chem. 8, 219–224 (1842) 15. Joule, J.P.: On the effect of magnetism upon the dimensions of iron and steel bars. Philos. Mag. (Third Series) 30, 76–87 (1847) 16. Joule, J.P.: On the effect of magnetism upon the dimensions of iron and steel bars. Philos. Mag.(Third Series) 30, 225–241 (1847) 17. Sander, D.: The correlation between mechanical stress and magnetic anisotropy in ultrathin films. Rep. Prog. Phys. 62, 809–858 (1999) 18. Clark, A.E., Belson, H.S.: Giant room-temperature magnetostrictions in TbFe2 and DyFe2 . Phys. Rev. B 5(9), 3642–3644 (1972) 19. Clark, A.E.: Magnetostrictive Rare Earth-Fe2 Compounds. Ferromagnetic Materials, vol. 1. Nort-Holland Publishing Company, Amsterdam (1980) 20. Clark, A.E., Restorff, J.B., Wun-Fogle, M., Lograsso, T.A., Schlagel, D.L.: Magnetostrictive properties of body-centered cubic Fe-Ga and Fe-Ga-Al alloys. Magn. IEEE Trans. 36(5), 3238–3240 (2000) 21. Clark, A.E., Wun-Fogle, M.: Modern magnetostrictive materials: classical and nonclassical alloys. Proc. SPIE 4699, 421–436 (2002) 22. Engdahl, G.: Chapter 1 – Physics of Giant magnetostriction. In: Engdahl, G. (ed.) Handbook of Giant Magnetostrictive Materials. Electromagnetism, pp. 1–125. Academic, San Diego (2000) 23. Wassermann, E.F.: Invar: moment-volume instabilities in transition metals and alloys. In: Buschow, K.H.J., Wohlfarth, E.P. (eds.) Ferromagnetic Materials, vol. 5, pp. 237–322. NortHolland Publishing Company, Amsterdam (1990) 24. Hausch, G.: Magnetovolume effects in invar alloys: spontaneous and forced volume magnetostriction. Phys. Status Solidi (A) 18(2), 735–740 (1973) 25. Wassermann, E.F.: The Invar problem. J. Magn. Magn. Mater. 100(1–3), 346–362 (1991) 26. Takenaka, K., Takagi, H.: Magnetovolume effect and negative thermal expansion in MnCuGeN. Mater. Trans. 47, 471–474 (2006) 27. Nagaoka, H., Honda, K.: Sur l’aimantation et la magnétostriction des aciers au nickel. J. Phys. Theor. Appl. 3(1), 613–620 (1904) 28. Khmelevskyi, S., Mohn, P.: Magnetostriction in Fe-based alloys and the origin of the Invar anomaly. Phys. Rev. B 69, 140404 (2004) 29. Shiga, M.: Invar alloys. Curr. Opin. Solid State Mater. Sci. 1(3), 340–348 (1996) 30. Khmelevskyi, S., Turek, I., Mohn, P.: Large negative magnetic contribution to the thermal expansion in iron-platinum alloys: quantitative theory of the Invar effect. Phys. Rev. Lett. 91, 037201 (2003) 31. Villari, E.: Intorno alle modificazioni del momento magnetico di una verga di ferro e di acciaio, prodotte per la trazione della medesima e pel passaggio di una corrente attraverso la stessa. Il Nuovo Cimento 20(1), 317–362 (1864) 32. Siegel, S., Quimby, S.L.: The variation of Young’s modulus with magnetization and temperature in nickel. Phys. Rev. 49, 663–670 (1936) 33. Hankemeier, S., Frömter, R., Mikuszeit, N., Stickler, D., Stillrich, H., Pütter, S., Vedmedenko, E.Y., Oepen, H.P.: Magnetic ground state of single and coupled permalloy rectangles. Phys. Rev. Lett. 103, 147204 (2009) 34. Berry, B.S., Pritchet, W.C.: Magnetic annealing and directional ordering of an amorphous ferromagnetic alloy. Phys. Rev. Lett. 34, 1022–1025 (1975) 35. Becker, R., Kornetzki, M.: Einige magneto-elastische Torsionsversuche. Zeitschrift für Physik 88(9–10), 634–646 (1934) 36. Hathaway, K.B., Clark, A.E., Teter, J.P.: Magnetomechanical damping in giant magnetostriction alloys. Metall. Mater. Trans. A 26(11), 2797–2801 (1995) 37. Wiedemann, G.: Über die Torsion und die Beziehung derselben zum Magnetismus. Ann. Phys. 182, 161–201 (1859)

    588

    D. Sander

    38. Drosdziok, S., Wessel, K.: A method for ultrasensitive magnetostriction measurement. Magn. IEEE Trans. 9(1), 56–59 (1973) 39. Schmoller, F.v.: Untersuchungen über den Matteucci-Effekt. Zeitschrift für Physik 93(1–2), 35–51 (1935) 40. Dimitropoulos, P.D., Avaritsiotis, J.N.: A micro-fluxgate sensor based on the Matteucci effect of amorphous magnetic fibers. Sensors Actuators A: Phys. 94(3), 165–176 (2001) 41. Szymczak, H.: From almost zero magnetostriction to giant magnetostrictive effects: recent results. J. Magn. Magn. Mater. 200(1–3), 425–438 (1999) 42. Ullakko, K., Huang, J.K., Kantner, C., O’Handley, R.C., Kokorin, V.V.: Large magnetic-fieldinduced strains in Ni2 MnGa single crystals. Appl. Phys. Lett. 69(13), 1966–1968 (1996) 43. O’Handley, R.C.: Model for strain and magnetization in magnetic shape-memory alloys. J. Appl. Phys. 83(6), 3263–3270 (1998) 44. Sozinov, A., Likhachev, A.A., Lanska, N., Ullakko, K.: Giant magnetic-field-induced strain in NiMnGa seven-layered martensitic phase. Appl. Phys. Lett. 80(10), 1746–1748 (2002) 45. Ye, M., Kimura, A., Miura, Y., Shirai, M., Cui, Y.T., Shimada, K., Namatame, H., Taniguchi, M., Ueda, S., Kobayashi, K., Kainuma, R., Shishido, T., Fukushima, K., Kanomata, T.: Role of electronic structure in the martensitic phase transition of Ni2 Mn1+x Sn1−x studied by hardX-ray photoelectron spectroscopy and ab initio calculation. Phys. Rev. Lett. 104, 176401 (2010) 46. Ibarra, M.R., Mahendiran, R., Marquina, C., García-Landa, B., Blasco, J.: Huge anisotropic magnetostriction in La1−x Srx CoO3−δ (x >∼ 0.3) Field-induced orbital instability. Phys. Rev. B 57, R3217–R3220 (1998) 47. Sato, K., Bartashevich, M.I., Goto, T., Kobayashi, Y., Suzuki, M., Asai, K., Matsuo, A., Kindo, K.: High-field magnetostriction of the spin-state transition compound LaCoO3 . J. Phys. Soc. Jpn. 77(2), 024601 (2008) 48. Ikuta, H., Hirota, N., Nakayama, Y., Kishio, K., Kitazawa, K.: Giant magnetostriction in Bi2 Sr2 CaCu2 O8 single crystal in the superconducting state and its mechanism. Phys. Rev. Lett. 70, 2166–2169 (1993) 49. Nye, J.F.: Physical Properties of Crystals. Oxford University Press, Oxford (1985) 50. Sander, D., Kirschner, J.: Non-linear magnetoelastic coupling in monolayers: experimental challenges and theoretical insights. Phys. Status Solidi (B) 248(10), 2389–2397 (2011) 51. Callen, E.R., Callen, H.B.: Static magnetoelastic coupling in cubic crystals. Phys. Rev. 129, 578–593 (1963) 52. Callen, E., Callen, H.B.: Magnetostriction, forced magnetostriction, and anomalous thermal expansion in ferromagnets. Phys. Rev. 139, A455–A471 (1965) 53. O’Handley, R.C., Sun, S.-W.: Strained layers and magnetoelastic coupling. J. Magn. Magn. Mater. 104–107(Part 3), 1717–1720 (1992) 54. O’Handley, R.C., Song, O.-S., Ballentine, C.A.: Determining thin-film magnetoelastic constants. J. Appl. Phys. 74, 6302–6307 (1993) 55. Koch, R., Weber, M., Rieder, K.H.: Magnetoelastic coupling of Fe at high stress investigated by means of epitaxial Fe(001) films. J. Magn. Magn. Mater. 159, L11–L16 (1996) 56. Kim, Y.K., Silva, T.J.: Magnetostriction characteristics of ultrathin permalloy films. Appl. Phys. Lett. 68(20), 2885–2886 (1996) 57. Wedler, G., Walz, J., Greuer, A., Koch, R.: Stress dependence of the magnetoelastic coupling constants B1 and B2 of epitaxial Fe(001). Phys. Rev. B 60(16), R11313–R11316 (1999) 58. Tian, Z., Sander, D., Kirschner, J.: Nonlinear magnetoelastic coupling of epitaxial layers of Fe, Co, and Ni on Ir(100). Phys. Rev. B 79, 024432 (2009) 59. Akulov, N.S.: Über die Anwendungen des Gesetzes ferromagnetischer Anisotropie zur Berechnung der Eigenschaften polykristallinischen Eisens. Zeitschrift für Physik 66(7–8), 533–542 (1930) 60. Callen, H.B., Goldberg, N.: Magnetostriction of polycrystalline aggregates. J. Appl. Phys. 36(3), 976–977 (1965) 61. Mason, W.P.: Derivation of magnetostriction and anisotropic energies for hexagonal, tetragonal, and orthorhombic crystals. Phys. Rev. 96(2), 302–310 (1954)

    11 Magnetostriction and Magnetoelasticity

    589

    62. Hubert, A., Unger, W., Kranz, J.: Messung der Magnetostriktionskonstanten des Kobalts als Funktion der Temperatur. Zeitschrift für Physik 224(1–3), 148–155 (1969) 63. Wu, R., Freeman, A.J.: First principles determination of magnetostriction in transition metals. J. Appl. Phys. 79, 6209–6212 (1996) 64. Hjortstam, O., Baberschke, K., Wills, J.M., Johansson, B., Eriksson, O.: Magnetic anisotropy and magnetostriction in tetragonal and cubic Ni. Phys. Rev. B 55(22), 15026–15032 (1997) 65. Sander, D.: The magnetic anisotropy and spin reorientation of nanostructures and nanoscale films. J. Phys.: Condens. Matter 16(20), R603 (2004) 66. Sander, D., Pan, W., Ouazi, S., Kirschner, J., Meyer, W., Krause, M., Müller, S., Hammer, L., Heinz, K.: Reversible H-induced switching of the magnetic easy axis in Ni/Cu(001) thin films. Phys. Rev. Lett. 93(24), 247203 (2004) 67. Marcus, P.M., Jona, F.: The strain tensor in a general pseudomorphic epitaxial film. J. Phys. Chem. Solids 55, 1513–1519 (1994) 68. Marcus, P.M., Jona, F.: Strains in epitaxial films: the general case. Phys. Rev. B 51(8), 5263– 5268 (1995) 69. Jona, F., Marcus, P.M.: The importance of strain analysis in studies of pseudomorphic growth. Surf. Rev. Lett. 4, 817–820 (1997) 70. Fähnle, M., Komelj, M.: Nonlinear magnetoelastic coupling coefficients from simultaneous measurements of magnetostrictive stress and anisotropy in epitaxial films. J. Magn. Magn. Mater. 220, L13–L17 (2000) 71. Komelj, M., Fähnle, M.: Nonlinear magnetelastic coupling coefficients in Fe from first principles. J. Magn. Magn. Mater. 220, L8–L12 (2000) 72. Fähnle, M., Komelj, M.: Second-order magnetoelastic effects: from the Dirac equation to the magnetic properties of ultrathin epitaxial films for magnetic thin-film applications. Zeitschrift für Metallkunde/Mater. Res. Adv. Tech. 93(10), 970–973 (2002) 73. Fähnle, M., Komelj, M., Wu, R.Q., Guo, G.Y.: Magnetoelasticity of Fe: possible failure of ab initio electron theory with the local-spin-density approximation and with the generalizedgradient approximation. Phys. Rev. B 65(14), 144436 (2002) 74. Komelj, M., Fähnle, M.: Shear-strain-related nonlinear magnetoelastic properties of epitaxial films. Phys. Rev. B 65(9), 092403 (2002) 75. Komelj, M., Fähnle, M.: Determination of the complete set of second-order magnetoelastic coupling constants on epitaxial films. Phys. Rev. B 65(21), 212410 (2002) 76. Komelj, M., Fähnle, M.: On the magnetoelastic contribution to the magnetic anisotropy of thin epitaxial Permalloy films: an ab initio study, J. Magn. Magn. Mater. 238(2–3), 125–128 (2002) 77. Komelj, M., Fähnle, M.: Nonlinear magnetoelastic behavior of the bcc phases of Co and Ni: importance of third-order contributions for bcc Ni. Phys. Rev. B 73(1), 012404 (2006) 78. Ekreem, N.B., Olabi, A.G., Prescott, T., Rafferty, A., Hashmi, M.S.J.: An overview of magnetostriction, its use and methods to measure these properties. J. Mater. Process. Technol. 191(1–3), 96–101 (2007) 79. Brizzolara, R.A., Colton, R.J.: Magnetostriction measurement using a tunneling-tip. J. Magn. Magn. Mater. 88, 343–350 (1990) 80. Costa, J.L., Nogués, J., Rao, K.V.: Direct measurement of magnetostrictive process in amorphous wires using a scanning tunneling microscope. J. Appl. Phys. 76, 7030 (1994) 81. White, G.K.: Measurement of thermal expansion at low temperatures. Cryogenics 1(3), 151– 158 (1961) 82. Brändli, G., Griessen, R.: Two capacitance dilatometers. Cryogenics 13(5), 299–302 (1973) 83. Tsuya, N., Arai, K.I., Ohmori, K., Shiraga, Y.: Magnetostriction measurement by three terminal capacitance method. Jpn. J. Appl. Phys. 13(11), 1808 (1974) 84. Pott, R., Schefzyk, R.: Apparatus for measuring the thermal expansion of solids between 1.5 and 380 K. J. Phys. E: Sci. Instrum. 16(5), 444 (1983) 85. Hüller, K., Sydow, M., Dietz, G.: Magnetic anisotropy, magnetostriction and intermediate range order in Co-P alloys. J. Magn. Magn. Mater. 53(3), 269–274 (1985)

    590

    D. Sander

    86. Weber, M., Koch, R., Rieder, K.H.: UHV cantilever beam technique for quantitative measurements of magnetization, magnetostriction, and intrinsic stress of ultrathin magnetic films. Phys. Rev. Lett. 73(8), 1166–1169 (1994) 87. Rotter, M., Müller, H., Gratz, E., Doerr, M., Loewenhaupt, M.: A miniature capacitance dilatometer for thermal expansion and magnetostriction. Rev. Sci. Instrum. 69(7), 2742–2746 (1998) 88. Schmiedeshoff, G.M., Lounsbury, A.W., Luna, D.J., Tracy, S.J., Schramm, A.J., Tozer, S.W., Correa, V.F., Hannahs, S.T., Murphy, T.P., Palm, E.C., Lacerda, A.H., Bud’ko, S.L., Canfield, P.C., Smith, J.L., Lashley, J.C., Cooley, J.C.: Versatile and compact capacitive dilatometer. Rev. Sci. Instrum. 77(12), 123907 (2006) 89. Küchler, R., Bauer, T., Brando, M., Steglich, F.: A compact and miniaturized high resolution capacitance dilatometer for measuring thermal expansion and magnetostriction. Rev. Sci. Instrum. 83(9), 095102 (2012) 90. Fawcett, E.: Magnetostriction of paramagnetic transition metals. I. Group 4 – Ti and Zr; Group 5 – V, Nb, and Ta; Group 6 – Mo and W. Phys. Rev. B 2(6), 1604–1613 (1970) 91. Restorff, J.B., Wun-Fogle, M., Hathaway, K.B., Clark, A.E., Lograsso, T.A., Petculescu, G.: Tetragonal magnetostriction and magnetoelastic coupling in Fe-Al, Fe-Ga, Fe-Ge, Fe-Si, FeGa-Al, and Fe-Ga-Ge alloys. J. Appl. Phys. 111(2)023905, 1–12 (2012) 92. Liniers, M., Madurga, V., Vázquez, M., Hernando, A.: Magnetostrictive torsional strain in transverse-field-annealed Metglas o˝ 2605. Phys. Rev. B 31(7), 4425–4432 (1985) 93. Mitra, A., Vázquez, M.: Measurement of the saturation magnetostriction constant of amorphous wires, J. Appl. Phys. 67, 4986–4988 (1990) 94. Yamasaki, J., Ohkubo, Y., Humphrey, F.B.: Magnetostriction measurement of amorphous wires by means os small-angle magnetization rotation. J. Appl. Phys. 67, 5472–5474 (1990) 95. Eastman, D.E.: Magneto-elastic coupling in RbMnF3 . Phys. Rev. 156(2), 645–654 (1967) 96. Smith, A.B.: A ferromagnetic resonance method for measuring magnetostriction constants. Rev. Sci. Instrum. 39, 378–385 (1967) 97. Krams, P., Hillebrands, B., Güntherodt, G., Oepen, H.P.: Magnetic anisotropies of ultrathin Co films on Cu(1 1 13) substrates. Phys. Rev. B 49(5), 3633–3636 (1994) 98. Premper, J., Sander, D., Kirschner, J.: A combined surface stress and magneto-optical Kerr effect measurement setup for temperatures down to 30 K and in fields of up to 0.7 T. Rev. Sci. Instrum. 83(7), 073904 (2012) 99. Dahmen, K., Ibach, H., Sander, D.: A finite element analysis of the bending of crystalline plates due to anisotropic surface and film stress applied to magnetoelasticity. J. Magn. Magn. Mater. 231(1), 74–84 (2001) 100. Dahmen, K., Lehwald, S., Ibach, H.: Bending of crystalline plates under the influence of surface stress — a finite element analysis. Surf. Sci. 446(1–2), 161–173 (2000) 101. Klokholm, E.: The measurement of magnetostriction in ferromagnetic thin films. IEEE Trans. Magn. MAG-12, 819–821 (1976) 102. Tam, A.C., Schroeder, H.: Precise measurement of a magnetostriction coefficient of a thin sofmagnetic film deposited on a substrate. J. Appl. Phys. 64, 5422–5424 (1988) 103. Betz, J., du Trémolet de Lacheisserie, E., Baczewski, L.T.: Magnetoelastic properties of nickel thin films. Appl. Phys. Lett. 68, 132–133 (1996) 104. Koch, R., Weber, M., Henze, E., Rieder. K.H.: Magnetization, magnetostriction and intrinsic stress of polycrystalline Fe films measured by a UHV cantilever beam technique. Surf. Sci. 331–333, 1398–1403 (1995) 105. Wu, R.: First principles determination of magnetic anisotropy and magnetostriction in transition metal alloys. In: Baberschke, K., Nolting, W., Donath, M. (eds.) Band-Ferromagnetism: Ground-State and Finite-Temperature Phenomena, pp. 60–71. Springer, Berlin/Heidelberg (2001) 106. Wu, R.: Theory of magnetocrystalline aniostropy and magnetoelasticity in transition metal system. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism, vol. 1. Wiley, New York (2007)

    11 Magnetostriction and Magnetoelasticity

    591

    107. Chappert, C., Bruno, P.: Magnetic anisotropy in metallic ultrathin films and related experiments on cobalt. J. Appl. Phys. 64, 5736–5741 (1988) 108. Bruno, P., Renard, J.-P.: Magnetic surface anisotropy of transition metal ultrathin films. Appl. Phys. A 49, 499–506 (1989) 109. de Jonge, W.J.M., Bloemen, P.J.H., den Broeder, F.J.A.: Experimental investigation of magnetic anisotropy. In: Bland, J., Heinrich, B. (eds.) Ultrathin Magnetic Structures I, pp. 65–90. Springer, Berlin (1994) 110. 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) 111. Skomski, R., Coey, J.M.D.: Permanent Magnetism. IOP Publishing, Bristol (1999) 112. Baberschke, K.: Anisotropy in magnetism. In: Baberschke, K., Nolting, W., Donath, M. (eds.) Band-Ferromagnetism: Ground-State and Finite-Temperature Phenomena, pp. 27–45. Springer, Berlin/Heidelberg (2001) 113. Burkert, T., Eriksson, O., James, P., Simak, S.I., Johansson, B., Nordström, L.: Calculation of uniaxial magnetic anisotropy energy of tetragonal and trigonal Fe, Co, Ni. Phys. Rev. B 69, 104426–1–104426–7 (2004) 114. Blügel, S., Bihlmayer, G.: Magnetism of low-dimensional systems: theory. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism, vol. 1, pp. 598–639. Wiley, New York (2007) 115. Wu, R., Chen, L.J., Shick, A., Freeman, A.J.: First-principles determinations of magnetocrystalline anisotropy and magnetostriction in bulk and thin-film transition metals. J. Magn. Magn. Mater. 177–181, 1216–1219 (1998) 116. Wang, H., Zhang, Y.N., Wu, R.Q., Sun, L.Z., Xu, D.S., Zhang, Z.D.: Understanding strong magnetostriction in Fe100−x Gax alloys. Sci. Rep. 3, 3521 (2013) 117. Bruno, P.: Tight-binding approach to the orbital magnetic moment and magnetocrystalline anisotropy of transition-metal monolayers. Phys. Rev. B 39(1), 865–868 (1989) 118. Weinberger, P.: Magnetic Anisotropies in Nanostructured Materials. CRC, Boca Raton (2008) 119. Wu, R., Chen, L., Freeman, A.J.: First principles determination of magnetostriction in bulk transition metals and thin films. J. Magn. Magn. Mater. 170(1–2), 103–109 (1997) 120. Cao, J.X., Zhang, Y.N., Ouyang, W.J., Wu, R.Q.: Large magnetostriction of Fe(1−x) Gex and its electronic origin: density functional study. Phys. Rev. B 80, 104414 (2009) 121. Cao, H., Gehring, P.M., Devreugd, C.P., Rodriguez, Rodriguez-Rivera, J.A., Li, J., Viehland, D.: Role of nanoscale precipitates on the enhanced magnetostriction of heat-treated galfenol (Fe1−x Gax ) alloys. Phys. Rev. Lett. 102, 127201 (2009) 122. Zhang, Y.N., Cao, J.X., Barsukov, I., Lindner, J., Krumme, B., Wende, H., Wu, R.Q.: Magnetocrystalline anisotropy of Fe-Si alloys on MgO(001). Phys. Rev. B 81, 144418 (2010) 123. Buck, S., Fähnle, M.: Ab initio calculation of the giant magnetostriction in terbium and erbium. Phys. Rev. B 57, R14044–R14047 (1998) 124. 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, 3272–3289 (1996) 125. Kanamori, J.: Theory of the magnetic properties of ferrous and cobaltous oxides, I. Prog. Theor. Phys. 17(2), 177–196 (1957) 126. Slonczewski, J.C.: Anisotropy and magnetostriction in magnetic oxides. J. Appl. Phys. 32(3), S253–S263 (1961) 127. Greenwald, S.: The antiferromagnetic structure deformations in CoO and MnTe. Acta Crystallogr. 6(5), 396–398 (1953) 128. Bozorth, R.M., Tilden, E.F., Williams, A.J.: Anisotropy and magnetostriction of some ferrites. Phys. Rev. 99, 1788–1798 (1955) 129. Sander, D., Phark, S.-H., Corbetta, M., Fischer, J.A., Oka, H., Kirschner, J.: The impact of structural relaxation on spin polarization and magnetization reversal of individual nano structures studied by spin-polarized scanning tunneling microscopy. J. Phys.: Condens. Matter 26(39), 394008 (2014)

    592

    D. Sander

    130. Brovko, O.O., Bazhanov, D.I., Meyerheim, H.L., Sander, D., Stepanyuk, V.S., Kirschner, J.: Effect of mesoscopic misfit on growth, morphology, electronic properties and magnetism of nanostructures at metallic surfaces. Surf. Sci. Rep. 69(4), 159–195 (2014) 131. Oka, H., Ignatiev, P.A., Wedekind, S., Rodary, G., Niebergall, L., Stepanyuk, V.S., Sander, D., Kirschner, J.: Spin-dependent quantum interference within a single magnetic nanostructure. Science 327(5967), 843–846 (2010) 132. Oka, H., Brovko, O.O., Corbetta, M., Stepanyuk, V.S., Sander, D., Kirschner, J.: Spinpolarized quantum confinement in nanostructures: scanning tunneling microscopy. Rev. Mod. Phys. 86, 1127–1168 (2014) 133. Wu, R., Freeman, A.J.: Spin-orbit induced magnetic phenomena in bulk metals and their surfaces and interfaces. J. Magn. Magn. Mater. 200(1–3), 498–514 (1999) 134. Gutjahr-Löser, T., Sander, D., Kirschner, J.: Magnetoelastic coupling in Ni and Fe monolayers on Cu(001). J. Appl. Phys. 87, 5920–5922 (2000) 135. Gutjahr-Löser, T., Sander, D., Kirschner, J.: Magnetoelastic coupling in Co thin films on W(001). J. Magn. Magn. Mater. 220, L1–L7 (2000) 136. Summers, E.M., Lograsso, T.A., Wun-Fogle, M.: Magnetostriction of binary and ternary Fe–Ga alloys. J. Mater. Sci. 42(23), 9582–9594 (2007) 137. Clark, A.E., Hathaway, K.B., Wun-Fogle, M., Restorff, J.B., Lograsso, T.A., Keppens, V.M., Petculescu, G., Taylor, R.A.: Extraordinary magnetoelasticity and lattice softening in bcc FeGa alloys. J. Appl. Phys. 93(10), 8621–8623 (2003) 138. Wu, R.: Origin of large magnetostriction in FeGa alloys. J. Appl. Phys. 91(10), 7358–7360 (2002) 139. Wang, H., Zhang, Y.N., Yang, T., Zhang, Z.D., Sun, L.Z., Wu, R.Q.: Ab initio studies of the effect of nanoclusters on magnetostriction of Fe1−x Gax alloys. Appl. Phys. Lett. 97(26), 262505 (2010) 140. Khachaturyan, A.G., Viehland, D.: Structurally heterogeneous model of extrinsic magnetostriction for Fe-Ga and similar magnetic alloys: part II. Giant magnetostriction and elastic softening. Metall. Mater. Trans. A 38(13), 2317–2328 (2007) 141. Bhattacharyya, S., Jinschek, J.R., Khachaturyan, A., Cao, H., Li, J.F., Viehland, D.: Nanodispersed DO3 -phase nanostructures observed in magnetostrictive Fe-19 % Ga Galfenol alloys. Phys. Rev. B 77, 104107 (2008) 142. Xing, Q., Lograsso, T.A.: Magnetic domains in magnetostrictive Fe–Ga alloys. Appl. Phys. Lett. 93(18), 182501–1–182501–3 (2008) 143. Cao, H., Gehring, P.M., Devreugd, C.P., Rodriguez-Rivera, J.A., Li, J., Viehland, D.: Role of nanoscale precipitates on the enhanced magnetostriction of heat-treated galfenol (Fe1−x Gax ) alloys. Phys. Rev. Lett. 102, 127201 (2009) 144. Wu, W., Liu, J., Jiang, C., Xu, H.: Giant magnetostriction in Tb-doped Fe83 Ga17 melt-spun ribbons. Appl. Phys. Lett. 103(26), 262403–1–262403–3 (2013) 145. O’Handley, R.C.: Physics of ferromagnetic amorphous alloys. J. Appl. Phys. 62(10), R15– R49 (1987) 146. Jiles: Magnetism and Magnetic Materials. Chapman&Hall, London (1994) 147. Herzer, G.: Modern soft magnets: amorphous and nanocrystalline materials. Acta Mater. 61(3), 718–734 (2013) 148. Fawcett, E.: Magnetostriction of paramagnetic transition metals. II. Group-VIII metals Ru, Rh, Pd, Ir, Pt, and their alloys. Phys. Rev. B 2, 3887–3890 (1970) 149. Kapitza, P.: The study of the magnetic properties of matter in strong magnetic fields. Part V. Experiments on magnetostriction in dia- and para-magnetic substances. Proc. R. Soc. Lond. A: Math. Phys. Eng. Sci. 135(828), 568–600 (1932) 150. Chandrasekhar, B.S., Fawcett, E.: Magnetostriction in metals. Adv. Phys. 20(88), 775–794 (1971) 151. Michenaud, J.P., Heremans, J., Shayegan, M., Haumont, C.: Magnetostriction of bismuth in quantizing magnetic fields. Phys. Rev. B 26, 2552–2559 (1982) 152. Rhyne, J.J., Legvold, S.: Magnetostriction of Tb single crystals. Phys. Rev. 138, A507–A514 (1965)

    11 Magnetostriction and Magnetoelasticity

    593

    153. Legvold, S., Alstad, J., Rhyne, J.: Giant magnetostriction in dysprosium and holmium single crystals. Phys. Rev. Lett. 10, 509–511 (1963) 154. del Moral, A., Arnaudas, J.I.: Magnetostriction of rare-earth random magnetic anisotropy spin glasses. Phys. Rev. B 39, 9453–9466 (1989) 155. del Moral, A., Algarabel, P.A., Arnaudas, J.I., Benito, L., Ciria, M., de la Fuente, C., GarcíaLanda, B., Ibarra, M.R., Marquina, C., Morellón, L., de Teresa, J.M.: Magnetostriction effects. J. Magn. Magn. Mater. 242–245, Part 2(0), 788–796 (2002) 156. Liu, J.H., Jiang, C.B., Xu, H.B.: Giant magnetostrictive materials. Sci. China Technol. Sci. 55(5), 1319–1326 (2012) Dirk Sander received his PhD from RWTH Aachen in 1992 and worked at the IBM T.J. Watson Research Center at Yorktown Heights, New York, before he joined the Max Planck Institute of Microstructure Physics, Halle, in 1993. He received his Habilitation from Halle University in 2000. His research interests include surface and film stress, magnetoelasticity, and spin-dependent phenomena and superconductivity on the nanoscale.

    Magnetoelectrics and Multiferroics Jia-Mian Hu

    12

    and Long-Qing Chen

    Contents Magnetoelectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Multiferroics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Evolving Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electric Field Switching of Magnetization in Multiferroic BiFeO3 : Status and Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Composite Multiferroics and Magnetoelectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Terminology and Exiting Reviews . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mechanisms and Application of Magnetoelectric Effects and Experimental Data of Magnetoelectric Coefficients in Composite Magnetoelectrics . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    596 598 598 602 607 607 609 614

    Abstract

    Magnetoelectrics and multiferroics can possess mutually coupled magnetic and ferroelectric order and thus have been utilized in exploring and designing many novel multifunctional devices such as sensors, transducers, and memories. This chapter presents a brief introduction to the terminology and classification of magnetoelectrics and multiferroics as well as the mechanisms underlying different types of magnetoelectric couplings. Both single-phase and composite materials

    J.-M. Hu () Department of Materials Science and Engineering, University of Wisconsin-Madison, Madison, WI, USA e-mail: [email protected] L.-Q. Chen Materials Research Institute, and Department of Materials Science and Engineering, The Pennsylvania State University, University Park, PA, 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_12

    595

    596

    J.-M. Hu and L.-Q. Chen

    are discussed. Experimental data showing the basic magnetic and ferroelectric properties of different types of single-phase multiferroics are collected. For composite magnetoelectrics, a relatively comprehensive experimental dataset of both the direct and inverse magnetoelectric coefficients is compiled.

    Magnetoelectrics Magnetoelectrics are a family of solid-state materials that enables mutual conversion between magnetic energy and electric energy in the absence of an electric current. This is based on the coupling between the electric polarization P and a magnetic field H, dP i = αijH dH j or between the magnetization M and an electric field E, μ0 dM i = αijE dE j , where αijH and αijE are the direct and inverse magnetoelectric coefficients, respectively, and μ0 is the vacuum permeability. In Landau theory, the magnetoelectric effect in a single-phase material is typically described by introducing an additional energy density (J/m3 ) term −α ij Ei Hj in the total free energy of the system. The magnetoelectric effect can be understood by analogy to other types of coupling effects in functional materials (Fig. 1). For example, piezoelectrics enable mutual conversion between mechanical energy and electric energy in the absence of an electric current, based on the piezoelectric effect – the coupling between the mechanical strain ε and E or between P and the mechanical stress σ . These coupling effects can find applications in devices such as sensors, actuators, memories, and many other devices that rely on solid-state energy

    Fig. 1 Coupling effects in functional materials that permit solid-state energy conversion, showing order parameters (blue) and conjugate fields (yellow), notably electric field E, magnetic field H, and temperature T. Other symbols represent coupling coefficients

    12 Magnetoelectrics and Multiferroics

    597

    conversion. Furthermore, devices based on these coupling effects are generally energy-efficient because the need for an electric current is obviated during energy conversion. The abovementioned P – H and M – E correlations were first proposed by Curie in 1894 based on symmetry considerations [1] and then termed “magnetoelectric” by Debye in 1926 [2]. However, in the 1920s–1950s, all attempts to experimentally observe such effects had failed, and some scientists believed that “no magnetoelectric effect can exist” (see a summary of these early works in a text book by O’Dell [3]). In 1957, Landau and Lifshtiz first [4] pointed out that the magnetoelectric effect is only allowed in time-asymmetric media (i.e., a time reversal transformation t → − t changes the sign of thermodynamic potential of a medium). Such broken time-reversal symmetry can be achieved extrinsically by moving a dielectric medium [5, 6], applying a constant magnetic field to a paramagnetic crystal [7], and appears naturally in a magnetically ordered (ferromagnetic, ferromagnetic, and antiferromagnetic) crystal. In 1960, Dzyaloshinskii [8] first pointed out that the magnetoelectric coupling should exist in the antiferromagnetic Cr2 O3 , which was experimentally confirmed by Astrov [9] and then others [10– 12]. The magnetoelectric effect was soon found in about 80 different magnetically ordered single-phase materials from the 1960s to the early 1970s. Examples include Ti2 O3 [13], Gd2-x Fex O3 (a ferromagnet) [14], a few compounds of boracites (e.g., Ni3 B7 O13 I [15], a ferroelectric ferromagnet), phosphates (e.g., LiCoPO4 [16, 17]), manganites (e.g., Ta2 Mn4 O9 [18]), and the PbFe0.5 Nb0.5 O3 solid solutions [19]. A detailed summary of these early works can be found in an article by Schmid [20]. Also in the early 1970s, the concept of composite magnetoelectrics/multiferroics and corresponding materials emerged, as will be discussed in Sect. “Composite Multiferroics and Magnetoelectrics”. The magnetoelectric effect is generally small in single-phase materials. For E in Cr O single crystal (the subscript z represents the instance, the peak value of αzz 2 3 Cartesian axis) was found to be 4.13 × 10−12 s/m [11], which suggests that applying an electric field of 106 V/m can only flip 4 or 5 spins out of one million spins in the antiferromagnetic lattice [21]. In 1968, Brown (who is best known for his contribution to the theory of micromagnetics [22]) and coworkers [23] proposed an upper bound for the magnetoelectric coefficients α ij in single-phase materials, αij2 ≤ ε0 μ0 εii μjj , where ε0 is the vacuum permittivity; the diagonalized tensors εii and μjj represent the relative permittivity and relative permeability, respectively. This equation naturally leads to the following two conclusions. First, magnetoelectric effects in ferroelectric (or ferromagnetic) materials that have large relative electric permittivity (or magnetic permeability) should also be large. Second, if a material is simultaneously ferroelectric and ferromagnetic, which is one type of “muliferroic” material according to the definition by Schmid in his 1994 paper “Multi-ferroic magnetoelectrics” [24], large magnetoelectric effects may emerge. Table 1 lists the magnetoelectric coefficients of several single-phase magnetoelectrics.

    598

    J.-M. Hu and L.-Q. Chen

    Table 1 Linear and nonlinear magnetoelectric coefficients in some single-phase magnetoelectrics

    Materiala Cr2 O3 (bulk single crystal) Cr2 O3 (thin-film) TbPO4 (bulk single crystal) Ba0.52 Sr2.48 Co2 Fe24 O41 (bulk single crystal) BiFeO3 (mono-domain bulk single crystal) TbMnO3 (bulk single crystal) TbMn2 O5 (bulk single crystal) Ni3 B7 O13 I (bulk single crystal)

    Linear magnetoelectric effect? Yes

    Testing condition M||Ec

    Testing temperature 265 K

    Reference [11]

    (4.6 ± 0.3) × 10−12 Yes 3 × 10−10 Yes

    M||E P||H

    250 K 4.2 K

    [25] [26]

    3.2 × 10−9 @10.5mT 1.1 × 10−10 @ 24 Td 1.1 × 10−9 @ 5 T

    No

    P||H

    305 K

    [27]

    No

    P⊥H

    4.2 K

    [28]

    No

    P⊥H

    9K

    [29]

    2 × 10−9 @ 1.1 T

    No

    P⊥H

    3K

    [30]

    1.6 × 10−9

    Yese

    P⊥H

    46 K

    [15]

    α ij (s/m)b 4.13 × 10−12

    a Bulk

    single crystal and thin-film materials are classified based on their method of growth direct magnetoelectric coefficient αijH = dP i /dH j and the inverse magnetoelectric coefficient αijE = μ0 dM i /dE j have the same unit of second/m in SI units; moreover, αijH = αijE in the case of linear magnetoelectric effect. In the case of nonlinear magnetoelectric effect, αijH and αijE would change with the driving magnetic and electric field, respectively, and we only list the peak values of αijH for these materials along with the testing magnetic field, because the data of nonlinear αijE are not available c The direction that the induced M was recorded along is parallel to the applied electric field E i j (i = j), likewise for the polarization Pi induced by applied magnetic field Hj d A large magnetic field is needed to break the spin cycloid such that the magnetoelectric effect can become appreciable. The data measured on the same sample indicates αijH 1.2%) biaxial compressive strain in thin-film form [63, 64]. Table 2 summarizes the spontaneous/saturation polarization , magnetization, and the onset temperatures for the ferroelectric order (TFE ) and ferromagnetic order (TC ) of these ferroelectric ferromagnets.

    Table 2 Properties of the ferroelectric-ferromagnet type single-phase multiferroics

    Material Ni3 B7 O13 I

    Type I or II multiferroics Type I

    BiMnO3

    Type I

    LaBiMnO3 Type I

    Polarization 0.076 μC cm−2 @46 K 0.15 μC cm−2 @87 Kb N/Ac

    Strained EuTiO3

    10 μC cm−2 @5 Kd

    a Saturating

    Type I

    Magnetization 3.26 μB per Ni atom @ 6 Ka 3.6 μB per Mn atom @5 K

    TFE 60 K

    TC 64 K

    Reference [15, 31, 65]

    450– 490 K

    105 K

    [57, 58, 66, 67]

    2.5 × 105 A m−1 @10 K 3 μB per Eu atom @ 1.8 Ke

    N/A

    90 K

    [61]

    250 K

    4K

    [63]

    magnetic moment polarization c Ferroelectricity demonstrated through piezoresponse force microscopy at 300 K d Spontaneous polarization e Spontaneous magnetic moment. μ : Bohr magneton B b Saturating

    12 Magnetoelectrics and Multiferroics

    601

    Ferroelectric Antiferromagnets Ferroelectric antiferromagnets are the most commonly reported single-phase multiferroics. These materials are grouped by their different origins of ferroelectricity [32, 33], as follows: • Solid solutions where the ferroelectric polarization arises from the 6 s lone pair of the A-site atom (A = Bi, Pb) in the ABO3 structure that distorts the unit cell (viz., lone pair-driven ferroelectricity). Examples include Pb(Fe1/2 Nb1/2 )O3 that was the first discovered material of this kind [68], and BiFeO3 (Sect. “Electric Field Switching of Magnetization in Multiferroic BiFeO3 : Status and Perspective”), the best known [69]. • Compounds where the ferroelectric polarization arises from the polar tilts and rotations of the anionic sublattice (viz., geometrically driven ferroelectricity), including hexagonal (h-) manganites RMnO3 (R = Sc, Y, In, or Dy-Lu, see, e.g., [70]), hexagonal LuFeO3 thin films [71], and some fluorides such as BaMF4 (M = Ni, Mn, Fe, Co, see, e.g., [72–74] and a review [75]). • Compounds where the ferroelectricity polarization is induced by spin ordering (viz., spin-driven ferroelectricity, see reviews [76–78]). Specifically, the polarization can arise from: (i) The noncollinear spin ordering through the inverse Dzyaloshinskii-Moriya interaction (asymmetric exchange coupling) [79, 80]. Examples include Cr2 BeO4 (the first material of this kind to be discovered) [81], orthorhombic (o-) rare earth (RE) manganites REMnO3 with a cycloidal spin spiral (Fig. 3a) (e.g., TbMnO3 [29]), and the CaMn7 O12 with a helical spin spiral (Fig. 3b) [82, 83]. (ii) The collinear spin ordering through Heisenberg (symmetric) exchange coupling [84, 85]. Examples include o-REMn2 O5 (e.g., TbMn2 O5 , the first material of this kind to be discovered [30]), GdMn2 O5 [86], and SmMn2 O5 [87]), and the Ising chain magnet Ca3 Co2-x Mnx O6 [88]. Note that ferroelectric polarization arising from such collinear spin ordering is generally Fig. 3 (a) Schematic of a cycloidal spin spiral, where the spin (orange arrows) rotates around an axis normal to the propagation wave vector Q. The induced polarization P is normal to both the spin rotation axis and the Q. (b) Schematic of a helical spin spiral where the spin rotates around the wave vector Q. The induced polarization P is parallel to the Q

    602

    J.-M. Hu and L.-Q. Chen

    larger than that those from noncollinear spin ordering. For example, the cycloidal spin spiral in o-TbMnO3 can be transformed into collinear spin ordering by applying pressure, leading to an order of magnitude increase in the spontaneous polarization [89]. (iii) The spin-dependent orbital hybridization between the 3d orbitals of magnetic transition metal (Co, Fe, Cr) and the 2p orbitals of ligand (O) via spin-orbit coupling [90]. Examples include triangular-lattice antiferromagnets such as CuFeO2 [91] and CuCrO2 [92] and a staggered antiferromagnet Ba2 CoGe2 O7 [93]. Ferroelectric antiferromagnets with lone pair-driven or geometrically driven ferroelectricity are Type I multiferroics, while those with spin-driven ferroelectricity belong to Type II. Table 3 summarizes the spontaneous/saturation polarization, magnetization, and the onset temperatures for the ferroelectric order (TFE ) and antiferromagnetic order (TAFM ) of some representative materials.

    Ferroelectric Ferrimagnets Ferroelectric ferrimagnets mainly include Fe3 O4 [103, 104], LuFe2 O4 [105], and Pr1-x Cax MnO3 [106]. In these materials, ferroelectricity was thought to arise from charge ordering ([107] is a brief review on this topic), which has been hotly debated. For example, reports have suggested that the LuFe2 O4 , a prototypical example of charge-ordering-driven ferroelectricity [105] with ferrimagnetic ordering, is not ferroelectric [108–110]. However, in a (LuFeO3 )9 /(LuFe2 O4 )1 superlattice, robust ferroelectricity has been demonstrated by out-of-plane piezoreponse force microscopy at room temperature, while an evident magnetic hysteresis loop has been directly recorded [111] at 250 K. Combining high-resolution transmission electron microscopy imaging with first-principles calculations led to the suggestion that the key ingredient for ferroelectricity is the severe rumpling imposed by the neighboring LuFeO3 . It is such rumpling that drives the ferrimagnetic LuFe2 O4 into a simultaneous ferroelectric state. Table 4 summarizes the spontaneous/saturation polarization, magnetization, and the onset temperature for the ferroelectric order (TFE ) and ferrimagnetism (TC ).

    Electric Field Switching of Magnetization in Multiferroic BiFeO3 : Status and Perspective Electric field switching of magnetization in single-phase multiferroics is generally weaker than it is in composite multiferroics/magnetoelectrics, partly because the magnitude of magnetization in these single-phase materials (see Tables 2, 3, and 4) is generally much smaller than that in ferromagnetic materials. However, many of these single-phase multiferroics are excellent test beds for exploring new phenomena and physics. In this section, we will briefly discuss the electric field switching of magnetization in BiFeO3 , which is perhaps the most intensively studied

    12 Magnetoelectrics and Multiferroics

    603

    Table 3 Properties of the ferroelectric-antiferromagnet-type single-phase multiferroics Type I or II multiferroics Type I

    Origin of ferroelectricity 6 s lone-pair

    BiFeO3

    Type I

    6 s lone-pair

    YMnO3

    Type I

    Geometrically driven

    Hexagonal Type I LuFeO3

    Geometrically driven

    BaNiF4

    Type I

    Geometrically driven

    Cr2 BeO4

    Type II

    Spin-driven (noncollinear spin ordering)

    TbMnO3

    Type II

    Spin-driven (cycloidal spin spiral) Spin-driven (noncollinear spin ordering)

    Material Pb(Fe1/2 Nb1/2 )O3

    CaMn7 O12 Type II

    TbMn2 O5 Type II

    Spin-driven (collinear spin ordering)

    Spontaneous polarization 11.5 μC cm−2 @RT in ceramics 100 μC cm−2 @RT in bulk single crytals, 55 μC cm−2 @RT in epitaxial thin films 5.5 μC cm−2 @RT in bulk single crystals 5 μC cm−2 @RT in bulk single crytalsa 6.7 μC cm−2 @RT in bulk single crystals 3 × 10−4 μC cm−2 @7 K @ 0.6 MV/md in ceramics 0.08 μC cm−2 @10 K in bulk single crystals 0.287 μC cm−2 @15 K in bulk single crystals,e 0.024 μC cm−2 @10 K in ceramics 0.04 μC cm−2 @3 K in bulk single crystals

    TFE 380 K

    TAFM 145–160 K (TN )

    Reference [94, 95]

    1098 K

    643 K (TN )

    [69, 96, 97]

    570– 990 K

    70–130 K (TN )

    [98, 99]

    1050 K

    TN ∼ 400 K in film, TWFM ∼ 100 Kb 50 K (TN )

    [71, 100, 101] [72, 75, 102]

    28 K (TN )

    [81]

    1320 K (melts first) 28 K

    27 K

    TN ∼ 41 K, [29] TCYCL ∼ 27 Kc

    90 K

    90 K (TN )

    [82]

    40 K

    40 K (TN )

    [30]

    (continued)

    604

    J.-M. Hu and L.-Q. Chen

    Table 3 (continued)

    Material TbMnO3 under high pressure (> 5 GPa) CuCrO2

    Type I or II multiferroics Type II

    Type II

    Origin of ferroelectricity Spin-driven (collinear spin ordering)

    Spontaneous polarization 1 μC cm−2 @5 K @ 5.2 GPa in bulk single crystals

    Spin-driven (p-d orbital hybridization)

    0.03 μC cm−2 @5 K @ 30 MV/md in bulk single crystals

    TFE 20 K

    TAFM 20 K (TE-AFM )f

    Reference [89]

    24 K

    24 K

    [92]

    a Estimated b Will

    from structural analyses, not through direct measurement experience a second phase transition to weak ferromagnetism via spin canting below 100 K,

    the TWFM c Below T CYCL ,

    the sinusoidal AFM ordering (emerging when TCYCL < T < TN ) will transform into a cycloidal spin spiral ordering. The latter induces ferroelectricity d Thus it does not represent spontaneous polarization, unlike others e It is still under debate whether CaMn O indeed exhibits such a surprisingly large spontaneous 7 12 polarization f Under high pressure (>5 GPa), the sinusoidal AFM ordering will transform into a collinear E-type AFM ordering, instead of transforming into the cycloidal spiral mentioned above (see temperaturepressure phase diagram in Ref. [89]) Table 4 Properties of the ferroelectric-antiferromagnet-type single-phase multiferroics

    Material Fe3 O4

    Type I or II multiferroics Type I

    (LuFeO3 )9 / Type I (LuFe2 O4 )1 b

    Origin of ferroelectricity Charge order Charge order

    Spontaneous polarization 5.0 μC cm−2 @ 4.2 K 6.0 μC cm−2 @ RT

    Spontaneous magnetization TFE TC 4.0 μB per 119–125 Ka 858 K Fe atom @ 125 K 2.0 μB per >700 K 281 K Fe atom @ 50 K

    Reference [103, 112] [111]

    a Ferroelectricity

    is thought to emerge below the Verwey transition temperature TV , below which the manganite transforms from a metal to insulator b The superlattices were grown by reactive-oxide molecular-beam epitaxy on (111) (ZrO2 )0.905 (Y2 O3 )0.095 (or 9.5 mol% yttria-stabilized zirconia, YSZ) substrates

    multiferroic material. Since a famous 2003 Science paper [96] that reports a large spontaneous polarization and an arguably large [113] spontaneous magnetization in BiFeO3 thin films, research efforts into BiFeO3 have remained the center stage for multiferroics and magnetoelectrics. In fact, there exist no less than seven (2006– 2018) comprehensive review articles that specifically discuss BiFeO3 [69, 114–120] including some very comprehensive ones [118–120]. Here we provide a briefing of

    12 Magnetoelectrics and Multiferroics

    605

    some of the seminal research efforts on the electric field control of magnetization in BiFeO3 .

    BiFeO3 Bulk Single Crystals BiFeO3 has a long-range antiferromagnetic spin cycloid with a period of about 64 nm [121, 122] in rhombohedral bulk crystals (Fig. 4a). Canted antiferromagnetic spins give rise a weak magnetic moment Mc via the Dzyaloshinskii-Moriya interaction [125, 126], demonstrated through both first-principles calculations [124] and experiments [127]. The local magnetic moment Mc , however, averages out over one cycloid period; thus the BiFeO3 bulk crystal does not exhibit a macroscopic magnetization. Remarkably, in high-quality single-domain BiFeO3 crystals, it has been experimentally demonstrated [123] that the cycloidal plane (within which the spin rotates, also defined as the magnetic easy plane) contains the spontaneous polarization along one of the 111 directions, the local net magnetic moment Mc , − → and the cycloidal propagation vector Q along one of the 110 directions. Notably, as the polarization is rotated by 71◦ through the application of an electric field, the

    Fig. 4 (a) Schematic of an antiferromagnetic spin cycloid. The canted antiferromagnetic spins (blue and green, representing two sublattices) rotate within the plane defined by the polarization P and the cycloidal propagation vector Q. They also induce a net magnetic moment Mc (orange) which is also within that plane and averaged out to zero over one cycloid period of λ = 64 nm. (b) Schematic of the magnetic easy plane (antiferromagnetic plane) containing the rotating spins (double-headed arrows), the P, and the Q in BiFeO3 bulk crystals. (c) Schematic of Gtype antiferromagnetism (left) and the 3D view of the Mc induced by canted antiferromagnetic spins via Dzyaloshinskii-Moriya interaction. (d) Relationship between the P, the Mc , and the antiferromagnetic vector L (parallel to the magnetic easy plane) in BiFeO3 epitaxial thin films in which the spin cycloid is destroyed by epitaxial strains or/and reduced thickness. (a, b), Redrawn based on Ref. [123]. (c) Redrawn based on Ref. [124]

    606

    J.-M. Hu and L.-Q. Chen

    magnetic easy plane also rotates (Fig. 4b), as demonstrated experimentally by two different research groups [128, 129]. If decorating the BiFeO3 single crystal with a soft magnetic thin film (such as Permalloy), the electric field-induced reorientation of magnetic easy plane can further switch the magnetic anisotropy of the soft magnetic thin film [128].

    Epitaxial BiFeO3 Thin Films The long-range spin cycloid in BiFeO3 can be transformed into homogenous antiferromagnetic ordering by applying a strong (>18 T) magnetic field [28, 130–132]. In epitaxial BiFeO3 thin films, it has long been suggested [133] that the epitaxial strain may destroy the spin cycloid and thus unleash the hidden weak ferromagnetism. Combined experimental and theoretical analyses [134] suggested that large tensile and compressive strain can transform the spin cycloid into homogenous pseudocollinear antiferromagnetic ordering and that a moderate tensile strain can stabilize a new type of spin cycloid with a propagation vector along one of the 110 directions. A relatively complete temperature-strain phase diagram for BiFeO3 can be found in [118]. Furthermore, in BiFeO3 thin film with pseudo-collinear G-type antiferromagnetic ordering (slightly canted spins, see Fig. 4c), where spin cycloid is destroyed by strain or possibly by the reduced thickness [69], first-principles calculations have predicted that the spontaneous polarization is perpendicular to the magnetic easy plane, which contains the canted antiferromagnetic moments MFe1 and MFe2 as well as their corresponding antiferromagnetic vector L = MFe1 -MFe2 and canted weak magnetization Mc = MFe1 + MFe2 (Mc and L are mutually orthogonal), as shown in Fig. 4d. In this case, as polarization is switched by 71◦ or 109◦ , both the magnetic easy plane and the Mc will be switched. Remarkably, the coupling between polarization and the magnetic easy plane has been experimentally demonstrated [135] by simultaneously visualizing the ferroelectric and antiferromagnetic domains using piezoresponse force microscopy and X-ray photoemission electron microscopy (X-PEEM), respectively. In contrast, a direct observation of polarization switchinginduced reorientation of Mc is still lacking. Nevertheless, the Mc reorientation has been indirectly demonstrated by the observation of local magnetization switching in a 2.5-nm-thick polycrystalline Co0.9 Fe0.1 film atop BiFeO3 [136–138], where the local magnetization in the Co0.9 Fe0.1 film and the Mc are locked through atomistic Heisenberg-type exchange coupling [139]. Notably, a full 180◦ reversal of local magnetization in the Co0.9 Fe0.1 film has been directly observed by X-PEEM [138] after the application and removal of an electric field across the Co0.9 Fe0.1 /BiFeO3 heterostructure. This was thought to be induced by a two-step 180◦ reversal of the Mc accompanying a unique two-step polarization switching at the BiFeO3 surface. The ability to electrically switch the magnetization in a magnetic film atop BiFeO3 offers interesting application as novel exchange coupling-mediated electric-write magnetic memories; see a perspective article [140] and a review on magnetoelectric devices for spintronics [141].

    12 Magnetoelectrics and Multiferroics

    607

    A Brief Future Perspective Electric Field Control of Spin Cycloid in BiFeO3 Thin Films: From the Control of Collinear Magnetism to Noncollinear Magnetism Electric field control of spin cycloid (and magnetic easy plane) has been experimentally demonstrated in BiFeO3 bulk crystals [123] but not yet in BiFeO3 thin films, although epitaxial strain-induced switching of spin cycloid in BiFeO3 thin films has been experimentally demonstrated [134]. The coupling among the oxygen octahedral tilts, polarization, and the spin cycloid, which has been discussed with phenomenological approaches [142, 143], awaits experimental investigation. Towards the Control of Magnetism in BiFeO3 Thin Films at the THz Frequency: New Opportunities with Ultrafast Stimuli Existing experimental reports on electric field control of magnetism (including spin cycloid [123, 128, 129], canted magnetization Mc [136–139], and spin waves [144]) in BiFeO3 all use a static or low-frequency electric field. It remains unclear how the magnetism or antiferromagnetism in BiFeO3 will respond to an ultrafast THz electric field, which can be created by applying a femtosecond laser pulse to the BiFeO3 via the photostriction effect [145]. Note that applying femtosecond laser excitation to ferroelectrics has already led to the discovery of several new phenomena that cannot be accessed through static/low-frequency electric fields, including, for example, the polarization oscillation [146, 147] and up to 0.5% transitional strains [148, 149] that are orders of magnitude larger than those obtained through linear inverse piezoelectric effect. Furthermore, in a heterostructure consisting of thin Ni film deposited onto BiFeO3 bulk substrates, it has been experimentally demonstrated [150] that an optically induced strain at the BiFeO3 surface can further modify the magnetic anisotropy of the Ni film, offering new opportunities as a strain-mediated optical control of magnetism.

    Composite Multiferroics and Magnetoelectrics Terminology and Exiting Reviews Composite multiferroics (also “multiferroic heterostructures” or “artificial multiferroics”) integrate magnetic and ferroelectric materials to produce magnetoelectric effects that are absent in either the magnetic or ferroelectric phase. Composite magnetoelectrics, strictly speaking, represent composites integrating magnetic and piezoelectric materials, enabling both the inverse (electric field control of magnetization) and direct (magnetic field control of electric polarization) magnetoelectric effects. However, composites integrating magnetic and plain-dielectric (such as MgO, GdOx ) or single-phase magnetoelectric (such as Cr2 O3 ) materials (see Fig. 5a), in which only the inverse magnetoelectric effect has so far been

    608

    J.-M. Hu and L.-Q. Chen

    Fig. 5 (a) Classifying composite magnetoelectrics based on the functionality of the constituent dielectric material, along with representative dielectric materials. The solid ellipses suggest the corresponding mechanisms of inverse magnetoelectric effects. EC, exchange coupling; FE-AF, ferroelectric antiferromagnet; ME-AF, magnetoelectric antiferromagnet. Redrawn based on Ref. [151] http://creativecommons.org/licenses/by/4.0/. (b) Relationship between composite multiferroics and composite magnetoelectrics by a loose definition, see discussion in the main text

    reported, can also be called as composite magnetoelectrics (also “magnetoelectric heterostructures”) by a loose definition. Thus, we may conclude that composite multiferroics represent a subset of composite magnetoelectrics (Fig. 5b). In contrast to single-phase magnetoelectric and multiferroic materials, magnetoelectric effects in these composite materials have no upper bound set by the geometric mean of the electric permittivity and magnetic permeability. Rather, the magnitudes of the magnetoelectric effects can be tailored by varying the choices of materials and designing the geometry and microstructure of each constituent materials and can reach the level of 10−5 s/m that is five orders of magnitude larger than the record in single-phase magnetoelectrics and multiferroics (i.e., ∼10−10 s/m in magnetoelectric TbPO4 [26, 152]). Furthermore, the magnetoelectric effects typically can emerge at room temperature. Overall, the design flexibility and the ability to achieve large magnetoelectric effects at room temperature make composite multiferroics/magnetoelectrics more attractive for potential device applications. There exist a series of comprehensive review articles on composite multiferroics [151, 153–159], brief perspectives [160–165], a few topical reviews on electric field control of magnetism [117, 166–170], and the magnetoelectric device applications [34, 141, 171–174]. Here, we will briefly discuss the mechanisms of magnetoelectric effects and the experimental data on magnetoelectric coefficients in composite magnetoelectrics as well as their potential device applications. For a more detailed discussion of the field, see Refs. [151, 159].

    12 Magnetoelectrics and Multiferroics

    609

    Mechanisms and Application of Magnetoelectric Effects and Experimental Data of Magnetoelectric Coefficients in Composite Magnetoelectrics Magnetoelectric effects enable a mutual conversion of magnetic and electric energy in the absence of an electric current, while the magnetoelectric coefficient represents a measure on the degree of energy conversion. In general, magnetoelectric effects in composite magnetoelectrics arise from the coupling among the four fundamental degrees of freedom (spin, charge, lattice, orbit) across the magnetic/ferroelectric (dielectric) interface through the exchange of magnetic, electric, and elastic energy or/and mass. From a phenomenological perspective, different mechanisms of magnetoelectric effects indicate different energy conversion pathways. For example, a strain-mediated magnetic field control of electric polarization (direct magnetoelectric effect) in magnetic/piezoelectric composites is achieved through a conversion of magnetic energy to elastic energy (via magnetostrictive effect in the magnetic phase) then to electric energy (via piezoelectric effect in the piezoelectric phase) and vice versa for the corresponding inverse magnetoelectric effect.

    Direct Magnetoelectric Effect The interfacial strain transfer mechanism mentioned above remains the only reported mechanism for a direct magnetoelectric effect in composite magnetoelectrics. Direct magnetoelectric effects are attractive mainly for applications in the form of low-cost and energy-efficient magnetic field sensors [171], also known as “magnetoelectric sensors” in the literature. Of particular interest is the use of magnetoelectric sensors, instead of the currently used superconducting quantum interference device (SQUID) sensors which are cumbersome and expensive, to detect weak (10 f T – 1 pT) and low-frequency (10−2 – 103 Hz) magnetic fields generated from the electrical activities of human organs [174]. This could lead to the development of a series of household devices for biomedical diagnosis, yet it remains a challenging goal because appropriate composite materials are still lacking. Specifically, Fig. 6a, b shows a relatively extensive experimental dataset of direct magnetoelectric voltage coefficient (α HV = E/(μ0 HAC )) that we compiled [151] for bulk and thin-film composite magnetoelectrics of different phase connectivity [175] (see schematics in Fig. 6c). The data of the α HV magnitudes are presented as a function of the frequency (fAC ) of the driving AC magnetic field μ0 HAC (Fig. 6a) and the magnitude of the bias magnetic field μ0 HDC (Fig. 6b). A high α HV under low fAC and small μ0 HDC is desirable for magnetoelectric sensor applications [151, 176]. However, as shown in Fig. 6a, there are no available experimental data of α HV in the frequency range 0 < fAC < 1 Hz. Inverse Magnetoelectric Effect The inverse magnetoelectric effect generally refers to the modulation of magnetism with an electric field, with potential device applications to energy-efficient electric-

    610

    J.-M. Hu and L.-Q. Chen

    Fig. 6 Complied experimental datasets showing the direct magnetoelectric voltage coefficientα HV as a function of (a), the frequency of the driving AC magnetic field, and (b), the magnitude of the bias magnetic field, both of which are necessary for the experimental measurement of α HV . Available data in both bulk and thin-film composites of different phase connectivity are compiled. Adapted from Ref. [151] http://creativecommons.org/licenses/by/4.0/. PC, particulate composites; HC, horizontal composites; VC, vertical composites (see schematics in (c))

    write magnetic-read memories [140, 177–179] and logic [180–184], electric field tunable radiofrequency or microwave devices [158], etc. For example, electric fielddriven magnetization switching in the free layer of a magnetic tunnel junction (MTJ) allows us to electrically toggle the electrical resistance of the entire junction between high and low value, representing the bit information “1” and “0,” respectively. The inverse magnetoelectric effect in composites magnetoelectrics may occur through different mechanisms, depending on the functionality of the constituent dielectric materials. Below we briefly discuss these mechanisms in composite magnetoelectrics with a “plain” dielectric, piezoelectric, ferroelectric, ferroelectric antiferromagnet, or a magnetoelectric antiferromagnet, respectively. We use a layered magnetoelectric (Fig. 7) for discussion. Plain Dielectric (Oxygen Ion Insulator). In the case of a “plain” dielectric with low concentration of ion-conducting defects (such as MgO and HfO2 , see Fig. 5a for representative materials), the inverse magnetoelectric effect typically occurs through the modulation of spin-polarized charge densities (ns ≈ nη) at the magnetic/dielectric interface as voltage is applied. Here n = ε0 εr E is the change of the screened charges densities with ε0 and εr denoting the vacuum and relative permittivity, respectively. The electric field-induced change of interface

    12 Magnetoelectrics and Multiferroics

    611

    Fig. 7 Schematics of inverse magnetoelectric effects in magnetoelectric heterostructures through electrically modulated (a) spin-polarized charge densities, (b) degree of interfacial oxidation via oxygen vacancy accumulation/depletion, (c) strain, and (d) interfacial exchange coupling. In (a), an enhanced (reduced) surface electron densities of the dielectric layer will reduce (enhance) the interfacial electron densities of the magnetic layer through electrostatic interaction, further shifting the Fermi level at the interface region of the magnet to a lower (higher) level, as shown in the top (bottom) panel. In (b), the MOx layer could affect the perpendicular magnetic anisotropy through the modulation of interfacial M-O orbital hybridization or/and imposing an interfacial exchange bias field, the details of which await further clarification. In (c), the electric field-induced strains are transferred to the overlaid magnet and modulate magnetization through magnetoelastic coupling. Different from the other mechanisms discussed here, strain-magnetization coupling is long-range. In (d), only one single antiferromagnetic domain with perpendicular sublattice magnetization (see arrows in the bottom layer) is shown for simplicity, which is also made largely based on a Cr2 O3 based magnetoelectric heterostructure. Details of inverse magnetoelectric effects in BiFeO3 -based or YMnO3 -based heterostructures may vary. (a–d) Reprinted with permission from Ref. [151] http://creativecommons.org/licenses/by/4.0/

    charge densities will shift the Fermi level at the interface, further changing the interfacial spin polarization P = (D↑ − D↓ )/(D↑ + D↓ ), where D↑ (D↓ ) indicates the densities of states for the spin-up (down) electrons, as schematically shown in Fig. 7a. Notably, inverse magnetoelectric effect through such mechanism, which we term as charge densities, can be directly observed [185–188] in a magnetic tunnel junction such as CoFeB(pinned layer)/MgO/CoFeB (free layer) without integrating additional dielectric materials. This is an important advantage for device applications. However, such charge density-mediated inverse magnetoelectric effect is generally very weak (see our compiled experimental dataset in Fig. 8a) and usually causes an electric field-controlled interface magnetic anisotropy of less than 0.1 pJ/m2 per unit electric field (1 V/m), which is too small to switch a magnetization with a sufficiently high thermal stability (>40 kB T) [189].

    612

    J.-M. Hu and L.-Q. Chen

    Fig. 8 (a) Compiled experimental datasets showing the inverse magnetoelectric coefficient α E (the slope) in magnetoelectric heterostructures, where the horizontal and vertical axes represent the driving electric field and the electrically induced change of magnetization, respectively. The data are classified by the corresponding mechanisms of inverse magnetoelectric effects as discussed in Fig. 7, including charge densities (CD), interfacial oxidation (IO), exchange coupling (EC), and strain transfer (ST). (b) Corresponding thickness of the constituent magnetic layer in a chronological order. (a–b) Reprinted with permission from Ref. [151] http://creativecommons.org/ licenses/by/4.0/

    Plain Dielectric (Oxygen Ion Conductor). If the dielectric layer is a good oxygen ion conductor (such as GdOx ) and if the magnetic layer (M = Co, Fe, CoFe) has a perpendicular magnetic anisotropy (PMA) that is sensitive to the degree of interface oxidation (i.e., the x in an interfacial MOx layer), it is then possible to tune the PMA by applying a voltage to tune the O2− concentration at the interface and hence the degree of interface oxidation. The inverse magnetoelectric effect through such mechanism, which we term as interfacial oxidation (see schematic in Fig. 7b), has been observed mainly in Co/GdOx heterostructures [190–193], and it can enable a giant electric field-controlled interface magnetic anisotropy of more than 10 pJ V−1 m−1 , over 100 times larger than that in the case of charge densities mechanism. This can be also seen from the approximately 100 times larger inverse magnetoelectric coefficient as shown in our complied experimental dataset (10−8 s/m vs. 10−10 s/m, see Fig. 8a). However, the speed of the inverse magnetoelectric effect through such an interfacial oxidation mechanism may be low compared to that through the charge density mechanism, because of intrinsically slow oxygen ion migration. In addition, the back-and-forth oxygen-ion migration between the upper and lower surfaces of the dielectric may accelerate dielectric degradation. Thus, the potential device application of such interface oxidation mechanism remains an open question. Piezoelectric or Ferroelectric Dielectric. If the dielectric layer is piezoelectric , magnetism can be tuned, through magnetoelastic coupling, by strain imparted by the piezoelectric layer as voltage is applied (see schematic in Fig. 7c), but it can

    12 Magnetoelectrics and Multiferroics

    613

    also be tuned through the charge density mechanism mentioned previously. If the piezoelectric layer is also ferroelectric, it is possible to obtain larger strains through ferroelectric domain switching or a structural phase transition and larger electrically induced change of interfacial charge densities due to the larger relative permittivity εr in the ferroelectric. Both will lead to larger inverse magnetoelectric effects. Furthermore, strain-mediated electric field-driven full magnetization reversal [194, 195] and unidirectional magnetic domain-wall motion [196] have recently been computationally demonstrated at room temperature, providing exciting application potential for electric-write magnetic-read memories. However, microfabrication of some piezoelectric or ferroelectric materials may not be compatible with Si CMOS (complementary metal-oxide-semiconductor) processing, and the fatigue problem of ferroelectrics is also a potential issue. Ferroelectric (or Magnetoelectric) Antiferromagnet. If the ferroelectric is also antiferromagnetic such as BiFeO3 , the modulation of magnetism in the magnetic layer can also be achieved, through Heisenberg-type exchange coupling, by electrically tuning the antiferromagnetic order through reorientation of the magnetic easy plane or/and the canted magnetization (see details in Sect. “Electric Field Switching of Magnetization in Multiferroic BiFeO3 : Status and Perspective”). Notably, exchange coupling-mediated electric field-driven 180◦ net magnetization reversal has been experimentally demonstrated at room temperature [138]. However, the high leakage current of BiFeO3 as well as its incompatibility with Si CMOS processing in terms of microfabrication (e.g., specific oxide substrates must be used for film deposition) undermines the application potential of BiFeO3 -based magnetoelectric heterostructures. Such an exchange coupling mechanism has also been observed in Cr2 O3 -based magnetoelectric heterostructures at room temperature [197]. Cr2 O3 is a magnetoelectric antiferromagnet, where the antiferromagnetic domains as well as the uncompensated surface magnetic moment can be reversed by applying an electric field (see schematic in Fig. 7d). Thus, the interfacial magnetic moments in the overlaid magnet can be tuned through exchange coupling. Cr2 O3 has excellent dielectric properties with high dielectric breakdown field of 109 V/m at room temperature, which is an advantage for potential device application, although it remains a challenge to achieve such dielectric properties in Cr2 O3 thin films. Moreover, the electric field switching of antiferromagnetic domains in Cr2 O3 requires simultaneous application of a static magnetic field to lift the timereversal symmetry. This will be a potential disadvantage for high-density device applications, as it could complicate the device design by increasing the noise level and possibly causing cross talk among neighboring device units. Among the different mechanisms discussed above, only the strain-magnetization coupling is long-range, while all the others are interface effects, consistent with our complied experimental dataset showing the thickness of the constituent magnetic layer in the case of different mechanisms (see Fig. 8b). Such long-range strainmagnetization coupling partially leads to the fact that strain-mediated inverse magnetoelectric effect is generally larger than the other types of effects, consistent with the experimental data shown in Fig. 8a. For more detailed discussions on these different mechanisms as well as remaining open questions, see He et al. [151].

    614

    J.-M. Hu and L.-Q. Chen

    Acknowledgments This work was supported by a start-up fund from the University of WisconsinMadison (J.-M.H.) and partially by the National Science Foundation under the grant no. DMR1744213 (Chen) and partially by the Army Research Office under the grant number W911NF-171-0462 (J.-M.H. and L.-Q.C.). The authors acknowledge Xin Zou for helping draw some of the schematics.

    References 1. Curie, P.: Sur la symétrie dans les phénomènes physiques, symétrie d’un champ électrique et d’un champ magnétique. J. Phys. Theor. Appl. 3(1), 393–415 (1894) 2. Debye, P.: Bemerkung zu einigen neuen Versuchen über einen magneto-elektrischen Richteffekt. Z. Phys. 36(4), 300–301 (1926) 3. O’Dell, T.H.: The Electrodynamics of Magneto-Electric Media. North-Holland Publishing Company, Amsterdam (1970) 4. Landau, L.D., Lifshitz, E.M.: Course of Theoretical Physics: Electrodynamics of Continuous Media. Pergamon Press, Oxford (1957).(English Transl. in 1960) 5. Röntgen, W.C.: Ueber die durch Bewegung eines im homogenen electrischen Felde befindlichen Dielectricums hervorgerufene electrodynamische Kraft. Ann. Phys. 271(10), 264–270 (1888) 6. Wilson, H.A.: On the electric effect of rotating a dielectric in a magnetic field. Philos. Trans. R. Soc. Lond. A. 204, 121–137 (1905) 7. Hou, S.L., Bloembergen, N.: Paramagnetoelectric effects in NiSO4 ·H2 O. Phys. Rev. 138(4A), A1218–A1226 (1965) 8. Dzyaloshinskii, I.: On the magneto-electrical effect in antiferromagnets. Sov. Phys. JETP. 10(3), 628–629 (1960) 9. Astrov, D.: The magnetoelectric effect in antiferromagnetics. Sov. Phys. JETP. 11(3), 708– 709 (1960) 10. Folen, V.J., Rado, G.T., Stalder, E.W.: Anisotropy of the magnetoelectric effect in Cr2 O3 . Phys. Rev. Lett. 6(11), 607–608 (1961) 11. Rado, G.T., Folen, V.J.: Observation of the magnetically induced magnetoelectric effect and evidence for antiferromagnetic domains. Phys. Rev. Lett. 7(8), 310–311 (1961) 12. Shtrikman, S., Treves, D.: Observation of the magnetoelectric effect in Cr2 O3 powders. Phys. Rev. 130(3), 986–988 (1963) 13. Al’Shin, B., Astrov, D.: Magnetoelectric effect in titanium oxide Ti2 O3 . Sov. Phys. JETP. 17(4), 809–811 (1963) 14. Rado, G.T.: Observation and possible mechanisms of magnetoelectric effects in a ferromagnet. Phys. Rev. Lett. 13(10), 335–337 (1964) 15. Ascher, E., Rieder, H., Schmid, H., Stössel, H.: Some properties of ferromagnetoelectric nickel-iodine boracite, Ni3 B7 O13 I. J. Appl. Phys. 37(3), 1404–1405 (1966) 16. Mercier, M., Gareyte, J., Bertaut, E.: Une nouvelle famille de corps magnetoelectriquesLiMPO4 (M= Mn, Co, Ni). Comptes Rendus. 264(13), 979 (1967) 17. Rivera, J.-P.: The linear magnetoelectric effect in LiCoPO4 revisited. Ferroelectrics. 161(1), 147–164 (1994) 18. Fischer, E., Gorodetsky, G., Hornreich, R.M.: A new family of magnetoelectric materials: A2 M4 O9 (A=Ta, Nb; M=Mn, Co). Solid State Commun. 10(12), 1127–1132 (1972) 19. Watanabe, T., Kohn, K.: Magnetoelectric effect and low temperature transition of PbFe0. 5Nb0. 5O3 single crystal. Phase Transit. Multinatl. J. 15(1), 57–68 (1989) 20. Schmid, H.: On a magnetoelectric classification of materials. Int. J. Magnetism. 4(4), 337–361 (1973) 21. Fiebig, M.: Revival of the magnetoelectric effect. J. Phys. D. Appl. Phys. 38(8), R123–R152 (2005)

    12 Magnetoelectrics and Multiferroics

    615

    22. Brown, W.F.: Micromagnetics: Interscience Tracts on Physics and Astronomy, 18. J. Wiley, New York, London (1963) 23. Brown, W.F., Hornreich, R.M., Shtrikman, S.: Upper bound on the magnetoelectric susceptibility. Phys. Rev. 168(2), 574–577 (1968) 24. Schmid, H.: Multi-ferroic magnetoelectrics. Ferroelectrics. 162(1–4), 317–338 (1994) 25. Borisov, P., Ashida, T., Nozaki, T., Sahashi, M., Lederman, D.: Magnetoelectric properties of 500-nm Cr2 O3 films. Phys. Rev. B. 93(17), 174415 (2016) 26. Rado, G.T., Ferrari, J.M., Maisch, W.G.: Magnetoelectric susceptibility and magnetic symmetry of magnetoelectrically annealed TbPO4 . Phys. Rev. B. 29(7), 4041–4048 (1984) 27. Chun, S.H., Chai, Y.S., Jeon, B.-G., Kim, H.J., Oh, Y.S., Kim, I., Kim, H., Jeon, B.J., Haam, S.Y., Park, J.-Y., Lee, S.H., Chung, J.-H., Park, J.-H., Kim, K.H.: Electric field control of nonvolatile four-state magnetization at room temperature. Phys. Rev. Lett. 108(17), 177201 (2012) 28. Tokunaga, M., Akaki, M., Ito, T., Miyahara, S., Miyake, A., Kuwahara, H., Furukawa, N.: Magnetic control of transverse electric polarization in BiFeO3 . Nat. Commun. 6, 5878 (2015) 29. Kimura, T., Goto, T., Shintani, H., Ishizaka, K., Arima, T., Tokura, Y.: Magnetic control of ferroelectric polarization. Nature. 426(6962), 55–58 (2003) 30. Hur, N., Park, S., Sharma, P.A., Ahn, J.S., Guha, S., Cheong, S.W.: Electric polarization reversal and memory in a multiferroic material induced by magnetic fields. Nature. 429(6990), 392–395 (2004) 31. Eerenstein, W., Mathur, N.D., Scott, J.F.: Multiferroic and magnetoelectric materials. Nature. 442(7104), 759–765 (2006) 32. Dong, S., Liu, J.-M., Cheong, S.-W., Ren, Z.: Multiferroic materials and magnetoelectric physics: symmetry, entanglement, excitation, and topology. Adv. Phys. 64(5–6), 519–626 (2015) 33. Fiebig, M., Lottermoser, T., Meier, D., Trassin, M.: The evolution of multiferroics. Nat. Rev. Mater. 1, 16046 (2016) 34. Scott, J.F.: Applications of magnetoelectrics. J. Mater. Chem. 22(11), 4567–4574 (2012) 35. Aizu, K.: Possible species of ferromagnetic, ferroelectric, and ferroelastic crystals. Phys. Rev. B. 2(3), 754–772 (1970) 36. Aizu, K.: Possible species of “ferroelastic” crystals and of simultaneously ferroelectric and ferroelastic crystals. J. Phys. Soc. Jpn. 27(2), 387–396 (1969) 37. Newnham, R.E.: Domains in minerals. Am. Miner. 59(9–10), 906–918 (1974) 38. Cheong, S.W., Mostovoy, M.: Multiferroics: a magnetic twist for ferroelectricity. Nat. Mater. 6(1), 13–20 (2007) 39. Ramesh, R., Spaldin, N.A.: Multiferroics: progress and prospects in thin films. Nat. Mater. 6(1), 21–29 (2007) 40. Tokura, Y.: Multiferroics – toward strong coupling between magnetization and polarization in a solid. J. Magn. Magn. Mater. 310(2), 1145–1150 (2007) 41. Martin, L., Crane, S.P., Chu, Y.H., Holcomb, M.B., Gajek, M., Huijben, M., Yang, C.H., Balke, N., Ramesh, R.: Multiferroics and magnetoelectrics: thin films and nanostructures. J. Phys. Condens. Matter. 20(43), 434220 (2008) 42. Spaldin, N.A., Cheong, S.-W., Ramesh, R.: Multiferroics: past, present, and future. Phys. Today. 63(10), 38–43 (2010) 43. Lawes, G., Srinivasan, G.: Introduction to magnetoelectric coupling and multiferroic films. J. Phys. D. Appl. Phys. 44(24), 243001 (2011) 44. Spaldin, N.A., Ramesh, R.: Advances in magnetoelectric multiferroics. Nat. Mater. 18(3), 203–212 (2019) 45. Khomskii, D.: Classifying multiferroics: mechanisms and effects. Physics. 2(20) (2009) 46. Kimura, T., Lawes, G., Ramirez, A.P.: Electric polarization rotation in a hexaferrite with longwavelength magnetic structures. Phys. Rev. Lett. 94(13), 137201 (2005) 47. Ishiwata, S., Taguchi, Y., Murakawa, H., Onose, Y., Tokura, Y.: Low-magnetic-field control of electric polarization vector in a helimagnet. Science. 319(5870), 1643–1646 (2008)

    616

    J.-M. Hu and L.-Q. Chen

    48. Chun, S.H., Chai, Y.S., Oh, Y.S., Jaiswal-Nagar, D., Haam, S.Y., Kim, I., Lee, B., Nam, D.H., Ko, K.-T., Park, J.-H., Chung, J.-H., Kim, K.H.: Realization of giant magnetoelectricity in helimagnets. Phys. Rev. Lett. 104(3), 037204 (2010) 49. Kitagawa, Y., Hiraoka, Y., Honda, T., Ishikura, T., Nakamura, H., Kimura, T.: Low-field magnetoelectric effect at room temperature. Nat. Mater. 9(10), 797–802 (2010) 50. Wang, F., Zou, T., Yan, L.-Q., Liu, Y., Sun, Y.: Low magnetic field reversal of electric polarization in a Y-type hexaferrite. Appl. Phys. Lett. 100(12), 122901 (2012) 51. Okumura, K., Haruki, K., Ishikura, T., Hirose, S., Kimura, T.: Multilevel magnetization switching by electric field in c-axis oriented polycrystalline Z-type hexaferrite. Appl. Phys. Lett. 103(3), 032906 (2013) 52. Chai, Y.S., Kwon, S., Chun, S.H., Kim, I., Jeon, B.-G., Kim, K.H., Lee, S.: Electrical control of large magnetization reversal in a helimagnet. Nat. Commun. 5, 4208 (2014) 53. Hirose, S., Haruki, K., Ando, A., Kimura, T.: Mutual control of magnetization and electrical polarization by electric and magnetic fields at room temperature in Y-type BaSrCo2−x Znx Fe11 AlO22 ceramics. Appl. Phys. Lett. 104(2), 022907 (2014) 54. Shen, S., Chai, Y., Sun, Y.: Nonvolatile electric-field control of magnetization in a Y-type hexaferrite. Sci. Rep. 5, 8254 (2015) 55. Kimura, T.: Magnetoelectric hexaferrites. Annu. Rev. Condens. Matter Phys. 3(1), 93–110 (2012) 56. Hill, N.A.: Why are there so few magnetic ferroelectrics? J. Phys. Chem. B. 104(29), 6694– 6709 (2000) 57. Sugawara, F., Iiida, S., Syono, Y., Akimoto, S.-I.: Magnetic properties and crystal distortions of BiMnO3 and BiCrO3 . J. Phys. Soc. Jpn. 25(6), 1553–1558 (1968) 58. Moreira dos Santos, A., Parashar, S., Raju, A.R., Zhao, Y.S., Cheetham, A.K., Rao, C.N.R.: Evidence for the likely occurrence of magnetoferroelectricity in the simple perovskite, BiMnO3 . Solid State Commun. 122(1), 49–52 (2002) 59. Kimura, T., Kawamoto, S., Yamada, I., Azuma, M., Takano, M., Tokura, Y.: Magnetocapacitance effect in multiferroic BiMnO3 . Phys. Rev. B. 67(18), 180401 (2003) 60. Eerenstein, W., Morrison, F.D., Scott, J.F., Mathur, N.D.: Growth of highly resistive BiMnO3 films. Appl. Phys. Lett. 87(10), 101906 (2005) 61. Gajek, M., Bibes, M., Fusil, S., Bouzehouane, K., Fontcuberta, J., Barthelemy, A.E., Fert, A.: Tunnel junctions with multiferroic barriers. Nat. Mater. 6(4), 296–302 (2007) 62. Shvartsman, V.V., Borisov, P., Kleemann, W., Kamba, S., Katsufuji, T.: Large off-diagonal magnetoelectric coupling in the quantum paraelectric antiferromagnet EuTiO3 . Phys. Rev. B. 81(6), 064426 (2010) 63. Lee, J.H., Fang, L., Vlahos, E., Ke, X., Jung, Y.W., Kourkoutis, L.F., Kim, J.-W., Ryan, P.J., Heeg, T., Roeckerath, M., Goian, V., Bernhagen, M., Uecker, R., Hammel, P.C., Rabe, K.M., Kamba, S., Schubert, J., Freeland, J.W., Muller, D.A., Fennie, C.J., Schiffer, P., Gopalan, V., Johnston-Halperin, E., Schlom, D.G.: A strong ferroelectric ferromagnet created by means of spin-lattice coupling. Nature. 466(7309), 954–958 (2010) 64. Fennie, C.J., Rabe, K.M.: Magnetic and electric phase control in epitaxial EuTiO3 from first principles. Phys. Rev. Lett. 97(26), 267602 (2006) 65. von Wartburg, W.: The magnetic structure of magnetoelectric nickel–iodine boracite Ni3 B7 O13 I. Phys. Status Solidi A. 21(2), 557–568 (1974) 66. Chiba, H., Atou, T., Syono, Y.: Magnetic and electrical properties of Bi1−x Srx MnO3 : holedoping effect on ferromagnetic perovskite BiMnO3 . J. Solid State Chem. 132(1), 139–143 (1997) 67. Chi, Z.H., Xiao, C.J., Feng, S.M., Li, F.Y., Jin, C.Q., Wang, X.H., Chen, R.Z., Li, L.T.: Manifestation of ferroelectromagnetism in multiferroic BiMnO3 . J. Appl. Phys. 98(10), 103519 (2005) 68. Smolenskii, G., Agranovskaya, A., Popov, S., Isupov, V.: New ferroelectrics of complex composition. 2. Pb2 Fe3+ NbO6 and Pb2 YbNbO6 . Soviet physics Tech. Phys. 3, 1981–1982 (1958)

    12 Magnetoelectrics and Multiferroics

    617

    69. Catalan, G., Scott, J.F.: Physics and applications of bismuth ferrite. Adv. Mater. 21(24), 2463– 2485 (2009) 70. Fiebig, M., Lottermoser, T., Frohlich, D., Goltsev, A.V., Pisarev, R.V.: Observation of coupled magnetic and electric domains. Nature. 419(6909), 818–820 (2002) 71. Wang, W., Zhao, J., Wang, W., Gai, Z., Balke, N., Chi, M., Lee, H.N., Tian, W., Zhu, L., Cheng, X., Keavney, D.J., Yi, J., Ward, T.Z., Snijders, P.C., Christen, H.M., Wu, W., Shen, J., Xu, X.: Room-temperature multiferroic hexagonal LuFeO3 films. Phys. Rev. Lett. 110(23), 237601 (2013) 72. Eibschütz, M., Guggenheim, H.J., Wemple, S.H., Camlibel, I., DiDomenico, M.: Ferroelectricity in BaM2+ F4 . Phys. Lett. A. 29(7), 409–410 (1969) 73. Fox, D.L., Scott, J.F.: Ferroelectrically induced ferromagnetism. J. Phys. C: Solid State Phys. 10(11), L329 (1977) 74. Fox, D.L., Tilley, D.R., Scott, J.F., Guggenheim, H.J.: Magnetoelectric phenomena in BaMnF4 and BaMn0.99 Co0.01 F4 . Phys. Rev. B. 21(7), 2926–2936 (1980) 75. Scott, J.F., Blinc, R.: Multiferroic magnetoelectric fluorides: why are there so many magnetic ferroelectrics? J. Phys. Condens. Matter. 23(11) (2011) 76. Yoshinori, T., Shinichiro, S., Naoto, N.: Multiferroics of spin origin. Rep. Prog. Phys. 77(7), 076501 (2014) 77. Tokura, Y., Seki, S.: Multiferroics with spiral spin orders. Adv. Mater. 22(14), 1554–1565 (2010) 78. Kimura, T.: Spiral magnets as magnetoelectrics. Annu. Rev. Mater. Res. 37(1), 387–413 (2007) 79. Katsura, H., Nagaosa, N., Balatsky, A.V.: Spin current and magnetoelectric effect in noncollinear magnets. Phys. Rev. Lett. 95(5), 057205 (2005) 80. Mostovoy, M.: Ferroelectricity in spiral magnets. Phys. Rev. Lett. 96(6), 067601 (2006) 81. Newnham, R.E., Kramer, J.J., Schulze, W.A., Cross, L.E.: Magnetoferroelectricity in Cr2 BeO4 . J. Appl. Phys. 49(12), 6088–6091 (1978) 82. Johnson, R.D., Chapon, L.C., Khalyavin, D.D., Manuel, P., Radaelli, P.G., Martin, C.: Giant improper ferroelectricity in the ferroaxial magnet CaMn7 O12 . Phys. Rev. Lett. 108(6), 067201 (2012) 83. Terada, N., Glazkova, Y.S., Belik, A.A.: Differentiation between ferroelectricity and thermally stimulated current in pyrocurrent measurements of multiferroic MMn7 O12 (M=Ca, Sr, Cd, Pb). Phys. Rev. B. 93(15), 155127 (2016) 84. Kagomiya, I., Matsumoto, S., Kohn, K., Fukuda, Y., Shoubu, T., Kimura, H., Noda, Y., Ikeda, N.: Lattice distortion at ferroelectric transition of YMn 2 O 5. Ferroelectrics. 286(1), 167–174 (2003) 85. Chapon, L.C., Blake, G.R., Gutmann, M.J., Park, S., Hur, N., Radaelli, P.G., Cheong, S.W.: Structural anomalies and multiferroic behavior in magnetically frustrated TbMn2 O5 . Phys. Rev. Lett. 93(17), 177402 (2004) 86. Lee, N., Vecchini, C., Choi, Y.J., Chapon, L.C., Bombardi, A., Radaelli, P.G., Cheong, S.W.: Giant tunability of ferroelectric polarization in GdMn2 O5 . Phys. Rev. Lett. 110(13), 137203 (2013) 87. Yahia, G., Damay, F., Chattopadhyay, S., Balédent, V., Peng, W., Elkaim, E., Whitaker, M., Greenblatt, M., Lepetit, M.B., Foury-Leylekian, P.: Recognition of exchange striction as the origin of magnetoelectric coupling in multiferroics. Phys. Rev. B. 95(18), 184112 (2017) 88. Choi, Y.J., Yi, H.T., Lee, S., Huang, Q., Kiryukhin, V., Cheong, S.W.: Ferroelectricity in an ising chain magnet. Phys. Rev. Lett. 100(4), 047601 (2008) 89. Aoyama, T., Yamauchi, K., Iyama, A., Picozzi, S., Shimizu, K., Kimura, T.: Giant spin-driven ferroelectric polarization in TbMnO3 under high pressure. Nat. Commun. 5, 4927 (2014) 90. Arima, T.-h.: Ferroelectricity induced by proper-screw type magnetic order. J. Phys. Soc. Jpn. 76(7), 073702 (2007) 91. Kimura, T., Lashley, J.C., Ramirez, A.P.: Inversion-symmetry breaking in the noncollinear magnetic phase of the triangular-lattice antiferromagnet CuFeO2 . Phys. Rev. B. 73(22), 220401 (2006)

    618

    J.-M. Hu and L.-Q. Chen

    92. Seki, S., Onose, Y., Tokura, Y.: Spin-driven ferroelectricity in triangular lattice antiferromagnets ACrO2 (A=Cu, Ag, Li, or Na). Phys. Rev. Lett. 101(6), 067204 (2008) 93. Murakawa, H., Onose, Y., Miyahara, S., Furukawa, N., Tokura, Y.: Ferroelectricity induced by spin-dependent metal-ligand hybridization in Ba2 CoGe2 O7 . Phys. Rev. Lett. 105(13), 137202 (2010) 94. Gao, X.S., Chen, X.Y., Yin, J., Wu, J., Liu, Z.G., Wang, M.: Ferroelectric and dielectric properties of ferroelectromagnet Pb(Fe1/2 Nb1/2 )O3 ceramics and thin films. J. Mater. Sci. 35(21), 5421–5425 (2000) 95. Howes, B., Pelizzone, M., Fischer, P., Tabares-munoz, C., Rivera, J.-P., Schmid, H.: Characterisation of some magnetic and magnetoelectric properties of ferroelectric Pb(Fe1/2 Nb1/2 )O3 . Ferroelectrics. 54(1), 317–320 (1984) 96. Wang, J., Neaton, J.B., Zheng, H., Nagarajan, V., Ogale, S.B., Liu, B., Viehland, D., Vaithyanathan, V., Schlom, D.G., Waghmare, U.V., Spaldin, N.A., Rabe, K.M., Wuttig, M., Ramesh, R.: Epitaxial BiFeO3 multiferroic thin film heterostructures. Science. 299(5613), 1719–1722 (2003) 97. Lebeugle, D., Colson, D., Forget, A., Viret, M.: Very large spontaneous electric polarization in BiFeO3 single crystals at room temperature and its evolution under cycling fields. Appl. Phys. Lett. 91(2), 022907–022903 (2007) 98. Yakel Jr., H.L., Koehler, W.C., Bertaut, E.F., Forrat, E.F.: On the crystal structure of the manganese(III) trioxides of the heavy lanthanides and yttrium. Acta Crystallogr. 16(10), 957– 962 (1963) 99. Bertaut, E.F., Mercier, M.: Structure magnetique de MnYO3 . Phys. Lett. 5(1), 27–29 (1963) 100. Magome, E., Moriyoshi, C., Kuroiwa, Y., Masuno, A., Inoue, H.: Noncentrosymmetric structure of LuFeO3 in metastable state. Jpn. J. Appl. Phys. 49(9S), 09ME06 (2010) 101. Disseler, S.M., Borchers, J.A., Brooks, C.M., Mundy, J.A., Moyer, J.A., Hillsberry, D.A., Thies, E.L., Tenne, D.A., Heron, J., Holtz, M.E., Clarkson, J.D., Stiehl, G.M., Schiffer, P., Muller, D.A., Schlom, D.G., Ratcliff, W.D.: Magnetic structure and ordering of multiferroic hexagonal LuFeO3 . Phys. Rev. Lett. 114(21), 217602 (2015) 102. Cox, D.E., Eibschütz, M., Guggenheim, H.J., Holmes, L.: Neutron diffraction study of the magnetic structure of BaNiF4 . J. Appl. Phys. 41(3), 943–945 (1970) 103. Kato, K., Iida, S., Yanai, K., Mizushima, K.: Ferrimagnetic ferroelectricity of Fe3 O4 . J. Magn. Magn. Mater. 31, 783–784 (1983) 104. Alexe, M., Ziese, M., Hesse, D., Esquinazi, P., Yamauchi, K., Fukushima, T., Picozzi, S., Gösele, U.: Ferroelectric switching in multiferroic magnetite (Fe3 O4 ) thin films. Adv. Mater. 21(44), 4452–4455 (2009) 105. Ikeda, N., Ohsumi, H., Ohwada, K., Ishii, K., Inami, T., Kakurai, K., Murakami, Y., Yoshii, K., Mori, S., Horibe, Y., Kito, H.: Ferroelectricity from iron valence ordering in the chargefrustrated system LuFe2 O4 . Nature. 436(7054), 1136–1138 (2005) 106. Lopes, A.M.L., Araújo, J.P., Amaral, V.S., Correia, J.G., Tomioka, Y., Tokura, Y.: New phase transition in the Pr1−x Cax MnO3 system: evidence for electrical polarization in charge ordered manganites. Phys. Rev. Lett. 100(15), 155702 (2008) 107. van den Brink, J., Khomskii, D.I.: Multiferroicity due to charge ordering. J. Phys. Condens. Matter. 20(43), 434217 (2008) 108. de Groot, J., Mueller, T., Rosenberg, R.A., Keavney, D.J., Islam, Z., Kim, J.W., Angst, M.: Charge order in LuFe2 O4 : an unlikely route to ferroelectricity. Phys. Rev. Lett. 108(18), 187601 (2012) 109. Niermann, D., Waschkowski, F., de Groot, J., Angst, M., Hemberger, J.: Dielectric properties of charge-ordered LuFe2 O4 revisited: the apparent influence of contacts. Phys. Rev. Lett. 109(1), 016405 (2012) 110. Lafuerza, S., García, J., Subías, G., Blasco, J., Conder, K., Pomjakushina, E.: Intrinsic electrical properties of LuFe2 O4 . Phys. Rev. B. 88(8), 085130 (2013) 111. Mundy, J.A., Brooks, C.M., Holtz, M.E., Moyer, J.A., Das, H., Rébola, A.F., Heron, J.T., Clarkson, J.D., Disseler, S.M., Liu, Z., Farhan, A., Held, R., Hovden, R., Padgett, E., Mao, Q., Paik, H., Misra, R., Kourkoutis, L.F., Arenholz, E., Scholl, A., Borchers, J.A., Ratcliff, W.D.,

    12 Magnetoelectrics and Multiferroics

    619

    Ramesh, R., Fennie, C.J., Schiffer, P., Muller, D.A., Schlom, D.G.: Atomically engineered ferroic layers yield a room-temperature magnetoelectric multiferroic. Nature. 537(7621), 523–527 (2016) 112. Liu, M., Hoffman, J., Wang, J., Zhang, J., Nelson-Cheeseman, B., Bhattacharya, A.: Non-volatile ferroelastic switching of the Verwey transition and resistivity of epitaxial Fe3 O4 /PMN-PT (011). Sci. Rep. 3, 1876 (2013) 113. Eerenstein, W., Morrison, F.D., Dho, J., Blamire, M.G., Scott, J.F., Mathur, N.D.: Comment on “Epitaxial BiFeO3 multiferroic thin film heterostructures”. Science. 307(5713), 1203 (2005) 114. Zavaliche, F., Yang, S.Y., Zhao, T., Chu, Y.H., Cruz, M.P., Eom, C.B., Ramesh, R.: Multiferroic BiFeO3 films: domain structure and polarization dynamics. Phase Transit. 79(12), 991–1017 (2006) 115. Chu, Y.H., Martin, L.W., Zhan, Q., Yang, P.L., Cruz, M.P., Lee, K., Barry, M., Yang, S.Y., Ramesh, R.: Epitaxial multiferroic BiFeO3 thin films: progress and future directions. Ferroelectrics. 354(1), 167–177 (2007) 116. Silva, J., Reyes, A., Esparza, H., Camacho, H., Fuentes, L.: BiFeO3 : a review on synthesis, doping and crystal structure. Integr. Ferroelectr. 126(1), 47–59 (2011) 117. Heron, J.T., Schlom, D.G., Ramesh, R.: Electric field control of magnetism using BiFeO3 based heterostructures. Appl. Phys. Rev. 1(2), 021303 (2014) 118. Sando, D., Barthélémy, A., Bibes, M.: BiFeO3 epitaxial thin films and devices: past, present and future. J. Phys. Condens. Matter. 26(47), 473201 (2014) 119. Yang, J.-C., He, Q., Yu, P., Chu, Y.-H.: BiFeO3 thin films: a playground for exploring electricfield control of multifunctionalities. Annu. Rev. Mater. Res. 45(1), 249–275 (2015) 120. Wu, J., Fan, Z., Xiao, D., Zhu, J., Wang, J.: Multiferroic bismuth ferrite-based materials for multifunctional applications: ceramic bulks, thin films and nanostructures. Prog. Mater. Sci. 84, 335–402 (2016) 121. Sosnowska, I., Neumaier, T.P., Steichele, E.: Spiral magnetic ordering in bismuth ferrite. J. Phys. C: Solid State Phys. 15(23), 4835 (1982) 122. Sosnowska, I., Zvezdin, A.K.: Origin of the long period magnetic ordering in BiFeO3 . J. Magn. Magn. Mater. 140, 167–168 (1995) 123. Lebeugle, D., Colson, D., Forget, A., Viret, M., Bataille, A.M., Gukasov, A.: Electric-fieldinduced spin flop in BiFeO3 single crystals at room temperature. Phys. Rev. Lett. 100(22), 227602 (2008) 124. Ederer, C., Spaldin, N.A.: Weak ferromagnetism and magnetoelectric coupling in bismuth ferrite. Phys. Rev. B. 71(6), 060401 (2005) 125. Dzialoshinskii, I.E.: Thermodynamic theory of weak ferromagnetism in antiferromagnetic substances. Sov. Phys. JETP. 5(6), 1259–1272 (1957) 126. Moriya, T.: Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120(1), 91–98 (1960) 127. Dmitrienko, V.E., Ovchinnikova, E.N., Collins, S.P., Nisbet, G., Beutier, G., Kvashnin, Y.O., Mazurenko, V.V., Lichtenstein, A.I., Katsnelson, M.I.: Measuring the Dzyaloshinskii-Moriya interaction in a weak ferromagnet. Nat. Phys. 10(3), 202–206 (2014) 128. Lebeugle, D., Mougin, A., Viret, M., Colson, D., Ranno, L.: Electric field switching of the magnetic anisotropy of a ferromagnetic layer exchange coupled to the multiferroic compound BiFeO3 . Phys. Rev. Lett. 103(25), 257601–257604 (2009) 129. Lee, S., Ratcliff II, W., Cheong, S.-W., Kiryukhin, V.: Electric field control of the magnetic state in BiFeO3 single crystals. Appl. Phys. Lett. 92(19), 192906 (2008) 130. Popov, Y.F., Zvezdin, A., Vorob’Ev, G., Kadomtseva, A., Murashev, V., Rakov, D.: Linear magnetoelectric effect and phase transitions in bismuth ferrite BiFeO3 . JETP Lett. 57, 69 (1993) 131. Ruette, B., Zvyagin, S., Pyatakov, A.P., Bush, A., Li, J.F., Belotelov, V.I., Zvezdin, A.K., Viehland, D.: Magnetic-field-induced phase transition in BiFeO3 observed by high-field electron spin resonance: cycloidal to homogeneous spin order. Phys. Rev. B. 69(6), 064114 (2004)

    620

    J.-M. Hu and L.-Q. Chen

    132. Tokunaga, M., Azuma, M., Shimakawa, Y.: High-field study of strong magnetoelectric coupling in single-domain crystals of BiFeO3 . J. Phys. Soc. Jpn. 79(6), 064713 (2010) 133. Bai, F., Wang, J., Wuttig, M., Li, J., Wang, N., Pyatakov, A.P., Zvezdin, A.K., Cross, L.E., Viehland, D.: Destruction of spin cycloid in (111)c -oriented BiFeO3 thin films by epitiaxial constraint: enhanced polarization and release of latent magnetization. Appl. Phys. Lett. 86(3), 032511–032513 (2005) 134. Sando, D., Agbelele, A., Rahmedov, D., Liu, J., Rovillain, P., Toulouse, C., Infante, I.C., Pyatakov, A.P., Fusil, S., Jacquet, E., Carrétéro, C., Deranlot, C., Lisenkov, S., Wang, D., Le Breton, J.M., Cazayous, M., Sacuto, A., Juraszek, J., Zvezdin, A.K., Bellaiche, L., Dkhil, B., Barthélémy, A., Bibes, M.: Crafting the magnonic and spintronic response of BiFeO3 films by epitaxial strain. Nat. Mater. 12(7), 641–646 (2013) 135. Zhao, T., Scholl, A., Zavaliche, F., Lee, K., Barry, M., Doran, A., Cruz, M.P., Chu, Y.H., Ederer, C., Spaldin, N.A., Das, R.R., Kim, D.M., Baek, S.H., Eom, C.B., Ramesh, R.: Electrical control of antiferromagnetic domains in multiferroic BiFeO3 films at room temperature. Nat. Mater. 5(10), 823–829 (2006) 136. Chu, Y.-H., Martin, L.W., Holcomb, M.B., Gajek, M., Han, S.-J., He, Q., Balke, N., Yang, C.-H., Lee, D., Hu, W., Zhan, Q., Yang, P.-L., Fraile-Rodriguez, A., Scholl, A., Wang, S.X., Ramesh, R.: Electric-field control of local ferromagnetism using a magnetoelectric multiferroic. Nat. Mater. 7(6), 478–482 (2008) 137. Heron, J.T., Trassin, M., Ashraf, K., Gajek, M., He, Q., Yang, S.Y., Nikonov, D.E., Chu, Y.H., Salahuddin, S., Ramesh, R.: Electric-field-induced magnetization reversal in a ferromagnetmultiferroic heterostructure. Phys. Rev. Lett. 107(21), 217202 (2011) 138. Heron, J.T., Bosse, J.L., He, Q., Gao, Y., Trassin, M., Ye, L., Clarkson, J.D., Wang, C., Liu, J., Salahuddin, S., Ralph, D.C., Schlom, D.G., Iniguez, J., Huey, B.D., Ramesh, R.: Deterministic switching of ferromagnetism at room temperature using an electric field. Nature. 516(7531), 370–373 (2014) 139. Zhou, Z., Trassin, M., Gao, Y., Gao, Y., Qiu, D., Ashraf, K., Nan, T., Yang, X., Bowden, S.R., Pierce, D.T., Stiles, M.D., Unguris, J., Liu, M., Howe, B.M., Brown, G.J., Salahuddin, S., Ramesh, R., Sun, N.X.: Probing electric field control of magnetism using ferromagnetic resonance. Nat. Commun. 6, 6082 (2015) 140. Bibes, M., Barthélémy, A.: Multiferroics: towards a magnetoelectric memory. Nat. Mater. 7(6), 425–426 (2008) 141. Fusil, S., Garcia, V., Barthélémy, A., Bibes, M.: Magnetoelectric devices for spintronics. Annu. Rev. Mater. Res. 44(1), 91–116 (2014) 142. Popkov, A.F., Kulagin, N.E., Soloviov, S.V., Sukmanova, K.S., Gareeva, Z.V., Zvezdin, A.K.: Cycloid manipulation by electric field in BiFeO3 : coupling between polarization, octahedral rotation, and antiferromagnetic order. Phys. Rev. B. 92(14), 140414 (2015) 143. Xue, F., Li, L., Britson, J., Hong, Z., Heikes, C.A., Adamo, C., Schlom, D.G., Pan, X., Chen, L.-Q.: Switching the curl of polarization vectors by an irrotational electric field. Phys. Rev. B. 94(10), 100103 (2016) 144. Rovillain, P., de Sousa, R., Gallais, Y., Sacuto, A., Méasson, M.A., Colson, D., Forget, A., Bibes, M., Barthélémy, A., Cazayous, M.: Electric-field control of spin waves at room temperature in multiferroic BiFeO3 . Nat. Mater. 9(12), 975–979 (2010) 145. Kundys, B.: Photostrictive materials. Appl. Phys. Rev. 2(1), 011301 (2015) 146. Kittel, C.: Domain boundary motion in ferroelectric crystals and the dielectric constant at high frequency. Phys. Rev. 83(2), 458–458 (1951) 147. Qi, T., Shin, Y.-H., Yeh, K.-L., Nelson, K.A., Rappe, A.M.: Collective coherent control: synchronization of polarization in ferroelectric PbTiO3 by shaped THz fields. Phys. Rev. Lett. 102(24), 247603 (2009) 148. Wen, H., Chen, P., Cosgriff, M.P., Walko, D.A., Lee, J.H., Adamo, C., Schaller, R.D., Ihlefeld, J.F., Dufresne, E.M., Schlom, D.G., Evans, P.G., Freeland, J.W., Li, Y.: Electronic origin of ultrafast photoinduced strain in BiFeO3 . Phys. Rev. Lett. 110(3), 037601 (2013) 149. Daranciang, D., Highland, M.J., Wen, H., Young, S.M., Brandt, N.C., Hwang, H.Y., Vattilana, M., Nicoul, M., Quirin, F., Goodfellow, J., Qi, T., Grinberg, I., Fritz, D.M., Cammarata,

    12 Magnetoelectrics and Multiferroics

    621

    M., Zhu, D., Lemke, H.T., Walko, D.A., Dufresne, E.M., Li, Y., Larsson, J., Reis, D.A., Sokolowski-Tinten, K., Nelson, K.A., Rappe, A.M., Fuoss, P.H., Stephenson, G.B., Lindenberg, A.M.: Ultrafast photovoltaic response in ferroelectric nanolayers. Phys. Rev. Lett. 108(8), 087601 (2012) 150. Iurchuk, V., Schick, D., Bran, J., Colson, D., Forget, A., Halley, D., Koc, A., Reinhardt, M., Kwamen, C., Morley, N.A., Bargheer, M., Viret, M., Gumeniuk, R., Schmerber, G., Doudin, B., Kundys, B.: Optical writing of magnetic properties by remanent photostriction. Phys. Rev. Lett. 117(10), 107403 (2016) 151. Hu, J.-M., Duan, C.-G., Nan, C.-W., Chen, L.-Q.: Understanding and designing magnetoelectric heterostructures guided by computation: progresses, remaining questions, and perspectives. npj Comput. Mater. 3(1), 18 (2017) 152. Rivera, J.P.: A short review of the magnetoelectric effect and relatedexperimental techniques on single phase (multi-) ferroics. Eur. Phys. J. B. 71(3), 299 (2009) 153. Nan, C.W., Bichurin, M.I., Dong, S.X., Viehland, D., Srinivasan, G.: Multiferroic magnetoelectric composites: historical perspective, status, and future directions. J. Appl. Phys. 103(3), 031101 (2008) 154. Wang, Y., Hu, J., Lin, Y., Nan, C.-W.: Multiferroic magnetoelectric composite nanostructures. NPG Asia Mater. 2(2), 61–68 (2010) 155. Srinivasan, G.: Magnetoelectric composites. Annu. Rev. Mater. Res. 40(1), 153–178 (2010) 156. Vaz, C.A.F., Hoffman, J., Ahn, C.H., Ramesh, R.: Magnetoelectric coupling effects in multiferroic complex oxide composite structures. Adv. Mater. 22(26–27), 2900–2918 (2010) 157. Ma, J., Hu, J., Li, Z., Nan, C.-W.: Recent progress in multiferroic magnetoelectric composites: from bulk to thin films. Adv. Mater. 23(9), 1062–1087 (2011) 158. Sun, N.X., Srinivasan, G.: Voltage control of magnetism in multiferroic heterostructures and devices. Spine. 02(03), 1240004 (2012) 159. Hu, J.-M., Chen, L.-Q., Nan, C.-W.: Multiferroic heterostructures integrating ferroelectric and magnetic materials. Adv. Mater. 28(1), 15–39 (2016) 160. Velev, J.P., Jaswal, S.S., Tsymbal, E.Y.: Multi-ferroic and magnetoelectric materials and interfaces. Phil. Trans. R. Soc. A. 369(1948), 3069–3097 (2011) 161. Yu, P., Chu, Y.H., Ramesh, R.: Oxide interfaces: pathways to novel phenomena. Mater. Today. 15(7–8), 320–327 (2012) 162. Bibes, M.: Nanoferronics is a winning combination. Nat. Mater. 11(5), 354–357 (2012) 163. Vaz, C.A.F., Urs, S.: Artificial multiferroic heterostructures. J. Mater. Chem. C. 1(41), 6731 (2013) 164. Hu, J.-M., Shu, L., Li, Z., Gao, Y., Shen, Y., Lin, Y.-H., Chen, L.-Q., Nan, C.-W.: Film sizedependent voltage-modulated magnetism in multiferroic heterostructures. Phil. Trans. R. Soc. A. 372(2009), 20120444 (2014) 165. Hu, J.-M., Nan, C.-W., Chen, L.-Q.: Perspective: voltage control of magnetization in multiferroic heterostructures. Natl. Sci. Rev. (2019). https://doi.org/10.1093/nsr/nwz047 166. Vaz, C.A.F.: Electric field control of magnetism in multiferroic heterostructures. J. Phys. Condens. Matter. 24(33), 333201 (2012) 167. Liu, M., Sun, N.X.: Voltage control of magnetism in multiferroic heterostructures. Phil. Trans. R. Soc. A. 372(2009), 20120439 (2014) 168. Ramesh, R.: Electric field control of ferromagnetism using multi-ferroics: the bismuth ferrite story. Phil. Trans. R. Soc. A. 372(2009), 20120437 (2014) 169. Matsukura, F., Tokura, Y., Ohno, H.: Control of magnetism by electric fields. Nat. Nanotechnol. 10(3), 209–220 (2015) 170. Taniyama, T.: Electric-field control of magnetism via strain transfer across ferromagnetic/ferroelectric interfaces. J. Phys. Condens. Matter. 27(50), 504001 (2015) 171. Wang, Y., Li, J., Viehland, D.: Magnetoelectrics for magnetic sensor applications: status, challenges and perspectives. Mater. Today. 17(6), 269–275 (2014) 172. Nikonov, D.E., Young, I.A.: Benchmarking spintronic logic devices based on magnetoelectric oxides. J. Mater. Res. 29(18), 2109–2115 (2014)

    622

    J.-M. Hu and L.-Q. Chen

    173. Ortega, N., Ashok, K., Scott, J.F., Ram, S.K.: Multifunctional magnetoelectric materials for device applications. J. Phys. Condens. Matter. 27(50), 504002 (2015) 174. Hu, J.-M., Nan, T., Sun, N.X., Chen, L.-Q.: Multiferroic magnetoelectric nanostructures for novel device applications. MRS Bull. 40(09), 728–735 (2015) 175. Newnham, R.E., Skinner, D.P., Cross, L.E.: Connectivity and piezoelectric-pyroelectric composites. MRS Bull. 13(5), 525–536 (1978) 176. Lage, E., Kirchhof, C., Hrkac, V., Kienle, L., Jahns, R., Knochel, R., Quandt, E., Meyners, D.: Exchange biasing of magnetoelectric composites. Nat. Mater. 11(6), 523–529 (2012) 177. Hu, J.-M., Li, Z., Wang, J., Nan, C.W.: Electric-field control of strain-mediated magnetoelectric random access memory. J. Appl. Phys. 107(9), 093912 (2010) 178. Pertsev, N.A., Kohlstedt, H.: Resistive switching via the converse magnetoelectric effect in ferromagnetic multilayers on ferroelectric substrates. Nanotechnology. 21(47), 475202 (2010) 179. Hu, J.-M., Li, Z., Chen, L.-Q., Nan, C.-W.: High-density magnetoresistive random access memory operating at ultralow voltage at room temperature. Nat. Commun. 2, 553 (2011) 180. Binek, C., Doudin, B.: Magnetoelectronics with magnetoelectrics. J. Phys. Condens. Matter. 17(2), L39 (2005) 181. Chen, X., Hochstrat, A., Borisov, P., Kleemann, W.: Magnetoelectric exchange bias systems in spintronics. Appl. Phys. Lett. 89(20), 202508–202503 (2006) 182. Hu, J.-M., Li, Z., Lin, Y.H., Nan, C.W.: A magnetoelectric logic gate. Phys. Status Solidi RRL. 4(5–6), 106–108 (2010) 183. Hrkac, G., Dean, J., Allwood, D.A.: Nanowire spintronics for storage class memories and logic. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 369(1948), 3214–3228 (2011) 184. Lei, N., Devolder, T., Agnus, G., Aubert, P., Daniel, L., Kim, J.-V., Zhao, W., Trypiniotis, T., Cowburn, R.P., Chappert, C., Ravelosona, D., Lecoeur, P.: Strain-controlled magnetic domain wall propagation in hybrid piezoelectric/ferromagnetic structures. Nat. Commun. 4, 1378 (2013) 185. Kanai, S., Yamanouchi, M., Ikeda, S., Nakatani, Y., Matsukura, F., Ohno, H.: Electric fieldinduced magnetization reversal in a perpendicular-anisotropy CoFeB-MgO magnetic tunnel junction. Appl. Phys. Lett. 101(12), 122403 (2012) 186. Wang, W.-G., Li, M., Hageman, S., Chien, C.L.: Electric-field-assisted switching in magnetic tunnel junctions. Nat. Mater. 11(1), 64–68 (2012) 187. Shiota, Y., Nozaki, T., Bonell, F., Murakami, S., Shinjo, T., Suzuki, Y.: Induction of coherent magnetization switching in a few atomic layers of FeCo using voltage pulses. Nat. Mater. 11(1), 39–43 (2012) 188. Grezes, C., Ebrahimi, F., Alzate, J.G., Cai, X., Katine, J.A., Langer, J., Ocker, B., Khalili Amiri, P., Wang, K.L.: Ultra-low switching energy and scaling in electric-field-controlled nanoscale magnetic tunnel junctions with high resistance-area product. Appl. Phys. Lett. 108(1), 012403 (2016) 189. Miwa, S., Suzuki, M., Tsujikawa, M., Matsuda, K., Nozaki, T., Tanaka, K., Tsukahara, T., Nawaoka, K., Goto, M., Kotani, Y., Ohkubo, T., Bonell, F., Tamura, E., Hono, K., Nakamura, T., Shirai, M., Yuasa, S., Suzuki, Y.: Voltage controlled interfacial magnetism through platinum orbits. Nat. Commun. 8, 15848 (2017) 190. Bauer, U., Emori, S., Beach, G.S.D.: Voltage-controlled domain wall traps in ferromagnetic nanowires. Nat. Nanotechnol. 8(6), 411–416 (2013) 191. Bi, C., Liu, Y., Newhouse-Illige, T., Xu, M., Rosales, M., Freeland, J.W., Mryasov, O., Zhang, S., te Velthuis, S.G.E., Wang, W.G.: Reversible control of co magnetism by voltage-induced oxidation. Phys. Rev. Lett. 113(26), 267202 (2014) 192. Bauer, U., Yao, L., Tan, A.J., Agrawal, P., Emori, S., Tuller, H.L., van Dijken, S., Beach, G.S.D.: Magneto-ionic control of interfacial magnetism. Nat. Mater. 14, 174–181 (2015) 193. Gilbert, D.A., Grutter, A.J., Arenholz, E., Liu, K., Kirby, B.J., Borchers, J.A., Maranville, B.B.: Structural and magnetic depth profiles of magneto-ionic heterostructures beyond the interface limit. Nat. Commun. 7, 12264 (2016) 194. Hu, J.-M., Yang, T., Wang, J., Huang, H., Zhang, J., Chen, L.-Q., Nan, C.-W.: Purely electricfield-driven perpendicular magnetization reversal. Nano Lett. 15(1), 616–622 (2015)

    12 Magnetoelectrics and Multiferroics

    623

    195. Wang, J.J., Hu, J.M., Ma, J., Zhang, J.X., Chen, L.Q., Nan, C.W.: Full 180◦ magnetization reversal with electric fields. Sci. Rep. 4, 7507 (2014) 196. Hu, J.-M., Yang, T., Momeni, K., Cheng, X., Chen, L., Lei, S., Zhang, S., Trolier-McKinstry, S., Gopalan, V., Carman, G.P., Nan, C.-W., Chen, L.-Q.: Fast magnetic domain-wall motion in a ring-shaped nanowire driven by a voltage. Nano Lett. 16(4), 2341–2348 (2016) 197. He, X., Wang, Y., Wu, N., Caruso, A.N., Vescovo, E., Belashchenko, K.D., Dowben, P.A., Binek, C.: Robust isothermal electric control of exchange bias at room temperature. Nat. Mater. 9(7), 579–585 (2010)

    Hu is Assistant Professor of Materials Science and Engineering at University of Wisconsin (UW)-Madison. He received his Ph.D. from Tsinghua University in Materials Science and Engineering in 2013 and joined the faculty at UW-Madison in 2018. His current research activities focus on mesoscale computational modelling of ferroic materials and heterostructures and computation-guided device designs based on these materials.

    Chen is Hamer Professor of Materials Science and Engineering at Penn State. He received his Ph.D. from MIT in Materials Science and Engineering in 1990 and joined the faculty at Penn State in 1992. He has published over 600 papers in the area of computational microstructure evolution and multiscale modeling of structural metallic alloys, functional oxides, and energy materials.

    Magnetism and Superconductivity

    13

    Ilya M. Eremin, Johannes Knolle, and Roderich Moessner

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Paramagnetic Limit and Nonuniform FFLO Superconducting State . . . . . . . . . . . . . . . . . . . . Interplay of Zeeman Field and Spin-Orbit Interaction: Spinless Fermions and a Route Towards Topological Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ising Superconductors: Interplay of Magnetic Field and Spin-Triplet Channels . . . . . . . . . . . Superconductivity in the Presence of Antiferromagnetic Order . . . . . . . . . . . . . . . . . . . . . . . . Phenomenological Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Microscopic Consideration: Important Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ferromagnetic Superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    626 628 635 638 640 640 643 648 651 652

    Abstract

    The interplay of magnetism and superconductivity is one of the most striking features of the quantum mechanical description of solids. Originally thought to be mutually exclusive, both phenomena in fact show interesting coexistence, whose investigation remains a fascinating topic in contemporary condensed

    I. M. Eremin () Institut für Theoretische Physik III, Ruhr-Universität Bochum, Bochum, Germany e-mail: [email protected] J. Knolle Blackett Laboratory, Imperial College London, London, UK e-mail: [email protected] R. Moessner Max-Planck Institut für Physik komplexer Systeme, 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_14

    625

    626

    I. M. Eremin et al.

    matter physics. In this short review, we have aimed to provide some introduction to this exciting field of research.

    Introduction Magnetism has played an important role in our understanding of the collective properties of matter for many decades, starting with the exact solutions of the Ising model in one and two dimensions by Ising and Onsager, via the development of the renormalization group all the way to the discovery of topological states of matter in the form of spin liquids. Superconductivity has played a similarly central role for the field of macroscopic quantum phenomena since its discovery at the dawn of low-temperature physics by Kammerlingh Onnes and its explanation, much later, in the BCS theory of Bardeen, Cooper, and Schrieffer. The latter being based on a phononic mechanism, the two fields for a long time appeared to be at best unconnected, and at worst mutually exclusive. While resolutions of this competition started to be discussed already half a century ago, the discovery of high-temperature superconductors thrust the relationship between superconductivity and a magnetism at the very center of condensed matter physics 30 years ago. Understanding this relationship remains a challenge to this day. The appearance of long-range magnetic order is often associated with the localization of electrons, whereas in the superconducting state “the opposite” happens as electrons are paired and can then flow without resistance. Therefore, a priori one may think that these two phenomena are mutually exclusive. Indeed, in bulk systems, the uniform spinsinglet superconductivity and ferromagnetic order do not coexist. Vitaly Ginzburg in 1957 demonstrated the problem of coexistence of magnetism and superconductivity by considering the interaction of a phenomenological superconducting order parameter with a vector potential A of the magnetic field [1]. After the formulation of the BCS theory in 1957 [2], it became also clear that superconductivity in the spin-singlet state could also be destroyed by an exchange mechanism. In this case, the exchange field in a magnetically ordered state tends to align spins of Cooper pairs in the same direction, thus preventing a Cooper pairing. This is the so-called paramagnetic effect [3]. Furthermore, Anderson and Suhl in 1959 demonstrated that ferromagnetic ordering is unlikely to appear in the spin-singlet superconducting phase [4]. The main reason for this is the suppression of the zero-wave-vector component of the electronic paramagnetic susceptibility (known as the Pauli susceptibility) in a metal in the presence of superconductivity. In such a situation, the energy for ferromagnetic ordering decreases and, instead of ferromagnetic order, nonuniform magnetic ordering may appear, known by now as cryptoferromagnetic. Furthermore, Larkin and Ovchinnikov and, independently, Fulde and Ferrell showed that, in a pure ferromagnetic superconductor at low temperature, superconductivity can only appear as a nonuniform solution [5, 6], known by now as a LOFF (or FFLO) state. The 1977 discovery of ternary rare-earth compounds ReRh4 B4 and ReMo6 X8 (here, Re = rare-earth element and X = S, Se) provided the first experimental

    13 Magnetism and Superconductivity

    627

    evidence of magnetism and superconductivity coexisting in stoichiometric compounds (see for a review [7]). It turned out that in these and in many other systems, superconductivity coexists with antiferromagnetic order, and the Neel temperature, TN , can be either larger or smaller than the superconducting temperature, Tc , allowing to study their coexistence at both limits Tc >> TN and Tc 100 Å, or like in unconventional superconductors the Cooper pairs do not feel the effective exchange field due to the particular momentum dependence of the superconducting gap, as happens in the extended s-wave or nodal d-wave gap in iron-based superconductors and heavy-fermion superconductors, respectively. Moreover, in iron-based superconductors, layered cuprates, and many heavyfermion systems, antiferromagnetism and superconductivity are strongly coupled, even if they do not necessarily coexist. This indicates that the magnetic fluctuations are also likely the driving force for the Cooper pairing instability in these systems. One of the strong hints towards a common origin of antiferromagnetism and superconductivity in these compounds is the particular behavior of the superconducting order parameter. The short-range antiferromagnetic fluctuations are usually peaked at the characteristic antiferromagnetic wave vector, QAF . This provides a very strong source of repulsion for the fermions, a priori inhibiting superconducting Cooper pairing. To minimize this repulsion and to make it effectively attractive in the Cooper channel within the BCS theory, the superconducting order parameter Δ acquires a sign change for this wave vector, Δk = −Δk+QAF [13, 14, 15]. For the iron-based superconductors, the overall symmetry of the gap remains anisotropic s-wave, while for the heavy-fermion superconductor CeMIn5 and high-Tc cuprates, this condition yields dx 2 −y 2 -wave symmetry of the total order parameter. Furthermore, in such a setting, a term appears in the free energy coupling antiferromagnetic order and superconductivity, which can enhance the regions of phase space where they coexist [16]. By contrast, the first ferromagnetic superconductors, UGe2 [17] and URhGe [18], were discovered only relatively recently. These systems have spin-triplet Cooper pairing, which allows the coexistence of superconducting order with ferromagnetism more directly than in the singlet case. Indeed, superconductivity in these systems appears below the Curie temperature of the ferromagnetic phase, Θ; this makes the spin-singlet wave function of the Cooper pairs very improbable. Nevertheless, the coexistence of singlet superconductivity with ferromagnetism can be achieved in artificially fabricated layered ferromagnet/superconductor (F/S) systems. Due to the proximity effect, the Cooper pairs can penetrate into the

    628

    I. M. Eremin et al.

    ferromagnetic layer and induce superconductivity there. It is possible then to study the interplay between superconductivity and ferromagnetism by varying the layer thicknesses [19]. Note that the superconducting state appearing inside the ferromagnet has a lot in common with the FFLO state including the oscillatory behavior of the superconducting wave function. Another interesting aspect of coexistence of superconductivity and magnetism is the presence of the spin-orbit coupling, which breaks the spin-rotational invariance of the system already in the paramagnetic state. This topic currently attracts a lot of attention due to the phenomenon known as topological superconductivity. Although it may formally also appear without magnetism, it represents an interesting example of coexistence of spin-triplet (or more correctly odd-parity) Cooper pairing in the presence of both spin-orbit coupling and the Zeeman exchange field [20, 21, 22]. This chapter is organized to give a simple and conceptual view on the interplay of magnetism and superconducting order. After an outline of conventional Cooper pairing, explaining the incompatibility of ferromagnetic and superconducting order, we proceed with the examples of FFLO superconductors. Special emphasis is then given to the microscopic coexistence of unconventional superconducting order and antiferromagnetism. Here, we describe the coexistence of AF and SC states using iron-pnictide superconductors as a prototypical example. Finally, we describe the spin-triplet superconductivity in ferromagnetic superconductors.

    Paramagnetic Limit and Nonuniform FFLO Superconducting State To gain an understanding of the paramagnetic effect, observe that the critical field to destroy the superconducting state, Hp , at zero temperature may be found from a comparison of two energy scales, the superconducting condensation energy, ΔEs , and the energy gain due to the spin polarization of electrons in the normal state, ΔEp . The spin polarization of the electron gas changes its energy in the magnetic field in the normal state by ΔEp = −χp

    H2 , 2

    (1)

    where χp = 2μ2B N(0) is the paramagnetic (Pauli) spin susceptibility of the normal metal, μB is the Bohr magneton, N(0) is the density of electron states at the Fermi level, and the electron g-factor is equal to 2. At the same time, the spin polarization is absent in a spin-singlet superconductor, and the BCS condensation energy is given by ΔEs = −N(0)

    Δ20 , 2

    (2)

    where Δ0 is the superconducting gap at zero temperature (in the BCS theory, its value is given by 1.76kB Tc ). Comparing the two energy scales, one finds the

    13 Magnetism and Superconductivity

    629

    paramagnetic limit at T = 0, known as the Chandrasekhar-Clogston limit [23, 24]: Δ0 . Hp (0) = √ 2μB

    (3)

    Note that this field would represent a first-order phase transition from a normal to a superconducting state. However, Fulde and Ferrell [6] and Larkin and Ovchinnikov [5] predicted the appearance of the nonuniform superconducting state with a sinusoidal modulation of the superconducting order parameter at the scale of the superconducting coherence length, ξs , which is now called FFLO state. In this FFLO state, the Cooper pairs have a finite momentum compared with zero momentum in conventional superconductors. The critical field of the second-order transition into the FFLO state appears above the first-order transition line into a uniform superconducting state [3]. At T = 0, it is H F F LO = 0.755Δ0 /μB , whereas Hp ∼ 0.7Δ0 /μB . The FFLO state only appears in the temperature interval 0 < T < T + (see Fig. 1) and is sensitive to impurities as it is a state with finite momentum [25]. In real situations, the paramagnetic effect has to be considered in addition to the orbital effects induced by the magnetic field. The relative importance of the orbital and paramagnetic effects in the suppression of the superconductivity is described by the Maki parameter [3]

    α=

    orb (0) √ Hc2 m ∗ Δ0 ≈ 2 , Hp (0) m0 εF

    (4)

    Fig. 1 (a) Schematic H − T phase diagram of three-dimensional spin-singlet superconductors in the pure Pauli limit without orbital effect. Below T + = 0.56Tc , the transition becomes first order [3]. The dashed line represents the metastable transition line.(b) H − T phase diagram in the presence of the orbital effect. The transitions from the normal state to the FFLO and BCS states are of second order (solid line), while the transition from the FFLO state to the BCS state is of first order (thick solid line). The figure was reproduced with permission from [26]. Copyright 2007 The Physical Society of Japan

    630

    I. M. Eremin et al.

    where m∗ is the effective mass of the conduction electron and m0 is the free electron mass. Gruenberg and Gunther examined the stability of the FFLO state against the orbital effect for the spin-singlet three-dimensional superconductors [26]. The FFLO state can exist at finite temperatures if α is larger than 1.8 and its region narrows considerably from that in the absence of the orbital effect. The typical phase diagram indicating the formation of the nonuniform FFLO state is shown in Fig. 1. The appearance of a modulation in the superconducting order parameter is related to the Zeeman splitting of the electron’s energy level under a magnetic field (or exchange field if we are talking about ferromagnets) as illustrated in Fig. 2. In the absence of the Zeeman field, a Cooper pair is formed by two electrons with opposite momenta and spins, +kF ⇑ and −kF , ⇓. The resulting momentum of the Cooper pair will be 0. Because of the Zeeman splitting, the Fermi momentum of the electron with spin ↑ will shift from kF to k↑ = kF + δkF , where δkF = −μB H /vF and vF is the Fermi velocity. Similarly, the Fermi momentum of an electron with spin ↓ will shift from −kF to k↓ = −kF + δkF . The resulting momentum of the Cooper pair will then be given by k↑ + k↓ = 2δkF , which implies that the space modulation of the superconducting order parameter has a resulting wave vector q = 2δkF . In contrast to one dimension, it is not possible to choose the single wave vector δkF in two- [27] and three-dimensional superconductors [6, 5], which compensates the Zeeman splitting for all electrons on the Fermi surface, as δkF depends on the direction of vF , which yields the existence of the paramagnetic limit, which is absent in one dimension. Observe that the critical field for a nonuniform state at T = 0 is always higher than a uniform one. When T > μB H , at finite temperature, the smearing of the electron distribution function near the Fermi energy decreases the difference of energies between nonuniform and uniform states. From microscopic calculations, at T > T + = 0.56Tc , the uniform superconducting phase

    (b)

    q

    ky

    ky

    (a)

    q

    kx

    kx

    Fig. 2 Schematic illustration of the pairing states in two dimensions. (a) BCS pairing state (k, ↑; −k, ↓). (b) FFLO pairing state (k, ↑; −k + q, ↓) with q = 2δkF . The inner and outer circles represent the Fermi surface of the spin-down and spin-up bands, respectively. The dashed circle represents the initial Fermi surface

    13 Magnetism and Superconductivity

    631

    is always favored [3]. Observe that in real systems the Maki parameter is usually much less than unity; thus, the influence of the paramagnetic effect is negligibly small in most superconductors. However, in quasi-2D layered superconductors (for parallel fields) and heavy-fermion superconductors such as CeCoIn5 , αM is largely enhanced owing to large m∗ /m0 values, and thus the superconductivity may be limited by the Pauli paramagnetic effect. We discuss this in more detail below. It should be stressed that in superconductors with small Fermi energies, a large Δ0 /εF leads to the enhancement of αM , as was recently argued in some of the iron-based superconductors such as KFe2 As2 [28] and FeSe [45]. The formation of the FFLO phase can be also described in the framework of the generalized Ginzburg-Landau expansion [19] b F = a|ψ|2 + γ |∇ψ|2 + |ψ|4 , 2

    (5)

    where ψ is the superconducting order parameter and the coefficient a is positive for T > Tc and vanishes at T = Tc . Below Tc a is negative and the minimum in the free energy functional F occurs for the uniform superconducting state |ψ|2 = −a b where b > 0. In a magnetic field, all coefficients depend on the energy of the Zeeman splitting, μb H . In addition, the orbital effects of the magnetic field can be included via gauge invariant substitution, ∇ → ∇ − (2ie/c)A. The FFLO state occurs due to the fact that the coefficient γ changes its sign at the point (H + , T + ) of the phase diagram; see Fig. 1a. A negative sign of γ means that the minimum of the functional does not correspond to a uniform state, and a spatial variation of the order parameter decreases the energy of the system. To describe this properly, one has to add a higher-order derivative term in the expansion (5), and the overall form is now given by F = a(H, T )|ψ|2 + γ (H, T )|∇ψ|2 +

    η(H, T ) 2 2 b(H, T ) 4 |∇ ψ| + |ψ| . 2 2

    (6)

    Its minimization results in the equation that determines the critical temperature of the second-order phase transition into a superconducting state aψ − γ Δψ +

    η 2 Δ ψ = 0, 2

    (7)

    where a = α(T − Tc (H )), where Tc (H ) is the critical temperature of the transition into the uniform superconducting state. In the case of γ < 0, we search for the nonuniform solution ψ = ψ0 eiqr . This yields the following equation: a = −γ q 2 −

    η 4 q . 2

    (8)

    Observe that for γ < 0, the maximum critical temperature corresponds to the finite value of the modulation vector q02 = −γ /η and the corresponding transition

    632

    I. M. Eremin et al.

    temperature FFLO state TcF F LO (H ) with the coefficient a =  F F LO into the nonuniform  2 α Tc (H ) − Tc (H ) = γ /2η. This temperature is higher than the critical temperature Tc (H ) of the uniform state. Thus, the appearance of an FFLO state may simply be interpreted as a sign change of the gradient term in the GinzburgLandau functional. Observe that a more detailed analysis of the FFLO state using the generalized Ginzburg-Landau functional gives a sinusoidal modulation of the order parameter [29, 30]. Formally, the generalized Ginzburg-Landau functional describes then a new type of superconductor with very different properties, and the theory of superconductivity must be redone on the basis of this functional, and the detailed analysis can be found in Ref. [19]. Concerning the experimental realization of the FFLO phase in the spin-singlet superconductors, one has to bear in mind that there are several necessary conditions. For example, this state may occur only in strongly type-II superconductors with very large Ginzburg-Landau parameter, κ = λ/ξ , and large Maki parameter (here, λ is a penetration depth and ξ is a superconducting coherence length), such that the upper critical field approaches the paramagnetic limit. In addition, this superconductor has to be in a clean limit, ξ 0 denotes the strength of the spin-orbit coupling and h ≥ 0 is the Zeeman energy arising from a magnetic field applied in the zdirection. Overall, Hwire refers to an electron-doped semiconducting wire such as InAs or InSb with strong Rashba spin-orbit coupling, in the limit where only the lowest transverse subband is relevant. The pairing term HΔ models the proximity effect arising from the adjacent s-wave superconductor like Nb or Al. The red and blue curves in Fig. 5a illustrate the band structure of Hwire for h = 0. Due to spin-orbit coupling, the blue and red parabolae respectively correspond to electronic states whose spin aligns along y and −y. The magnetic field lifts the crossing between these parabolae at k = 0, producing band energies

    636

    I. M. Eremin et al.

    Fig. 5 (a) Electronic band structure for the interface between a wire and a conventional s-wave superconductor, shown in the inset, without (red and blue curves) and with external magnetic field (black curves). (b) Shows the kinetic energy in the Kitaev toy model with spinless fermions. The superconducting gap opens a bulk gap except at the chemical potential values μ = ±t. For |μ| < t, the system forms a topological weak-pairing phase. In this case, the model, written in terms of Majorana fermions, couples at adjacent lattice sites, leaving two “unpaired” Majorana degrees of freedom at the ends of the chain. Figures taken with permission from [20]. Copyright 2012 by IOP Publishing Ltd.

    ± (k) =

     k2 − μ ± (αk)2 + h2 2m

    (12)

    shown by the solid black curves in Fig. 5a. When the Fermi level resides within this field-induced gap as shown in the figure, the fermions in the wire appear to be effectively spinless. Focusing on the spinless regime and assuming Δ μ2 + Δ2 (topological criterion), the Hamiltonian can be brought into the simple form describing spinless fermions that hop on an N -site chain and exhibit long-range-ordered p-wave superconductivity: HKitaev = −μ

    

    cx† cx −

    1  † tcx cx+1 + Δe−iφ cx cx+1 + h.c. . 2 x

    (13)

    Here, μ is again the chemical potential, t > 0 the nearest-neighbor hopping, Δ the p-wave pairing amplitude, and φ the corresponding superconducting phase. The chain bulk properties can be studied by imposing periodic boundary conditions on the system. In momentum space, the Hamiltonian, HKitaev , can be brought to a diagonal form with quasiparticle energy

    13 Magnetism and Superconductivity

    637

    Ebulk (k) =

    εk2 + |Δ˜ k |2

    (14)

    where εk = −t cos k − μ is the kinetic energy and Δ˜ k = −iΔeiφ sin k is the pairing amplitude in momentum space. Ebulk (k) allows gapless bulk excitations when the chemical potential is finetuned to μ = t or −t where the Fermi level respectively coincides with the top and bottom of the conduction band. The gap closure at these isolated μ values reflects the p-wave nature of the pairing required by the Pauli exclusion principle. In particular, Δ˜ k is an odd function of k, and Cooper pairing at k = 0 or k = ±π is prohibited, thereby leaving the system gapless at the Fermi level when μ = ±t (see Fig. 5). The physics of the chain is different in the two gapped regimes with |μ| > t and |μ| < t. The first case connects smoothly to the trivial vacuum (upon taking μ → ∞) where no fermions are present. By contrast, in the latter, a partially filled band acquires a gap due to p-wave pairing. The new physics associated with this state can be most simply accessed by decomposing the spinless fermion operators cx in the original Kitaev Hamiltonian in terms of two Majorana fermions via cx =

     e−iφ/2  γB,x + iγA,x , 2

    (15)

    † where the γ operators obey the Majorana fermion relation γα,x = γα,x . In this basis, the Hamiltonian, HKitaev , reads

     μ  1 + iγB,x γA,x 2 N

    HKitaev = −

    x=1



    i 4

    N −1 



    (Δ + t) γB,x γA,x+1 + (Δ − t) γA,x γB,x+1 .

    (16)

    x=1

    To access the topologically nontrivial phase, one sets μ = 0 and t = Δ = 0. Here, the Hamiltonian simplifies to

    HKitaev = −

    N −1 it  γB,x γA,x+1 2

    (17)

    x=1

    which couples Majorana fermions only at adjacent lattice sites. However, as there are two Majorana fermions per site, the ends of the chain (for the open boundary conditions) now support “unpaired” zero-energy Majorana modes γ1 = γA,1 and γ2 = γB,N . These can be combined into an ordinary – although highly nonlocal – fermion that costs zero energy and therefore produces a twofold ground-state degeneracy. Note the stark difference from conventional gapped superconductors,

    638

    I. M. Eremin et al.

    where typically there exists a unique ground state with even parity so that all electrons can form Cooper pairs. The appearance of localized zero-energy Majorana end states and the associated ground-state degeneracy arise because the chain forms a topological phase while the vacuum bordering the chain is trivial. For the purpose of our review, it is important that the fermions have to be spinless. This property ensures that a single zeroenergy Majorana mode resides at each end of the chain in its topological phase. Suppose that instead spinful fermions, initially without spin-orbit interactions, form a p-wave superconductor. In this case, spin merely doubles the degeneracy for every eigenstate of the Hamiltonian, so that when |μ| < t, each end supports two Majorana zero modes, or equivalently one ordinary fermionic zero mode. Unless special symmetries are present, these ordinary fermionic states will move away from zero energy upon including perturbations such as spin-orbit coupling. Note that despite the intense search for intrinsic topological superconductors, the practical realization of Majorana fermions appears to be easiest in the heterostructure described above with recent significant advances towards their experimental realizations [53, 54, 55, 56, 57, 58]. This promises further exciting developments on the interplay of magnetism and superconductivity in the field of topological superconductors. To study the subject of topological superconductivity in depth, the reader is invited to look into special reviews on the subject [20, 22, 53].

    Ising Superconductors: Interplay of Magnetic Field and Spin-Triplet Channels In recent years, two-dimensional (2D) superconductivity became an active field of research as a result of technological advances in the fabrication of heterostructures, made from the van der Waals materials [59]. These systems are comprised of one-toseveral atomically thin monolayers exfoliated on substrates. Many of the properties of the bulk persist down to the monolayer limit, especially in those systems where the electronic structure is already quasi-two-dimensional one in the bulk samples. For example, both bulk and monolayer NbSe2 are charge density wave metallic superconductors [60,61]. However, due to the different effective dimensionality, the monolayers react differently to the applied fields and often lack the inversion center, present in the bulk system [62, 63]. These non-centrosymmetric superconducting monolayers having in-plane mirror σh symmetry are called Ising superconductors. Due to the time-reversal symmetry, the state |k ↑ (|k ↓) is degenerate with | − k ↓ (| − k ↑). When the lattice breaks parity, the spin-orbit coupling (SOC) causes the spin splitting of Bloch states with the typical energy difference of Δso , and the typical Fermi surface is shown in Fig. 6. As a result of the spin-orbit splitting, the probability amplitude of the Cooper pair to be in a state |k ↑; −k ↓ differs from the corresponding amplitude for the state |k ↓; −k ↑. Correspondingly, the parityeven singlets and parity-odd triplets |Ψs,t  ∝ |k ↑; −k ↓ ∓ |k ↓; −k ↑ coexist [64, 65].

    13 Magnetism and Superconductivity

    639

    Fig. 6 Electronic structure of monolayer metallic transition-metal dichalcogenides as adapted from Ref. [62]. Schematic representation of the crystal structure of 2Ha-MX2 viewed along [100] direction for M = (Nb, Ta) and X = (S, Se) with 1H (monolayer). Spin-projected Fermi surface of monolayer TaS2 and NbSe2 as computed by density functional theory (DFT) are also shown [62]. The variation of the shading and curve thickness refer to the magnitude of spin splitting in the valence band due to spin-orbit coupling. Figures are taken from Ref. [62] using Creative Commons CC BY license

    Apart from inducing singlet-triplet mixing, the SOC makes the superconducting state robust against the in-plane Zeeman field B. Because of the negligible thickness of the monolayer, orbital-limiting effects do not contribute, and the only way a magnetic field can affect the electronic states is via the paramagnetic effect, discussed above. The large SOC enhances the critical in-plane field Bc well beyond the Pauli limit. This has been studied theoretically [66, 65, 67, 68, 69, 70, 71] and demonstrated experimentally [60, 72, 73, 61, 74, 75, 76, 62]. Till now, theoretical analyses of the superconducting pairing mechanism in monolayer NbSe2 [71, 77] have relied on model descriptions of superconductivity in materials that lack inversion symmetry based on the band structure calculated from first principles. There is also a lack of consistency between first-principles descriptions of superconductivity in bulk NbSe2 and experimental results. For example, first-principles calculations overestimate Tc in bulk NbSe2 and isostructural NbS2 and the zero-temperature gap [78]. Most recently, it was demonstrated that bulk and monolayer NbSe2 are close to a magnetic instability [79], and repulsive as well as spin-fluctuation-induced interactions cannot be most likely neglected when addressing superconductivity in NbSe2 [80]. This perspective on the role of magnetism in monolayer NbSe2 will also be crucial to un- derstand and control the superconducting properties of monolayer NbSe2 in the presence of an external magnetic field or with heterostructures between monolayer NbSe2 and magnetic materials.

    640

    I. M. Eremin et al.

    Superconductivity in the Presence of Antiferromagnetic Order Phenomenological Description Similar to the case of the FFLO state, the competition between superconductivity and antiferromagnetism near their finite-temperature phase transitions can be described in terms of a Ginzburg-Landau theory of coupled order parameters without referring to microscopic details; see, for example, [16]. The free energy functional in the spatially homogeneous case is given by FAF +SC (ψ, M) =

    as bs δ am 2 bm 4 |ψ|2 + |ψ|4 + |ψ|2 M2 + M + M , 2 4 2 2 4

    (18)

    where ψ and M refer to the superconducting (SC) and antiferromagnetic (AFM) order parameter, respectively, and the integration over d-dimensional space is implicitly assumed. As above, ψ is a complex order parameter, characterized by an amplitude and a phase, while M is a three-component vector encoding magnetic ordering. The quadratic coefficients am = am,0 (T − TN,0 ) and as = as,0 (T − Ts,0 ) change sign at TN,0 and Tc,0 , which denote the Neel and SC transition temperatures without order-parameter competition. The leading term of the order-parameter coupling is characterized by the coefficient δ, where the sign of δ > 0 reflects that orders compete with each other. We consider the situation where the transitions for δ = 0 are second order and the quartic coefficients bm and bs are positive. This free energy functional accounts for various phase diagrams for competing and coexisting AFM and SC orders. In particular, consider TN,0 (p) and Tc,0 (p) vary as a function of a physical parameter p, which typically is pressure, magnetic field, or doping (electron density and/or disorder). In the case when both transitions meet at p = p∗ , i.e., T ∗ = Tc,0 (p∗ ) = Tm,0 (p∗ ), one has a multicritical point (p∗ , T ∗ ) in the phase diagram. In the vicinity of this multicritical point, a simultaneous expansion of the order parameters is allowed. The mean-field analysis of Eq. (18) near (p ∗ , T ∗ ) depends on whether δ 2 > bs bm or δ 2 < bs bm . We are interested in the situation when AFM and SC orders compete (the physically more realistic situation) δ > 0 and one can define a dimensionless quantity [16] δ g=√ −1 bm bs

    .

    (19)

    √ For g < 0 (0 < δ < bm bs ), (p∗ , T ∗ ) is a tetra-critical point where the two second-order phase lines cross, leading to a regime in the phase diagram where simultaneous AFM and SC orders coexist homogeneously. This is shown in Fig. 7a; both phases compete but do not exclude √ each other. On the other hand, if g > 0 (δ > bm bs ), the competition between two phases is sufficiently strong such that both phases are separated by a first-order transition that terminates at the bi-critical point (p∗ , T ∗ ). Notice that if the parameter p jumps

    13 Magnetism and Superconductivity

    641

    Fig. 7 Schematic phase diagrams (p, T ) for competing AFM and SC orders, after [16]. Here, p is a generic physical parameter and T is temperature. Solid and dashed lines denote secondorder or first-order phase transitions. For g < 0, there is a tetra-critical point and a region of homogeneous coexistence (a). For g > 0, there is a bi-critical point, shown in (b) and (c). If p changes discontinuously across the first-order transition, there will be a region of heterogeneous coexistence for p1 < p < p2 (b), while its conjugate variable hp changes continuously and the phase diagram has only one first-order line (c). Figure taken with permission from [16]. Copyright 2010 by the American Physical Society

    discontinuously from p1 to p2 at the first-order transition, there is an intermediate regime p1 < p < p2 of heterogeneous phase coexistence (see Fig. 7b). A sharp line of first-order transitions occurs if one considers the phase diagram as a function of hp , the variable that is thermodynamically conjugate to p (see Fig. 7c). Note that critical fluctuations will change the universal exponents near the critical temperatures and the slopes of the phase lines, but they do not change the generic behavior shown in Fig. 7 (see [81, 82]). Both order parameters can be finite simultaneously for g < 0, and this regime is often referred to as coexistence of AF and SC, referring to microscopic coexistence of SC and AFM orders. This should not be confused with phase coexistence in the thermodynamic sense. The area in Fig. 7a below the tetra-critical point

    642

    I. M. Eremin et al.

    is a single thermodynamic phase characterized by two order parameters that are simultaneously finite. Similarly, the tetra-critical point is not a point where four phases coexist, which would not be allowed by the Gibbs phase rule, but a point where the system is in a single phase and both order parameters are infinitesimal simultaneously. Below the bi-critical point, coexistence of thermodynamic phases only occurs for p1 < p < p2 , where macroscopic AFM and SC regions occur together in the sample. Usually, one uses the term homogeneous coexistence of AFM and SC orders below the tetra-critical point to refer to coexisting order, and heterogeneous coexistence below the bi-critical point to refer to coexistence of phases. From the Ginzburg-Landau expression, it is easy to obtain the temperature dependence of the magnetic moment in the SC phase in the case of homogeneous coexistence M2 (T ) =

    am,0 bs (TN,0 − T ) + as,0 δ(T − Tc,0 ) . √ bm bs − δ

    (20)

    Without phase competition, both order parameters decrease as function of increasing temperature, dM2 /dT 2 < 0 and dψ 2 /dT 2 < 0. However, once as,0 δ > am,0 bs , one finds dM2 /dT 2 > 0 in the superconducting state. Thus, below Tc , the ordered moment decreases with decreasing temperature and the same condition implies a back bending of the antiferromagnetic phase boundary upon entering the superconducting state, a feature which is indeed often observed experimentally [16]. The considerations above are quite general and refer to conventional and unconventional superconducting states coexisting with antiferromagnetic order either of itinerant or purely localized nature. Nevertheless, experimentally, both the AFM and SC orders homogeneously coexist more often in the case of unconventional superconductors. To understand this in detail requires a microscopic description. However, one should bear in mind that in unconventional superconductors, the superconducting gap often has opposite signs on the different sections of the Fermi surface. In this case, the staggered magnetization, M, can be viewed as a mediator of an intrinsic Josephson coupling between those regions. Specifically for the case of the iron-based superconductors, which are believed to have opposite signs of the superconducting gaps, ψ1 and ψ2 on the electron and hole Fermi surface pockets, this Josephson interband coupling can be written as [16] EJ ∝ M2 |ψ1 ||ψ2 | cos θ,

    (21)

    where θ is a relative phase between the SC order parameters of the two Fermi surface sheets. Once the phase difference is π (i.e., the superconducting gaps have opposite phases), this term contributes to the reduction of the generally positive value of δ in Eq. (18). Therefore, the situation of Fig. 7a appears more likely for these unconventional superconductors.

    13 Magnetism and Superconductivity

    643

    Microscopic Consideration: Important Aspects The interplay between AFM and SC has been investigated in many contexts and has a long history [83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 16, 94, 95, 96, 97, 98]. In the following, we are not going to review these works in detail but concentrate on exposing the most important aspects. The first important aspect of superconductivity in the presence of translational symmetry breaking AFM order is the folding of the electronic structure and formation of new bands below TN , as illustrated in Fig. 8. The details depend on the particular model under consideration, but the main features can be clarified for the simplest case of the single-band model. In particular, consider the single-band Hubbard Hamiltonian on a square lattice H =−

    

    † ti,j ciσ cj σ + U

    i,j,σ

     i

    ni↑ ni↓ − μ

    

    (22)

    niσ ,

    i,σ

    † creates an electron on site i with spin σ . The interaction term U denotes where ciσ the energy cost associated with having two electrons on the same site. In reciprocal space, the Hamiltonian reads

    H =

     kσ

    † k ckσ ckσ +

    U   † † c c c−k+qσ ckσ , 2N σ k σ −k +qσ

    (b)

    (a)

    (23)

    k,k ,q

    (c)

    ky

    QAF

    kx

    kx

    kx

    Fig. 8 Evolution of the Fermi surface topology for the commensurate AF state based on the single-band model for layered cuprates. Due to the breaking of translational symmetry at the antiferromagnetic momentum QAF M = (π, π ), the original energy dispersion, εk , in the paramagnetic metallic state folds with the one shifted by the antiferromagnetic momentum εk+QAFM . The original large Fermi surface (a) in the first Brillouin zone gets reconstructed as an antiferromagnetic gap, W , opens on the “hot spots” εk = εk+QAFM (b). For intermediate W , the antiferromagnetic state is still metallic, and the residual Fermi surface consists of small hole pockets, centered on (±π/2, ±π/2), and electron pockets, centered on (±π, 0), (0, ±π )

    644

    I. M. Eremin et al.

    with k = −2t[cos(kx ) + cos(ky )] − 4t cos(kx ) cos(ky ) − μ.

    (24)

    The parameter −t is the energy gain corresponding to hopping between neighboring sites, and −t denotes the energy gain by hopping to next-nearest-neighbor sites. The doping level of the system is controlled by changing the chemical potential μ. The interaction between two electrons is first treated in the HartreeFock approximation giving rise to AF ordering of the spins. We therefore consider the mean-field Hamiltonian HSDW =

      k

    σ

    † (ckσ

    † ck+Qσ )

     σ W c k kσ , σ W k+Q ck+Qσ

    (25)

    † † U  where W = − N k [ck+Q↑ ck↑  − ck+Q↓ ck↓ ] is the AF order parameter. Diagonalization of the mean-field Hamiltonian leads to the following energy  ± α,β spectrum: Ek = k+ ± (k− )2 + W 2 , k± = k 2k+Q . The magnetic gap equation is usually solved self-consistently for a given set of values of the hopping integrals t = 1, t , the Coulomb repulsion U , and the doping. The electronic structure in the AF state folds and yields the new Fermi surface topology, illustrated in Fig. 8. The problem immediately becomes a two-band one as there are now two types of fermionic bands, α and β. These give rise to electron and hole pockets. In this regard, it is then more practical to consider the Cooper pairing in terms of the α and β fermions, obtained after diagonalization of the mean-field AF Hamiltonian. In particular, for the magnetization pointing along the z-direction, the quasiparticle operators are related to the bare electron operators by the transformation:

    ckσ = uk αkσ + vk βkσ , ck+Qσ = sign(σ )[vk αkσ − uk βkσ ].

    (26) (27)

    Using the new basis allows to account naturally for the umklapp Cooper pairing, ck,↑ c−k−QAFM ,↓ , absent in the paramagnetic state, and the renormalization of the chemical potential in the AF state, μ. In addition, one has to bear in mind that the AF state breaks spin-rotational invariance, which yields parity (momentum in the reduced BZ and its inversion) as the only good quantum number of the Cooper pairs. The Cooper pairing has to be considered between quasiparticles residing † † † † in the same band, i.e., αkσ α−kσ  and βkσ β−kσ . In addition, one has to take into account the possibility of anomalous pairings between fermions belonging to † † different pockets, i.e., αkσ β−kσ . The complete mean-field Hamiltonian includes then the pair scattering processes between normal and anomalous gaps, as well as scatterings between anomalous gaps. Such a pairing has been considered in the context of the iron-based superconductors [96] and layered cuprates [98] as it

    13 Magnetism and Superconductivity

    645

    couples linearly to the normal intraband gaps with a coupling constant proportional to the SDW order parameter. † † In principle, anomalous pairs of the form αkσ β−kσ  involve fermions far from the Fermi level for T < TN , since the two bands are gapped by the antiferromagnetic β gap |Ekα − Ek | ≥ 2W . Nevertheless, the anomalous gaps become sizeable due to the coupling to the normal intraband pairs and exist only in the AFM background, since the coupling between normal and anomalous gaps is proportional to W . A detailed analysis [98] of the interaction Hamiltonian reveals, however, that even † † † † parity intraband gaps, i.e., αk↑ α−k↓ −αk↓ α−k↑ , couple to even parity anomalous interband gaps. In the iron-pnictide study of Ref. [96], the reported anomalous gap was dubbed spin triplet and coupled to an even parity singlet intraband gap. However, one has to bear in mind that in this case the triplet gap is actually of even parity, which is allowed due to the band index. Nevertheless, the spinsinglet and spin-triplet gaps do coexist in the AFM background both having the same parity. As the spin-rotational and the time-reversal symmetry for a given sublattice is broken, one could even find the so-called s + it even parity state, as shown in Fig. 9. However, to avoid confusion, we believe that it may be better to classify the superconducting gaps by parity rather than by spin quantum numbers, as spin-rotational symmetry (and to some extent time-reversal symmetry) is explicitly broken, while the parity is conserved. So far, the interactions yielding superconductivity have not been specified. They are often considered on a phenomenological level. At the same time, the most interesting situation is when the Cooper pairing arises from the same interaction which drives the AFM transition. In the single-band Hubbard model, the source of Cooper pairing is the spin and charge fluctuations. In particular, higher-order interactions in U generate superconductivity on top of the AFM order through

    Fig. 9 The spin structure of the gap function in an AFM background for the iron-based superconductor and the arrows refer to the relative phase of the intra- and interband order parameter: (a) pure spin-singlet (s) even parity s +− -wave state, (b) pure spin-triplet (t) even parity s ++ -state, and (c) s + it state with π/2 phase difference between s- and t-components as adopted from Ref. [96]. Figure taken with permission from [86]. Copyright 2014 by the American Physical Society

    646

    I. M. Eremin et al.

    longitudinal and transverse spin fluctuations, following the original proposals of Refs. [99,90]. Since U connects the bare electrons, the diagrammatics are performed in terms of the bare electron Green’s functions, while the Cooper pairing is better thought of to take place between the quasiparticles of the AF state. Transverse and longitudinal spin fluctuations give rise to fundamentally different interactions. Inspection of the interaction vertex formulated in real space [98] shows that the charge and longitudinal interaction vertices give rise to no spin flips whereas the transverse interaction does. In the latter channel, there are the gapless Goldstone modes of the AFM phase, which give rise to a divergent interaction potential between the bare electrons. However, when we consider pairing between the quasiparticles of the AF state, this divergence is removed by the coherence factors as noted in earlier works [100, 101]. In the case of pairing between opposite spin electrons, spin flip processes are possible, and the effective interaction is mediated both by longitudinal and transverse spin fluctuations. If pairing occurs between same spin electrons, only longitudinal spin fluctuations contribute and the pairing potential does not include the bare Coulomb repulsion U since this acts only between opposite spin electrons. The interaction Hamiltonian is formulated in terms of the SDW quasiparticles, and in line with earlier work [90, 100], one finds the interactions in the longitudinal and transverse channel individually, with the transverse part of the interaction stated as a spin flip vertex explicitly. The details of this procedure was elaborated in Ref. [98]. The complete doping evolution of the three leading superconducting order parameters, dx 2 −y 2 , g, and p -wave, was discussed in Ref. [98] (see Fig. 10). On the hole-doped side, the near degeneracy of the dx 2 −y 2 and g-wave solutions, which was observed very close to half filling, is split when hole doping is increased and the dx 2 −y 2 solution becomes clearly dominant. In the case of electron doping, the d-wave solution is strongest very close to half filling even though this is not where

    Fig. 10 Phase diagram of coexisting unconventional superconductivity and antiferromagnetic order, adapted from Ref. [98]. (a) Three leading superconducting instabilities as a function of filling for the three largest eigenvalues of the linearized gap equation. The AFM region is shown by the green area. (b) The critical temperature, Tc = 1.13c e−1/λ , with the energy cutoff set to c = 0.25 t and t = 400 meV. The superconducting instability is dx 2 −y 2 at all fillings. Figure taken with permission from [98]. Copyright 2016 by the American Physical Society

    13 Magnetism and Superconductivity

    647

    the longitudinal and transverse pairing potentials achieve their maximum strengths. The reason is that in the limit where electron pockets are small, the structure of the intrapocket pairing potentials is purely attractive and this strongly supports a d-wave solution. At critical electron doping for which W → 0, the Fermi arcs just touch the magnetic zone boundary, and as a consequence, nesting by Q on the paramagnetic side is rapidly weakened upon increased electron doping. There is a jump in the magnitudes of the eigenvalues for the three leading gap symmetries in the paramagnetic state. Their ordering, however, remains the same as in the SDW phase, with the p solution, which in the paramagnetic phase takes the simpler form [cos(kx ) − cos(ky )] sin(kx ), the least favorable. The jump in the eigenvalues is due to the weak nesting properties of the Fermi surface on the paramagnetic side. In fact, the paramagnetic Fermi surface, which is a hole pocket centered at (π, π ), is roughly circular, thereby preventing nesting not only at Q but also at any q-vector. As a result, spin-fluctuation-mediated superconductivity rapidly dies off. It is interesting that the arguments on the stability of the dx 2 −y 2 -wave symmetry on the AF background change if the Cooper pairing is mediated by an exchange interaction which involves only nearest-neighbor sites, while the discussed approach above deals with the on-site Coulomb repulsion and its renormalization in the Cooper channel at higher orders. The similarity of the treatment of the coexistence phase in Refs. [102] and [98] allows for a direct comparison of the interaction Hamiltonians. In the simplified case of only hole pockets at the Fermi surface, the interaction Hamiltonian reads generally Hhole =

     k,k σ

    † † Γ (k, k )βkσ β−kσ β−k σ βk σ ,

    (28)

    where in the t − J -like model without double occupancy constraint, employed in Ref. [102], the effective interaction takes the form Γt−J (k, k ) = −

    J (k − k ) 2 [m (k, k ) + l 2 (k, k )] 2

    −J (k + k )[n2 (k, k ) + p2 (k, k )],

    (29)

    with J (q) = J [cos(qx ) + cos(qy )] and J > 0. In the Hubbard model we have +− z ΓHub (k, k ) = Γk,k ± 2Γk,k ,

    (30)

    with the longitudinal and transverse effective interactions. The lower sign of Eq. (30) belongs to the triplet channel, and this is the source of an effective intrapocket repulsion on the hole-doped side. On the contrary, the effective interaction as stated in Eq. (29) gives rise to a purely attractive triplet potential for k and k residing on the same hole pocket, and therefore the p-wave solution would indeed appear to be the dominating instability. Note also that the model of Ref. [102] cannot be considered strictly as a strong-coupling limit of the single-band Hubbard model

    648

    I. M. Eremin et al.

    as it does not include a constraint for no double occupancies of the fermions explicitly. This could play an important role as the original strong-coupling study of unconventional superconductivity driven by the spin waves, studied within the t − J model with the constraint of no double occupancies [103], does find the dx 2 −y 2 wave symmetry of the superconducting gap to be the only stable solution. Finally, let us mention another interesting phase, which is probably a result of the induced spin density wave (Q-phase) in the d−wave (nodal) superconductivity in an external magnetic field, observed in CeCoIn5 for fields in the basal plane. This phase exists only once the superconductivity is present and gets rapidly suppressed when the field is turned away from this plane (see [31] for a review and references therein). The SDW modulation wave vector in this phase was found to be pinned to the lattice and points along the nodal direction. This was considered as evidence that electronic nesting plays a role in the formation of the SDW order. Its origin in CeCoIn5 is not fully understood at present, and several microscopic theories have been developed [104, 105, 106, 107, 108, 109, 110, 111]. In the first approach, the appearance of the magnetic phase appears as an instability towards an FFLO state of modulated d-wave superconductivity at high fields close to the upper critical field. Superconductivity in this FFLO state features a long-wavelength modulation. Magnetism occurs in the FFLO nodal regions as Andreev bound states and is thus a consequence of the modulated nature of superconductivity [104]. In the other approaches, the magnetism arises for different reasons either as a result of normal Fermi pockets around the nodal regions of the d-wave superconducting gap [108, 109] or due to confined nature of the vortex structure [110]. In another interpretation, the spin excitations in the dx 2 −y 2 superconductor condense at high fields into the ground state, thereby creating a novel superconducting state [111]. The debate which of these approaches best describes the SDW phase in CeCoIn5 is still ongoing. The coexistence of magnetic and superconducting order parameters has made it difficult to unambiguously identify the nature of the Q-phase and its origin. On a phenomenological level, it is an important observation that the onset of the Q-phase inside the superconducting phase occurs through a second-order phase transition [112]. This is in contrast to many of the FFLO-based scenarios that predict a first-order transition that separates a uniform d-wave phase from a high-field modulated d-wave phase. Clearly, further analysis is called for.

    Ferromagnetic Superconductors Coexistence of superconductivity and ferromagnetism has recently been discovered [17, 18, 113] in several uranium-based compounds, UGe2 , URhGe, and UCoGe. This points towards an unconventional spin-triplet Cooper pairing mechanism in these systems. In the first two compounds, the Curie temperature, TC , is higher than their critical temperatures for superconductivity Tsc by more than an order of magnitude, while in UCoGe, TC /Tsc = 4 at ambient pressure (see Fig. 11). Also, the upper critical field at low temperatures greatly exceeds the paramagnetic

    13 Magnetism and Superconductivity

    649

    Fig. 11 Temperature-pressure phase diagram of several ferromagnetic superconductors, (a) UGe2 , (b) URhGe, and (c) UCoGe. FM, SC, and PM stand for ferromagnetic, superconducting, and paramagnetic phases, respectively; the data points are taken from Ref. [114]. Originally published in J. Phys. Soc. Jpn. 83, 061011 (2014). © 2014 The Author(s)

    limit field in the first three compounds [114]. Further ferromagnetism does not suppress superconductivity with triplet pairing, and no traces of spatial modulation of magnetic moment directions on a scale smaller than the coherence length have been revealed [18]. It is therefore natural to assume that these ferromagnetic superconductors are triplet superconductors similar to superfluid phases of 3 He. One has to keep in mind, however, that unlike liquid helium, which is a completely isotropic neutral Fermi liquid, superconductivity in ferromagnetic superconductors develops in strongly anisotropic ferromagnetic metals. In particular, all these materials have an orthorhombic structure with the magnetic moment oriented either along the a-axis like it is in UGe2 or along the c-axis in URhGe and UCoGe. The magnetic moments are mostly concentrated around uranium ions, and at T → 0, they are equal to 1.4μB in UGe2 [115], 0.4μB in URhGe [116], and 0.07μB in UCoGe [117]. Although these values are much smaller than the moment per uranium atom as deduced from the uniform susceptibilities above TC , this is still not sufficient to treat uranium compounds as completely itinerant ferromagnets. Instead, they are dual localized and itinerant ferromagnets. The interaction between conduction electrons by means of spin waves in a system of localized moments is considered to be the most plausible pairing mechanism. Models of this type have been previously applied to the superconducting antiferromagnet UPd2 Al3 [118] and to reentrant superconduc- tivity in the ferromagnetic URhGe [119]. In application to the orthorhombic ferromagnets, the general structure of the order parameter, dictated by symmetry, was proposed by [120, 121]. Subsequently, the microscopic formulation based on the pairing interaction due to exchange of magnetization fluctuations in orthorhombic ferromagnets with strong magnetic anisotropy was also recently formulated [122, 123]. In the following, we quickly outline the details of this consideration assuming Tsc < TC . In the ferromagnetic state, the simplest model system consists of two bands, ε↓ (k) and ε↑ (k), for the spin-down and spin-up bands, respectively. The spin-triplet superconducting state arising in a ferromagnetic metal consists of spin-up, spindown, and zero-spin Cooper pairs described by the vector order parameter

    650

    I. M. Eremin et al.

    Δ(k, r) = Δ↑ (k, r)| ↑↑ + Δ0 (k, r)(| ↑↓ + | ↓↑) + Δ↓ (k, r)| ↓↓,

    (31)

    where Δ↑ (k, r), Δ0 (k, r), and Δ↓ (k, r) are the spin-up, zero-spin, and spin-down amplitudes of the superconducting order parameter, depending on the Cooper pair center of gravity coordinate r and the momentum k of pairing electrons. Equivalently, one can write Δ(k, r) = (d(k, r)σ )iσ y ,

    (32)

    with σ = (σ x , σ y , σ z ) being the Pauli matrices and the complex vector d(k, r) =

     1 −Δ↑ (k, r)(ˆx + i yˆ ) + Δ↓ (k, r)(ˆx − i yˆ ) + Δ0 (k, r)ˆz, 2

    (33)

    and xˆ , yˆ , and zˆ are the unit vectors along the corresponding coordinate axes. Most of the ferromagnetic superconductors are ferromagnetic orthorhombic crystals with a strong spin-orbit coupling fixing the spontaneous magnetization along one of the symmetry axes chosen as the z-direction. Superconducting states with different critical temperatures are described by the basis functions of the different irreducible representations of the symmetry group of the normal state of the crystal [124]. There are only two different representations, dubbed A and B of the group GF M [120, 121]. The resulting vector order parameters are five-component dA (k, r) and four-component dB (k, r). In particular, the symmetry considerations [120] give the following form: ↑ ΔA (k, r) = kˆx ηx↑ (r) + i kˆy ηy↑ (r) ↓ ΔA (k, r) = kˆx ηx↓ (r) + i kˆy ηy↓ (r)

    (34)

    Δ0A (k, r) = kˆz ηz0 (r) and ↑ ΔB (k, r) = kˆz ζz↑ (r) ↓ ΔB (k, r) = kˆz ζz↓ (r)

    (35)

    Δ0B (k, r) = kˆx ζx0 (r) + i kˆy ζy0 (r) ˆ It is where kˆx , kˆy , and kˆz are the components of the unit momentum vector k. important to note that the five-component order parameter of the A state and the four-component order parameter of the B state include by symmetry the zero-spin components. In other words, they are not equal-spin-pairing states consisting of Cooper pairs with opposite spins, as one would expect naively for the ferromagnetic state. This property originates from the spin-orbit coupling. However, although

    13 Magnetism and Superconductivity

    651

    the pairing amplitudes for the zero-spin components can indeed arise due to the spin-orbit terms in the ferromagnet gradient energy, they were shown on the basis of microscopic considerations to be small [125]. Therefore, one can in the first approximation ignore this amplitude Δ0 in the A and B states. In this case, one deals with a two-band superconducting state similar to the A2 state of superfluid 3 He [126]. Microscopically, the interaction between two electrons driving the Cooper pairing is assumed to be due to the attraction of one electron by the magnetic polarization cloud of the other. The pairing of electrons in a ferromagnetic metal occurs in an anisotropic medium due to polarization of the electron liquid and the localized moments. This theory in application to the ferromagnetic superconductors was elaborated in the context of ferromagnetic superconductors [123, 125]. It is important to mention that the results of the microscopic considerations seem to agree with the pure symmetry considerations, yet the experimental verification of the symmetry of the superconducting order parameter is still absent. Part of the complication is that in the ferromagnetic superconductor, there is an internal field Hint acting on the electron charges even in the absence of an external field. The internal magnetic field in all uranium ferromagnets is larger than the lower critical field Hc1 . Hence, the Meissner state is absent and the superconductor is in an Abrikosov vortex state with spatially inhomogeneous distributions of the order parameter and the internal magnetic field, which complicates the analysis. Nevertheless, a comprehensive state-of-the-art review on the subject can be found in [125].

    Conclusions The interplay of magnetism and superconductivity is one of the most striking features of the quantum mechanical description of solids. Originally thought to be mutually exclusive, both phenomena in fact show interesting coexistence, whose investigation remains a fascinating topic in contemporary condensed matter physics. In this short review, we have aimed to provide some introduction to this exciting field of research. We have first analyzed the FFLO state, which appears in the spin-singlet superconducting state in the particular case when the Pauli-limiting field, Hp , is lower than the expected upper critical field, Hc2 , determined from orbital effects. As a result, there appears a modulation in the superconducting order parameter, which is related to the Zeeman splitting of the electron’s energy level under a magnetic field (or exchange field for ferromagnets) acting on electron spins. Experimentally, such a state is seen in some organic superconductors and more recently in some iron-based superconductors. We then analyzed the interplay between spin-orbit coupling and an external, or an internal exchange, magnetic field acted in proximity to conventional superconductors as a platform for the formation of topological superconductivity with (localized) surface zero-energy excitations. Due to particlehole symmetry in a superconductor within the energy scale of the superconducting gap, these bound states are described in terms of Majorana fermions. Another recent

    652

    I. M. Eremin et al.

    material advance is provided by the two-dimensional superconductors with broken inversion symmetry, which together with a sizeable spin-orbit coupling gives rise to a mixture of spin-singlet and spin-triplet superconductivity. Although this effect is known previously for bulk superconducting, its presence in these 2D materials (known as Ising superconductivity) offers another interesting line of research, which will be studied extensively in the future. We have then described the peculiarities of coexistence between antiferromagnetic order and spin-singlet superconductivity. As their coexistence is not restructured by any symmetry arguments, there are multiple examples ranging from conventional to unconventional superconductors, where both orders are adjacent on the phase diagram. Furthermore, there is a strong belief that antiferromagnetism and spin-singlet superconductivity with sign changing gaps on parts of the Fermi surface (denoted as unconventional superconductors, as the mechanism of the Cooper pairing is not due to electron-phonon interaction) are a result of the very same underlying magnetic interaction and therefore naturally appear together in those systems. The examples are heavy-fermion systems, layered cuprates, and iron-based superconductors. Finally, we have also given a brief introduction to the field of ferromagnetic superconductors, which in contrast to the FFLO state are uniform spin-triplet superconductors on a background of the orthorhombic ferromagnetic state. These fields of research are very dynamic, and the most recent progress in this area should serve as both a theoretical and experimental guide in the search for novel systems exhibiting cooperative behavior of magnetism and superconductivity.

    References 1. Ginzburg, V.L.: Zh.Eksp. Teor. Fiz. 31, 202 (1956) [Sov. Phys. JETP 4, 153 (1957)] 2. Bardeen, J., Cooper, L.N., Schrieffer, J.R.: Phys. Rev. 106, 162 (1957) 3. Saint-James, D.D., Sarma, D., Thomas, E.J.: Type II Superconductivity. Pergamon, New York (1969) 4. Anderson, P.W., Suhl, H.: Phys. Rev. 116, 898 (1959) 5. Larkin, A.I., Ovchinnikov, Y.N.: Zh. Eksp. Teor. Fiz. 47, 1136 (1964) [Sov. Phys. JETP 20, 762 (1965)] 6. Fulde, P., Ferrell, R.A.: Phys. Rev. 135, A550 (1964) 7. Maple, M.B., Fisher, O. (eds.): Superconductivity in Ternary Compounds II, Topics in Current Physics. Springer, Berlin (1982) 8. Müller, K.-H., Narozhnyi, V.N.: Rep. Prog. Phys. 64, 943 (2001) 9. Pfleiderer, C.: Rev. Mod. Phys. 81, 1551 (2009) 10. Stewart, G.R.: Rev. Mod. Phys. 83, 1589 (2011) 11. Johnston, D.C.: Adv. Phys. 59, 803 (2010) 12. Hirschfeld, P.J., Korshunov, M.M., Mazin, I.I.: Rep. Prog. Phys. 74, 124508 (2011) 13. Mazin, I.I., Schmalian, J.: Phys. C 469, 614 (2009) 14. Scalapino, D.J.: Rev. Mod. Phys. 84, 1383 (2012) 15. Chubukov, A.V.: Annual Rev. Cond. Matt. Phys. 3, 57 (2012) 16. Fernandes, R.M., Schmalian, J.: Phys. Rev. B 82, 014521 (2010) 17. Saxena, S.S., Agarwal, P., Ahilan, K., Grosche, F.M., Haselwimmer, R.K.W., Steiner, M.J., Pugh, E., Walker, I.R., Julian, S.R., Monthoux, P., Lonzarich, G.G., Huxley, A., Sheikin, I., Braithwaite, D., Flouquet, J.: Nature (London) 406, 587 (200)

    13 Magnetism and Superconductivity

    653

    18. Aoki, D., Huxley, A., Ressouche, E., Braithwaite, D., Flouquet, J., Brison, J.-P., Lhotel, E., Paulsen, C.: Nature (London) 413, 613 (2001) 19. Buzdin, A.I.: Rev. Mod. Phys. 77, 935 (2005) 20. Alicea, J.: Rep. Prog. Phys. 75, 076501 (2012) 21. Stanescu, T.D., Tewari, S.: J. Phys. Condens. Matter 25, 233201 (2013) 22. Sato, M., Ando, Y: Rep. Prog. Phys. 80, 076501 (2017) 23. Chandrasekhar, B.S.: Appl. Phys. Lett. 1, 7 (1962) 24. Clogston, A.M.: Phys. Rev. Lett. 9, 266 (1962) 25. Aslamazov, L.G.: Zh. Eksp. Teor. Fiz. 55, 1477 (1968) [Sov. Phys. JETP 28, 773 (1969)] 26. Matsuda, Y., Shimahara, H.: J. Phys. Soc. Jpn. 76, 051005 (2007) 27. Bulaevskii, L.N.: Zh. Eksp. Teor. Fiz. 64, 2241 (1973) [Sov. Phys. JETP 37, 1133 (1973)] 28. Cho, C.-w., Yang, J.H., Yuan, N.F.Q., Shen, J., Wolf, T., Lortz, R.: Phys. Rev. Lett. 119, 217002 (2017) 29. Buzdin, A.I., Kachkachi, H.: Phys. Lett. A 225, 341 (1999) 30. Houzet, M., Meurdesoif, Y., Coste, O., Buzdin, A.: Phys. C 316, 89 (1999) 31. Kenzelmann, M.: Rep. Prog. Phys. 80 034501 (2017) 32. Jerome, D., Mazaud, A., Ribault, M., Bechgaard, K.: J. Phys. Lett. 41 L95 (1980) 33. Urayama, H., Yamochi, H., Saito, G., Nozawa, K., Sugano, T., Kinoshita, M., Sato, S., Oshima, K., Kawamoto, A., Tanaka, J.: Chem. Lett. 17, 55 (1988) 34. Williams, J.M., Kini, A.M., Wang, H.H., Carlson, K.D., Geiser, U., Montgomery, L.K., Pyrka, G.J., Watkins, D.M., Kommers, J.M.: Inorg. Chem. 29, 3272 (1990) 35. Kini, A.M., Geiser, U., Wang, H.H., Carlson, K.D., Williams, J.M., Kwok, W.K., Vandervoort, K.G., Thompson, J.E., Stupka, D.L.A.: Inorg. Chem. 29 2555 (1990) 36. Taylor, O.J., Carrington, A., Schlueter, J.A.: Phys. Rev. Lett. 99, 057001 (2007) 37. Wright, J.A., Green, E., Kuhns, P., Reyes, A., Brooks, J., Schlueter, J., Kato, R., Yamamoto, H., Kobayashi, M., Brown, S.E.: Phys. Rev. Lett. 107, 087002 (2001) 38. Singleton, J., Symington, J.A., Nam, M.-S., Ardavan, A., Kurmoo, M., Day, P.: J. Phys. Condens. Matter 12, L641 (2000) 39. Lortz, R., Wang, Y., Demuer, A., Böttger, P.H.M., Bergk, B., Zwicknagl, G., Nakazawa, Y., Wosnitza, J.: Phys. Rev. Lett. 99, 187002 (2007) 40. Bergk, B., Demuer, A., Sheikin, I., Wang, Y., Wosnitza, J., Nakazawa, Y., Lortz, R.: Phys. Rev. B 83 064506 (2011) 41. Mayaffre, H., Krämer, S., Horvatic, M., Berthier, C., Miyagawa, K., Kanoda, K., Mitrovic, V.F.: Nat. Phys. 10, 928 (2014) 42. Wosnitza, J.: Ann. Phys. (Berlin) 2017, 1700282 (2017) 43. Beyer, R., Bergk, B., Yasin, S., Schlueter, J.A., Wosnitza, J.: Phys. Rev. Lett. 109, 027003 (2012) 44. Kasahara, S., Watashige, T., Hanaguri, T., Kohsaka, Y., Yamashita, T., Shimoyama, Y., Mizukami, Y., Endo, R., Ikeda, H., Aoyama, K., Terashima, T., Uji, S., Wolf, T., Löhneysen, H.V., Shibauchi, T., Matsuda, Y.: Proc. Natl. Acad. Sci. USA 111, 16309–16313 (2014) ˇ 45. Kasahara, S., Sato, Y., Licciardello, S., Culo, M., Arsenijevi´c, S., Ottenbros, T., Tominaga, T., Böker, J., Eremin, I., Shibauchi, T., Wosnitza, J., Hussey, N.E., Matsuda, Y.: Phys. Rev. Lett. 24, 107001 (2020) 46. Gurevich, A.: Phys. Rev. B 82, 184504 (2010) 47. Jin, B.: Physica (Amsterdam) C 468, 2378 (2008) 48. Beenakker, C.W.J.: Annu. Rev. Condens. Matter Phys. 4 113 (2013) 49. Nayak, C., Simon, S.H., Stern, A., Freedman, M., Das Sarma, S.: Rev. Mod. Phys. 80, 1083 (2008) 50. Lutchyn, R.M., Sau, J.D., Das Sarma, S.: Phys. Rev. Lett. 105, 077001 (2010) 51. Oreg, Y., Refael, G., von Oppen, F.: Phys. Rev. Lett. 105, 177002 (2010) 52. Kitaev, A.Y.: Usp. Fiz. Nauk (Suppl.) 171 (10) (2001) [A.Y. Kitaev, Physics-Uspekhi 44, 131 (2001)] 53. Lutchyn, R.M., Bakkers, E.P.A.M., Kouwenhoven, L.P., Krogstrup, P., Marcus, C.M., Oreg, Y.: Nat. Rev. Mater. 3, 52 (2018)

    654

    I. M. Eremin et al.

    54. Zhang, H., Liu, C., Gazibegovic, S., et al.: Nature 556, 74 (2018) 55. Sestoft, J.E., Kanne, T., Norskov Gejl, A., von Soosten, M., Yodh, J.S., Sherman, D., Tarasinski, B., Wimmer, M., Johnson, E., Deng, M., Nygard, J., Sand Jespersen, T., Marcus, C.M., Krogstrup, P.: Phys. Rev. Materials 2, 044202 (2018) 56. Fornieri, A., Whiticar, A.M., Setiawan, F., Marin, E.P., Drachmann, A.C.C., Keselman, A., Gronin, S., Thomas, C., Wang, T., Kallaher, R., Gardner, G.C., Berg, E., Manfra, M.J., Stern, A., Marcus, C.M., Nichele, F.: Nature 569, 89 (2019) 57. Jäck, B., Xie, Y., Li, J., Jeon, S., Andrei Bernevig, B., Yazdani, A.: Science 364, 1255 (2019) 58. Manna, S., Wei, P., Xie, Y., Law, K.T., Lee, P.A., Moodera, J.S.: Proc. Nat. Acad. Sci. 117, 8775 (2020) 59. Gibertini, M., Koperski, M., Morpurgo, A.F., Novoselov, K.S.: Nature Nanotechnology 14, 408 (2019) 60. Ugeda, M.M., Bradley, A.J., Zhang, Y., Onishi, S., Chen, Y., Ruan, W., Ojeda-Aristizabal, C., Ryu, H., Edmonds, M.T., Tsai, H.-Z., Riss, A., Mo, S.-K., Lee, D., Zettl, A., Hussain, Z., Shen, Z.-X., Crommie, M.F.: Nat. Phys. 12, 92 (2016) 61. Dvir, T., Massee, F., Attias, L., Khodas, M., Aprili, M., Quay, C.H.L., Steinberg, H.: Nat. Commun. 9, 598 (2018) 62. de la Barrera, S.C., Sinko, M.R., Gopalan, D.P., Sivadas, N., Seyler, K.L., Watanabe, K., Taniguchi, T., Tsen, A.W., Xu, X., Xiao, D., Hunt, B.M.: Nat. Commun. 9, 1427 (2018) 63. Smidman, M., Salamon, M.B., Yuan, H.Q., Agterberg, D.F.: Rep. Prog. Phys. 80, 036501 (2017) 64. Gor’kov, L.P., Rashba, E.I.: Phys. Rev. Lett. 87, 037004(2001) 65. Frigeri, P.A., Agterberg, D.F., Koga, A., Sigrist, M.: Phys.Rev.Lett. 92, 097001 (2004) 66. Bulaevskii, L.N., Guseinov, A.A., Rusinov, A.I.: Zh. Eksp.Teor. Fiz. 71, 2356 (1976); [Sov. Phys. JETP 44, 1243 (1976)] 67. Samokhin, K.V.: Phys. Rev. B 78, 224520 (2008) 68. Ilic, S., Meyer, J.S., Houzet, M.: Phys. Rev. Lett. 119, 117001 (2017) 69. Möckli, D., Khodas, M.: Phys. Rev. B 98, 144518 (2018) 70. Möckli, D., Khodas, M.: Phys. Rev. B 99, 180505(R) (2019) 71. Möckli, D., Khodas, M.: Phys. Rev. B 101, 014510 (2020) 72. Xi, X., Wang, Z., Zhao, W., Park, J.-H., Law, K.T., Berger, H., Forro, L., Shan, J., Mak, K.F.: Nat. Phys. 12, 139 (2016) 73. Saito, Y., Nakamura, Y., Bahramy, M.S., Kohama, Y., Ye, J., Kasahara, Y., Nakagawa, Y., Onga, M., Tokunaga, M., Nojima, T., Yanase, Y., Iwasa, Y.: Nat. Phys. 12, 144 (2016) 74. Liu, Y., Wang, Z., Zhang, X., Liu, C., Liu, Y., Zhou, Z., Wang, J., Wang, Q., Liu, Y., Xi, C., Tian, M., Liu, H., Feng, J., Xie, X.C., Wang, J.: Phys. Rev. X 8, 021002 (2018) 75. Sohn, E., Xi, X., He, W.-Y., Jiang, S., Wang, Z., Kang, K., Park, J.-H., Berger, H., Forro, L., Law, K.T., Shan, J., Mak, K.F.: Nat. Mater. 17, 504 (2018) 76. Nakata, Y., Sugawara, K., Ichinokura, S., Okada, Y., Hitosugi, T., Koretsune, T., Ueno, K., Hasegawa, S., Takahashi, T., Sato, T.: npj 2D Mater. Appl. 2, 12 (2018) 77. He, W.-Y., Zhou, B.T., He, J.J., Yuan, N.F., Zhang, T., Law, K.: Commun. Phys. 1, 1 (2018) 78. Heil, C., Ponce, S., Lambert, H., Schlipf, M., Margine, E.R., Giustino, F.: Phys. Rev. Lett 119, 087003 (2017) 79. Wickramaratne, D., Khmelevskyi, S., Agterberg, D.F., Mazin, I.I.: arXiv:2005.05497 80. Shaffer, D., Kang, J., Burnell, F.J., Fernandes, R.M.: arXiv:1904.05470 81. Kosterlitz, J.M., Nelson, D.R., Fisher, M.E.: Phys. Rev. B 13, 412 (1976) 82. Aharony, A.: J. Stat. Phys. 110, 659 (2003) 83. Baltensperger, W., Strassler, S.: Phys. Kondens. Mater. 1, 20 (1963) 84. Bulaevskii, L.N., Rusinov, A.I., Kulic, M.: J. Low Temp. Phys. 39, 255 (1980) 85. Nass, M.J., Levin, K., Grest, G.S.: Phys. Rev. Lett. 46, 614 (1981) 86. Machida, K.: J. Phys. Soc. Jpn. 50, 2195 (1981) 87. Gulacsi, M., Gulacsi, Z.S.: Phys. Rev. B 33, 6147 (1986) 88. Gabovich, A.M., Shpigel, A.S.: J. Phys. F: Met. Phys. 14, 3031 (1984) 89. Kato, M., Machida, K.: Phys. Rev. B 37, 1510 (1988)

    13 Magnetism and Superconductivity

    655

    90. Schrieffer, J.R., Wen, X.G., Zhang, S.C.: Phys. Rev. B 39, 11663 (1989) 91. Amici, A., Thalmeier, P., Fulde, P.: Phys. Rev. Lett. 84, 1800 (2000) 92. Das, T., Markiewicz, R.S., Bansil, A.: Phys. Rev. B 77, 134516 (2008) 93. Parker, D., Vavilov, M.G., Chubukov, A.V., Mazin, I.I.: Phys. Rev. B 80, 100508(R) (2009) 94. Vorontsov, A.B., Vavilov, M.G., Chubukov, A.V.: Phys. Rev. B 81, 174538 (2010) 95. Knolle, J., Eremin, I., Schmalian, J., Moessner, R.: Phys. Rev. B 84, 180510(R) (2011) 96. Hinojosa, A., Fernandes, R.M., Chubukov, A.V.: Phys. Rev. Lett. 113, 167001 (2014) 97. Schmiedt, J., Brydon, P.M.R., Timm, C.: Phys. Rev. B 89, 054515 (2014) 98. Roemer, A.T., Eremin, I., Hirschfeld, P.J., Andersen, B.M.: Phys. Rev. B 93, 174519 (2016) 99. Scalapino, D.J., Loh, E. Jr., Hirsch, J.E.: Phys. Rev. B 34, 8190 (1986) 100. Rowe, W., Eremin, I., Rømer, A.T., Andersen, B.M., Hirschfeld, P.J.: New J. Phys. 17, 023022 (2015) 101. Frenkel, D.M., Hanke, W.: Phys. Rev. B 42, 6711 (1990) 102. Lu, Y.-M., Xiang, T., Lee, D.-H.: Nat. Phys. 10, 634 (2014) 103. Kuchiev, M.Yu., Sushkov, O.P.: Phys. C 218, 197 (1993) 104. Yanase, Y., Sigrist, M.: J. Phys. Soc. Jpn. 78, 114715 (2009) 105. Yanase, Y., Sigrist, M.: J. Phys. Condens. Matter 23, 094219 (2011) 106. Yanase, Y., Sigrist, M.: J. Phys. Soc. Jpn. 80, SA005 (2011) 107. Yanase, Y., Sigrist, M.: J. Phys. Soc. Jpn. 80, 094702 (2011) 108. Kato, Y., Batista, C.D., Vekhter, I.: Phys. Rev. Lett. 107, 096401 (2011) 109. Kato, Y., Batista, C.D., Vekhter, I.: Phys. Rev. B 86, 174517 (2012) 110. Suzuki, K.M., Ichioka, M., Machida, K.: Phys. Rev. B 83, 140503 (2011) 111. Michal, V.P., Mineev, V.P.: Phys. Rev. B 84, 052508 (2011) 112. Tokiwa, Y., Bauer, E.D., Gegenwart, P.: Phys. Rev. Lett. 109, 116402 (2012) 113. Huy, N.T., Gasparini, A., de Nijs, D.E., Huang, Y., Klaasse, J.C.P., Gortenmulder, T., de Visser, A., Hamann, A., Görlach, T., Löhneysen, H.V.: Phys. Rev. Lett. 99, 067006 (2007) 114. Aoki, D., Flouquet, J.: J. Phys. Soc. Jpn. 83 061011 (2014) 115. Kernavanois, N., Grenier, B., Huxley, A., Ressouche, E., Sanchez, J.P., Flouquet, J.: Phys. Rev. B 64 174509 (2001) 116. Prokes, K., Gukasov, A., Takabatake, T., Fujita, T., Sechovsky, V.: Acta Phys. Polon. B 34, 1473 (2003) 117. Prokes, K., de Visser, A., Huang, Y.K., Fak, B., Ressouche, E.: Phys. Rev. B 81, 180407(R) (2010) 118. McHale, P., Fulde, P., Thalmeier, P.: Phys. Rev. B 70, 014513 (2004) 119. Hattori, K., Tsunetsugu, H.: Phys. Rev. B 87 064501 (2013) 120. Mineev, V.P.: Phys. Rev. B 66, 134504 (2002) 121. Mineev, V.P., Champel, T.: Phys. Rev. B 69, 144521 (2004) 122. Mineev, V.P.: Phys. Rev. B 83, 064515 (2011) 123. Mineev, V.P.: Phys. Rev. B 90, 064506 (2014) 124. Sigrist, M., Ueda, K.: Rev. Mod. Phys. 63, 239 (1991) 125. Mineev, V.P.: Phys. Uspechi 60, 121 (2017) 126. Ambegaokar, V., Mermin, N.D.: Phys. Rev. Lett. 30, 81 (1973) 127. J. Dod, in The Dictionary of Substances and Their Effects, Royal Society of Chemistry. (Available via DIALOG, 1999), 128. Ismer, J.-P., Eremin, I., Rossi, E., Morr, D.K., Blumberg, G.: Phys. Rev. Lett. 105, 037003 (2010) 129. Terashima, T., Kikugawa, N., Kiswandhi, A., Choi, E.-S., Brooks, J.S., Kasahara, S., Watashige, T., Ikeda, H., Shibauchi, T., Matsuda, Y., Wolf, T., Böhmer, A.E., Hardy, F., Meingast, C., Löhneysen, H.V., Uji, S.: Phys. Rev. B 90, 144517 (2014) 130. F. Ahn, I. Eremin, J. Knolle, V. B. Zabolotnyy, S. V. Borisenko, B. Büchner, and A. V. Chubukov, Phys. Rev. B 89, 144513, (2014)

    Part II Magnetic Materials

    Magnetism of the Elements

    14

    Plamen Stamenov

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Magnetism of Iron, Cobalt and Nickel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Band Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thin Films of Fe, Co and Ni . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Iron, Steels and Other Iron-Based Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Phase Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Manganese and Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Manganese . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Density Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Rare Earths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetism of the Rare Earths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Structures and Phase Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fermi-Level Spin Polarisation of the Magnetic Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Polarisation of the 3d Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Polarisation of the 4f Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Molecular Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Other Examples of Magnetic Order in the p- and d-Shell Elements . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    660 662 662 664 664 664 664 667 667 668 669 671 671 677 678 680 682 685 685 689 691

    Abstract

    This chapter presents the magnetic properties of the elements in relation to their magnetic structure, with emphasis on those that order magnetically in

    P. Stamenov () School of Physics and CRANN, Trinity College, University of Dublin, 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_15

    659

    660

    P. Stamenov

    bulk form – iron, cobalt, nickel, manganese, chromium, most of the rare earths and oxygen. All except oxygen are metals. The importance of spin polarisation at the Fermi level is illustrated by the three room-temperature ferromagnets: iron, cobalt and nickel. Manganese and chromium are atypical antiferromagnets; α-Mn is a multi-sublattice antiferromagnet, β-Mn is a spin liquid, and Cr exhibits an incommensurate spin density wave. A summary of the magnetism of the rare earth elements emphasises the effects of the crystal field, their moments and exchange integrals on their magnetic structures and phase transitions. Some magnetoelastic effects are discussed, and a detailed account of the spin polarisation of the heavy rare earths concludes the section. Finally, more exotic forms of magnetism such as the molecular antiferromagnetism of oxygen and defect-induced magnetism, exhibited by p and d-shell elements, such as carbon and ruthenium, are presented.

    Introduction This is a short tale of the very foundations of material magnetism – the properties of the magnetic elements. It is also a rather confined story; for as far as possible, it is restricted to the properties of the essentially pure elements in their bulk form, rather than their compounds and alloys. Crystal structure is a critical factor for determining the magnetism of a solid. A glance at the magnetic periodic table in Fig. 1 reveals that most of the elements possess a magnetic moment when they exist as free atoms. The moment is determined by Hund’s rules. Only those elements with filled s, p, d, or f shells or an electronic configuration such as p2 or d 4 , where L and S cancel, have no moment to call their own. But everything changes when they enter the community of a solid. Bonding with their neighbours often effaces all traces of their former identity and eliminates their electrons with unpaired spin, be it covalent or metallic. The moments survive to order ferromagnetically below a Curie point TC or antiferromagnetically below a Néel point TN in just 16 of the 88 stable elements. Only four of these order above room temperature, the ferromagnets (Fe, Co, Ni and Gd) and the antiferromagnet (Cr), as vividly illustrated on Fig. 1. The importance of crystal structure is underlined by the behaviour of different allotropes of the same element; α- or γ -Fe, α- or β-Mn, β- or γ -Ce, for example, each have quite different magnetic properties. A common, yet striking comparison can be made of the magnetic and superconducting periodic Table 2. Despite both superconductivity and ferromagnetic order being critical phenomena occurring below some well-defined temperature, superconductivity is twice as common amongst the pure element, as can be seen from an inspection of Fig. 2, where at least 32 pure elements have non-vanishing superconducting Tc , albeit at temperatures that are much lower than their magnetic counterparts. No element exhibits both types of order. This, of course, reflects the quantum mechanical origins of these two phenomena. While in the case of elemental superconductivity, the underpinning requirements are to have a metal, which the vast majority of elements are (having a partially filled valence band) and

    14 Magnetism of the Elements

    661

    Fig. 1 The magnetic periodic table. (After: [1])

    Fig. 2 The periodic table of the superconducting transition temperatures. (After: www. webelements.com)

    662

    P. Stamenov

    strong-enough electron-phonon coupling – something satisfied best by ‘bad’ metals that have relatively high room-temperature electrical resistivity. The requirements for magnetism are substantially different - namely, to have a partially filled 3d or 4f orbitals, with very few exceptions, which will be discussed later.

    The Magnetism of Iron, Cobalt and Nickel The magnetism of the 3d metals is a subject of discussion in many well-established textbooks [1]. The conventional approach is to start from a band picture of the two bands most strongly present at the Fermi level – 3d and 4s. While the dispersions of the 3d states are rather flat, with an effective mass components substantially bigger than m0 and the corresponding partial density of states (PDOS) has peaks at or near EF , the 4s states are substantially more mobile, with steeper dispersions and effective masses of about or less than m0 . The two families of states hybridise courtesy of the s-d exchange interactions, on the scale of couple of eV below and above EF . The s band dispersion starts somewhat lower, about −5 to −4 eV below the Fermi level, while the bandwidth and spin-splitting of the d-states are both of the order of 1 eV. Within the conventional Stoner picture for the occurrence of spontaneous ordering in metallic systems, the energy gain due to the exchange interactions upon polarising or spin-splitting the states at or near the Fermi level must exceed the energy cost of pushing what is to become the majority spin band below EF . This is traditionally expressed with the help of the Stoner criterion: SD (EF ) > 1

    (1)

    where S is the Stoner exchange parameter, valued at ∼1 eV for the d-states.

    Band Structures It is instructive to look into the detail of the DOS for the three primary 3d ferromagnets and analyse the differences that exist between them. In the case of iron, while the overall difference of the spin-up and spin-down DOS, integrated from EF down, and therefore the net available magnetic moment is the largest, the DOS in the vicinity of EF is actually the smallest. Courtesy of the large exchange splitting of the peak of the d states ∼2.8 eV, a lot of the available DOS for spin-down is unoccupied, while the peak of the spin-up states is pushed to about 1 eV below EF . This results in a reasonably high positive DOS polarisation P > 50%. The situation is different for Co, as of both the different crystal symmetry (hcp), rather than bcc, and the different total electron count. The exchange splitting of the peak of the d states is substantially smaller or about 2 eV. The purely DOS spin polarisation is negative, P ∼ −70%.

    14 Magnetism of the Elements

    663

    Out of the three primary elemental ferromagnets, Ni suffers the most of the high occupancy of d states, at 3d 8 . This minimises further the exchange splitting of the peak of the d states to approximately 1 eV and minimises further the net disbalance of the spin-up and spin-down occupied DOS, resulting in the smallest magnitudes of the net polarisation and Curie temperature in the tetrad. More surprising is the fact that the purely DOS polarisation is rather large and negative at P ∼ −80%, as the Fermi level is ‘pinned’ at the large peak of the d-states spin-down PDOS. This large negative polarisation has been the source of much controversy in the early efforts for matching theoretical predictions of the band structure with experiment (Fig. 3).

    Fig. 3 Calculated density of states for the three primary 3d ferromagnetic metals. (After: [2])

    664

    P. Stamenov

    Table 1 Magnetic properties of Iron, Co and Ni Element Fe Co Ni

    Structure BCC HCP FCC

    Density (kg/m3 ) 7874 8836 8902

    Lattice parameter (pm) 287 251 352

    Tc (K) 1044 1388 628

    Ms (MA/m) 1.71 1.45 0.49

    K1 ( kJ/m3 ) 48 530 −5

    λs (10−6 ) −7 −62 −34

    Magnetic Properties Thin Films of Fe, Co and Ni While thin films of the pure elements are rarely used for spin electronic applications, perhaps with the exception of epitaxial Fe, grown on MgO, the 3d random alloys and ordered phases are critical building blocks for modern devices. The reader is referred to  Chap. 16, “Metallic Magnetic Thin Films,” for details on these (Table 1).

    Iron, Steels and Other Iron-Based Alloys Phase Diagram Soft Steels Electrical Steel Silicon is a common ternary addition to low-carbon steel, which helps to suppress and delay the α → γ transition from ferrite to austenite, normally occurring about 912◦ C. As can be seen in the phase diagram in Fig. 5, the nominal phase border is at 3.8 at. %, for otherwise pure iron. About 6 at. % of Si are usually added, to account for leftover carbon, which substantially extends the austenitic stability region, potentially all the way down to 727◦ C (see Fig. 4). The discovery of the process is attributed to Robert Hadfield, around 1900, who had realised that the compositions around 6 at % Si are sufficiently ductile to hot-roll to thin sheets. The optimal thickness depends largely of the frequency band of application and the skin depth, typically at main frequencies of 50–60 Hz. The largest amounts (about 7 B$) are produced with thickness of about 300 μm. There are additional gains as Si lowers both the anisotropy and the magnetostriction in mild steel, while increasing its electrical resistivity and corrosion resistance. Goss Texture While in motor and generator applications isotropic microstructure is appropriate, as the direction of the conducted flux can change substantially during a cycle of the rotary machine (typically close to the ends of the pole pieces), in transformer applications, the direction of the conducted flux stays locally largely restrained onto a particular axis. This is, of course, the case away from zero magnetic induction B, when changes in magnetic induction result in Eddy currents being induced perpendicular to core lamination. Providing a crystallographic texture and associated magnetocrystalline anisotropy can, indeed, be beneficial to

    14 Magnetism of the Elements

    665

    Fig. 4 The carbon-iron phase diagram, at carbon concentrations below the cementite border. (After: www.himikatus.ru/art and www.asminternational.com, Binary Alloy Phase Diagrams, 1986, Ed. T. B. Massalski, ISBN: 978-0-87170-403-0)

    the minimisation of the core losses. The additional restriction provided by the anisotropy, especially active at small B values, helps to keep the Eddy current density low and eases the process of domain growth along the easy axis (in-plane). Permalloy Permalloy is an all-purpose soft magnetic material that has been called ‘the fourth ferromagnetic element’. It is a partially ordered f cc alloy of Fe and Ni, with approximate composition Fe20 Ni80 . The saturation polarisation is a little more than 1 Tesla, but the alloy owes its success to the fortuitous cancellation of both magnetostriction and intrinsic anisotropy at the same composition – a lucky coincidence in a binary alloy; both change sign at a composition close to 80% Ni. The soft magnetic properties are essentially impervious to strain. The soft magnetic properties are improved by annealing and by small alloy additions of Cu and Mo, to make alloys known as Supermalloy and MuMetal which can have permeabilities as high as 105 and are very effective for magnetic shields. Permendur The name permendur is reserved for the soft magnetic alloys of iron and cobalt with compositions close to 50/50. It can contain up to 1% carbon and undergoes order-disorder transition at 730◦ C, as illustrated in Fig. 6. Modern compositions often have additional alloying elements, such as vanadium and silicon, used to increase the machinability by increasing the eutectoid temperature and

    666

    P. Stamenov

    1700

    0

    10

    20

    Weight Percent Silicon 30 40 50

    60

    70

    80

    90 100

    1538°C

    1500

    L 1410°C

    1394°C

    Temperature °C

    1300

    23.5

    19.5 (γ Fe)

    1212

    1203°C

    29.8

    3.8

    β

    1200

    1100

    1212°C

    1220°C 67 73.5

    50.8

    α2

    965°C 28.2

    1207°C

    ζα

    1060°C (α Fe)

    900

    1414°C

    982°C 70 70.5

    η

    912°C

    937°C

    825°C 770°C

    700

    ε

    α1

    (Si)

    ζβ

    Magnetic Trans.

    500 0 Fe

    10

    20

    30

    40 50 60 Atomic Percent Silicon

    70

    80

    90

    100 Si

    Fig. 5 The iron-silicon phase diagram. (After: www.himikatus.ru/art and www.asminternational. com, Binary Alloy Phase Diagrams, 1986, Ed. T. B. Massalski, ISBN: 978-0-87170-403-0)

    improving the ductility. The magnetic transition temperature can be as high as 965◦ C for slightly iron-rich compositions. Permendur has the highest saturation polarisation of all commercial soft magnetic alloys, with saturation induction in the range of 2.3–2.4 Tesla. It finds applications in specific motors and generators and rather importantly as the material of choice for the manufacturing of pole pieces of laboratory electromagnets. Hard Steels Special Steels Magnetic and Anti-magnetic Stainless Steel Stainless steels are generally separated into three main classes, based on their underlying crystalline structure – austenitic (f cc), ferritic (bcc) and martensitic (bct). Out of these only the first one can be classified as essentially non-magnetic or anti-magnetic. This is due to the austenitic structure being antiferromagnetic, with an ordering temperature, which is strongly dependent on the amount and type of the additive(s) used to stabilise the γ -phase, with Cr and Ni being the most commonly used ones, in concentrations of up to 20%. The transitions generally occur at

    14 Magnetism of the Elements

    1700

    0

    10

    20

    667

    30

    Weight Percent Iron 40 50 60

    70

    80

    90

    L 1500

    100

    1538°C

    1495°C (δFe) ∼33%,1476°C 1394°C

    Temperature °C

    1300

    (αCo,γFe)

    1121°C

    1100

    Ma Tra g. ns.

    900

    ∼55%,985°C ∼75

    ∼27

    912°C Ma Tra g. ns.

    (αFe) ∼50%,730°C

    770°C

    700 α1

    500

    0 Co

    10

    20

    30

    40 50 60 Atomic Percent Iron

    70

    80

    90

    100 Fe

    Fig. 6 The Cobalt-Iron phase diagram. (After: www.himikatus.ru/art and www.asminternational. com, Binary Alloy Phase Diagrams, 1986, Ed. T. B. Massalski, ISBN: 978-0-87170-403-0)

    cryogenic temperatures (5–50 K) and create significant problems for the application of stainless steels in the construction of sensitive magnetometers and their use in components designed for operation in high magnetic fields. The judicious choice of composition and thermal treatment allows for the delay of both the Neel transition and the transformation to the equilibrium martensitic or ferritic phase to be delayed, with rather large amount of super-cooling, all the way down to He temperatures [3].

    Manganese and Chromium Manganese Bulk manganese metal is a close to cubic antiferromagnet with rather complex ¯ atomic and magnetic structures. Above its TN = 95 K, α-Mn adopts a cubic I 43m with 58 atoms per unit cell. Below the transition temperature, it distorts tetragonally and adopts a non-collinear magnetic structure. The unusual set of properties originates in the competition between the tendencies to maximise bond strength and spin moment as dictated by Hund’s rules. Short Mn-Mn distances result in antiferromagnetic exchange and low moments and generally tend to quench magnetism.

    668

    P. Stamenov

    Fig. 7 The magnetic structure of α-Mn. (After: [4])

    Crystal Structure The crystal structure of α-Mn may be represented as a closed-packed intermetallic compound, containing strongly magnetic (MnI) and (MnII), the weaker (MnIII), and the even weaker (MnIV). Low-temperature magnetic moments are, respectively, 2.8, 1.8 and 0.5 μB . Atomic moments and exchange interactions in manganese and its alloys are very sensitive to the Mn-Mn separation (Fig. 4). Magnetic Structure The non-collinear magnetic structure is a result of the weakly magnetic (MnIV) being arranged on the triangular faces of polyhedra and exhibiting a frustrated antiferromagnetic coupling, somewhat similar to that of the triangular antiferromagnets. The other magnetic moments are rotated slightly from their otherwise collinear antiferromagnetic orientation as a consequence of their finite interaction with the (MnIV) [4] (Fig. 7).

    Chromium Crystal Structure Paramagnetic chromium has a relatively simple bcc crystal structure, with an a = 2.884 Å. This is, however, altered below its TN = 311 K, by effects such as exchange

    14 Magnetism of the Elements

    669

    striction. Since the details of the atomic order are in this case dependent on the spin order, it is useful to understand better the underlying reasons for the existence of the so-called charge density wave (CDW) state. Magnetic Structure Chromium is the prototypical 3d itinerant antiferromagnet. The underlying structure is that of an alternating (001) spin planes, with the ones containing the corner atoms opossing the ones passing through the body centring ones. Should this spin structure have been perfectly commensurate, it would have been described by [5] (Fig. 8):  comm = 2π [001] = 2π [001] = {0, 0, 1} Q aCr Λ

    (2)

    This would imply that for the commensurate structure and in thermodynamic equilibrium, three different domains would co-exist with the spins aligned along the three main crystallographic directions. Reality is somewhat more complicated, with an incommensurate linearly polarised spin density wave (I-SDW) structure, being observed below the Néel point. This is a sinusoidal modulation of the magnetic moment density of the type:    ± · r μ  (r ) = μ  0 sin Q

    (3)

    with μ0 being about 0.5 μB at helium temperatures. The incommensurate wave vector is defined as (Fig. 8):  ± = 2π Q

    

    1 1 − aCr ΛSDW

     [001] =

    2π (1 − δ) aCr

    (4)

    Both the commensurate and incommensurate states have been the subjects of very detailed experimental and theoretical investigations, imploying a variety of techniques, including neutron scattering and ultrasound attenuation, in order to characterise both the order parameters and their influence on the electronic structure. A detailed account of the state of investigations on elemental chromium [6] and its alloys [7] is given by Fawcett and collaborators.

    Spin Density Waves The microscopic theory of the formation of the commensurate SDW has been the subject of many debates, but the underlying principle seems to be related to

    670

    P. Stamenov

    Fig. 8 The magnetic structure of α-Cr and a typical spin density wave. (After: [5])

    the nesting properties of the Fermi surface, in this case [6]. The nesting implies that the electron and hole Fermi surfaces are rather similar and can be essentially  In a situation, similar to the one superimposed by translation by a nesting vector Q. appearing in some rare earths, this leads to anomalous electron-magnon scattering, with both static and dynamic consequences. The SDW wave vector Q therefore is,  in most cases, very close to Q. Commensurate SDWs Highly defective Cr, resulting, for example, from a cold-working process, can display rather high Neel transition (up to 475 K) to a commensurate AFM structure. This is attributed to the build-up of strain-induced dislocations [6]. It is worth noting that in the commensurate AFM state, the PDOS of the d states in chromium is not completely unpolarised. The small differences in the local environment on the α and β sites should result in a rather small, yet potentially observable Fermi level polarisation as indicated on Fig. 9. It probably comes to a little surprise that some finite polarisation (P ∼ 5%) can be recovered in experiments involving nano- and micro-indentation-based contact formation, such as PCAR. Incommensurate SDWs

    14 Magnetism of the Elements

    671

    Fig. 9 The d-states DOS of Cr in the commensurate spin density wave regime. (After: [6])

    Rare Earths Magnetism of the Rare Earths The properties of the rare earth elements in their metallic form are determined primarily by the relative localisation of the 4f electrons, whose states resemble closely the ones of the free atoms and ions, and forming the majority contribution to the magnetic moments. The 5d and 6s states, on the other hand, delocalise and serve a dual purpose, as quasi-free carriers and as mediators of the exchange interactions. This separation, although not absolute, is a good starting point for the discussion of the properties of the 4f states. Their interactions with the surroundings can be classified into two classes – single-ion interactions, which are independent on each

    672

    P. Stamenov

    crystal site i and contribute energy sums to the Hamiltonian, which are independent on the different ionic residues, and two-ion interactions, which involve sums over pairs of ions, indexed by i and j . Crystal Field The most important single-ion contribution is the effect of the crystal field. This is effectively the influence of the charge distribution around an ionic site with a particular local point symmetry, resulting in the large magnetic anisotropies, characteristic of the rare earths, even in their metallic form. In first-order terms, the crystal field contribution to the potential energy can be written as:  vcf (r) =

    dR

    eρ(R) |r − R|

    (5)

    where ρ(R) is the charge density distribution created by the surrounding ionic residues (both electronic and nuclear parts). If the overlap of the 4f electronic cloud can be considered small, the solution of the Laplace’s equation may be written in terms of an expansion over spherical harmonics of the type: vcf (r) =

    

    l  Am l r Ylm (r )

    (6)

    lm

    with the coefficients of the expansion Am l , given by the appropriate integrals of the type:  ρ(R) m 4π = (−1) (7) Am dR l+1 Yl−m (R ) l 2l + 1 R Therefore, the interaction is represented by a set of multipoles of the 4f electronic distribution, interacting with the components of the electric field present at the site. The situation gets somewhat more complicated, when the overlap between the 4f distribution and the neighbouring charges is not small. Of course, mathematically, vcf (r) can still be expanded into spherical harmonics, obeying the right local point symmetry; however, the coefficients are given by the above relatively simple integrals and are generally not scaling with r l , either. For the 4f electrons, the multipole contributions with l > 6 vanish. This however does not significantly simplify the task of evaluating the remaining terms. For the cases, when the crystal field splitting remains small, when compared to the spin-orbit splitting, one can take J to be a good quantum number and write the sums and integrals in terms of the J operators. Hcf =

     i

    lm

    l Am l αl r

    2l + 1  Olm (Ji ) 4π

    (8)

    lm (J) are defined via the spherical harmonics, projected The Racah operators O over Cartesian coordinates, by replacing r by J and symmetrising. Convenient tables

    14 Magnetism of the Elements

    673

    of these are readily available in the literature. The convenience in using the Racah operators, however, is not complete, as they transform as tensor operators and behave under rotations like spherical harmonics. It is, therefore, rather customary in the field to introduce the Stephens operators Olm (J), so that the crystal field part of the Hamiltonian can be written as [8]: Hcf =

     i

    Blm Olm (Ji )

    (9)

    lm

    The first few of the Stephens operators can be expressed in terms of X ≡ J (J +1) and J± ≡ Jx ± iJy as:  1 2 J+ + J−2 2 1 O21 = (Jz Jx + Jx Jz ) 2

    O22 =

    O20 = 3Jz2 − X Despite the apparent simplicity of the expression for the crystal field, in terms of the Stephens operators, the coefficients Blm are still quite difficult to estimate with an adequate accuracy, due to the non-spherical shapes of the charge distributions and the existence of substantial charge redistribution and shielding effects. Two practical approaches are very often deployed. One is to consider an adjustable magnitude point-charge model for the charge distribution (compensating for the electronic screening). The other is to treat BlM simply as experimental parameters, fixed based on experimental data. For the rather popular hexagonal case, the crystal field can be specified by fixing the values of four independent parameters:

    Hcf =

    

    ⎡ i⎣

    

    ⎤ Bl0 Ol0 (Ji ) + B66 O66 (Ji )⎦

    (10)

    l=2,4,6

    For effectively all rare earths, these parameters have been determined experimentally, taking into consideration data available from a number of experimental techniques, with the resulting multiplet splitting typically being in the range of 10ns to 100-ths of meV. Magnetoelastic Effects When the lattice is strained, both the crystal field and the two ion terms of the Hamiltonian are affected. This results in a substantial magnetoelastic coupling energy and, courtesy of the modest elastic constants of the rare earth metals, to the possibility of having large additional equilibrium strain and/or large magnetostriction. As the strain scaling of the elastic energy is quadratic and the one of the magnetoelastic energy is linear, balance situations may occur, indeed, for non-zero values of equilibrium strain. Within the pure rare earths and their alloys is, indeed, where we

    674

    P. Stamenov

    Table 2 Magnetic properties of the rare earths Element La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

    Structure HCP HCP HCP HCP HCP RH BCC HCP HCP HCP HCP HCP HCP FCC HCP

    Density (kg/m3 ) 6146 6689 6773 7008 7264 7520 5244 7901 8230 8551 8795 9066 9321 6966 9841

    Lattice parameter (pm) 377 368 367 366 365 363 458 363 361 359 358 356 354 549 351

    Tc,N (K) – 13.7* 0.05* 19.9* – 106* 90.4* 293 220 89 20 20 32 – –

    μs (μB ) – 0.6 2.7 2.2 – 0.13 5.1 7.63 9.34 10.33 10.34 9.1 7.14 – –

    K2 (kJ/m3 ) – – – – – – – – 27 31 10 −19 – – –

    λs (10−3 ) – – – – – – – – 4.2 2.2 0.75 −0.25 – – –

    * the highest of multiple transition temperatures

    find the largest magnetostriction coefficients; see, for example, Table 2. An example of great practical importance is the alloy system Terfenol-D (Tbx Dy1−x Fe2 ) , for x ∼ 0.3 , which exhibits magnetostriction coefficient λs > 2 · 10−3 . Before leaving strain behind, it should be mentioned that the magnetoelastic effects can, indeed, be considered to be strain-dependent renormalisations of the crystal field parameters. There is, however, also a dynamic side to the coupling, which results in the crossing and anti-crossing of the dispersion relations of magnons and phonons, in particular high-symmetry directions, and can result in the formation of combined excitations, sometime referred to as elasto-magnons. Out of the two-ion couplings, the most important one to consider is the indirect exchange, by courtesy of which the ions are coupled via the conduction electrons. The general form of this interaction can be written as [8]: 2 Hsf (i) = − N

    

     drI (r − R)Si · s(r) = −

    drHi (r) · μ(r)

    (11)

    with N being the number of ions, s(r) is the conduction electron spin density and the magnitude of the exchange integral I is determined by the geometric overlap between the charge distributions of the 4f and the conduction states. The same expression also defines an inhomogeneous effective magnetic field Hi (r) acting on the conduction electron moment density μ(r) = 2μB s(r). Employing the standard approach of the generalised susceptibility theory, a spin located at Ri contributes to the moment μiα like:

    14 Magnetism of the Elements

    μiα (r) =

    675

       1  dr χαβ r − r Hiβ (r ) V

    (12)

    β

    or in vector notation μi = χˆ · Hi . In order to try and exploit the periodicity of the crystal lattice, it is convenient to define the forward and backward Fourier transforms of the generalised non-local conduction electro-susceptibility as: 

    χ (q) =

    1 V

    χ (r) =

    V (2π )3

    drχ (r)e−iq·r  drχ (q)eiq·r

    which is valid within a scalar approximation and written this way for simplicity. Moments and Exchange Integrals The additional term in the Hamiltonian, corresponding to the f − f coupling thorough the conduction electrons, is then simplified, defining I (q) to be the Fourier transform of the exchange integral to: Hff = −

    1 JS (ij )Si · Sj 2

    (13)

    ij

    taking the form of the familiar Heisenberg exchange, with JS (ij ) computed as: JS (ij ) =

    1  JS (q)eiq·(Ri −Rj ) N q

    (14)

    where JS (q) is effectively the projected exchange integral in momentum space: JS (q) =

    V |I (q)|2 χ (q) Nμ2B

    (15)

    Within the famous RKKY (Rudderman-Kittel-Kasuya-Yosida) approximation, the non-local interaction between the conduction electron spins s(r) is replaced by a local approximation as: Hsf = −I0 δ(r − R)s · S

    (16)

    which allows for an approximate estimation of the interaction H (ij ) as: H (ij ) ∼

     1 1 m∗  2kF Rij cos(2kF Rij ) − sin(2kF Rij ) Si · Sj 4 4 (Rij ) h¯

    (17)

    676

    P. Stamenov

    Fig. 10 The indirect exchange interaction coupling strength, as deduced from neutron scattering measurements of magnon dispersions in Pr. (After: [8] and [9])

    where kF = q(EF ) is the Fermi wavevector for the conduction electrons, m∗ is the conduction electron effective mass and Rij = Ri − Rj . The form of this interaction is decaying oscillatory, with it covering typical distances of up to a couple of nanometre. The appearance of the so-called Kohn anomalies in J (q), due to the resonance matrix elements for interaction of filled and empty states in the immediate vicinity of the Fermi surface, separated by q gets Fourier transformed into Friedel oscillations in J (R). The long periodicity of these is associated with the smallness of the difference in energy and momenta of the above conduction states and is critical for the interpretation of both the magnetic structures of the rare earths and their magnetic and electronic excitations. A rather famous example of the verification of this type of dependence is shown on Fig. 10 and is based on the processing of neutron data for the low-temperature magnon excitations in praseodymium [9]. Yet, more generally, for cases, where L is large, a large orbital component of the moment is expected, which can lead to anisotropic contributions to the effective Heisenberg exchange Hamiltonian. A more general form of the interaction is therefore required, with different interaction constants coupling the different effective dipole moments of the 4f electrons: Hdd = −

    1 Jαβ (ij )Jiα Jjβ 2 ij

    (18)

    14 Magnetism of the Elements

    677

    or in other words an anisotropic dipole-dipole coupling. The determination of all relevant matrix components of the Jˆ tensor, without taking advantage of all possible symmetry cancellations, is generally an insurmountable task. The smallest in magnitude part of the Hamiltonian is due to interactions between the f and s states with the nuclear spins I  Hhf = A Ii · Ji (19) i

    with the hyperfine coupling constant A being of the order of several micro-electron volts, in most cases, and the interaction represented with an effective contact form. While generally quite negligible, at sufficiently low temperatures, this term can influence the magnetic ordering of the rare earths. A particular example is the ground-state order in Pr. Magnetic data from [8]. Anisotropy data from [10].

    Magnetic Structures and Phase Transitions The basic modes of the magnetic structures, which can be found in hexagonal systems and specifically in the heavy rare earths are illustrated on Fig. 11. Fig. 11 Basic magnetic structures of the heavy rare earths. The moments, within each hexagonal layer, are parallel to each other. (After: [8])

    678

    P. Stamenov

    Fermi-Level Spin Polarisation of the Magnetic Elements As of the critical dependence of the functionality of spintronic devices on the Fermi level spin polarisation, the problem of its quantification for an arbitrary material has occupied both theorists and experimentalists for the best part of half a century. Multiple definitions exist, which can be summarised, for the most commonly used experimental methods, following reference [11]. The appropriate power n and the nature of the energy and directional averaging of the corresponding velocities v and density of states for the main quantisation axis D are different, depending on the method of choice for the interrogation of the electronic system. The bare DOS polarisation is almost impossible to access (the case of n = 0), although some types of photoemission experiments do come close. The energy resolution with which P (E) can be probed by means of photoemission is limited to about 100 meV, by the availability of high-resolution electron analysers and corresponding flexible excitation synchrotron beam-lines. This limited resolution close to EF does not usually permit direct comparisons with other methods. As a rule of thumb, transport experiments, which depend on either spin-dependent tunnelling or quasiballistic transport of electrons, would sample a form of the averaging for n ∼ 1, with a relatively narrow interval of energies 105 K/s, while for bulk amorphous materials, the critical cooling rate may be as low as 10 K/s. The elements used to create amorphous, bulk amorphous, and nanocrystalline soft magnetic alloys have many commonalities. As shown in Fig. 1, all three groups are made up of a majority of 3d magnetic transition metals, especially Fe, Co, and Ni, which give the desired magnetization.

    696

    J. P. Liu et al.

    Fig. 1 Elemental makeup of typical nanocomposite soft magnetic alloys with four major components: magnetic transition metals (MTM), early transition metals (ETM), metalloid/post-transition metals (PTM), and late transition metals (LTM)

    Glass-forming ability is imparted by the metalloid and post-transition metal elements in most amorphous and bulk amorphous alloys. This is largely due to the partitionless transformation to the amorphous phase that can occur directly from the melt at eutectic compositions near pure MTM. Early transition elements are used in some amorphous and bulk amorphous and in all nanocrystalline soft magnetic alloys. The ETMs are necessary as a deterrent to grain coarsening in nanocrystalline alloys, and they are usually accompanied by late transition metal alloying elements, which provide nucleation sites during the early stages of crystallization [2, 3]. The conventional melt spinning technique uses an induction coil to melt ingots of a desired composition and then expels the melt through an orifice in the crucible onto a rapidly rotating wheel, giving effective quench rates of up to 106 K/s. Using this conventional technique, alloys with compositions near deep eutectics are routinely formed into amorphous ribbons 20 to 50 microns in thickness, a few millimeters to many centimeters in width, and meters in length. Bulk amorphous alloys can be formed at lower cooling rates by casting, sometimes with a flux to prevent oxidation and improve the maximum casting diameter. Nanocrystalline alloys can be produced using an amorphous alloy as precursor, but most studies have been carried out on melt-spun ribbons with post-quench annealing to optimize alloy performance. The final microstructure consists of nanocrystalline grains surrounded by a thin amorphous matrix phase (1–2 nm in width). This gives the optimal microstructure with 70–80 vol% crystallinity. More details about the processing of amorphous, bulk amorphous, and nanocrystalline alloys can be found in reviews [4–6]. The MTM composition has a major influence on the magnetization, magnetostriction, magnetocrystalline anisotropy, and Curie temperature. Adjusting the ratios of Co, Ni, and Fe influences the crystallization behavior and magnetic performance. While Fe-based alloys have the great advantages of low cost and strong performance in terms of magnetization and low losses, improvements can be made by adjusting the composition to increase the operating temperature and saturation magnetization. The Curie temperature of the amorphous phase (TCam ) is an upper bound on the operating temperature of all three classes of alloys. In amorphous alloys, the lack of long-range periodic order tends to minimize the local magnetocrystalline anisotropy, leaving magnetoelastic anisotropy and long-range dipolar anisotropy to dominate. In nanocrystalline alloys, exchange averaging helps to reduce the effective magnetocrystalline anisotropy, leaving the magnetostriction as the major influence on the core losses. Residual stress in

    15 Metallic Magnetic Materials

    697

    Fig. 2 Magnetocrystalline anisotropy (K1 ) energy density for BCC Fe single crystal (dark blue) as a function of angle from the [001] direction when viewed from the [100] projection and the effective magnetocrystalline anisotropy (light green) for nanostructured BCC Fe (where N grains are exchange coupled in a volumetric exchange length (Lex 3 ))

    the alloy couples magnetoelastically with the magnetization resulting in increased losses. This can be avoided by choosing compositions where the magnetostrictive coefficient (λ) is near zero. The remarkable reduction in coercivity when grain sizes are reduced to the nanocrystalline regime was first described by Herzer [7]. Based on the idea that randomly oriented grains are exchange coupled through the residual amorphous matrix, a random anisotropy model shows that the coercivity is proportional to the grain size to the sixth power. This is possible when the exchange correlation length (Lex ) is larger than the structural correlation length (grain size, D), an effect referred to as exchange softening which results from exchange averaged anisotropy. This is illustrated in Fig. 2, where the integrated anisotropy energy is unchanged by the averaging but the fluctuations (which are more important to soft magnetic properties) are significantly reduced. The exchange length indicates the minimum length scale over which the atomic moments must remain aligned due to exchange forces. The magnitude of this √ fundamental magnetic material parameter can be found by Lex = A/K1 , where A is the exchange stiffness and K1 is the magnetocrystalline anisotropy constant. A value of 35 nm was calculated for Fe-Si-based nanocrystalline alloys. Since D ∼10 nm is much smaller than Lex , the magnetic moments in each individual grain cannot relax into the local easy direction dictated by the grain orientation. This results in averaging of the local magnetocrystalline anisotropy over the exchange correlation volume (Lex 3 ). Using the random anisotropy model, the effective magnetocrystalline anisotropy, K1 , representing the material response √ can be determined as K = K1 / N, with N being the number of grains within

    698

    J. P. Liu et al.

    the exchange correlation length. The value of N in a cubic volume with sides Lex can be estimated by the relation N = (Lex /D)3 . Using these equations and the definition of Lex , the effective anisotropy can be determined in terms of the crystalline material parameters, K1 , A, and D: K = K14 · D 6 /A3 . The exchange softening described by this model breaks down as the grains become decoupled, especially when the Curie temperature of the amorphous intergranular phase is exceeded. This effective anisotropy is proportional to the coercivity and inversely proportional to the permeability of nanocrystalline alloys. Due to the small value of the averaged anisotropy, amorphous and bulk amorphous alloys have relatively small coercivity (again proportional to K1 ) [8–11]. In these materials, the magnetoelastic anisotropy dominates the coercivity. An unfortunate correlation between the square of magnetization and the magnetostrictive coefficient is found for amorphous alloys. This effect is absent in nanocrystalline alloys due to the two-phase microstructure which can be adapted to provide near-zero values of magnetostrictive coefficient together with moderate magnetization [12]. There are a number of advantages that amorphous and nanocrystalline soft magnetic alloys have over conventional large-grained materials, as summarized in Fig. 3 and Table 1. Among the advantages, an important one is their better performance at switching frequencies above 1 kHz. This is due in part to (a) the smaller interaction of domain walls with the nanocrystalline grains and (b) their higher resistivity. The hysteresis loss is present at all frequencies and proportional to frequency. Controlling the microstructure and composition of the alloy can lead to reduction of hysteretic losses. As the switching frequency is increased, eddy currents in the alloy enhance the core losses. Reducing lamination thickness and increasing the resistivity of the alloy help to limit the eddy currents. At high frequencies (especially above 10 kHz), dynamic eddy current losses dominate. These eddy currents are localized at the domain walls and can be minimized by refinement of the domain structure in the alloys.

    Alnicos In the first section, iron–cobalt-based alloys featured as soft magnetic materials. Here, iron–cobalt-based nanostructures feature as hard magnets. Alnico magnets have been developed, based on ferromagnetic Alnico alloys discovered by T. Mishima in 1932 [17]. They are composed of Fe, Co, Ni, and Al with additions of Cu, Ti, and Nb. It is generally recognized that high coercivity Hc in Alnico magnets depends on the shape anisotropy of the ferromagnetic Fe-Co precipitates (α 1 phase) finely dispersed in a nonmagnetic Ni-Al matrix (α 2 -phase). This dispersion is caused by spinodal decomposition (SD) through an appropriate heat treatment [18–27]. The most widely used Alnico alloys are Alnico 5–7, 8, and 9 (Table 2). Micrographies showing typical microstructure of a commercial Alnico 9 magnet are shown in Fig. 4 where the high-aspect-ratio, aligned iron–cobalt needles can be seen [27]. The longitudinal microstructure (a) exhibits highly elongated α 1 rods (>400 nm) which produce a strong shape anisotropy. The transverse (b) consists of the Fe-Co-rich α 1 -phases (brighter mosaic tile), Ni-Al-rich α 2 -phases

    15 Metallic Magnetic Materials

    699

    Fig. 3 The magnetic permeability at 1 kHz plotted against the saturation induction (μ0 Ms ) for soft magnetic materials [8–11]

    (darker matrix), and Cu-rich phases (white spots located between the corners of α 1 -phase). The transverse microstructure of Alnico 8 is similar to that of Alnico 9, but the longitudinal α 1 -phase is elongated in two orthogonal by the applied field [27]. The magnets can be grouped into two main categories: isotropic and anisotropic, distinguished by the different alloy crystallization processes and thermomagnetic treatments.

    Material Fe49 Co49 V2 Soft Fe Fe 3.2%Si Fe67 Co18 B14 Si1 Metglas 2605CO Fe81 B14 Si3 C2 Metglas 2605SA1 Ni50 Fe50 Permalloy Fe77 P7 B13 Nb2 Cr Senntix Fe73.5 Si13.5 Nb3 B9 Cu Finemet Fe40 Ni38 Mo4 B18 Metglas 2826MB Co70 Fe5 Ni2 B3 Si15 Metglas 2705M Ni78 Fe17 Mo5 Supermalloy 800 600 100–800

    0.88

    0.77

    0.65–0.82

    0.25–0.64

    0.8

    0.4

    0.6–2.5

    80

    4–20

    1.23–1.35

    70

    1.4–1.6

    0.1

    2.5

    600

    1.56

    Hc (A/m) 16–398 4–80 6 3.5

    1.31

    μmax (103 ) 5–50 10–50 0.5–5 50

    Js (T) 2.4 2.16 2.0 1.8

    60

    136

    160

    110

    40–50

    130

     (10−8  m) 27 10 48 123

    673

    638

    626

    843

    556

    753

    668

    TC (K) 1203 1044 1018 658







    0.01

    0.65 W/m3

    0.33

    0.08

    Pcm (W/kg) 1.10 20 0.84 0.66

    2–3

    0

    12

    0–2

    18

    27

    λs (ppm) 60 −2 7 35

    793

    683

    780

    795

    783

    703

    Tx1 (K)

    Table 1 Magnetic properties of commonly used soft magnetic materials [13–16]. Metglas is amorphous, Senntix is bulk amorphous, and Finemet is nanocrystalline. The saturation magnetization for alloys marked with “†” exceeds 1 T. Core loss (Pcm ) is reported for 60 Hz sinusoidal switching at 1T maximum induction amplitude. Js is the magnetic polarization at saturation; μmax is the maximum permeability; Tx1 is the primary crystallization temperature

    700 J. P. Liu et al.

    A B A A A A A A A B B B B

    1 2 3 4 5 6 7 8 9 10 11 12 13

    5–7 8 9 9 9h 9 9 9 9 8 8 8 8

    Type

    Composition (wt.%, balanced Fe) Co Ni Al Cu 24.3 14.0 8.2 2.3 40.1 13.0 7.1 3.0 35.4 12.1 7.0 3.2 36 13.5 7.2 3.0 40 14 8.4 3.0 34 – – – 36 – – – 38 – – – 40 – – – 38 – – – 38 – – – 38 – – – 38 Others 1.0 Nb 6.5 Ti 5.0 Ti, 0.5 Nb 6.05 Ti 8.2 Ti – – – – 6.5 Ti 7.5 Ti 7.5 Ti, 0.5 Nb 8.0, 2.0 Nb

    Note: “A” field-treated, oriented grain; “B” field-treated, random grain; “–” unlisted or unknown

    Characteristics

    No.

    Table 2 Nominal composition and magnetic properties of cast Alnico 8 and 9 magnets [28, 29] Magnetic properties Br(T) Hc (kA/m) 1.35 58.9 0.74 151.2 1.12 109.5 1.08 125 0.86 160.8 1.14 117.0 1.13 126.0 1.08 132.0 1.04 144.0 0.88 147.0 0.80 152.0 0.77 183.0 0.61 209.3

    (BH)max (kJ/m3 ) 60.0 41.0 81.0 90.0 75.6 93.5 104.3 97.9 93.4 47.8 46.2 44.6 34.0

    15 Metallic Magnetic Materials 701

    702

    J. P. Liu et al.

    Fig. 4 Microstructure images of a commercial Alnico 9 magnet (a) longitudinal, (b) transverse [27]. (With kind permission from Elsevier)

    Much work has been done on the composition, targeted at enhancing magnetic properties, especially coercivity. The early studies on the relationship of columnar crystallization conditions to compositions obtained a Br of 0.89–1.0 T, Hc of 141.7– 167.2 kA/m, and (BH)max of 87.6–93.1 kJ/m3 in the grain-oriented alloy containing about 39 wt.% Co and 7–8 wt.% Ti [30]. By varying the Co and Ti content and optimizing the microstructure, the magnetic properties of Alnico 9 were improved as shown by No. 4–5 [29] in Table 2. Alnico 9 magnets are obtained by grain orientation of Alnico 8 alloys with higher Co content [28]. By increasing the Co content from 34 to 40 wt.%, as shown by No. 6–9 [28] in Table 2, Br decreases, while Hc increases. A record (BH)max for Alnico of 104.3 kJ/m3 was achieved in the magnet with 36 wt.% Co. Several studies have confirmed that the addition of Ti effectively improves the coercivity by increasing the local shape anisotropy constant [31], altering the lattice mismatch between α 1 - and α 2 -phases [32], or narrowing the miscibility gap between α 1 - and α 2 -phases [33]. Simultaneously increasing Co from 24 to 39 wt.% and Ti from 0 to 7.5 wt.% resulted in an increase of Hc from 56.6 to 166.4 kA/m [31]. The addition of Nb showed a similar effect to that seen in Ti [34]. The addition of 1.5 wt% Nb increased Hc by ∼ 24 kA/m. Moreover, Nb was found to help the growth of columnar grains and suppress the precipitation of undesirable γ phase [35]. Co-addition of Ti and Nb can significantly improve the coercivity, as shown by No. 10–13 in Table 2, with a record Hc of 209.3 kA/m. Although Ti can improve the coercivity, it makes the columnar crystallization of Alnico 9 difficult due to fine-grained size. Therefore, a small amount of sulfur, selenium, or phosphorus can be added to increase grain size and improve columnar crystal growth [36]. Cu addition is necessary to separate the well-formed mosaic Fe-Co structure by forming a Cu-rich phase [37], as shown in Fig. 4b. The Cu-rich phase forms as rods of 2

    15 Metallic Magnetic Materials

    703

    Table 3 Magnetic properties of HP, HIP, and CMS samples [41] Method HP HIP CMS-1 CMS-2

    Process 810 ◦ C, 10 min 1250 ◦ C, 4 h 1250 ◦ C, 4 h 1250 ◦ C, 8 h

    Br (T) 0.91 0.94 0.88 1.01

    Hc (kA/m) 142.9 146.9 134.1 134.4

    (BH)max (kJ/m3 ) 45.4 47.0 39.8 51.7

    Grain size(μm) 90 50 30 330

    nm diameter between α 1 -phase regions [27]. Cu was also thought to increase the diffusion rate of atoms in the α 2 -phase [33]. Alnico magnets are manufactured using one of two processes: casting by conventional foundry methods and powder metallurgical processing. The grain orientation of cast Alnico 8 is random, whereas Alnico 9 with a perfect columnar grain orientation is obtained by directional solidification. All the Alnico magnets can be made from a powdered mixture of chemical elements or pre-alloyed powder between 1250 ◦ C and 1350 ◦ C, but the grain orientation of magnets is random [38, 39]. The magnetic alignment of cast and sintered magnets is achieved by thermomagnetic treatment. Gas atomization (GA) is an efficient technique to make Alnico powder [39]. The powder is oval in shape and has an average size of 119 m. [40]. The powder can be consolidated by techniques including hot pressing (HP), hot isostatic pressing (HIP), and compression molding with subsequent sintering (CMS) [39, 41]. Table 3 shows the magnetic properties and grain size of HP, HIP, and CMS samples made from the GA powder [41]. These magnets showed comparable magnetic properties to cast magnets with the same composition. Injection-molded magnets are prepared by mixing as-milled Alnico 8 powder with a binder [38]. The technique may allow sintered magnets to be made with complex shapes. A typical heat treatment process is shown in Fig. 5, where the alloys are first solutionized (homogenized) at 1250 ◦ C for at least 30 min and then quenched to obtain a single α-phase. Subsequently, the alloys are heat treated in a magnetic field of 1.0–1.5 T, between 810 ◦ C and 840 ◦ C for 10–20 min to initiate the spinodal decomposition. As a result, the α-phase separates into α 1 - and α 2 -phases, leading to an initial coercivity. The applied magnetic field makes the Fe-Co particles to preferentially grow in the field direction. A tempering treatment is performed at lower temperatures for 15–30 h in two or more steps to promote chemical diffusion between the α 1 - and α 2 -phases. These processes further increase the magnetic anisotropy of the α 1 -phase, leading to enhanced coercivity. The effects of thermomagnetic treatment at different conditions on microstructure and magnetic properties have been extensively investigated [20, 24, 26, 34, 37, 42, 43]. Iwama’s studies on thermomagnetic processing of Alnico magnets established a linear relationship between a square of the wavenumber of the modulated structure and the annealing temperature [44]. The relationship was consistent with Cahn’s theoretical predictions on spinodal decomposition [18]. More recent studies confirm that achieving the highest Hc is closely related to

    704

    J. P. Liu et al.

    Fig. 5 Schematic of the typical heat treatment for Alnico alloys

    ATOMIC MOMENT IN BOHR MAGNETONS

    3.0

    Fe – V Fe – Cr

    2.5

    Fe – Ni Fe – Co Ni – Co

    2.0

    Ni – Cu Ni – Zn Ni – V

    1.5

    Ni – Cr Ni – Mn

    1.0

    Co – Cr

    0.5

    Co – Mn PURE METALS

    0 Cr 6

    Mn 7

    Fe 8

    Co 9

    Ni 10

    Cu 11

    ELECTRON CONCENTRATION, C

    Fig. 6 Average atomic moments of binary alloys of the iron-group elements. (Reprinted figure with permission from [47] by the American Physical Society)

    optimizing the microstructures [26]. Chu’s studies explained that there exist two splittings of Fe-Co particles during magnetic annealing which leads to a decrease of mean diameter of Fe-Co particles and thus increases coercivity [42]. High magnetic fields (up to 10 T) strongly influence the modulated (or mosaic) microstructure [43] but did not improve Hc .

    15 Metallic Magnetic Materials

    705

    Intermetallic Compounds of d-d and d-p Types Binary intermetallic compounds show a great variety of magnetic properties. Itinerant magnetism with non-integer magnetic moments was first observed and explained in binary alloys of late 3d metals by Slater and Pauling [45, 46]. A few years later, Bozorth demonstrated that the band picture becomes much more diverse, as depicted in Fig. 6, if one does not restrict the alloying elements to direct neighbors in the periodic table [47]. Magnetic ordering is not restricted to alloys of the magnetic 3d elements. This section is the result of a thorough literature search for magnetically ordered binary intermetallic d-d- and d-p-type compounds and alloys. The available compounds with the type of magnetic order, the ordering temperature, and the value of the magnetic moment are listed in Table 4. In the case of polymorphs, we include the magnetic characteristics of all known crystal structures. We restrict ourselves to alloys with metallic character and do not include alloys with rare-earth or actinide elements.

    Magnetic Shape Memory Alloys and Compounds Magnetic shape memory (MSM) alloys and compounds are ferromagnets that exhibit large strains under an applied magnetic field due to martensitic phase transformation [173–175]. Currently, the materials with the best MSM properties are the near-stoichiometric Ni2 MnGa compounds (a Heusler alloy) with five- and seven-layer stacked martensite structures that show magnetic-field-induced strains (MFIS) of 6.0% and 9.5% in a field less than 800 kA/m, respectively. This exceeds the normal magnetostriction based on magnetoelasticity by orders of magnitude. The origin of the MFIS in MSM alloys is due to magnetically induced twin boundary motion in the martensitic phase. The MSM operation temperature is between 150 and 330 K depending on composition. These characteristics render the MSM alloys promising functional materials for the development of magneto-mechanically controlled actuators or dampers. The composition, crystal structure and structural/magnetic transition temperatures, and field-induced strain in single-crystalline Ni-Mn-Ga alloys are given in Table 5. The characteristics and performance of NiMn-Ga alloys in other forms of polycrystals, composites, and thin films are collected in Table 6. In the second part of this section, the multifunctional MSM systems such as NiMn-(In, Sn, Sb) and Ni-Co-Mn-Ti Heusler alloys and hexagonal MM’X compounds are addressed. Magnetic-field-induced reversible martensitic transformation has been achieved in newly developed alloys, where a metamagnetic transition occurs from weak magnetic martensite (austenite) to ferromagnetic austenite (martensite) by applying a magnetic field. These alloys have the following advantages: (i) shape recovery can be obtained in polycrystals, (ii) high blocking stress can be obtained, and (iii) multifunctionality with large magnetocaloric effect and magnetoresistance

    F

    F F AF F F

    TiCo3 TiCo2

    (Ni,Ti) σ -phase α (V, Fe) (V, α-Co)

    (V, Ni)

    σ -phase CrBe12 (α-Cr, Mn) σ -phase (α-Cr, Fe)

    Ti-Co

    Ti-Ni V-Fe

    V-Ni

    Cr-Be Cr-Mn Cr-Fe

    V-Co

    Sc3 In ScCo2 TiBe2 TiFe2

    Sc-In Sc-Co Ti-Be Ti-Fe

    Type of order F F F P AF AF F F F AF F F F F

    Compound ScFe2

    System Sc-Fe

    52 50 300–800 9–47 1045–0

    1.64

    10 0–65 424–323 0.5 0.54 0.84 0.9 0.25 >0.2 8.0 0.38 0.15 – – – – –

    γ *(J/mol·K2 ) 1.6 0.26 1.6 0.35 0.8 4.2 6.4 FL AF AF AF AF AF AF AF AF QCP QCP NFL QCP NFL

    Type FL FL FL FL FL FL FL – 9.7 5.0 3.4 15 1.1 0.4 2.8 0.65 – – 0.07 – 0.31

    TN (K) – – – – – – –

    Local moment QPT Itinerant QPT Field-tuned QCP Local moment QPT Also quadrupolar ordering at TQP = 0.5 K

    Low carrier semimetal

    Additional ordering at 1.2 K

    Heavy-fermion due to quadrupolar fluctuations T = Fe, Co, Rh, Os, Ir

    Metamagnetic transition at 8 T

    Remarks First heavy-fermion compound

    [578] [579] [580] [581] [582] [583] [584] [585] [586] [587] [588] [589] [590] [591]

    Ref. [569] [570] [571, 572] [573, 574] [575] [576] [577]

    Table 30 Basic properties of non-superconducting heavy-fermion compounds. The crystal structure, year of discovery, the γ *-value expressed in J/molf-atom K2 , type of ground state, and Néel temperature, TN , are tabulated. Types are FL, Fermi liquid; AF, antiferromagnet; QCP, quantum critical point; and NFL, non-Fermi liquid

    15 Metallic Magnetic Materials 775

    Structure Tetragonal

    Cubic Hexagonal

    Tetragonal Tetragonal Tetragonal Cubic Tetragonal Orthorhombic Tetragonal Hexagonal Hexagonal Tetragonal

    Compound CeCu2 Si2

    UBe13 UPt3

    URu2 Si2 CeCoIn5 CeIrIn5 PrOs4 Sb12 Ce2 CoIn8 β-YbAlB4 Ce2 PdIn8 UNi2 Al3 UPd2 Al3 CePt3 Si

    1986 2001 2001 2002 2002 2008 2009 1991 1991 2004

    1984 1984

    Year 1979

    0.18 0.29 0.72 0.35 0.50 0.15 1.00 0.12 0.15 0.39

    1.1 0.45

    γ * (J/mol·K2 ) 1.0

    SC SC SC SC SC SC SC AF, SC AF, SC AF, SC

    SC SC

    Type SC

    0.8 2.3 0.4 1.85 0.4 0.08 0.65 1.0 2.0 0.75

    0.85 0.55

    Tc (K) 0.5

    – – – – – – – 4.2 14.2 2.2

    – –

    TN (K) –

    Small moment AF Large moment AF Non-centrosymmetric structure

    NFL compound

    Two SC transitions at 1.75 K and 1.85 K

    Remarks First heavy-fermion superconductor; close to AF order with TN = 0.8 K Two SC transitions when doped with Th Two SC transitions Tc + = 0.49 K and Tc − = 0.43 K; weak AF order at TN = 5 K Hidden order at 17.5 K

    [599] [600] [601] [602, 603] [604] [605] [606] [607] [608] [609]

    [594, 595] [596–598]

    Ref. [592, 593]

    Table 31 Basic properties of heavy-fermion superconductors. The crystal structure, year of discovery, γ *-value, type of ground state, superconducting transition temperature, Tc , and the Néel temperature, TN , are tabulated. SC, superconductor at ambient pressure; SC(p), superconductor under pressure; AF, antiferromagnet; FM, ferromagnet. For SC(p) compounds, the maximum Tc and corresponding pressure are listed. The γ *-value is expressed in J/molf-atom ·K2

    776 J. P. Liu et al.

    Tetragonal Tetragonal Tetragonal Tetragonal Cubic Tetragonal Tetragonal

    Tetragonal

    Orthorhombic Orthorhombic Monoclinic Orthorhombic

    Ce3 PtIn11 CeCu2 Ge2 CePd2 Si2 CeRh2 Si2 CeIn3 CeRhIn5 CeRhSi3

    CeIrSi3

    UGe2 URhGe UIr UCoGe

    1991 2001 2004 2007

    2006

    2015 1992 1996 1996 1997 2000 2005

    0.03 0.16 0.05 0.06

    0.11

    1.2 0.2 0.06 0.08 0.14 0.42 0.12

    FM, SC(p) FM, SC FM, SC(p) FM, SC

    AF, SC(p)

    AF, SC AF, SC(p) AF, SC(p) AF, SC(p) AF, SC(p) AF, SC(p) AF, SC(p)

    0.81.2 GPa 0.25 0.12.7 GPa 0.6

    1.6 2.5 GPa

    0.32 0.6510 GPa 0.42.8 GPa 0.350.8 GPa 0.182.5 GPa 2.11.9 GPa 1.12.0 GPa

    – – – –

    5

    2 4.1 10 36 10.1 3.8 1.6 Non-centrosymmetric structure Non-centrosymmetric structure Coexistence FM and SC; TCurie Coexistence FM and SC; TCurie Coexistence FM and SC; TCurie Coexistence FM and SC; TCurie = 53 K = 9.5 K = 46 K = 3.0 K

    [620] [621] [622] [623]

    [619]

    [610] [611, 612] [613][614] [615] [616] [617] [618]

    15 Metallic Magnetic Materials 777

    778

    J. P. Liu et al.

    Another fascinating aspect of heavy-fermion physics is the superconducting ground state that develops out of a strongly interacting electron liquid. The discovery in 1979 of superconductivity at Tc ∼ 0.5 K in the heavy-electron compound CeCu2 Si2 (γ * ∼ 1000 mJ/molK2 ) was a surprise. Whereas magnetic interactions that lead to the strongly renormalized Fermi liquid state are normally detrimental for superconductivity, in CeCu2 Si2 , the 4f band-like electrons participate in the superconducting state. The discovery of superconductivity in the heavy 5f -electron compounds UBe13 and UPt3 initiated the whole new research field of unconventional superconductivity [566]. In traditional superconductors, superconductivity is explained by the standard Bardeen–Cooper–Schrieffer model in which spinsinglet Cooper pairs are mediated by the exchange of phonons. However, for heavy fermion: (i) the superconducting state is described by a gap function with line or point nodes, rather than a full gap over the whole Fermi surface, and (ii) Cooper pairs are mediated by (anti)ferromagnetic spin fluctuations rather than by phonons. Nowadays, heavy-fermion materials provide a vast, fruitful playground to investigate the interplay of magnetic order and complex superconducting order parameters [567, 568]. An overview of the most important heavy-fermion superconductors is presented in Table 31. These can be divided into three main groups: heavy-electron compounds with (i) a superconducting ground state, (ii) coexistence of superconductivity (under pressure) and antiferromagnetic order, and (iii) coexistence of superconductivity (under pressure) and ferromagnetic order. UBe13 and UPt3 belong to the first group and served as prototypes to shape the field of unconventional superconductivity.

    References 1. Gutfleisch, O., Willard, M.A., Brück, E., Chen, C.H., Sankar, S., Liu, J.P.: Magnetic materials and devices for the 21st century: stronger, lighter, and more energy efficient. Adv. Mater. 23, 821–842 (2011) 2. Ayers, J., Harris, V., Sprague, J., Elam, W., Jones, H.: On the formation of nanocrystals in the soft magnetic alloy Fe 73.5 Nb 3 Cu 1 Si 13.5 B 9. Acta Mater. 46, 1861–1874 (1998) 3. Hono, K., Ping, D., Ohnuma, M., Onodera, H.: Cu clustering and Si partitioning in the early crystallization stage of an Fe 73.5 Si 13.5 B 9 Nb 3 Cu 1 amorphous alloy. Acta Mater. 47, 997–1006 (1999) 4. McHenry, M.E., Willard, M.A., Laughlin, D.E.: Amorphous and nanocrystalline materials for applications as soft magnets. Prog. Mater. Sci. 44, 291–433 (1999) 5. Inoue, A., Takeuchi, A.: Recent development and application products of bulk glassy alloys. Acta Mater. 59, 2243–2267 (2011) 6. Willard, M.A., Daniil, M.: Nanocrystalline soft magnetic alloys two decades of progress. In: Buschow, K.H.J. (ed.) Handbook of Magnetic Materials, vol. 21, pp. 173–342 (2013) 7. Herzer, G.: Chapter 3 Nanocrystalline soft magnetic alloys. In: Handbook of Magnetic Materials, vol. 10, pp. 415–462. Elsevier (1997) 8. Willard, M.A., Daniil, M.: Nanostructured soft magnetic materials. In: Nanoscale Magnetic Materials and Applications, pp. 373–397. Springer (2009) 9. Willard, M., Laughlin, D., McHenry, M., Thoma, D., Sickafus, K., Cross, J.O., et al.: Structure and magnetic properties of (Fe0. 5Co0. 5) 88Zr7B4Cu1 nanocrystalline alloys. J. Appl. Phys. 84, 6773–6777 (1998)

    15 Metallic Magnetic Materials

    779

    10. Yoshizawa, Y., Yamauchi, K., Yamane, T., Sugihara, H.: Common mode choke cores using the new Fe-based alloys composed of ultrafine grain structure. J. Appl. Phys. 64, 6047–6049 (1988) 11. Suzuki, K., Kikuchi, M., Makino, A., Inoue, A., Masumoto, T.: Changes in microstructure and soft magnetic properties of an Fe 86 Zr 7 B 6 Cu 1 amorphous alloy upon crystallization. Mater. Trans. JIM. 32, 961–968 (1991) 12. Knipling, K., Daniil, M., Willard, M.: Nanocrystalline Fe88− 2xCoxNixZr7B4Cu1 alloys: soft magnets for vehicle electrification technologies. J. Appl. Phys. 117, 172611 (2015) 13. Chen, C.: Magnetism and Metallurgy of Soft Magnetic Materials. Dover Publications, New York (1986) 14. O’handley, R.C.: Modern Magnetic Materials: Principles and Applications, vol. 830622677. Wiley, New York (2000) 15. M. Inc., Alloy Specification Sheets. 16. Matsumoto, H., Urata, A., Yamada, Y., Inoue, A.: Novel Fe97-x-y Px By Nb2 Cr1 glassy alloys with high magnetization and low loss characteristics for inductor core materials. IEEE Trans. Magn. 46, 373–376 (2010) 17. Mishima, T.: Nickel-aluminum steel for permanent magnets. Stahl und Eisen. 53, 79 (1931) 18. Cahn, J.W.: Magnetic aging of spinodal alloys. J. Appl. Phys. 34, 3581–3586 (1963) 19. Cahn, J.W.: Phase separation by spinodal decomposition in isotropic systems. J. Chem. Phys. 42, 93–99 (1965) 20. Iwama, Y., Takeuchi, M.: Spinodal decomposition in AlNiCo 8 magnet alloy. T. Jpn. I. Met. 15, 371–377 (1974) 21. Iwama, Y., Takeuchi, M., Iwata, M.: New determination of magnetic anisotropy constants of Alnico magnet alloys. J. Phys. Colloq. 32, C1-556–C1-557 (1971) 22. Cronk, E.R.: Recent developments in high-energy Alnico alloys. J. Appl. Phys. 37, 1097– 1100 (1966) 23. McCaig, M.: Phase changes in high coercivity alloys. Z. Angew. Phys. 21, 66 (1966) 24. Stanek, M., Leonowicz, M.: Investigations of thermo-magnetic treatment of Alnico 8 alloy. Arch. Metall. Mater. 55, 571–577 (2010) 25. McCurrie, R.: Ferromagnetic Materials: Structure and Properties Academic, vol. 4, (1994) 26. Zhou, L., Miller, M.K., Dillon, H., Palasyuk, A., Constantinides, S., McCallum, R.W., et al.: Role of the applied magnetic field on the microstructural evolution in Alnico 8 alloys. Metall. Mater. Trans. E. 1, 27–35 (2014) 27. Zhou, L., Miller, M.K., Lu, P., Ke, L., Skomski, R., Dillon, H., et al.: Architecture and magnetism of Alnico. Acta Mater. 74, 224–233 (2014) 28. Liu, T., Li, W., Zhu, M., Guo, Z., Li, Y.: Effect of Co on the thermal stability and magnetic properties of Alnico 8 alloys. J. Appl. Phys. 115, 17A751 (2014) 29. Sun, Y., Zhao, J., Liu, Z., Xia, W., Zhu, S., Lee, D., et al.: The phase and microstructure analysis of Alnico magnets with high coercivity. J. Magn. Magn. Mater. 379, 58–62 (2015) 30. Hoffmann, A., StÄblein, H.: Investigations of high-coercivity Alnico alloys. IEEE Trans. Magn. 6, 225–230 (1970) 31. Takeuchi, M., Iwama, Y.: Effects of titanium upon magnetic anisotropy and coercivity in alnico magnet alloys. T. Jpn. I. Met. 17, 489–496 (1976) 32. Mason, J.J., Ashall, D.W., Dean, A.V.: The structure of Ni-Co-Al-Ti-Cu-Fe permanent magnets. IEEE Trans. Magn. 6, 191–194 (1970) 33. Hao, S.M., Ishida, K., Nishizawa, T.: Role of alloying elements in phase decomposition in Alnico magnet alloys. Metall. Trans. A. 16, 179–185 (1985) 34. Bronner, C., Haberer, J.-P., Planchard, E., Sauze, J.: Contribution to the study of the influence of Nb on the magnetic properties of Alnico alloys containing 4.5–6.5 % Ti. Cobalt. 46, 15–22 (1970) 35. Pramanik, S., Rao, V., Mohanty, O.: Effect of niobium on the directional solidification and properties of Alnico alloys. J. Mater. Sci. 28, 1237–1244 (1993) 36. Palmer, D., Shaw, S.: Production of columnar castings in high-coercivity magnet alloys containing 5–10 Per Cent Titanium. Cobalt. 43, 63–72 (1969)

    780

    J. P. Liu et al.

    37. Zhou, L., Tang, W., Ke, L., Guo, W., Poplawsky, J.D., Anderson, I.E., et al.: Microstructural and magnetic property evolution with different heat-treatment conditions in an Alnico alloy. Acta Mater. 133, 73–80 (2017) 38. Zlatkov, B., Bavdek, U., Nikolic, M., Aleksic, O., Danninger, H., Gierl, C., et al.: Magnetic properties of Alnico 8 sintered magnets produced by powder injection moulding. PIM Int. 3, 58–63 (2009) 39. Anderson, I., Kassen, A., White, E., Zhou, L., Tang, W., Palasyuk, A., et al.: Novel prealloyed powder processing of modified Alnico 8: Correlation of microstructure and magnetic properties. J. Appl. Phys. 117, 17D138 (2015) 40. Song, C., Han, B., Li, Y.: A study on Alnico permanent magnet powders prepared by atomization. Cailiao Kexue Yu Jishu(J. Mater. Sci. Technol.)(China). 20, 347–349 (2004) 41. Tang, W., Zhou, L., Kassen, A., Palasyuk, A., White, E., Dennis, K., et al.: New Alnico magnets fabricated from pre-alloyed gas atomization powder through diverse consolidation techniques. In: Proceedings of Conference INTERMAG, pp. 1–1 (2015) 42. Chu, W., Fei, W., Li, X., Yang, D., Wang, J.: Evolution of Fe–Co rich particles in Alnico 8 alloy thermomagnetically treated at 800◦ C. Mater. Sci. Tech. 16, 1023–1028 (2000) 43. Sun, X., Chen, C., Yang, L., Lv, L., Atroshenko, S., Shao, W., et al.: Experimental study on modulated structure in Alnico alloys under high magnetic field and comparison with phasefield simulation. J. Magn. Magn. Mater. 348, 27–32 (2013) 44. Iwama, Y., Takeuchi, M.: Spinodal decomposition in AlNiCo 8 magnet alloy. Trans. Jpn. Ins. Met. 15, 371–377 (1974) 45. Slater, J.C.: Electronic Structure of Alloys. J. Appl. Phys. 8, 385–390 (1937) 46. Pauling, L.: The nature of the interatomic forces in metals. Phys. Rev. 54, 899 (1938) 47. Bozorth, R.: Atomic moments of ferromagnetic alloys. Phys. Rev. 79, 887 (1950) 48. Ikeda, K., Nakamichi, T., Yamada, T., Yamamoto, M.: Ferromagnetism in Fe2Sc with the Hexagonal MgZn2-Type Structure. J. Phys. Soc. Jpn. 36, 611–611 (1974) 49. Nevitt, M., Kimball, C., Preston, R.: Variation of 57Fe isomer shift with atomic size in Laves phases and associated quadrupole and magnetic hyperfine fields. In: Proceedings of International Conference on Magnetism, Nottingham September 1964, p. 137 (1964) 50. Ikeda, K., Gschneider, K.: Disappearance of the heat capacity peak of Sc 3 In around the Curie temperature in high magnetic fields. J. Magn. Magn. Mater. 22, 207–211 (1981) 51. Yamada, H., Tohyama, T., Shimizu, M.: Metamagnetic transition of ScCo2 and LuCo2. J. Phys. F: Met. Phys. 17, L163 (1987) 52. Acker, F., Fisk, Z., Smith, J.L., Huang, C.Y.: Enhanced paramagnetism of TiBe2 and ferromagnetic transitions in TiBe2-xCux. J. Magn. Magn. Mater. 22, 250–256 (1981) 53. Brückner, W., Kleinstück, K., Schulze, G.E.R.: Mössbauer study of the laves phase Ti1xFe2+x. Phys. Status Solidi A. 1, K1–K4 (1970) 54. Brückner, W., Perthel, R., Kleinstück, K., Schulze, G.E.R.: Magnetic Properties of ZrFe2 and TiFe2 within Their Homogeneity Range. Phys. Status Solidi B. 29, 211–216 (1968) 55. Aoki, Y.: Magnetic properties of the Intermetallic Compound with the Cu3Au-type structure in cobalt-titanium alloy system. J. Phys. Soc. Jpn. 28, 1451–1456 (1970) 56. Aoki, Y., Nakamichi, T., Yamamoto, M.: Magnetic behavior of two laves phases in cobalttitanium alloy system. J. Phys. Soc. Jpn. 21, 565–566 (1966) 57. Bozorth, R.: Ferromagnetism (The Bell Telephone Laboratories Series). Van Nostrand, New York (1951) 58. Read, D.A., Thomas, E.H., Forsythe, J.B.: Evidence of itinerant electron ferromagnetism in sigma phase alloys. J. Phys. Chem. Solids. 29, 1569–1572 (1968) 59. Nevitt, M.V., Aldred, A.T.: Ferromagnetism in V-Fe and Cr-Fe alloys. J. Appl. Phys. 34, 463– 468 (1963) 60. Nevitt, M.V.: Curie Temperatures of Binary and Ternary Sigma Phases. University of Illinois at Urbana-Champaign (1954) 61. Wolcott, N., Falge Jr., R.: Ferromagnetism of CrBe $ sub 1$. National Bureau of Standards, Washington, DC (1968)

    15 Metallic Magnetic Materials

    781

    62. Wachtel, E., Bartelt, C.: Measurements of the susceptibility in the chromium-manganesesystem. Z. Metallkd. 55, 29–36 (1964) 63. Martin, J., Downie, D.: Heat capacities of transition-metal alloys 1. Sigma-phase alloy Co 0.435 Cr 0.565. J. Chem. Thermodyn. 15, 691–699 (1983) 64. Kussmann, A., Muller, K., Raub, E.: Thermomagnetic measurements on alloys of the platinum-group metals with chromium. Zeitschrift für Metallkunde. 59, 859–863 (1968) 65. Pickart, S., Nathans, R.: Neutron diffraction investigation of Pt-based alloys of the first transition series. J. Appl. Phys. 34, 1203–1204 (1963) 66. Bauer, A., Regnat, A., Blum, C.G., Gottlieb-Schönmeyer, S., Pedersen, B., Meven, M., et al.: Low-temperature properties of single-crystal CrB 2. Phys. Rev. B. 90, 064414 (2014) 67. Schoop, L., Hirschberger, M., Tao, J., Felser, C., Ong, N., Cava, R.: Paramagnetic to ferromagnetic phase transition in lightly Fe-doped Cr 2 B. Phys. Rev. B. 89, 224417 (2014) 68. Coldea, M., Cri¸san, M., Néda, A., Pop, I.: NMR, magnetic susceptibility and specific heat of CrAl 7. J. Phys. Chem. Solids. 35, 1095–1098 (1974) 69. Susner, M.A., Parker, D.S., Sefat, A.: Importance of doping and frustration in itinerant Fedoped Cr 2 Al. J. Magn. Magn. Mater. 392, 68–73 (2015) 70. Lee, C.-G., Youn, K.-T., Fukamichi, K.: Antiferromagnetic Cr-Ga layers with a high Neel temperature for biasing spin valves. IEEE Trans. Magn. 36, 2902–2904 (2000) 71. Ohsugi, I.J., Kojima, T., Nishida, I.A.: Temperature dependence of the magnetic susceptibility of a CrSi 2 single crystal. Phys. Rev. B. 42, 10761 (1990) 72. Hsu, H.-F., Tsai, P.-C., Lu, K.-C.: Single-crystalline chromium silicide nanowires and their physical properties. Nanoscale Res. Lett. 10, 1–8 (2015) 73. Ghimire, N., McGuire, M.A., Parker, D.S., Sales, B.C., Yan, J.-Q., Keppens, V., et al.: Complex itinerant ferromagnetism in noncentrosymmetric Cr 11 Ge 19. Phys. Rev. B. 85, 224405 (2012) 74. Browne, J., Liddell, P., Street, R., Mills, T.: An investigation of the antiferromagnetic transition of CrN. Phys. Status Solidi A. 1, 715–723 (1970) 75. Wu, W., Cheng, J., Matsubayashi, K., Kong, P., Lin, F., Jin, C., et al.: Superconductivity in the vicinity of antiferromagnetic order in CrAs. Nat. Commun. 5 (2014) 76. Shinohara, T., Watanabe, H.: Nuclear Magnetic Resonance in Cr 3 As 2. J. Phys. Soc. Jpn. 21, 2076–2076 (1966) 77. Ishimoto, K., Okonogi, M., Ohoyama, K., Nakajima, K., Ohashi, M., Yamauchi, H., et al.: Anisotropic exchange interaction in Cr 2 As. Phys. B. 213, 336–338 (1995) 78. Snow, A.: Magnetic moment orientation and thermal expansion of antiferromagnetic CrSb. Rev. Mod. Phys. 25, 127 (1953) 79. Shimada, K., Saitoh, T., Namatame, H., Fujimori, A., Ishida, S., Asano, S., et al.: Photoemission study of itinerant ferromagnet Cr 1− δ Te. Phys. Rev. B. 53, 7673 (1996) 80. Matsui, M., Ido, T., Sato, K., Adachi, K.: Ferromagnetism and Antiferromagnetism in Co–Mn Alloy. J. Phys. Soc. Jpn. 28, 791–791 (1970) 81. Acet, M., John, C., Wassermann, E.: Magnetism and structural stability in CoMn alloys. J. Appl. Phys. 70, 6556–6558 (1991) 82. Paoletti, A., Ricci, F., Passari, L.: Magnetization and State of Order in MnNi3. J. Appl. Phys. 37, 3236–3239 (1966) 83. Coey, J.M.: Magnetism and Magnetic Materials. Cambridge University Press, Cambridge (2010) 84. Hori, T., Nakagawa, Y., Sakurai, J.: Magnetization and magnetic structure of Mn-Zn and MnZn-Ga alloys of CsCl-type structure. J. Phys. Soc. Jpn. 24, 971–976 (1968) 85. Leavey, C., Stewart, J., Rainford, B., Hillier, A.: Magnetic ground states and spin dynamics of β-Mn1− xRux alloys. J. Phys.: Condens. Matter.19, 145288 (2007) 86. Krén, E., Kádár, G., Tarnóczi, T.: Atomic and magnetic order in Mn 2 Pd 3. Phys. Lett. A. 25, 56–57 (1967) 87. Caron, L., Hudl, M., Höglin, V., Dung, N., Gomez, C.P., Sahlberg, M., et al.: Magnetocrystalline anisotropy and the magnetocaloric effect in Fe 2 P. Phys. Rev. B. 88, 094440 (2013)

    782

    J. P. Liu et al.

    88. Cadeville, M.: Proprietes magnetiques des diborures de manganese et de chrome: MnB 2 et CrB 2. J. Phys. Chem. Solids. 27, 667–670 (1966) 89. Neov, S., Legrand, E.: Neutron diffraction study of the magnetic structure of Mn3B4. Phys. Status Solidi B. 49, 589–596 (1972) 90. Lundquist, N., Myers, H., Westin, R.: The paramagnetic properties of the monoborides of V, Cr, Mn, Fe, Co and Ni. Philos. Mag. 7, 1187–1195 (1962) 91. Park, J., Hong, Y., Bae, S., Lee, J., Jalli, J., Abo, G., et al.: Saturation magnetization and crystalline anisotropy calculations for MnAl permanent magnet. J. Appl. Phys. 107, 09A731 (2010) 92. Bither, T.A., Cloud, W.H.: Magnetic tetragonal δ phase in the Mn–Ga binary. J. Appl. Phys. 36, 1501–1502 (1965) 93. Tsuboya, I., Sugihara, M.: Magnetic properties of ζ phase in Mn-Ga system. J. Phys. Soc. Jpn. 18, 1096–1096 (1963) 94. Tsuboya, I., Sugihara, M.: The magnetic properties of ε phase in Mn–Ga system. J. Phys. Soc. Jpn. 18, 143–143 (1963) 95. Kharel, P., Huh, Y., Al-Aqtash, N., Shah, V., Sabirianov, R., Skomski, R., et al.: Structural and magnetic transitions in cubic Mn3Ga. J. Phys.: Condens. Matter. 26, 126001 (2014) 96. Zhang, Q., Li, D., Cui, W., Li, J., Zhang, Z.: Magnetic properties and spin-glass-like behavior in stoichiometric Mn (3) In compound. J. Appl. Phys. 106, 113915 (2009) 97. Karen, P., Fjellvag, H., Kjekshus, A., Andresen, A.: On the phase relations and structural and magnetic properties of the stable manganese carbides Mn23C6, Mn5C2 and Mn7C3. Acta Chem. Scand. 45, 549–557 (1991) 98. Mühlbauer, S., Binz, B., Jonietz, F., Pfleiderer, C., Rosch, A., Neubauer, A., et al.: Skyrmion lattice in a chiral magnet. Science. 323, 915–919 (2009) 99. Yamada, N., Maeda, K., Usami, Y., Ohoyama, T.: Magnetic properties of intermetallic compound Mn11Ge8. J. Phys. Soc. Jpn. 55, 3721–3724 (1986) 100. Kim, Y., Kim, E.J., Choi, K., Han, W.B., Kim, H.-S., Yoon, C.S.: Magnetocaloric effect of Mn 5+ xGe 3− x alloys. J. Alloys Compd. 620, 164–167 (2015) 101. Timoschuk, V., Rozenberg, E.: Magnetic phase transitions of ζ-Mn 5 Ge 2 in magnetic fields up to 25 T. Solid State Commun. 78, 531–534 (1991) 102. Wachtel, E., Henig, E.-T.: Magnetic properties of germanium and germanium- manganese alloys in the solid and liquid states. Z. Metallkd. 60, 316–321 (1969) 103. Yasuk¯ochi, K., Kanematsu, K., Ohoyama, T.: Magnetic Properties of Intermetallic Compounds in Iron-Germanium System: Fe1. 67Ge and FeGe2. J. Phys. Soc. Jpn. 16, 429–433 (1961) 104. Duan, T., Ren, W., Liu, W., Li, S., Liu, W., Zhang, Z.: Magnetic anisotropy of singlecrystalline Mn3Sn in triangular and helix-phase states. Appl. Phys. Lett. 107, 082403 (2015) 105. Mazet, T., Recour, Q., Malaman, B.: Neutron diffraction and S 119 n Mössbauer spectroscopy study of Mn 3 Sn 2. Phys. Rev. B. 81, 174427 (2010) 106. Felcher, G.: Magnetic structure of MnP. J. Appl. Phys. 37, 1056–1058 (1966) 107. Bacmann, M., Fruchart, D., Chenevier, B., Fruchart, R., Puertolas, J., Rillo, C.: Magnetic phase diagram of the (Fe 1− x Mn x) 2 P system. J. Magn. Magn. Mater. 83, 313–314 (1990) 108. Leitão, J., Xinmin, Y., Caron, L., Brück, E.: Magnetostructural study of the (Mn, Fe) 3 (P, Si) system. J. Alloys Compd. 520, 52–58 (2012) 109. Nascimento, F.C., dos Santos, A.O., de Campos, A., Gama, S., Cardoso, L.P.: Structural and magnetic study of the MnAs magnetocaloric compound. Mater. Res. 9, 111–114 (2006) 110. Kanomata, T., Goto, T., Ido, H.: Magnetic phase transitions in the Fe2As-Mn2As system. J. Phys. Soc. Jpn. 43, 1178–1184 (1977) 111. Takei, W., Cox, D., Shirane, G.: Magnetic structures in the MnSb-CrSb system. Phys. Rev. 129, 2008 (1963) 112. Shirakawa, K., Ido, H.: Magnetic transition of the Mn2Sb-Mn2As system. J. Phys. Soc. Jpn. 40, 666–673 (1976) 113. Adachi, K., Sato, K., Takeda, M.: Magnetic properties of some compounds with pyrite structure. J. Appl. Phys. 39, 900–900 (1968)

    15 Metallic Magnetic Materials

    783

    114. Massalski, T., Okamoto, H., Subramanian, P., Kacprzak, L.: Binary Alloys Phase Diagrams, vol. II. ASM International, Materials Park (1990) 2001 115. Brando, M., Duncan, W., Moroni-Klementowicz, D., Albrecht, C., Grüner, D., Ballou, R., et al.: Logarithmic Fermi-liquid breakdown in NbFe 2. Phys. Rev. Lett. 101, 026401 (2008) 116. Nikitin, S., Myalikgulyev, G., Tishin, A., Annaorazov, M., Asatryan, K., Tyurin, A.: The magnetocaloric effect in Fe 49 Rh 51 compound. Phys. Lett. A. 148, 363–366 (1990) 117. Klemmer, T., Hoydick, D., Okumura, H., Zhang, B., Soffa, W.: Magnetic hardening and coercivity mechanisms in L1 0 ordered FePd ferromagnets. Scr. Metall. Mater. 33, 1793– 1805 (1995) 118. Ikeda, K.: Ferromagnetism in hexagonal and cubic Fe 2 Hf compounds (1977) 119. Duijn, H., Brück, E., Menovsky, A., Buschow, K., De Boer, F., Coehoorn, R., et al.: Magnetic and transport properties of the itinerant electron system Hf1− xTaxFe2. J. Appl. Phys. 81, 4218–4220 (1997) 120. Liu, Y., Jiang, Y., Zhang, X., Wang, Y., Zhang, Y., Liu, H., et al.: Structural and magnetic properties of the ordered FePt 3, FePt and Fe 3 Pt nanoparticles. J. Solid State Chem. 209, 69–73 (2014) 121. Zeng, H., Li, J., Liu, J.P., Wang, Z.L., Sun, S.: Exchange-coupled nanocomposite magnets by nanoparticle self-assembly. Nature. 420, 395–398 (2002) 122. Kakeshita, T., Takeuchi, T., Fukuda, T., Tsujiguchi, M., Saburi, T., Oshima, R., et al.: Giant magnetostriction in an ordered Fe3Pt single crystal exhibiting a martensitic transformation. Appl. Phys. Lett. 77, 1502–1504 (2000) 123. Kneller, E., Khan, Y.: The phase Fe sub 2 B. Z. Metallkd. 78, 825–835 (1987) 124. Dobrzynski, L., Petrillo, C., Sacchetti, F.: Iron magnetic moments and spin-density asphericity in Fe 3 (Al x Si 1− x) alloys. Phys. Rev. B. 42, 1142 (1990) 125. Kawamiya, N., Adachi, K.: Magnetic structure of Fe 3 Ga studied by neutron diffraction. Trans. Jpn. Inst. Met. 23, 296–302 (1982) 126. Taheri, P., Barua, R., Hsu, J., Zamanpour, M., Chen, Y., Harris, V.: Structure, magnetism, and magnetostrictive properties of mechanically alloyed Fe 81 Ga 19. J. Alloys Compd. 661, 306–311 (2016) 127. Mendez, J., Ekuma, C., Wu, Y., Fulfer, B., Prestigiacomo, J., Shelton, W., et al.: Competing magnetic states, disorder, and the magnetic character of Fe 3 Ga 4. Phys. Rev. B. 91, 144409 (2015) 128. Kobeissi, M., Hutchings, J., Appleyard, P., Thomas, M., Booth, J.: Mössbauer studies of the alloys Fe3Ga4,(Fe1-xTix) 3Ga4 and (Fe1-yCry) 3Ga4. J. Phys.: Condens. Matter. 11, 6251 (1999) 129. Matsushita, M., Matsushima, Y., Ono, F.: Anomalous structural transformation and magnetism of Fe–Ga alloys. Phys. B. 405, 1154–1157 (2010) 130. Sluchanko, N., Glushkov, V., Demishev, S., Menovsky, A., Weckhuysen, L., Moshchalkov, V.: Crossover in magnetic properties of FeSi. Phys. Rev. B. 65, 064404 (2002) 131. Adams, T., Mühlbauer, S., Neubauer, A., Münzer, W., Jonietz, F., Georgii, R., et al.: Skyrmion lattice domains in Fe1− xCoxSi. J. Phys. Conf. Ser., 032001 (2010) 132. Yasuk¯ochi, K., Kanematsu, K., Ohoyama, T.: Magnetic properties of intermetallic compounds in manganese-Tin system: Mn3. 67Sn, Mn1. 77Sn, and MnSn2. J. Phys. Soc. Jpn. 16, 1123– 1130 (1961) 133. Yu, X., Kanazawa, N., Onose, Y., Kimoto, K., Zhang, W., Ishiwata, S., et al.: Near roomtemperature formation of a skyrmion crystal in thin-films of the helimagnet FeGe. Nat. Mater. 10, 106–109 (2011) 134. Felcher, G., Jorgensen, J.: Magnetic structures of monoclinic FeGe. J. Phys. C: Solid State Phys. 16, 6281 (1983) 135. Beckman, O., Carrander, K., Lundgren, L., Richardson, M.: Susceptibility measurements and magnetic ordering of hexagonal FeGe. Phys. Scr. 6, 151 (1972) 136. Drijver, J., Sinnema, S., Van der Woude, F.: Magnetic properties of hexagonal and cubic Fe3Ge. J. Phys. F: Met. Phys. 6, 2165 (1976)

    784

    J. P. Liu et al.

    137. Venturini, G., Malaman, B., Le Caër, G., Fruchart, D.: Low-temperature magnetic structure of FeSn 2. Phys. Rev. B. 35, 7038 (1987) 138. Häggström, L., Ericsson, T., Wäppling, R., Chandra, K.: Studies of the magnetic structure of FeSn using the Mössbauer effect. Phys. Scr. 11, 47 (1975) 139. Malaman, B., Fruchart, D., Le Caër, G.: Magnetic properties of Fe3Sn2. II. Neutron diffraction study (and Mossbauer effect). J. Phys. F.: Metal Phys. 8, 2389 (1978) 140. Sales, B.C., Saparov, B., McGuire, M.A., Singh, D.J., Parker, D.S.: Ferromagnetism of Fe3Sn and alloys. Sci. Rep. 4 (2014) 141. Felcher, G., Smith, F., Bellavance, D., Wold, A.: Magnetic structure of iron monophosphide. Phys. Rev. B. 3, 3046 (1971) 142. Andersson, Y., Rundqvist, S., Beckman, O., Lundgren, L., Nordblad, P.: Properties of Fe2 P crystals prepared from a liquid copper medium. Phys. Status Solidi A. 49, K153–K156 (1978) 143. Broddefalk, A., Granberg, P., Nordblad, P., Liu, H.-p., Andersson, Y.: Magnetocrystalline anisotropy of (Fe1-xCox) 3P. J. Appl. Phys. 83, 6980–6982 (1998) 144. Selte, K., Kjekshus, A., Andresen, A.: Magnetic structure and properties of FeAs. Acta Chem. Scand. 26, 3101–3113 (1972) 145. Önnerud, P., Andersson, Y., Tellgren, R., Ericsson, T., Nordblad, P., Krishnamurthy, A., et al.: A magnetic structure investigation of tetragonal (Fe 1− x Co x) 2 As. J. Magn. Magn. Mater. 147, 346–354 (1995) 146. Petrovic, C., Kim, J.W., Bud’ko, S.L., Goldman, A., Canfield, P.C., Choe, W., et al.: Anisotropy and large magnetoresistance in the narrow-gap semiconductor FeSb 2. Phys. Rev. B. 67, 155205 (2003) 147. Kom˛edera, K., Jasek, A., Błachowski, A., Ruebenbauer, K., Krzto´n-Maziopa, A.: Magnetic anisotropy in FeSb studied by 57 Fe Mössbauer spectroscopy. J. Magn. Magn. Mater. 399, 221–227 (2016) 148. Buschow, K.: Intermetallic compounds of rare-earth and 3d transition metals. Rep. Prog. Phys. 40, 1179 (1977) 149. Handbook of Magnetic Materials, vol. 4, p. 211, Elseviers Science (1988) 150. Fujii, H., Pourarian, F., Wallace, W.: Appearance of spontaneous ferromagnetism in nonstoichiometric ZrCo 2. J. Magn. Magn. Mater. 24, 93–96 (1981) 151. Pareti, L., Solzi, M., Paoluzi, A.: Magnetocrystalline anisotropy of the 3d sublattice in the cubic intermetallic system Zr6Co23− xMx (M= Fe, Ni). J. Appl. Phys. 73, 2941–2947 (1993) 152. Burzo, E., Grössinger, R., Hundegger, P., Kirchmayr, H., Krewenka, R., Mayerhofer, O., et al.: Magnetic properties of ZrCo5. 1− xFex alloys. J. Appl. Phys. 70, 6550–6552 (1991) 153. Buschow, K.: Differences in magnetic properties between amorphous and crystalline alloys. J. Appl. Phys. 53, 7713–7716 (1982) 154. Crangle, J., Parsons, D.: The magnetization of ferromagnetic binary alloys of cobalt or nickel with elements of the palladium and platinum groups. Proc. Roy. Soc., 509–519 (1960) 155. Shen, Y., Turgut, Z., Horwath, J., Huang, M.: Bulk nanocomposite LaCo5/LaCo13 magnets. J. Appl. Phys. 109, 07A765 (2011) 156. Franse, J., Gerdsdorf, R.: Magnetic properties of metals-3d, 4d and 5d elements, alloys and compounds (Landolt-Börnstein, New Series vol 19a) ed HPJ Wijn. Springer, Berlin (1986) 157. Menzinger, F., Paoletti, A.: Magnetic moments and unpaired-electron densities in Co Pt 3. Phys. Rev. 143, 365 (1966) 158. Cadeville, M., Dahmani, C., Kern, F.: Magnetism and spatial order in Ni-Pt and Co-Pt alloys. J. Magn. Magn. Mater. 54, 1055–1056 (1986) 159. Sanchez, J., Mora-Lopez, ´ J., Leroux, C., Cadeville, M.: Chemical and magnetic ordering in CoPt. J. Phys. Colloques. 49, C8-107–C8-108 (1988) 160. Onnerud, P., Andersson, Y., Tellgren, R., Nordblad, P.: The ferromagnetic structure of hexagonal (Fe1-xCox) 2As. Solid State Commun. 101, 271–275 (1997) 161. Krumbügel-Nylund, A.: Thesis, Orsay, (1974) 162. Paccard, D., Pauthenet, R.: Proprietes Cristallographiques Et Magnetiques Des Alliages De Formule TNI3 Dans Laquelle T Designe Un Metal De Terre Rare Ou Lyttrium. C. R. Heb. Séances Acad. Sci. Ser. B. 264, 1056–1967

    15 Metallic Magnetic Materials

    785

    163. Gignoux, D., Lemaire, R., Molho, P., Tasset, F.: Onset of magnetism in the yttrium-nickel compounds: II. Very weak itinerant ferromagnetism in YNi3. J. Magn. Magn. Mater. 21, 307–315 (1980) 164. Lemaire, R., Paccard, D., Pauthenet, R., Schweizer, J.: Magnetic behavior of cobalt and of nickel in compounds with rare earth metals. J. Appl. Phys. 39, 1092–1093 (1968) 165. Gignoux, D., Lemaire, R., Molho, P.: Onset of magnetism in the yttrium-nickel compounds: I. Collective electron metamagnetism in Y2Ni17. J. Magn. Magn. Mater. 21, 119–124 (1980) 166. Sadron, S.: Ferromagnetic moments of the elements and the periodic system. Ann. Phvs. 17, 371 (1932) 167. Tazuke, Y., Nakabayashi, R., Murayama, S., Sakakibara, T., Goto, T.: Magnetism of R2Ni7 and RNi3 (R= Y, La, Ce). Phys. B. 186, 596–598 (1993) 168. Buschow, K.: Magnetic properties of La 2 Ni 7 and its hydride. J. Magn. Magn. Mater. 40, 224–226 (1983) 169. Kawamiya, N., Adachi, K.: Magnetic properties of ordered and disordered Ni 1− x Fe x Pt. Trans. Jpn. Inst. Met. 16, 327–332 (1975) 170. Parra, R., Cable, J.: Neutron study of magnetic-moment distribution in Ni-Pt alloys. Phys. Rev. B. 21, 5494 (1980) 171. Marian, V.: Ferromagnetic Curie points and the absolute saturation of some nickel alloys. Ann. Phys. 7, 459–527 (1937) 172. De Boer, F., Schinkel, C., Biesterbos, J., Proost, S.: Exchange-enhanced paramagnetism and weak ferromagnetism in the Ni3Al and Ni3Ga phases; Giant moment inducement in FeDoped Ni3Ga. J. Appl. Phys. 40, 1049–1055 (1969) 173. Heczko, O., Scheerbaum, N., Gutfleisch, O.: Magnetic shape memory phenomena. In: Nanoscale Magnetic Materials and Applications, pp. 399–439. Springer (2009) 174. Tickle, R., James, R.: Magnetic and magnetomechanical properties of Ni 2 MnGa. J. Mag.Mag. Mater. 195, 627–638 (1999) 175. Planes, A., Mañosa, L., Acet, M.: Magnetocaloric effect and its relation to shape-memory properties in ferromagnetic Heusler alloys. J. Phys.: Condens. Matter. 21, 233201 (2009) 176. Ullakko, K., Huang, J., Kantner, C., O’handley, R., Kokorin, V.: Large magnetic-field-induced strains in Ni2MnGa single crystals. Appl. Phys. Lett. 69, 1966–1968 (1996) 177. Wu, G., Yu, C., Meng, L., Chen, J., Yang, F., Qi, S., et al.: Giant magnetic-field-induced strains in Heusler alloy NiMnGa with modified composition. Appl. Phys. Lett. 75, 2990–2992 (1999) 178. Yu, C., Wang, W., Chen, J., Wu, G., Yang, F., Tang, N., et al.: Magnetic-field-induced strains and magnetic properties of Heusler alloy Ni52Mn23Ga25. J. Appl. Phys. 87, 6292–6294 (2000) 179. Murray, S.J., Marioni, M., Allen, S., O’handley, R., Lograsso, T.A.: 6% magnetic-fieldinduced strain by twin-boundary motion in ferromagnetic Ni–Mn–Ga. Appl. Phys. Lett. 77, 886–888 (2000) 180. Likhachev, A., Ullakko, K.: Magnetic-field-controlled twin boundaries motion and giant magneto-mechanical effects in Ni–Mn–Ga shape memory alloy. Phys. Lett. A. 275, 142–151 (2000) 181. Heczko, O., Sozinov, A., Ullakko, K.: Giant field-induced reversible strain in magnetic shape memory NiMnGa alloy. IEEE Trans. Magn. 36, 3266–3268 (2000) 182. Wang, W., Wu, G., Chen, J., Gao, S., Zhan, W., Wen, G., et al.: Intermartensitic transformation and magnetic-field-induced strain in Ni52Mn24. 5Ga23. 5 single crystals. Appl. Phys. Lett. 79, 1148–1150 (2001) 183. Sozinov, A., Likhachev, A., Lanska, N., Ullakko, K.: Giant magnetic-field-induced strain in NiMnGa seven-layered martensitic phase. Appl. Phys. Lett. 80, 1746–1748 (2002) 184. Liang, T., Jiang, C., Xu, H.: Temperature dependence of transformation strain and magneticfield-induced strain in Ni 51 Mn 24 Ga 25 single crystal. Mater. Sci. Eng., A. 402, 5–8 (2005) 185. Jiang, C., Wang, J., Xu, H.: Temperature dependence of the giant magnetostrain in a NiMnGa magnetic shape memory alloy. Appl. Phys. Lett. 86, 252508 (2005)

    786

    J. P. Liu et al.

    186. Callaway, J., Sehitoglu, H., Hamilton, R., Aslantas, K., Miller, N., Maier, H., et al.: Magnetic shape memory in Ni2MnGa as influenced by applied stress. Appl. Phys. Lett. 89, 221905 (2006) 187. Popov, A., Belozerov, E., Sagaradze, V., Pecherkina, N., Kabanova, I., Gaviko, V., et al.: Martensitic transformations and magnetic-field-induced strains in Ni50Mn50− x Gax alloys. Physi Met Metallogr. 102, 140–148 (2006) 188. Pagounis, E., Chulist, R., Szczerba, M., Laufenberg, M.: Over 7% magnetic field-induced strain in a Ni-Mn-Ga five-layered martensite. Appl. Phys. Lett. 105, 052405 (2014) 189. Pagounis, E., Chulist, R., Szczerba, M., Laufenberg, M.: High-temperature magnetic shape memory actuation in a Ni–Mn–Ga single crystal. Scr. Mater. 83, 29–32 (2014) 190. Scheerbaum, N., Hinz, D., Gutfleisch, O., Müller, K.-H., Schultz, L.: Textured polymer bonded composites with Ni–Mn–Ga magnetic shape memory particles. Acta Mater. 55, 2707– 2713 (2007) 191. Kauffmann-Weiss, S., Scheerbaum, N., Liu, J., Klauss, H., Schultz, L., Mäder, E., et al.: Reversible magnetic field induced strain in Ni2MnGa-polymer-composites. Adv. Eng. Mater. 14, 20–27 (2012) 192. Scheerbaum, N., Heczko, O., Liu, J., Hinz, D., Schultz, L., Gutfleisch, O.: Magnetic fieldinduced twin boundary motion in polycrystalline Ni–Mn–Ga fibres. New J. Phys. 10, 073002 (2008) 193. Boonyongmaneerat, Y., Chmielus, M., Dunand, D.C., Müllner, P.: Increasing magnetoplasticity in polycrystalline Ni-Mn-Ga by reducing internal constraints through porosity. Phys. Rev. Lett. 99, 247201 (2007) 194. Chmielus, M., Zhang, X., Witherspoon, C., Dunand, D., Müllner, P.: Giant magnetic-fieldinduced strains in polycrystalline Ni–Mn–Ga foams. Nat. Mater. 8, 863–866 (2009) 195. Pötschke, M., Weiss, S., Gaitzsch, U., Cong, D., Hürrich, C., Roth, S., et al.: Magnetically resettable 0.16% free strain in polycrystalline Ni–Mn–Ga plates. Scr. Mater. 63, 383–386 (2010) 196. Gaitzsch, U., Romberg, J., Pötschke, M., Roth, S., Müllner, P.: Stable magnetic-field-induced strain above 1% in polycrystalline Ni–Mn–Ga. Scr. Mater. 65, 679–682 (2011) 197. Gaitzsch, U., Pötschke, M., Roth, S., Rellinghaus, B., Schultz, L.: A 1% magnetostrain in polycrystalline 5M Ni–Mn–Ga. Acta Mater. 57, 365–370 (2009) 198. Liang, T., Jiang, C., Xu, H., Liu, Z., Zhang, M., Cui, Y., et al.: Phase transition strain and large magnetic field induced strain in Ni 50.5 Mn 24 Ga 25.5 unidirectionally solidified alloy. J. Magn. Magn. Mater. 268, 29–32 (2004) 199. Kohl, M., Agarwal, A., Chernenko, V., Ohtsuka, M., Seemann, K.: Shape memory effect and magnetostriction in polycrystalline Ni–Mn–Ga thin film microactuators. Mater. Sci. Eng., A. 438, 940–943 (2006) 200. Liu, G., Chen, J., Liu, Z., Dai, X., Wu, G., Zhang, B., et al.: Martensitic transformation and shape memory effect in a ferromagnetic shape memory alloy: Mn2NiGa. Appl. Phys. Lett. 87, 262504 (2005) 201. Liu, J., Zheng, H., Xia, M., Huang, Y., Li, J.: Martensitic transformation and magnetic properties in Heusler CoNiGa magnetic shape memory alloys. Scr. Mater. 52, 935–938 (2005) 202. Morito, H., Oikawa, K., Fujita, A., Fukamichi, K., Kainuma, R., Ishida, K.: Enhancement of magnetic-field-induced strain in Ni–Fe–Ga–Co Heusler alloy. Scr. Mater. 53, 1237–1240 (2005) 203. Kainuma, R., Imano, Y., Ito, W., Sutou, Y., Morito, H., Okamoto, S., et al.: Magnetic-fieldinduced shape recovery by reverse phase transformation. Nature. 439, 957–960 (2006) 204. Monroe, J., Karaman, I., Basaran, B., Ito, W., Umetsu, R., Kainuma, R., et al.: Direct measurement of large reversible magnetic-field-induced strain in Ni–Co–Mn–In metamagnetic shape memory alloys. Acta Mater. 60, 6883–6891 (2012) 205. Morito, H., Fujita, A., Oikawa, K., Ishida, K., Fukamichi, K., Kainuma, R.: Stress-assisted magnetic-field-induced strain in Ni–Fe–Ga–Co ferromagnetic shape memory alloys. Appl. Phys. Lett. 90, 062505 (2007)

    15 Metallic Magnetic Materials

    787

    206. Morito, H., Oikawa, K., Fujita, A., Fukamichi, K., Kainuma, R., Ishida, K.: A large magneticfield-induced strain in Ni–Fe–Mn–Ga–Co ferromagnetic shape memory alloy. J. Alloys Compd. 577, S372–S375 (2013) 207. Omori, T., Watanabe, K., Umetsu, R., Kainuma, R., Ishida, K.: Martensitic transformation and magnetic field-induced strain in Fe–Mn–Ga shape memory alloy. Appl. Phys. Lett. 95, 082508 (2009) 208. Sugiyama, M., Oshima, R., Fujita, F.E.: Martensitic transformation in the Fe–Pd alloy system. Trans. Jpn. Ins. Met. 25, 585–592 (1984) 209. James, R.D., Wuttig, M.: Magnetostriction of martensite. Philos. Mag. A. 77, 1273–1299 (1998) 210. Kakeshita, T., Fukuda, T.: Giant magnetostriction in Fe3Pt and FePd ferromagnetic shapememory alloys. Mater. Sci. For. 394, 531–536 (2002) 211. Sakon, T., Takaha, A., Obara, K., Dejima, K., Nojiri, H., Motokawa, M., et al.: Magneticfield-induced strain of shape-memory alloy Fe3Pt studied by a capacitance method in a pulsed magnetic field. Jpn. J. Appl. Phys. 46, 146 (2007) 212. Fukuda, T., Kakeshita, T.: Giant magnetic field induced strain in ferromagnetic shape memory alloys and its condition. Mater. Sci. Technol. 24, 890–895 (2008) 213. Liu, E., Zhu, W., Feng, L., Chen, J., Wang, W., Wu, G., et al.: Vacancy-tuned paramagnetic/ferromagnetic martensitic transformation in Mn-poor Mn1-xCoGe alloys. Europhys. Lett. 91, 17003 (2010) 214. Liu, E., Wang, W., Feng, L., Zhu, W., Li, G., Chen, J., et al.: Stable magnetostructural coupling with tunable magnetoresponsive effects in hexagonal ferromagnets. Nat. Commun. 3, 873 (2012) 215. Zhang, C., Shi, H., Ye, E., Nie, Y., Han, Z., Wang, D.: Magnetostructural transition and magnetocaloric effect in MnCoGe–NiCoGe system. J. Alloys Compd. 639, 36–39 (2015) 216. Wei, Z.Y., Liu, E.K., Li, Y., Xu, G.Z., Zhang, X.M., Liu, G.D., et al.: Unprecedentedly wide Curie-temperature windows as phase-transition design platform for tunable magnetomultifunctional materials. Adv. Electron. Mater. 1 (2015) 217. Fujita, A., Fukamichi, K., Gejima, F., Kainuma, R., Ishida, K.: Magnetic properties and large magnetic-field-induced strains in off-stoichiometric Ni-Mn-Al Heusler alloys. Appl. Phys. Lett. 77, 3054 (2000) 218. Liu, Z., Zhang, M., Cui, Y., Zhou, Y., Wang, W., Wu, G., et al.: Martensitic transformation and shape memory effect in ferromagnetic Heusler alloy Ni2FeGa. Appl. Phys. Lett. 82, 424–426 (2003) 219. Krenke, T., Acet, M., Wassermann, E.F., Moya, X., Mañosa, L., Planes, A.: Ferromagnetism in the austenitic and martensitic states of Ni− Mn− In alloys. Phys. Rev. B. 73, 174413 (2006) 220. Zhang, M., Brück, E., de Boer, F.R., Wu, G.: Magnetic, martensitic transformation, magnetostriction and shape memory effect in Co50Ni20Ga30 melt-spun ribbons. J. Phys. D: Appl. Phys. 38, 1361 (2005) 221. Li, Z., Xu, K., Yang, H., Zhang, Y., Jing, C.: Magnetostrain and magnetocaloric effect by field-induced reverse martensitic transformation for Pd-doped Ni45Co5Mn37In13 Heusler alloy. J. Appl. Phys. 117, 223904 (2015) 222. Jing, C., Wang, X., Liao, P., Li, Z., Yang, Y., Kang, B., et al.: Martensitic phase transition, inverse magnetocaloric effect, and magnetostrain in Ni50Mn37-xFexIn13 Heusler alloys. J. Appl. Phys. 114, 063907 (2013) 223. Kainuma, R., Imano, Y., Ito, W., Morito, H., Sutou, Y., Oikawa, K., et al.: Metamagnetic shape memory effect in a Heusler-type Ni43Co7Mn39Sn11 polycrystalline alloy. Appl. Phys. Lett. 88, 192513 (2006) 224. Liao, P., Jing, C., Zheng, D., Li, Z., Kang, B., Deng, D., et al.: Tuning martensitic transformation, large magnetoresistance and strain in Ni 50− x Fe x Mn 36 Sn 14 Heusler alloys. Solid State Commun. 217, 28–33 (2015) 225. Chernenko, V., Barandiarán, J., L’vov, V., Gutiérrez, J., Lázpita, P., Orue, I.: Temperature dependent magnetostrains in polycrystalline magnetic shape memory Heusler alloys. J. Alloys Compd. 577, S305–S308 (2013)

    788

    J. P. Liu et al.

    226. Li, Z., Jing, C., Zhang, H., Qiao, Y., Cao, S., Zhang, J., et al.: A considerable metamagnetic shape memory effect without any prestrain in Ni46Cu4Mn38Sn12 Heusler alloy. J. Appl. Phys. 106, 083908 (2009) 227. Zhu, W., Liu, E., Feng, L., Tang, X., Chen, J., Wu, G., et al.: Magnetic-field-induced transformation in FeMnGa alloys. Appl. Phys. Lett. 95, 222512–222511 (2009) 228. Wei, Z., Liu, E., Chen, J., Li, Y., Liu, G., Luo, H., et al.: Realization of multifunctional shapememory ferromagnets in all-d-metal Heusler phases. Appl. Phys. Lett. 107, 022406 (2015) 229. Felser, C., Fecher, G.H., Balke, B.: Spintronics: a challenge for materials science and solidstate chemistry. Angew. Chem. Int. Ed. 46, 668–699 (2007) 230. Graf, T., Felser, C., Parkin, S.S.: Simple rules for the understanding of Heusler compounds. Prog. Solid State Chem. 39, 1–50 (2011) 231. Villar, P., Calvert, L.: Pearson’s Handbook of Crystallographic Data for Intermetallic Phases, 2nd edn. ASM International, Materials Park (1996) 232. Kandpal, H.C., Felser, C., Seshadri, R.: Covalent bonding and the nature of band gaps in some half-Heusler compounds. J. Phys. D: Appl. Phys. 39, 776 (2006) 233. Pan, Y., Nikitin, A., Bay, T., Huang, Y., Paulsen, C., Yan, B., et al.: Superconductivity and magnetic order in the noncentrosymmetric half-Heusler compound ErPdBi. Europhys. Lett. 104, 27001 (2013) 234. Nikitin, A., Pan, Y., Mao, X., Jehee, R., Araizi, G., Huang, Y., et al.: Magnetic and superconducting phase diagram of the half-Heusler topological semimetal HoPdBi. J. Phys.: Condens. Matter. 27, 275701 (2015) 235. De Groot, R., Mueller, F., Van Engen, P., Buschow, K.: New class of materials: half-metallic ferromagnets. Phys. Rev. Lett. 50, 2024 (1983) 236. Felser, C., Wollmann, L., Chadov, S., Fecher, G.H., Parkin, S.S.: Basics and prospective of magnetic Heusler compounds. APL Mater. 3, 041518 (2015) 237. Felser, C., Wollmann, L., Chadov, S., Fecher, G.H., Parkin, S.S.P.: Heusler alloys. In: Felser, C., Hirohata, A. (eds.) Springer Series in Materials Science, vol. 222, pp. 37–48. Springer International Publishing (2016) 238. Nishino, Y., Kato, M., Asano, S., Soda, K., Hayasaki, M., Mizutani, U.: Semiconductorlike behavior of electrical resistivity in Heusler-type Fe 2 VAl compound. Phys. Rev. Lett. 79, 1909 (1997) 239. Galanakis, I., Dederichs, P., Papanikolaou, N.: Slater-Pauling behavior and origin of the halfmetallicity of the full-Heusler alloys. Phys. Rev. B. 66, 174429 (2002) 240. Felser, C., Fecher, G.H.: Spintronics: From Materials to Devices. Springer Science & Business Media (2013) 241. Kandpal, H.C., Ksenofontov, V., Wojcik, M., Seshadri, R., Felser, C.: Electronic structure, magnetism and disorder in the Heusler compound Co2TiSn. J. Phys. D: Appl. Phys. 40, 1587 (2007) 242. Liu, H.-x., Kawami, T., Moges, K., Uemura, T., Yamamoto, M., Shi, F., et al.: Influence of film composition in quaternary Heusler alloy Co2 (Mn, Fe) Si thin films on tunnelling magnetoresistance of Co2 (Mn, Fe) Si/MgO-based magnetic tunnel junctions. J. Phys. D: Appl. Phys. 48, 164001 (2015) 243. Jourdan, M., Minár, J., Braun, J., Kronenberg, A., Chadov, S., Balke, B., et al.: Direct observation of half-metallicity in the Heusler compound Co2MnSi. Nat. Commun. 5 (2014) 244. Wurmehl, S., Fecher, G.H., Kandpal, H.C., Ksenofontov, V., Felser, C., Lin, H.-J., et al.: Geometric, electronic, and magnetic structure of Co 2 FeSi: Curie temperature and magnetic moment measurements and calculations. Phys. Rev. B. 72, 184434 (2005) 245. Kübler, J., Fecher, G., Felser, C.: Understanding the trend in the Curie temperatures of Co 2-based Heusler compounds: Ab initio calculations. Phys. Rev. B. 76, 024414 (2007) 246. Wurmehl, S., Kandpal, H.C., Fecher, G.H., Felser, C.: Valence electron rules for prediction of half-metallic compensated-ferrimagnetic behaviour of Heusler compounds with complete spin polarization. J. Phys.: Condens. Matter. 18, 6171 (2006) 247. Wollmann, L., Chadov, S., Kübler, J., Felser, C.: Magnetism in cubic manganese-rich Heusler compounds. Phys. Rev. B. 90, 214420 (2014)

    15 Metallic Magnetic Materials

    789

    248. Galanakis, I., Özdogan, ˇ K., Akta¸s, B., Sa¸ ¸ sinoglu, ˇ E.: Ferrimagnetism and antiferromagnetism in half-metallic Heusler alloys. Phys. Status Solidi A. 205, 1036–1039 (2008) 249. Wollmann, L., Chadov, S., Kübler, J., Felser, C.: Magnetism in tetragonal manganese-rich Heusler compounds. Phys. Rev. B. 92, 064417 (2015) 250. Kübler, J., William, A., Sommers, C.: Formation and coupling of magnetic moments in Heusler alloys. Phys. Rev. B. 28, 1745 (1983) 251. Ouardi, S., Fecher, G.H., Stinshoff, R., Felser, C., Kubota, T., Mizukami, S., et al.: Stoichiometry dependent phase transition in Mn-Co-Ga-based thin films: from cubic in-plane, soft magnetized to tetragonal perpendicular, hard magnetized. Appl. Phys. Lett. 101 (2012) 252. Balke, B., Fecher, G.H., Winterlik, J., Felser, C.: Mn3Ga, a compensated ferrimagnet with high Curie temperature and low magnetic moment for spin torque transfer applications. Appl. Phys. Lett. 90, 2504 (2007) 253. Coey, J.: Permanent magnets: plugging the gap. Scr. Mater. 67, 524–529 (2012) 254. Winterlik, J., Chadov, S., Gupta, A., Alijani, V., Gasi, T., Filsinger, K., et al.: Design scheme of new tetragonal Heusler compounds for spin-transfer torque applications and its experimental realization. Adv. Mater. 24, 6283–6287 (2012) 255. Siewert, M., Gruner, M., Dannenberg, A., Chakrabarti, A., Herper, H., Wuttig, M., et al.: Designing shape-memory Heusler alloys from first-principles. Appl. Phys. Lett. 99, 191904 (2011) 256. Suits, J.C.: New magnetic compounds with Heusler and Heusler-related structures. Phys. Rev. B. 14, 4131 (1976) 257. Kurt, H., Rode, K., Stamenov, P., Venkatesan, M., Lau, Y.-C., Fonda, E., et al.: Cubic Mn 2 Ga thin films: crossing the spin gap with ruthenium. Phys. Rev. Lett. 112, 027201 (2014) 258. Galanakis, I., Sa¸ ¸ sıo˘glu, E.: High TC half-metallic fully-compensated ferrimagnetic Heusler compounds. Appl. Phys. Lett. 99, 052509 (2011) 259. Nayak, A.K., Nicklas, M., Chadov, S., Khuntia, P., Shekhar, C., Kalache, A., et al.: Design of compensated ferrimagnetic Heusler alloys for giant tunable exchange bias. Nat. Mater. 14, 679–684 (2015) 260. Yoon, S., Booth, J.: Structural and magnetic properties of Fe 3− x Mn x Si alloys. Phys. Lett. A. 48, 381–382 (1974) 261. Mohn, P., Supanetz, E.: Spin ordering in Fe3-x Mn x Si Heusler alloys. Philos. Mag. B. 78, 629–636 (1998) 262. Meshcheriakova, O., Chadov, S., Nayak, A., Rößler, U., Kübler, J., André, G., et al.: Large noncollinearity and spin reorientation in the novel Mn 2 RhSn Heusler magnet. Phys. Rev. Lett. 113, 087203 (2014) 263. Buschow, K.: The crystal structure of Th2Co7. Acta Crystallogr. Sect. B. 26, 1389–1392 (1970) 264. Buschow, K.: Structural and magnetic characteristics of Th–Co and Th–Fe compounds. J. Appl. Phys. 42, 3433–3437 (1971) 265. Bartashevich, M., Deryagin, A., Kudrevatykh, N., Tarasov, E.: High-temperature metamagnetism of the Y2CO7H6 hydride. Zhurnal Eksperimentalnoi I Teoreticheskoi Fiziki. 84, 1140–1144 (1983) 266. Andreev, A., Bartashevich, M.: AV Der3’agin, tnd Ye, N. Tarasov. Phys. Met. Metallogr. 62, 66 (1986) 267. Kuz’min, M., Skokov, K., Radulov, I., Schwöbel, C., Foro, S., Donner, W., et al.: Magnetic anisotropy of La2Co7. J. Appl. Phys. 118, 053905 (2015) 268. Buschow, K.: Magnetic properties of CeCo3, Ce2Co7 and CeNi3 and their ternary hydrides. J. Less-Common Met. 72, 257–263 (1980) 269. Buschow, K.H.J.: Colloq. Int. CNRS, Les Elements de Terres Rares 1969, 1, 101 (1970) 270. Ray, A.E., Biermann, A.T., Hammer, R.S., Davison, J.E.: Cobalt. 4, 103 (1973) 271. Bartashevich, M., Andreev, A., Tarasov, E., Goto, T., Yamaguchi, M.: Magnetic properties and spontaneous magnetostriction of a Sm2Co7 single crystal. Phys. B. 183, 369–378 (1993)

    790

    J. P. Liu et al.

    272. Bartashevich, M., Goto, T., Yamaguchi, M.: Field induced magnetic phase transition and magnetostriction in ErCo 3, HoCo 3 and Nd 2 Co 7 single crystals. J. Magn. Magn. Mater. 111, 83–89 (1992) 273. Lemaire, R.: Magnetic properties of the intermetallic compounds of cobalt with the rare earth metals and yttrium. Cobalt, 201–211 (1966) 274. Andreev, A., Tarasov, E., Deryagin, A., Zadvorkin, S.: Influence of symmetry of the crystalline structure on the magnetic anisotropy of Tb2Co7. Phys. Status Solidi A. 71, K245– K247 (1982) 275. Ostertag, W.: The crystal structure of Er2Co7 and other rare earth-cobalt compounds R2Co7 (R = Gd, Tb, Dy, Ho, Tm, Lu, Y). J. Less-Common Met. 13, 385–390 (1967) 276. Andreev, A., Zadvorkin, S., Tarasov, E.: Thermal expansion anomalies and spontaneous magnetostriction in Tm 2 Co 7. J. Alloys Compd. 189, 187–190 (1992) 277. Parker, F., Oesterreicher, H.: Magnetic properties of La2Ni7. J. Less-Common Met. 90, 127– 136 (1983) 278. Lemaire, R., Paccard, D., Pauthenet, R.: Structural and – magnetic properties of alloys of composition T 2 NI 7, where T is a rare-earth metal or yttrium. Compt. Rend. 265(23), 1280– 1282 (1967) 279. Virkar, A.V., Raman, A.: Crystal structures of AB3 and A2B7 rare earth-nickel phases. J. Less-Common Met. 18, 59–66 (1969) 280. Buschow, K., Van Der Goot, A.: The crystal structure of rare-earth nickel compounds of the type R 2 Ni 7. J. Less-Common Met. 22, 419–428 (1970) 281. Bhattacharyya, A., Giri, S., Majumdar, S.: Successive magnetic transitions and low temperature magnetocaloric effect in RE2Ni7 (RE=Dy, Ho). J. Magn. Magn. Mater. 323, 1484–1489 (2011) 282. Tsvyashchenko, A.V.: High-pressure synthesis of YB3CO, YB3NI compounds. J. LessCommon Met. 118, 103–107 (1986) 283. Palenzona, A., Cirafici, S.: The equilibrium diagram of the Th-Ni system. J. Less-Common Met. 142, 311–317 (1988) 284. Schneider, G., Landgraf, F., Villas-Boas, V., Bezerra, G., Missell, F., Ray, A.: New stable phase in the binary Fe-Nd system. Mater. Lett. 8, 472–476 (1989) 285. Moreau, J.M., Paccard, L., Nozieres, J.P., Missell, F.P., Schneider, G., Villas-Boas, V.: A new phase in the Nd-Fe system: Crystal structure of Nd5Fe17. J. Less Common Met. 163, 245–251 (1990) 286. Nozieres, J., Rechenberg, H.: Magnetic properties of Nd5Fe17. Solid State Commun. 79, 21–24 (1991) 287. Stadelmaier, H., Schneider, G., Henig, E.-T., Ellner, M.: Magnetic Fe17R5 in the Fe-Nd and Fe (-Ti)-Sm systems, and other phases in Fe-Nd. Mater. Lett. 10, 303–309 (1991) 288. Maruyama, F., Amako, Y.: Magnetic properties of Sm 5 (Fe 1− x T x) 17 (T= Ti and V) compounds. J. Alloys Compd. 474, 1–3 (2009) 289. Saito, T.: High coercivity in Sm 5 Fe 17 melt-spun ribbon. J. Alloys Compd. 440, 315–318 (2007) 290. Saito, T., Ichihara, M.: Synthesis and magnetic properties of Sm 5 Fe 17 hard magnetic phase. Scr. Mater. 57, 457–460 (2007) 291. Yelon, W., Luo, H., Chen, M., Missell, F.: Magnetic and crystallographic structure of Nd5Fe17. J. Appl. Phys. 85, 5693–5695 (1999) 292. Chu, Z., Yelon, W., Missell, F., Murakami, R.: Site occupancy of Sm in (Nd1− xSmx) 5 (Fe1− yTiy) 17. J. Appl. Phys. 87, 6704–6706 (2000) 293. Saito, T.: Annealing of amorphous Sm5Fe17 melt-spun ribbon. Mater. Trans. 49, 1446–1450 (2008) 294. Stadelmaier, H., Schneider, G., Ellner, M.: A CaCu5-type iron-neodymium phase stabilized by rapid solidification. J. Less-Common Met. 115, L11–L14 (1986) 295. Moze, O., Pareti, L., Paoluzi, A., Buschow, K.: Magnetic structure and anisotropy of Ga-and Al-substituted LaCo 5 and YCo 5 intermetallics. Phys. Rev. B. 53, 11550 (1996)

    15 Metallic Magnetic Materials

    791

    296. Alameda, J., Givord, D., Lemaire, R., Lu, Q.: Co energy and magnetization anisotropies in RCo5 intermetallics between 4.2 K and 300 K. J. Appl. Phys. 52, 2079–2081 (1981) 297. Alameda, J., Deportes, J., Givord, D., Lemaire, R., Lu, Q.: Large magnetization anisotropy in uniaxial YCo5 intermetallic. J. Magn. Magn. Mater. 15, 1257–1258 (1980) 298. Barbara, B., Uehara, M.: Anisotropy and coercivity in SmCo 5-based compounds. IEEE Trans. Magn. 12, 997–999 (1976) 299. Gubbens, P., Van der Kraan, A.: Magnetic properties of ThFe 5. J. Magn. Magn. Mater. 9, 349–354 (1978) 300. Cadieu, F., Cheung, T., Wickramasekara, L., Aly, S.: Magnetic properties of a metastable Sm-Fe phase synthesized by selectively thermalized sputtering. J. Appl. Phys. 55, 2611–2613 (1984) 301. Alameda, J., Givord, D., Lemaire, R., Lu, Q., Palmer, S., Tasset, F.: Reduced 4f-moment of the Nd ground state in NdCo5. Le Journal de Physique Colloques. 43, C7-133–C7-139 (1982) 302. Schweizer, J., Tasset, F.: Polarised neutron study of the RCo5 intermetallic compounds. I. The cobalt magnetisation in YCo5. J. Phys. F: Met. Phys. 10, 2799 (1980) 303. Decrop, B., Deportes, J., Lemaire, R.: Magnetic structure of HoCo 5 below room temperature. J. Less-Common Met. 94, 199–203 (1983) 304. Coroian, N., Klosek, V., Isnard, O.: The influence of substituting Si for Co on the magnetic properties of PrCo 5. J. Alloys Compd. 427, 5–10 (2007) 305. Barthem, V., Gignoux, D., Nait-Saada, A., Schmitt, D., Creuzet, G.: Magnetic and magnetoelastic properties of PrNi 5 single crystal. Phys. Rev. B. 37, 1733 (1988) 306. Coldea, M., Andreica, D., Bitu, M., Crisan, V.: Spin fluctuations in YNi 5 and CeNi 5. J. Magn. Magn. Mater. 157, 627–628 (1996) 307. Grechnev, G., Logosha, A., Panfilov, A., Kuchin, A., Vasijev, A.: Effect of pressure on the magnetic properties of YNi5, LaNi5, and CeNi5. Low Temp. Phys. 37, 138 (2011) 308. Burzo, E., Ursu, I.: Paramagnetic resonance and magnetic measurements on GdNi 5 compound. Solid State Commun. 9, 2289–2292 (1971) 309. Barthem, V., Gignoux, D., Naitsaada, A., Schmitt, D., Takeuchi, A.Y.: Magnetic-properties of the hexagonal NDNI5 and NDCU5 Compounds. J. Magn. Magn. Mater. 80, 142–148 (1989) 310. Gubbens, P., Van Der Kraan, A., Buschow, K.: Crystal-field effects and magnetic behavior in R Ni 5 and R Co 5+ x rare-earth compounds. Phys. Rev. B. 39, 12548 (1989) 311. Gignoux, D., Nait-Saada, A., De La Bâthie, R.P.: Magnetic properties of TbNi5 and HoNi5 single crystals. Le Journal de Physique Colloques. 40, C5-188–C5-190 (1979) 312. Sada, A.N.: Ph.D. thesis, University of Grenoble (1980) 313. Givord, D., Laforest, J., Schweizer, J., Tasset, F.: Temperature dependence of the samarium magnetic form factor in SmCo5. J. Appl. Phys. 50, 2008–2010 (1979) 314. Buschow, K., Van Diepen, A., De Wijn, H.: Crystal-field anisotropy of Sm 3+ in SmCo 5. Solid State Commun. 15, 903–906 (1974) 315. Laforest, J.: Ph.D. thesis, Université de Grenoble (1981) 316. Sankar, S., Rao, V., Segal, E., Wallace, W., Frederick, W., Garrett, H.: Magnetocrystalline anisotropy of Sm Co 5 and its interpretation on a crystal-field model. Phys. Rev. B. 11, 435 (1975) 317. Von Ranke, P., Mota, M., Grangeia, D., Carvalho, A.M.G., Gandra, F., Coelho, A., et al.: Magnetocaloric effect in the R Ni 5 (R= Pr, Nd, Gd, Tb, Dy, Ho, Er) series. Phys. Rev. B. 70, 134428 (2004) 318. Skokov, K.P., Pastushenkov, Y.G., Koshkid’ko, Y.S., Shütz, G., Goll, D., Ivanova, T.I., et al.: Magnetocaloric effect, magnetic domain structure and spin-reorientation transitions in HoCo5 single crystals. J. Magn. Magn. Mater. 323, 447–450 (2011) 319. Givord, D., Laforest, J., Lemaire, R.: Polarized neutron study of the itinerant electron metamagnetism in ThCo5. J. Appl. Phys. 50, 7489–7491 (1979) 320. Givord, D., Laforest, J., Lemaire, R.: Magnetic transition in ThCo5 due to change of Comoment. Phys. B. 86–88, Part 1, 204–206 (1977) 321. Ballou, R., Shimizu, M., Voiron, J.: Pressure effects on the metamagnetic transition of ThCo 5. J. Magn. Magn. Mater. 84, 23–28 (1990)

    792

    J. P. Liu et al.

    322. Velge, W., Buschow, K.: Magnetic and crystallographic properties of some rare earth cobalt compounds with CaZn5 structure. J. Appl. Phys. 39, 1717–1720 (1968) 323. Wernick, J., Geller, S.: Transition element–rare earth compounds with Cu5Ca structure. Acta Crystallogr. 12, 662–665 (1959) 324. Klein, H.P., Menth, A., Perkins, R.S.: Magnetocrystalline anisotropy of light rare-earth cobalt compounds. Phys. B. 80, 153–163 (1975) 325. Ibarra, M., Morellon, L., Algarabel, P., Moze, O.: Single-ion competing magnetic anisotropies in Pr x Nd 1− x Co 5 intermetallic compounds. Phys. Rev. B. 44, 9368 (1991) 326. Tatsumoto, E., Okamoto, T., Fujii, H., Inoue, C.: Saturation magnetic moment and crystalline anisotropy of single crystals of light rare earth cobalt compounds RCo5. Le Journal de Physique Colloques. 32, C1-550–C1-551 (1971) 327. Decrop, B.: Ph.D. thesis, INPG Grenoble (1982) 328. Lemaire, R., Schweizer, J.: Structures magnétiques des composés intermétalliques CeCo5 et TbCo5. J. Phys. 28, 216–220 (1967) 329. Ohkoshi, M., Kobayashi, H., Katayama, T., Hirano, M., Tsushima, T.: Spin reorientation in DyCo 5. Phys. B. 86, 195–196 (1977) 330. Buschow, K., Velge, W.: Permanent magnetic materials of rare earth – cobalt compounds. Philips Research Lab. and Metallurgical Lab, Eindhoven (1969) 331. Givord, D., Laforest, J., Lemaire, R., Lu, Q.: Cobalt magnetism in RCo5-intermetallics: onset of 3d magnetism and magnetocrystalline anisotropy (r=rare earth or Th). J. Magn. Magn. Mater. 31–34, Part 1, 191–196 (1983) 332. Van Diepen, A., Buschow, K., Van Wieringen, J.: Study of the Mössbauer effect, magnetization, and crystal structures of the pseudobinary compounds ThCo5− 5xFe5x and ThNi5− 5xFe5x. J. Appl. Phys. 43, 645–650 (1972) 333. Ballou, R., Barthem, V., Gignoux, D.: Crystal field effects in the hexagonal SmNi5 compound. Phys. B. 149, 340–344 (1988) 334. Oliver, F., West, K., Cohen, R., Buschow, K.: Mossbauer effect of 151Eu in EuNi5, EuMg2 and their hydrides. J. Phys. F: Met. Phys. 8, 701 (1978) 335. De Jesus, V., Oliveira, I., Riedi, P., Guimaraes, A.: 155,157 Gd NMR study of Gd–Ni intermetallic compounds. J. Magn. Magn. Mater. 212, 125–137 (2000) 336. Mulder, F., Thiel, R., Buschow, K.: 155 Gd Mössbauer effect and magnetic properties of aluminium-and gallium-substituted GdCu 5 and GdNi 5. J. Alloys Compd. 190, 77–82 (1992) 337. Haldar, A., Dhiman, I., Das, A., Suresh, K., Nigam, A.: Magnetic, magnetocaloric and neutron diffraction studies on TbNi 5− x M x (M= Co and Fe) compounds. J. Alloys Compd. 509, 3760–3765 (2011) 338. Mulders, A., Kaiser, C., Harker, S., Gubbens, P., Amato, A., Gygax, F., et al.: Muon location and muon dynamics in DyNi 5. Phys. Rev. B. 67, 014303 (2003) 339. Barthem, V., Gignoux, D., Schmitt, D., Creuzet, G.: Magnetic and magnetoelastic properties of the hexagonal TmNi 5 compound. J. Magn. Magn. Mater. 78, 56–66 (1989) 340. Zhang, F., Gignoux, D., Schmitt, D., Franse, J., Kayzel, F., Kim-Ngan, N., et al.: Crystalline electric field and high field magnetization in ErNi 5 single crystal. J. Magn. Magn. Mater. 130, 108–114 (1994) 341. Hodges, J., Bonville, P., Ocio, M.: Magnetic properties of YbNi5 from 170Yb Mössbauer and magnetisation measurements. Eur. Phys. J. B. 57, 365–370 (2007) 342. Knyazev, Y.V., Lukoyanov, A., Kuz’min, Y.I., Kuchin, A.: Influence of copper impurities on the evolution of the electronic structure and optical spectra of the LuNi5 compound. Phys. Solid State. 57, 866–870 (2015) 343. Florio, J.V., Baenziger, N., Rundle, R.: Compounds of thorium with transition metals. II. Systems with iron, cobalt and nickel. Acta Crystallogr. 9, 367–372 (1956) 344. Narasimhan, K., Do-Dinh, C., Wallace, W., Hutchens, R.: Magnetic properties of the ThCo5− xNix system. J. Appl. Phys. 46, 4961–4964 (1975) 345. Aubert, G., Gignoux, D., Hennion, B., Michelutti, B., Saada, A.N.: Bulk magnetization study of a DyNi 5 single crystal. Solid State Commun. 37, 741–743 (1981)

    15 Metallic Magnetic Materials

    793

    346. Boucherle, J., Givord, D., Laforest, J., Schweizer, J., Tasset, F.: Determination of exchange and crystal field effects in Sm alloys by polarized neutron diffraction. Le Journal de Physique Colloques. 40, C5-180–C5-182 (1979) 347. Lu, Q.: PhD Thesis, Université de Grenoble (1981). 348. Ibarra, M., Morellon, L., Algarabel, P., Moze, O.: A determination of the crystal electric field and exchange parameters of Pr 3+ and Nd 3+ ions in RCo 5 intermetallics. J. Magn. Magn. Mater. 104, 1149–1151 (1992) 349. Decrop, B., Deportes, J., Givord, D., Lemaire, R., Chapert, J.: Study of the magnetization reorientation in HoCo5. J. Appl. Phys. 53, 1953–1955 (1982) 350. Kren, E., Schweizer, J., Tasset, F.: Polarized-neutron-diffraction study of magnetic moments in yttrium-cobalt alloys. Phys. Rev. 186, 479 (1969) 351. Buschow, K.: The crystal structures of the rare-earth compounds of the form R 2 Ni 17, R 2 Co 17 and R 2 Fe 17. J. Less-Common Met. 11, 204–208 (1966) 352. Givord, D., Givord, F., Lemaire, R., James, W.J., Shah, J.: Evidence of disordered substitutions in the “Th 2 Ni 17-type” structure. Exact structure determination of the Th-Ni, Y-Ni and Er-Co compounds. J. Less-Common Met. 29, 389–396 (1972) 353. Isnard, O., Hautot, D., Long, G.J., Grandjean, F.: A structural, magnetic, and Mössbauer spectral study of Dy2Fe17 and its hydrides. J. Appl. Phys. 88, 2750–2759 (2000) 354. Christensen, A.N., Hasell, R.G.: Acta Chem. Scand A. 34, 6 (1980) 355. Averbuch-Pouchot, M., Chevalier, R., Deportes, J., Kebe, B., Lemaire, R.: Anisotropy of the magnetization and of the iron hyperfine field in R 2 Fe 17 compounds. J. Magn. Magn. Mater. 68, 190–196 (1987) 356. Burzo, E., Chelkowski, A., Kirchmayr, H.: Magnetic Properties of Metals: Compounds Between Rare Earth Elements and 3d, 4d or 5d Elements HPJ Wijn, Landolt-Börnstein, New Series, Group III, vol. 19, (1990) 357. Givord, D., Lemaire, R.: Magnetic transition and anomalous thermal expansion in R 2 Fe 17 compounds. IEEE Trans. Magn. 10, 109–113 (1974) 358. Givord, D., Lemaire, R.: Ferromagnetic-Helimagnetic Transition in the Compounds LuFe 9. 5 and Ce 2 Fe 17. Compt. Rend. 274, 1166–1169 (1972) 359. Prokhnenko, O., Ritter, C., Arnold, Z., Isnard, O., Kamarád, J., Pirogov, A., et al.: Neutron diffraction studies of the magnetic phase transitions in Ce2Fe17 compound under pressure. J. Appl. Phys. 92, 385–391 (2002) 360. Kamarád, J., Prokhnenko, O., Prokeš, K., Arnold, Z.: Magnetization and neutron diffraction studies of Lu2Fe17 under high pressure. J. Phys.: Condens. Matter. 17, S3069 (2005) 361. Kreyssig, A., Chang, S., Janssen, Y., Kim, J., Nandi, S., Yan, J., et al.: Crystallographic phase transition within the magnetically ordered state of Ce 2 Fe 17. Phys. Rev. B. 76, 054421 (2007) 362. Janssen, Y., Chang, S., Kreyssig, A., Kracher, A., Mozharivskyj, Y., Misra, S., et al.: Magnetic phase diagram of Ce 2 Fe 17. Phys. Rev. B. 76, 054420 (2007) 363. Prokhnenko, O., Kamarád, J., Prokeš, K., Arnold, Z., Andreev, A.: Helimagnetism of Fe: High Pressure Study of an Y 2 Fe 17 Single Crystal. Phys. Rev. Lett. 94, 107201 (2005) 364. Kamarád, J., Prokhnenko, O., Prokeš, K., Arnold, Z., Andreev, A.: Pressure induced helimagnetism in Fe-based (Y 2 Fe 17, Lu 2 Fe 17) intermetallic compounds. J. Magn. Magn. Mater. 310, 1801–1803 (2007) 365. Prokhnenko, O., Ritter, C., Medvedeva, I., Arnold, Z., Kamarád, J., Kuchin, A.: Neutron diffraction study of Lu 2 Fe 17 under high pressure. J. Magn. Magn. Mater. 258, 564–566 (2003) 366. Burzo, E.: F. Givord CR hebd. séanc. Acad. Sci. Paris B. 271, 1159 (1970) 367. Elemans, J., Buschow, K.: The Magnetic Structure of Tm2Fe17. Phys. Status Solidi A. 24, K125–K127 (1974) 368. Isnard, O., Andreev, A., Kuz’min, M., Skourski, Y., Gorbunov, D., Wosnitza, J., et al.: High magnetic field study of the Tm 2 Fe 17 and Tm 2 Fe 17 D 3.2 compounds. Phys. Rev. B. 88, 174406 (2013)

    794

    J. P. Liu et al.

    369. Deportes, J., Kebe, B., Lemaire, R.: Hyperfine field anisotropy in RE-Fe compounds. J. Magn. Magn. Mater. 54, 1089–1090 (1986) 370. Koyama, K., Fujii, H., Canfield, P.: Magnetocrystalline anisotropy of a Nd 2 Fe 17 single crystal. Phys. B. 226, 363–369 (1996) 371. Clausen, K., Nielsen, O.V.: Magnetic anisotropy in single crystals of Ho2Co17 and Ho2Fe17. J. Magn. Magn. Mater. 23, 237–240 (1981) 372. Isnard, O., Miraglia, S., Guillot, M., Fruchart, D.: High field magnetization measurements of Sm2Fe17, Sm2Fe17N3, Sm2Fe17D5, and Pr2Fe17, Pr2Fe17N3. J. Appl. Phys. 75, 5988– 5993 (1994) 373. Kou, X., De Boer, F., Grössinger, R., Wiesinger, G., Suzuki, H., Kitazawa, H., et al.: Magnetic anisotropy and magnetic phase transitions in R 2 Fe 17 with R= Y, Ce, Pr, Nd, Sm, Gd, Tb, Dy, Ho, Er, Tm and Lu. J. Magn. Magn. Mater. 177, 1002–1007 (1998) 374. Isnard, O., Fruchart, D.: Magnetism in Fe-based intermetallics: relationships between local environments and local magnetic moments. J. Alloys Compd. 205, 1–15 (1994) 375. Long, G.J., Isnard, O., Grandjean, F.: A Mössbauer spectral study of the magnetic properties of Ho2Fe17 and Ho2Fe17D3. 8. J. Appl. Phys. 91, 1423–1430 (2002) 376. Schweizer, J., Tasset, F.: Proceedings of the International Conference on Magnetism, 1973 (1974) 377. Ohashi, K.: Present and future of Sm 2 Co 17 magnets. J. Jpn. Inst. Met. (2012) 378. Chin, G.: New magnetic alloys. Science. 208, 888–894 (1980) 379. Chin, G., Wernick, J.: Magnetic Materials, Bulk. In: Encyclopedia of Chemical Technology, vol. 14, pp. 646–686. Wiley, New York (1981) 380. Strnat, K.: In: Wohlfarth, E.P., Buschow, K.H.J. (eds.) Ferromagnetic Materials, vol. 4. Elsevier Science Publishers BV (1988) 381. Narasimhan, K., Wallace, W., Hutchens, R., Greedan, J.: Magnetic anisotropy of R2Co17 compounds (R= Er, Tm, Yb). In: American Institute of Physics Conference Series, pp. 1212– 1216 (1974) 382. Deryagin, A., Kudrevatykh, N.: Magnetic anisotropy of single crystals of intermetallic R2Co17 (R= Tb, Dy, Ho, Lu) compounds. Phys. Status Solidi A. 30, K129–K133 (1975) 383. Deryagin, A., Kudrevatykh, N., Baskhov, Y.: Proceedings of the International Conference on Magnetism, 1973 (1974) 384. Tereshina, E., Andreev, A.: Crystal structure and magnetic properties of Lu 2 Co 17− x Six single crystals. Intermetallics. 18, 641–648 (2010) 385. Moze, O., Caciuffo, R., Gillon, B., Kayzel, F.: Magnetization density in Er 2 Co 17. J. Magn. Magn. Mater. 104, 1394–1396 (1992) 386. Le Caer, G., Malaman, B., Isnard, O., Soubeyroux, J., Fruchart, D., Jacobs, T., et al.: Magnetic characterisation of the ternary carbide ThFe11C x (1.5≤ x≤ 2) by57Fe Mössbauer spectroscopy. Hyperfine Interact. 77, 221–234 (1993) 387. Brouha, M., Buschow, K.: Pressure dependence of the Curie temperature of intermetallic compounds of iron and rare-earth elements, Th, and Zr. J. Appl. Phys. 44, 1813–1816 (1973) 388. Hautot, D., Long, G.J., Grandjean, F., Isnard, O.: Mössbauer spectral study of the magnetic properties of Ce 2 Fe 17 H x (x= 0, 1, 2, 3, 4, and 5). Phys. Rev. B. 62, 11731 (2000) 389. Dan’kov, S.Y., Ivtchenko, V., Tishin, A., Gschneidner Jr., K., Pecharsky, V.: Magnetocaloric effect in GdAl2 and Nd2Fe17. In: Advances in Cryogenic Engineering Materials, pp. 397– 404. Springer (2000) 390. Chen, H., Zhang, Y., Han, J., Du, H., Wang, C., Yang, Y.: Magnetocaloric effect in R 2 Fe 17 (R= Sm, Gd, Tb, Dy, Er). J. Magn. Magn. Mater. 320, 1382–1384 (2008) 391. Isnard, O., Miraglia, S., Fruchart, D., Guillot, M.: High field magnetization study of the Gd 2 Fe 17 H x system. IEEE Trans. Magn. 30, 4969–4971 (1994) 392. Kuz’min, M., Skourski, Y., Skokov, K., Müller, K.-H., Gutfleisch, O.: Determining anisotropy constants from a first-order magnetization process in Tb 2 Fe 17. Phys. Rev. B. 77, 132411 (2008) 393. Zhao, T., Lee, T., Pang, K., Lee, J.: High-field magnetization processes in Tb 2 Fe 17 and Er 2 Fe 17. J. Magn. Magn. Mater. 140, 1009–1010 (1995)

    15 Metallic Magnetic Materials

    795

    394. Nikitin, S., Tereshina, I., Tereshina, E., Suski, W., Drulis, H.: The effect of hydrogen on the magnetocrystalline anisotropy of R 2 Fe 17 and R (Fe, Ti) 12 (R= Dy, Lu) compounds. J. Alloys Compd. 451, 477–480 (2008) 395. Franse, J., Radwanski, R., Buschow, K.: Handbook of Magnetic Materials, vol. 7, p. 307. North-Holland, Amsterdam (1993) 396. Gubbens, P.C.M.: Ph.D. thesis, Delft University Press (1977) 397. Andreev, A., Deryagin, A., Zadvorkin, S., Kudrevatykh, N., Moskalev, V., Levitin, R., et al.: Physics of Magnetic Materials, p. 21. Kalinin University, Kalinin, USSR (1985) 398. Buschow, K.: Note on the structure and occurrence of ytterbium transition metal compounds. J. Less-Common Met. 26, 329–333 (1972) 399. Tishin, A.M., Spichkin, Y.I.: The Magnetocaloric Effect and Its Applications. CRC Press (2003) 400. Bouchet, G., Laforest, J., Lemaire, R., Schweizer, J.: Comptes Rendus Hebdomadaires des Seances de l’Academie des Sciences, Serie B: Sci. Phys. (1966) 401. Khan, Y.: On the crystal structures of the R2Co17 intermetallic compounds. Acta Crystallogr. Sect. B. 29, 2502–2507 (1973) 402. Hirosawa, S., Wallace, W.: Effect of Substitution of Zr and Pr on magnetic properties of R 2 Co 17 (R= Er, Yb). J. Magn. Magn. Mater. 30, 238–242 (1982) 403. Merches, M., Wallace, W., Craig, R.: Magnetic and structural characteristics of some 2: 17 rare earth-cobalt systems. J. Magn. Magn. Mater. 24, 97–105 (1981) 404. Herbst, J., Croat, J., Yelon, W.: Structural and magnetic properties of Nd2Fe14B. J. Appl. Phys. 57, 4086–4090 (1985) 405. Shoemaker, C.B., Shoemaker, D., Fruchart, R.: The structure of a new magnetic phase related to the sigma phase: iron neodymium boride Nd2Fe14B. Acta Crystallogr. Sect. C: Cryst. Struct. Commun. 40, 1665–1668 (1984) 406. Givord, D., Li, H., Moreau, J.: Magnetic properties and crystal structure of Nd 2 Fe 14 B. Solid State Commun. 50, 497–499 (1984) 407. Coey, J.M.D.: In: Coey, J.M.D. (ed.) Chap. 1 in Rare-Earth Iron Permanent Magnets. Oxford Science Publishing, Oxford (1996) 408. Herbst, J.: R 2 Fe 14 B materials: intrinsic properties and technological aspects. Rev. Mod. Phys. 63, 819 (1991) 409. Burzo, E.: Permanent magnets based on R-Fe-B and R-Fe-C alloys. Rep. Prog. Phys. 61, 1099 (1998) 410. Franse, J.J.M., Radwanski, R.J.: In: Coey, J.M.D. (ed.) Chap. 2 in Rare-Earth Iron Permanent Magnets. Oxford Science Publishing, Oxford (1996) 411. Franse, J.J.M., Radwanski, R.J.: In: Buschow, K.H.J. (ed.) Chap. 5 in Handbook of Magnetic Materials, vol. 7, p. 307. Elsevier B.V (1993) ISBN: 978-0-444-89853-1. 412. Givord, D., Li, H., de La Bathie, R.P.: Magnetic properties of Y 2 Fe 14 B and Nd 2 Fe 14 B single crystals. Solid State Commun. 51, 857–860 (1984) 413. Fruchart, R., l’Héritier, P., De Reotier, P.D., Fruchart, D., Wolfers, P., Coey, J., et al.: Mossbauer spectroscopy of R2Fe14B. J. Phys. F: Met. Phys. 17, 483 (1987) 414. Isnard, O., Yelon, W., Miraglia, S., Fruchart, D.: Neutron-diffraction study of the insertion scheme of hydrogen in Nd2Fe14B. J. Appl. Phys. 78, 1892–1898 (1995) 415. Givord, D., Li, H., Tasset, F.: Polarized neutron study of the compounds Y2Fe14B and Nd2Fe14B. J. Appl. Phys. 57, 4100–4102 (1985) 416. Wolfers, P., Obbade, S., Fruchart, D., Verhoef, R.: Precise crystal and magnetic structure determinations. Part I: a neutron diffraction study of Nd 2 Fe 14 B at 20 K. J. Alloys Compd. 242, 74–79 (1996) 417. Wolfers, P., Bacmann, M., Fruchart, D.: Single crystal neutron diffraction investigations of the crystal and magnetic structures of R 2 Fe 14 B (R= Y, Nd, Ho, Er). J. Alloys Compd. 317, 39–43 (2001) 418. Obbade, S., Wolfers, P., Fruchart, D., Argoud, R., Muller, J., Palacios, E.: A precise crystal structure determination. Part II: an X-ray four-circle study of Nd 2 Fe 14 B at 20 and 290 K. J. Alloys Compd. 242, 80–84 (1996)

    796

    J. P. Liu et al.

    419. Herbst, J., Fuerst, C., Yelon, W.: Neutron powder diffraction study of Tb2Fe14B. J. Appl. Phys. 73, 5884–5886 (1993) 420. Drebov, N., Martinez-Limia, A., Kunz, L., Gola, A., Shigematsu, T., Eckl, T., et al.: Ab initio screening methodology applied to the search for new permanent magnetic materials. New J. Phys. 15, 125023 (2013) 421. Liebs, M., Hummler, K., Fähnle, M.: Ab-initio calculation of the effective exchange couplings in rare-earth—transition-metal intermetallics. J. Magn. Magn. Mater. 124, 239–242 (1993) 422. Andreev, A., Bartashevich, M.: Magnetic properties of Th 2 Fe 14 B and its hydride. J. LessCommon Met. 167, 107–111 (1990) 423. Sinnema, S., Franse, J., Radwanski, R., Buschow, K., de Mooij, D.: Magnetic measurements on R2Fe14B and R2Co14B compounds in high fields. Le Journal de Physique Colloques. 46, C6-301–C6-304 (1985) 424. Verhoef, R., Franse, J.J.M., Menovsky, A.A., Radwanski, R.J., Ji, S., Yang, F., et al.: High-field magnetization measurements on R2FE14B single-crystals. J. Phys. 49, 565–566 (1988) 425. Li, H.S., Gavigan, J.P., Cadogan, J.M., Givord, D., Coey, J.M.D.: A study of exchange and crystalline electric-field interactions in ND2CO14B – comparison with ND2FE14B. J. Magn. Magn. Mater. 72, L241–L246 (1988) 426. Sagawa, M., Fujimura, S., Yamamoto, H., Matsuura, Y., Hirosawa, S.: Magnetic properties of rare-earth-iron-boron permanent magnet materials. J. Appl. Phys. 57, 4094–4096 (1985) 427. Hirosawa, S., Tokuhara, K., Yamamoto, H., Fujimura, S., Sagawa, M., Yamauchi, H.: Magnetization and magnetic anisotropy of R2Co14B and Nd2 (Fe1− x Cox) 14B measured on single crystals. J. Appl. Phys. 61, 3571–3573 (1987) 428. Itoh, T., Hikosaka, K., Takahashi, H., Ukai, T., Mori, N.: Anisotropy energies for Y2Fe14B and Nd2Fe14B. J. Appl. Phys. 61, 3430–3432 (1987) 429. Burlet, P., Coey, J.M.D., Gavigan, J.P., Givord, D., Meyer, C.: A note on exchange and crystalfield interactions in R2FE14B compounds – YB2FE14B. Solid State Commun. 60, 723–727 (1986) 430. Meyer, C., Gavigan, J.P., Czjzek, G., Bornemann, H.J.: A study of crystal fields in YB2FE14B bY YB-174 Mossbauer-spectroscopy. Solid State Commun. 69, 83–86 (1989) 431. Luong, N.H., Thuy, N.P., Tai, L.T., Hien, T.D.: Rare earth magnetocrystalline anisotropy in R2Fe14B (R= Sm, Er, Tm, Yb). Phys. Status Solidi A. 111, 591–595 (1989) 432. Kou, X., Grössinger, R., Müller, H., Buschow, K.: Anomalous 3d anisotropy of R 2 Fe 14 C and R 2 Fe 14 B compounds. J. Magn. Magn. Mater. 101, 349–351 (1991) 433. Dennis, K.W., Laabs, F.C., Cook, B.A., Harringa, J.L., Russell, A.M., McCallum, R.W.: Observations of multi-phase microstructures in R-2(Fe1-xCox)(14)B where R = Nd or Dy. J. Magn. Magn. Mater. 231, L33–L37 (2001) 434. Leroux, D., Vincent, H., Lheritier, P., Fruchart, R.: Crystallographic and magnetic studies of ND2CO14B and Y2CO14B. J. Phys. 46, 243–247 (1985) 435. Roux, D.L.: D. Le Roux, Ph.D. thesis, University Grenoble France (1986) 436. Buschow, K.H.J., Demooij, D.B., Sinnema, S., Radwanski, R.J., Franse, J.J.M.: Magnetic and crystallographic properties of ternary rare-earth compounds of the type R2CO14B. J. Magn. Magn. Mater. 51, 211–217 (1985) 437. Jacobs, T.H., Denissen, C.J.M., Buschow, K.H.J.: Note on the magnetic-properties of CE2FE14C. J. Less-Common Met.153, L5–L8 (1989) 438. Denissen, C.J.M., Demooij, B.D., Buschow, K.H.J.: Spin reorientation in ND2FE14C. J. Less-Common Met. 142, 195–202 (1988) 439. Hellwig, C., Girgis, K., Schefer, J., Buschow, K.H.J., Fischer, P.: Crystal and magneticstructure of the permanent-magnet materiaL TB2FE14C. J. Less-Common Met. 169, 147–156 (1991) 440. Deboer, F.R., Verhoef, R., Zhang, Z.D., Demooij, D.B., Buschow, K.H.J.: Magneticproperties of ND2FE14C and some related pseudoternary compounds. J. Magn. Magn. Mater. 73, 263–266 (1988)

    15 Metallic Magnetic Materials

    797

    441. Deboer, F.R., Huang, Y.K., Zhang, Z.D., Demooij, D.B., Buschow, K.H.J.: Magnetic and crystallographic properties of ternary rare-earth compounds of the type R2FE14C. J. Magn. Magn. Mater. 72, 167–173 (1988) 442. Obbade, S., Isnard, O., Miraglia, S., Fruchart, D., L’heritier, P., Lazaro, F., et al.: Hydrogenation, crystal structure and magnetic ordering of R 2 Fe 14 C (R≡ Sm, Er, Tm). J. Less-Common Met. 168, 321–328 (1991) 443. Hellwig, C., Girgis, K., Fischer, P., Buschow, K.H.J., Schefer, J.: Crystal and magneticstructure of the permanent magnetic material TM2FE14C. J. Alloys Compd. 184, 175–185 (1992) 444. Hellwig, C., Girgis, K., Schefer, J., Buschow, K.H.J., Fischer, P.: Crystal and magneticstructure of the permanent-magnet material LU2FE14C. J. Less-Common Met. 163, 361–368 (1990) 445. Fuerst, C.D., Herbst, J.F.: Formation of R2FE14C compounds (R = Y, CE) by rapid solidification. J. Appl. Phys. 69, 7727–7730 (1991) 446. Coey, J.M.D., Li, H.S., Gavigan, J.P., Cadogan, J.M., Hu, B.P.: In: Mitchell, I.V., Coey, J.M.D., Harris, I.R., Hanitsch, R. (eds.) Concerted European Action on Magnets, p. 76. Elsevier Applied Science Publishing, London (1989) 447. Wolfers, P., Bacmann, M., Fruchart, D.: Single crystal neutron diffraction investigations of the crystal and magnetic structures of R2Fe14B (R=Y, Nd, Ho, Er). J. Alloys Compd. 317–318, 39–43 (2001) 448. Kido, G., Kato, H., Yamada, M., Nakagawa, Y., Hirosawa, S., Sagawa, M.: Magnetization anomaly of Sm 2 Fe 14 B single crystal in high magnetic fields. J. Phys. Soc. Jpn. 56, 4635– 4636 (1987) 449. Buschow, K.: New developments in hard magnetic materials. Rep. Prog. Phys. 54, 1123 (1991) 450. Bogé, M., Coey, J., Czjzek, G., Givord, D., Jeandey, C., Li, H., et al.: 3d-4f magnetic interactions and crystalline electric field in the R 2 Fe 14 B compounds: magnetization measurements and Mössbauer study of Gd 2 Fe 14 B. Solid State Commun. 55, 295–298 (1985) 451. Loewenhaupt, M., Sosnowska, I.: Exchange and crystal fields in R2Fe14B studied by inelastic neutron scattering. J. Appl. Phys. 70, 5967–5971 (1991) 452. Ito, M., Yano, M., Dempsey, N.M., Givord, D.: Calculations of the magnetic properties of R 2 M 14 B intermetallic compounds (R= rare earth, M= Fe, Co). J. Magn. Magn. Mater. 400, 379–383 (2016) 453. Allemand, J., Letant, A., Moreau, J., Nozieres, J., De La Bathie, R.P.: A new phase in Nd 2 Fe 14 B magnets. Crystal structure and magnetic properties of Nd 6 Fe 13 Si. J. Less-Comm. Met. 166, 73–79 (1990) 454. Weitzer, F., Leithe-Jasper, A., Rogl, P., Hiebl, K., Rainbacher, A.: G. experience frustration with regard to the coupling with the Wiesinger, J. Friedl, FE Wagner. J. Appl. Phys. 75, 7745 (1995) 455. Schrey, P., Velicescu, M.: Influence of Sn additions on the magnetic and microstructural properties of Nd-Dy-Fe-B magnets. J. Magn. Magn. Mater. 101, 417–418 (1991) 456. Kramer, M., O’connor, A., Dennis, K., McCallum, R., Lewis, L., Tung, L., et al.: Origins of coercivity in the amorphous alloy Nd-Fe-Al. IEEE Trans. Magn. 37, 2497–2499 (2001) 457. Weitzer, F., Leithe-Jasper, A., Rogl, P., Hiebl, K., Noël, H., Wiesinger, G., et al.: Magnetism of (Fe, Co)-based alloys with the La 6 Co 11 Ga 3-type. J. Solid State Chem. 104, 368–376 (1993) 458. Schobinger-Papamantellos, P., Buschow, K., Ritter, C.: Magnetic ordering of the Nd 6 Fe 13− x Ga 1+ x (x= 0, 1) and Pr 6 Fe 13− x Ga 1+ x (x= 0, 1) compounds: a neutron diffraction study. J. Alloys Compd. 359, 10–21 (2003) 459. Nagata, Y., Kamonji, M., Kurihara, M., Yashiro, S., Samata, H., Abe, S.: Magnetism and transport properties of Nd 6 Fe 13− x Al 1+ x crystals. J. Alloys Compd. 296, 209–218 (2000)

    798

    J. P. Liu et al.

    460. Kennedy, S., Wu, E., Wang, F., Zhang, P., Yan, Q.: Neutron diffraction study of the magnetic structure of Pr 6 Fe 11 Al 3. Phys. B. 276, 622–623 (2000) 461. De Groot, C., Buschow, K., De Boer, F.: Magnetic properties of R 6 Fe 13− x M 1+ x compounds and their hydrides. Phys. Rev. B. 57, 11472 (1998) 462. Schobinger-Papamantellos, P., Buschow, K., De Groot, C., De Boer, F., Ritter, C., Fauth, F., et al.: On the magnetic ordering of R 6 Fe 13 X compounds. J. Alloys Compd. 280, 44–55 (1998) 463. Isnard, O., Long, G.J., Hautot, D., Buschow, K., Grandjean, F.: A neutron diffraction and Mössbauer spectral study of the magnetic spin reorientation in Nd6Fe13Si. J. Phys.: Condens. Matter. 14, 12391 (2002) 464. Schobinger-Papamantellos, P., Buschow, K., De Groot, C., De Boer, F., Ritter, C.: Magnetic ordering of the R 6 Fe 13 Sn (R= Nd, Pr) compounds studied by neutron diffraction. J. Magn. Magn. Mater. 218, 31–41 (2000) 465. De Groot, C., De Boer, F., Buschow, K., Hautot, D., Long, G.J., Grandjean, F.: Magnetic and Mössbauer spectral properties of the compound Nd 6 Fe 13 Au. J. Alloys Compd. 233, 161–164 (1996) 466. Hu, B.-p., Coey, J., Klesnar, H., Rogl, P.: Crystal structure, magnetism and 57Fe Mössbauer spectra of ternary RE6Fe11Al3 and RE6Fe13Ge compounds. J. Magn. Magn. Mater. 117, 225–231 (1992) 467. Ruzitschka, R., Reissner, M., Steiner, W., Rogl, P.: Investigation of magnetic order in RE 6 Fe 13 X (RE= Nd, Pr; X= Pd, Sn, Si). J. Magn. Magn. Mater. 242, 806–808 (2002) 468. Iranmanesh, P., Tajabor, N., Pourarian, F.: Magnetostriction effect of Co substitution in the Nd 6 Fe 13 Si intermetallic compound. Intermetallics. 42, 180–183 (2013) 469. Iranmanesh, P., Tajabor, N., Roknabadi, M.R., Pourarian, F., Brück, E.: Influence of Co substitution on magnetic properties and thermal expansion of Nd 6 Fe 13 Si intermetallic compound. Intermetallics. 19, 682–687 (2011) 470. Schobinger-Papamantellos, P., Buschow, K., De Groot, C., De Boer, F., Böttger, G., Ritter, C.: Magnetic ordering of Pr6Fe13Si and Nd6Fe13Au studied by neutron diffraction. J. Phys.: Condens. Matter. 11, 4469 (1999) 471. Leithe-Jasper, A., Skomski, R., Qi, Q., Coey, J., Weitzer, F., Rogl, P.: Hydrogen in intermetallic compounds (RE= Pr, Nd; X= Ag, Au, Si, Ge, Sn, Pb). J. Phys.: Condens. Matter. 8, 3453 (1996) 472. Bodak, O., Stepien-Damm, J., Galdecka, E.: Phase equilibria in the ternary systems Pr–Fe–Bi and Sm–Fe–Bi. J. Alloys Compd. 298, 195–197 (2000) 473. Hautot, D., Long, G.J., Grandjean, F., Buschow, K.: A comparative Mössbauer spectral study of the electronic and magnetic properties of Nd 6 Fe 13 Ag and Nd 6 Fe 13 AgH 13. J. Alloys Compd. 388, 159–167 (2005) 474. Wang, F., Zhang, P., Shen, B.-g., Yan, Q., Gong, H.: Transport properties of R6Fe11Al3 compounds (R= La, Nd). J. Appl. Phys. 87, 6043–6045 (2000) 475. Jonen, S., Rechenberg, H.: Magnetoresistance effects at the metamagnetic transition in R6Fe14− xAlx (R= Nd, La). J. Appl. Phys. 85, 4448–4450 (1999) 476. Coey, J., Qi, Q., Knoch, K., Leithe-Jasper, A., Rogl, P.: Hydrogen induced metamagnetism in R 6 Fe 13 X compounds. J. Magn. Magn. Mater. 129, 87–97 (1994) 477. Yartys, V., Denys, R., Bulyk, I., Delaplane, R., Hauback, B.: Powder neutron diffraction study of Nd 6 Fe 13 GaD 12.3 with a filled Nd 6 Fe 13 Si-type structure. J. Alloys Compd. 312, 158–164 (2000) 478. Yartys, V.A., de Boer, F.R., Buschow, K.H.J., Ouladdiaf, B., Brinks, H.W., Hauback, B.C.: Crystallographic and magnetic structure of Pr6Fe13AuD13. J. Alloys Compd. 356, 142–146 (2003) 479. Schobinger-Papamantellos, P., Ritter, C., Buschow, K.: On the magnetic ordering of Nd 6 Fe 13− x Al 1+ x (x= 1–3) and La 6 Fe 11 Al 3 compounds. J. Magn. Magn. Mater. 260, 156–172 (2003) 480. Leithe-Jasper, A., Rogl, P., Wiesinger, G., Rainbacher, A., Hatzl, R., Forsthuber, M.: A Sn 119 Mössbauer study of RE 6 M 13 Sn (RE= La, Pr, Nd; M= Fe, Co) and their hydrides. J. Magn. Magn. Mater. 170, 189–200 (1997)

    15 Metallic Magnetic Materials

    799

    481. Nesbitt, E., Wernick, J., Corenzwit, E.: Magnetic moments of alloys and compounds of iron and cobalt with rare earth metal additions. J. Appl. Phys. 30, 365–367 (1959) 482. Hubbard, W.M., Adams, E., Gilfrich, J.: Magnetic moments of alloys of gadolinium with some of the transition elements. J. Appl. Phys. 31, S368–S369 (1960) 483. Hoffer, G., Strnat, K.: Magnetocrystalline anisotropy of YCo 5 and Y 2 Co 17. IEEE Trans. Magn. 2, 487–489 (1966) 484. Strnat, K.J., Hoffer, G.: Air Force Materials Lab. Technical Report AFML-TR-65-446, Dayton (1965) 485. Strnat, K., Hoffer, G., Olson, J., Ostertag, W., Becker, J.: A family of new cobalt-base permanent magnet materials. J. Appl. Phys. 38, 1001–1002 (1967) 486. Das, D.K.: Twenty million energy product samarium-cobalt magnet. IEEE Trans. Magn. 5, 214–216 (1969) 487. Benz, M., Martin, D.: Cobalt-samarium permanent magnets prepared by liquid phase sintering. Appl. Phys. Lett. 17, 176–177 (1970) 488. Ray, A., Strnat, K.: Research and development of rare earth-transition metal alloys as permanent-magnet materials. Semiannual interim technical report, 1 January–30 June 1972. Dayton University, Ohio. Research Institute (1971–1973) 489. Ray, A.E., Strnat, K.J.: In: Savitsky, E.I. (ed.) 7th Rare Earth Metals Conference, p. 75. A.A. Baikov Institute of Meals, Moscow (1972) 490. Mildrum, H., Hartings, M., Strnat, K., Tront, J.: Magnetic properties of the intermetallic phases Sm $ sub 2$(Co, Fe) $ sub 17$. In: AIP Conference Proceedings No. 10, pp. 618– 622 (1973) 491. Senno, H., Tawara, Y.: Permanent-magnet properties of Sm-Ce-Co-Fe-Cu alloys with compositions between 1-5 and 2-17. IEEE Trans. Magn. 10, 313–317 (1974) 492. Ojima, T., Tomizawa, S., Yoneyama, T., Hori, T.: Magnetic properties of a new type of rare-earth cobalt magnets Sm 2 (Co, Cu, Fe, M) 17. IEEE Trans. Magn. 13, 1317–1319 (1977) 493. Mishra, R.K., Thomas, G., Yoneyama, T., Fukuno, A., Ojima, T.: Microstructure and properties of step aged rare earth alloy magnets. J. Appl. Phys. 52, 2517–2519 (1981) 494. Massalski, T.B.: In: Massalski, T.B. (ed.) Binary Alloy Phase Diagrams, vol. 2, 2nd edn, p. 1241. ASM International, Materials Park (1990) 495. Khan, Y.: Proceedings of the 11th Rare Earth Research Conference, vol. II, pp. 652–661 (1974) 496. Buschow, K., Van der Goot, A.: Intermetallic compounds in the system samarium-cobalt. J. Less Common Met. 14, 323–328 (1968) 497. Strnat, J.: Rare earth – cobalt permanent magnets. In: Ferromagnetic Materials – A Handbook on the Properties of Magnetically Ordered Substances, p. 154, 186, 195. North-Holland Physics Publishing (1988) 498. Campbell, P.: Permanent Magnet Materials and Their Application, p. 51. Cambridge University Press, Cambridge (1996) 499. Liu, S.: 18th International Workshop on High Performance Magnets, p. 691 (2004) 500. Wang, Y., Li, Y., Rong, C., Liu, J.P.: Sm–Co hard magnetic nanoparticles prepared by surfactant-assisted ball milling. Nanotechnology. 18, 465701 (2007) 501. Cui, B., Li, W., Hadjipanayis, G.: Formation of SmCo 5 single-crystal submicron flakes and textured polycrystalline nanoflakes. Acta Mater. 59, 563–571 (2011) 502. Poudyal, N., Liu, J.P.: Advances in nanostructured permanent magnets research. J. Phys. D: Appl. Phys. 46, 043001 (2012) 503. Hu, D., Yue, M., Zuo, J., Pan, R., Zhang, D., Liu, W., et al.: Structure and magnetic properties of bulk anisotropic SmCo 5/α-Fe nanocomposite permanent magnets prepared via a bottom up approach. J. Alloys Compd. 538, 173–176 (2012) 504. Yue, M., Zuo, J., Liu, W., Lv, W., Zhang, D., Zhang, J., et al.: Magnetic anisotropy in bulk nanocrystalline SmCo5 permanent magnet prepared by hot deformation. J. Appl. Phys. 109, 07A711 (2011) 505. Shen, Y., Leontsev, S.O., Turgut, Z., Lucas, M.S., Sheets, A.O., Horwath, J.C.: Effect of soft phase on magnetic properties of bulk Sm – Co/ – Fe nanocomposite magnets. IEEE Trans. Magn. 49, 3244–3247 (2013) 506. Croat, J.J., Herbst, J.F., Lee, R.W., Pinkerton, F.E.: High-energy product Nd-Fe-B permanent magnets. Appl. Phys. Lett. 44, 148–149 (1984) 507. Croat, J.J., Herbst, J.F., Lee, R.W., Pinkerton, F.E.: Pr-Fe and Nd-Fe-based materials: A new class of high-performance permanent magnets (invited). J. Appl. Phys.55, 2078–2082 (1984) 508. Sagawa, M., Fujimura, S., Togawa, N., Yamamoto, H., Matsuura, Y.: New material for permanent magnets on a base of Nd and Fe. J. Appl. Phys. 55, 2083–2087 (1984) 509. Herbst, J.F., Croat, J.J., Pinkerton, F.E., Yelon, W.: Relationships between crystal structure and magnetic properties in Nd 2 Fe 14 B. Phys. Rev. B. 29, 4176 (1984) 510. Nd-Fe-B Phase Diagram, ASM Alloy Phase Diagrams Database, P. Villars, editor-inchief; H. Okamoto and K. Cenzual, section editors; http://www.asminternational.org, ASM International, Materials Park 511. Herbst, J.F.: ${\mathrm{R}}_{2}$${\mathrm{Fe}}_{14}$B materials: intrinsic properties and technological aspects. Rev. Mod. Phys. 63, 819–898 (1991) 512. Gutfleisch, O.: Controlling the properties of high energy density permanent magnetic materials by different processing routes. J. Phys. D: Appl. Phys. 33, R157 (2000) 513. Buschow, K.: Ferromagnetic Materials, vol. 3, 4, p. 1. North-Holland, Amsterdam (1988) 514. Sagawa, M., Hirosawa, S., Yamamoto, H., Fujimura, S., Matsuura, Y.: Nd–Fe–B permanent magnet materials. Jpn. J. Appl. Phys. 26, 785 (1987) 515. Mishra, R.K.: Microstructure of melt-spun Nd-Fe-B magnequench magnets. J. Magn. Magn. Mater. 54, 450–456 (1986) 516. Coey, J., Skomski, R.: New magnets from interstitial intermetallics. Phys. Scr. 1993, 315 (1993) 517. GschneidnerJr, K.A., Pecharsky, V., Tsokol, A.: Recent developments in magnetocaloric materials. Rep. Prog. Phys. 68, 1479 (2005) 518. Moya, X., Kar-Narayan, S., Mathur, N.: Caloric materials near ferroic phase transitions. Nat. Mater. 13, 439–450 (2014) 519. Moya, X., Defay, E., Heine, V., Mathur, N.D.: Too cool to work. Nat. Phys. 11, 202–205 (2015) 520. Nikitin, S., Myalikgulyev, G., Annaorazov, M., Tyurin, A., Myndyev, R., Akopyan, S.: Giant elastocaloric effect in FeRh alloy. Phys. Lett. A. 171, 234–236 (1992) 521. Stern-Taulats, E., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-L., Pramanick, S., et al.: Barocaloric and magnetocaloric effects in Fe 49 Rh 51. Phys. Rev. B. 89, 214105 (2014) 522. Stern-Taulats, E., Gràcia-Condal, A., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-L., et al.: Reversible adiabatic temperature changes at the magnetocaloric and barocaloric effects in Fe49Rh51. Appl. Phys. Lett. 107, 152409 (2015) 523. Pecharsky, V.K., Gschneidner Jr., K.A.: Giant magnetocaloric effect in Gd 5 (Si 2 Ge 2). Phys. Rev. Lett. 78, 4494 (1997) 524. Pecharsky, V., Gschneidner, K.: Phase relationships and crystallography in the pseudobinary system Gd 5 Si 4-Gd 5 Ge 4. J. Alloys Compd. 260, 98–106 (1997) 525. Morellon, L., Blasco, J., Algarabel, P., Ibarra, M.: Nature of the first-order antiferromagneticferromagnetic transition in the Ge-rich magnetocaloric compounds Gd 5 (Si x Ge 1− x) 4. Phys. Rev. B. 62, 1022 (2000) 526. Pecharsky, V.K., Gschneidner Jr., K.A.: Tunable magnetic regenerator alloys with a giant magnetocaloric effect for magnetic refrigeration from {bold {approximately}} 20 to {bold {approximately}} 290K. Appl. Phys. Lett. 70, 3299 (1997) 527. Yuce, S., Barrio, M., Emre, B., Stern-Taulats, E., Planes, A., Tamarit, J.-L., et al.: Barocaloric effect in the magnetocaloric prototype Gd5Si2Ge2. Appl. Phys. Lett. 101, 1906 (2012) 528. Hu, F.-X., Shen, B.-G., Sun, J.-R., Cheng, Z.-H., Rao, G.-H., Zhang, X.-X.: Influence of negative lattice expansion and metamagnetic transition on magnetic entropy change in the compound LaFe11. 4Si1. 6. Appl. Phys. Lett. 78, 3675–3677 (2001) 529. Shen, B., Sun, J., Hu, F., Zhang, H., Cheng, Z.: Recent progress in exploring magnetocaloric materials. Adv. Mater. 21, 4545–4564 (2009)

    15 Metallic Magnetic Materials

    801

    530. Hu, F.X.: Magnetic properties and magnetic entropy change of Fe-based La(Fe,M)13 compounds and Ni-Mn-Ga alloys, Ph.D. thesis, (2002) 531. Hu, F., Shen, B., Sun, J., Wang, G., Cheng, Z.: Very large magnetic entropy change near room temperature. Appl. Phys. Lett. 80, 826–828 (2002) 532. Wang, F., Chen, Y., Wang, G., Shen, B.: Magnetic entropy in La (Fe, Si) 13 compounds. J. Phys. D: Appl. Phys. 36, 1–5 (2003) 533. Chen, Y.-f., Wang, F., Shen, B.-g., Hu, F.-x., Sun, J.-r., Wang, G.-j., et al.: Magnetic properties and magnetic entropy change of LaFe11. 5Si1. 5Hy interstitial compounds. J. Phys.: Condens. Matter. 15, L161 (2003) 534. Fujita, A., Fujieda, S., Hasegawa, Y., Fukamichi, K.: Itinerant-electron metamagnetic transition and large magnetocaloric effects in La (Fe x Si 1− x) 13 compounds and their hydrides. Phys. Rev. B. 67, 104416–104427 (2003) 535. Chen, Y.-F., Wang, F., Shen, B.-G., Wang, G.-J., Sun, J.-R.: Magnetism and magnetic entropy change of LaFe11. 6Si1. 4Cx (x= 0-0.6) interstitial compounds. J. Appl. Phys. 93, 1323–1325 (2003) 536. Shen, J., Gao, B., Zhang, H., Hu, F., Li, Y., Sun, J., et al.: Reduction of hysteresis loss and large magnetic entropy change in the NaZn13-type LaPrFeSiC interstitial compounds. Appl. Phys. Lett. 91, 142504–142504 (2007) 537. Manosa, L., González-Alonso, D., Planes, A., Barrio, M., Tamarit, J.-L., Titov, I.S., et al.: Inverse barocaloric effect in the giant magnetocaloric La–Fe–Si–Co compound. Nat. Commun. 2, 595 (2011) 538. Liu, D., Huang, Q., Yue, M., Lynn, J., Liu, L., Chen, Y., et al.: Temperature, magnetic field, and pressure dependence of the crystal and magnetic structures of the magnetocaloric compound Mn 1.1 Fe 0.9 (P 0.8 Ge 0.2). Phys. Rev. B. 80, 174415 (2009) 539. Tegus, O., Brück, E., Buschow, K., De Boer, F.: Transition-metal-based magnetic refrigerants for room-temperature applications. Nature. 415, 150–152 (2002) 540. Brück, E., Kamarad, J., Sechovsky, V., Arnold, Z., Tegus, O., de Boer, F.: Pressure effects on the magnetocaloric properties of MnFeP 1− xAsx. J. Magn. Magn. Mater. 310, e1008–e1009 (2007) 541. Thanh, D.C., Brück, E., Tegus, O., Klaasse, J., Gortenmulder, T., Buschow, K.: Magnetocaloric effect in MnFe (P, Si, Ge) compounds. J. Appl. Phys. 99, 08Q107 (2006) 542. Hu, F.-X., Shen, B.-g., Sun, J.-R.: Magnetic entropy change in Ni51. 5Mn22. 7Ga25. 8 alloy. Appl. Phys. Lett. 76, 3460 (2000) 543. Krenke, T., Duman, E., Acet, M., Wassermann, E.F., Moya, X., Mañosa, L., et al.: Inverse magnetocaloric effect in ferromagnetic Ni–Mn–Sn alloys. Nat. Mater. 4, 450–454 (2005) 544. Liu, J., Gottschall, T., Skokov, K.P., Moore, J.D., Gutfleisch, O.: Giant magnetocaloric effect driven by structural transitions. Nat. Mater. 11, 620–626 (2012) 545. Mañosa, L., González-Alonso, D., Planes, A., Bonnot, E., Barrio, M., Tamarit, J.-L., et al.: Giant solid-state barocaloric effect in the Ni-Mn-In magnetic shape-memory alloy. Nat. Mater. 9, 478–481 (2010) 546. Lu, B., Xiao, F., Yan, A., Liu, J.: Elastocaloric effect in a textured polycrystalline Ni-Mn-InCo metamagnetic shape memory alloy. Appl. Phys. Lett. 105, 161905 (2014) 547. Zhao, Y.-Y., Hu, F.-X., Bao, L.-F., Wang, J., Wu, H., Huang, Q.-Z., et al.: Giant negative thermal expansion in bonded MnCoGe-based compounds with Ni2In-type hexagonal structure. J. Am. Chem. Soc. 137, 1746–1749 (2015) 548. Wang, J.-T., Wang, D.-S., Chen, C., Nashima, O., Kanomata, T., Mizuseki, H., et al.: Vacancy induced structural and magnetic transition in MnCo1-xGe. Appl. Phys. Lett. 89, 2504 (2006) 549. Trung, N., Biharie, V., Zhang, L., Caron, L., Buschow, K., Brück, E.: From single-to doublefirst-order magnetic phase transition in magnetocaloric Mn1− xCrxCoGe compounds. Appl. Phys. Lett. 96, 162507 (2010) 550. Gercsi, Z., Hono, K., Sandeman, K.G.: Designed metamagnetism in CoMnGe 1− x P x. Phys. Rev. B. 83, 174403 (2011) 551. Wu, R.-R., Bao, L.-F., Hu, F.-X., Wu, H., Huang, Q.-Z., Wang, J., et al.: Sci. Rep. 5, 18027 (2015)

    802

    J. P. Liu et al.

    552. Kuhrt, C., Schittny, T., Barner, K.: Magnetic b-t phase-diagram of anion substituted mnas – magnetocaloric experiments. Phys. Status Solidi A. 91, 105–113 (1985) 553. Wada, H., Tanabe, Y.: Giant magnetocaloric effect of MnAs1-xSbx. Appl. Phys. Lett. 79, 3302–3304 (2001) 554. Wada, H., Matsuo, S., Mitsuda, A.: Pressure dependence of magnetic entropy change and magnetic transition in MnAs1-xSbx. Phys. Rev. B. 79 (2009) 555. Mosca, D.H., Vidal, F., Etgens, V.H.: Strain engineering of the magnetocaloric effect in MnAs epilayers. Phys. Rev. Lett. 101 (2008) 556. Stewart, G.: Heavy-fermion systems. Rev. Mod. Phys. 56, 755 (1984) 557. Flouquet, J.: On the heavy fermion road. Prog. Low Temp. Phys. 15, 139 (2005) 558. Coleman, P., Maple, B., Millis, A.: Special issue containing papers presented at the Institute for Theoretical Physics Conference on Non-Fermi Liquid Behaviour in Metals – Santa Barbara, USA, 17–21 June 1996 – Preface. J. Phys. Conden. Matter. 8, U3–U6 (1996) 559. Stewart, G.: Non-Fermi-liquid behavior in d-and f-electron metals. Rev. Mod. Phys. 73, 797 (2001) 560. Grewe, N., Steglich, F.: In: Gschneidner, K.A., Eyring, L. (eds.) Handbook on thePhysics and Chemistry of Rare Earths, vol. 14, pp. 343–474. Elsevier Science Publishers, Amsterdam (1991) 561. Ott, H.R., Fisk, Z.: In: Freeman, A.J., Lander, G.H. (eds.) Handbook on thePhysics and Chemistry ofActinides, vol. 5, pp. 85–225. Elsevier Science Publishers, Amsterdam (1987) 562. Doniach, S.: The Kondo lattice and weak antiferromagnetism. Phys. B+ C. 91, 231–234 (1977) 563. von Loehneysen, H., Rosch, A., Vojta, M., Woelfle, P.: Fermi-liquid instabilities at magnetic quantum phase transitions. Rev. Mod. Phys. 79, 1015–1075 (2007) 564. Coleman, P., Pepin, C., Si, Q.M., Ramazashvili, R.: How do Fermi liquids get heavy and die? J Phys. Conden. Matter. 13, R723–R738 (2001) 565. Si, Q.: Quantum criticality and global phase diagram of magnetic heavy fermions. Phys. Status Solidi B. 247, 476–484 (2010) 566. Sigrist, M., Ueda, K.: Phenomenological theory of unconventional superconductivity. Rev. Mod. Phys. 63, 239–311 (1991) 567. Pfleiderer, C.: Superconducting phases of f-electron compounds. Rev. Mod. Phys. 81, 1551– 1624 (2009) 568. White, B.D., Thompson, J.D., Maple, M.B.: Unconventional superconductivity in heavyfermion compounds. Phys. C. 514, 246–278 (2015) 569. Andres, K., Graebner, J.E., Ott, H.R.: 4F-virtual-bound-state formation in ceal-3 at lowtemperatures. Phys. Rev. Lett. 35, 1779–1782 (1975) 570. Mattens, W.C.M., Elenbaas, R.A., Boer, F.R.D.: Mixed-valence behavior in intermetallic compound ybcual. Commun. Phys. 2, 147–150 (1977) 571. Stewart, G.R., Fisk, Z., Wire, M.S.: New Ce heavy-fermion system – CeCu6. Phys. Rev. B. 30, 482–484 (1984) 572. Fujita, T., Satoh, K., Onuki, Y., Komatsubara, T.: Specific-heat of the dense kondo compound CeCu6. J. Magn. Magn. Mater. 47–8, 66–68 (1985) 573. Besnus, M.J., Kappler, J.P., Lehmann, P., Meyer, A.: Low-temperature heat-capacity, magnetization, resistivity of CeRu2SI2, with Y or La substitution. Solid State Commun.55, 779–782 (1985) 574. Haen, P., Flouquet, J., Lapierre, F., Lejay, P., Remenyi, G.: Metamagnetic-like transition in CeRu2SI2. J. Low Temp. Phys. 67, 391–419 (1987) 575. Lee, W., Shelton, R.: CePtSi: a new heavy-fermion compound. Phys. Rev. B. 35, 5369 (1987) 576. Havela, L., Sechovský, V., Svoboda, P., Diviš, M., Nakotte, H., Prokeš, K., et al.: Heavy fermion behavior of U2T2X compounds. J. Appl. Phys. 76, 6214–6216 (1994) 577. Yatskar, A., Beyermann, W., Movshovich, R., Canfield, P.: Possible correlated-electron behavior from quadrupolar fluctuations in PrInA g 2. Phys. Rev. Lett. 77, 3637 (1996) 578. Torikachvili, M., Jia, S., Mun, E., Hannahs, S., Black, R., Neils, W., et al.: Six closely related YbT2Zn20 (T= Fe, Co, Ru, Rh, Os, Ir) heavy fermion compounds with large local moment degeneracy. Proc. Natl. Acad. Sci. U.S.A. 104, 9960–9963 (2007)

    15 Metallic Magnetic Materials

    803

    579. Ott, H., Rudigier, H., Delsing, P., Fisk, Z.: Magnetic ground state of a heavy-electron system: U 2 Zn 17. Phys. Rev. Lett. 52, 1551 (1984) 580. Fisk, Z., Stewart, G., Willis, J., Ott, H., Hulliger, F.: Low-temperature properties of the heavyfermion system U Cd 11. Phys. Rev. B. 30, 6360 (1984) 581. Stewart, G., Fisk, Z., Smith, J., Willis, J., Wire, M.: New heavy-fermion system, NpBe 13, with a comparison to UBe 13 and PuBe 13. Phys. Rev. B. 30, 1249 (1984) 582. Ott, H., Rudigier, H., Felder, E., Fisk, Z., Batlogg, B.: Low-temperature state of U Cu 5: formation of heavy electrons in a magnetically ordered material. Phys. Rev. Lett. 55, 1595 (1985) 583. Lin, C., Teter, J., Crow, J., Mihalisin, T., Brooks, J., Abou-Aly, A., et al.: Observation of Magnetic-Field-Induced Superconductivity in a Heavy-Fermion Antiferromagnet: Ce Pb 3. Phys. Rev. Lett. 54, 2541 (1985) 584. Fisk, Z., Canfield, P., Beyermann, W., Thompson, J., Hundley, M., Ott, H., et al.: Massive electron state in YbBiPt. Phys. Rev. Lett. 67, 3310 (1991) 585. Kitazawa, H., Schank, C., Thies, S., Seidel, B., Geibel, C., Steglich, F.: A new antiferromagnetic heavy fermion compound: CePd 2 Al 3. J. Phys. Soc. Jpn. 61, 1461–1464 (1992) 586. Morosan, E., Bud’ko, S.L., Canfield, P.C., Torikachvili, M.S., Lacerda, A.H.: Thermodynamic and transport properties of RAgGe (R = Tb-Lu) single crystals. J. Magn. Magn. Mater. 277, 298–321 (2004) 587. Vonlohneysen, H., Pietrus, T., Portisch, G., Schlager, H.G., Schroder, A., Sieck, M., et al.: Non-fermi-liquid behavior in a heavy-fermion alloy at a magnetic instabilitY. Phys. Rev. Lett. 72, 3262–3265 (1994) 588. Raymond, S., Regnault, L.P., Kambe, S., Mignot, J.M., Lejay, P., Flouquet, J.: Magnetic correlations in Ce0.925La0.075Ru2Si2. J. Low Temp. Phys. 109, 205–224 (1997) 589. Trovarelli, O., Geibel, C., Mederle, S., Langhammer, C., Grosche, F.M., Gegenwart, P., et al.: YbRh2Si2: pronounced non-Fermi-liquid effects above a low-lying magnetic phase transition. Phys. Rev. Lett. 85, 626–629 (2000) 590. Custers, J., Gegenwart, P., Wilhelm, H., Neumaier, K., Tokiwa, Y., Trovarelli, O., et al.: The break-up of heavy electrons at a quantum critical point. Nature. 424, 524–527 (2003) 591. Paschen, S., Mueller, M., Custers, J., Kriegisch, M., Prokofiev, A., Hilscher, G., et al.: Quantum critical behaviour in Ce3Pd20Si6? J. Magn. Magn. Mater. 316, 90–92 (2007) 592. Steglich, F., Aarts, J., Bredl, C.D., Lieke, W., Meschede, D., Franz, W., et al.: Superconductivity in the presence of strong pauli paramagnetism – CeCu2SI2. Phys. Rev. Lett. 43, 1892–1896 (1979) 593. Stockert, O., Faulhaber, E., Zwicknagl, G., Stusser, N., Jeevan, H.S., Deppe, M., et al.: Nature of the A phase in CeCu2Si2. Phys. Rev. Lett. 92 (2004) 594. Ott, H.R., Rudigier, H., Fisk, Z., Smith, J.L.: UBe13 – an unconventional actinide superconductor. Phys. Rev. Lett. 50, 1595–1598 (1983) 595. Ott, H.R., Rudigier, H., Felder, E., Fisk, Z., Smith, J.L.: Influence of impurities and magneticfields on the normal and superconducting states of UBe13. Phys. Rev. B. 33, 126–131 (1986) 596. Stewart, G.R., Fisk, Z., Willis, J.O., Smith, J.L.: Possibility of coexistence of bulk superconductivity and spin fluctuations in UPt3. Phys. Rev. Lett. 52, 679–682 (1984) 597. Aeppli, G., Bucher, E., Broholm, C., Kjems, J.K., Baumann, J., Hufnagl, J.: Magnetic order and fluctuations in superconducting UPt3. Phys. Rev. Lett. 60, 615–618 (1988) 598. Fisher, R.A., Kim, S., Woodfield, B.F., Phillips, N.E., Taillefer, L., Hasselbach, K., et al.: Specific-heat of UPt3 – evidence for unconventional superconductivitY. Phys. Rev. Lett. 62, 1411–1414 (1989) 599. Palstra, T., Menovsky, A., Van den Berg, J., Dirkmaat, A., Kes, P., Nieuwenhuys, G., et al.: Superconducting and magnetic transitions in the heavy-fermion system U Ru 2 Si 2. Phys. Rev. Lett. 55, 2727 (1985) 600. Petrovic, C., Pagliuso, P.G., Hundley, M.F., Movshovich, R., Sarrao, J.L., Thompson, J.D., et al.: Heavy-fermion superconductivity in CeCoIn5 at 2.3 K. J. Phys.: Condens. Matter. 13, L337–L342 (2001)

    804

    J. P. Liu et al.

    601. Petrovic, C., Movshovich, R., Jaime, M., Pagliuso, P.G., Hundley, M.F., Sarrao, J.L., et al.: A new heavy-fermion superconductor CeIrIn5: A relative of the cuprates? Europhys. Lett. 53, 354–359 (2001) 602. Bauer, E.D., Frederick, N.A., Ho, P.C., Zapf, V.S., Maple, M.B.: Superconductivity and heavy fermion behavior in PrOs4Sb12. Phys. Rev. B. 65 (2002) 603. Vollmer, R., Faisst, A., Pfleiderer, C., von Lohneysen, H., Bauer, E.D., Ho, P.C., et al.: Lowtemperature specific heat of the heavy-fermion superconductor PrOs4Sb12. Phys. Rev. Lett. 90 (2003) 604. Chen, G.F., Ohara, S., Hedo, M., Uwatoko, Y., Saito, K., Sorai, M., et al.: Observation of superconductivity in heavy-fermion compounds of Ce2CoIn8. J. Phys. Soc. Jpn. 71, 2836– 2838 (2002) 605. Nakatsuji, S., Kuga, K., Machida, Y., Tayama, T., Sakakibara, T., Karaki, Y., et al.: Superconductivity and quantum criticality in the heavy-fermion system beta-YbAlB(4). Nat. Phys. 4, 603–607 (2008) 606. Kaczorowski, D., Pikul, A.P., Gnida, D., Tran, V.H.: Emergence of a superconducting state from an antiferromagnetic phase in single crystals of the heavy fermion compound Ce2PdIn8. Phys. Rev. Lett. 103 (2009) 607. Geibel, C., Thies, S., Kaczorowski, D., Mehner, A., Grauel, A., Seidel, B., et al.: A new heavy-fermion superconduCTOR – UNI2AL3. Z. Phys. B: Condens. Matter. 83, 305–306 (1991) 608. Geibel, C., Schank, C., Thies, S., Kitazawa, H., Bredl, C.D., Bohm, A., et al.: Heavy-fermion superconductivity at Tc=2K in the antiferromagnet UPd2AL3. Z. Phys. B: Condens. Matter. 84, 1–2 (1991) 609. Bauer, E., Hilscher, G., Michor, H., Paul, C., Scheidt, E.W., Gribanov, A., et al.: Heavy fermion superconductivity and magnetic order in noncentrosymmetric CePt3Si. Phys. Rev. Lett. 92 (2004) 610. Prokleska, J., Kratochvilova, M., Uhlirova, K., Sechovsky, V., Custers, J.: Magnetism, superconductivity, and quantum criticality in the multisite cerium heavy-fermion compound Ce3PtIn11. Phys. Rev. B. 92 (2015) 611. Jaccard, D., Behnia, K., Sierro, J.: Pressure-induced heavy fermion superconductivity of CeCu2Ge2. Phys. Lett. A. 163, 475–480 (1992) 612. Fisher, R.A., Emerson, J.P., Caspary, R., Phillips, N.E., Steglich, F.: The low-temperature specific-heat of CeCu2Ge2 at 0-kbar and 9.5-kbar. Phys. B. 194, 459–460 (1994) 613. Grosche, F.M., Julian, S.R., Mathur, N.D., Lonzarich, G.G.: Magnetic and superconducting phases of CePd2Si2. Phys. B. 223–24, 50–52 (1996) 614. Steeman, R.A., Frikkee, E., Helmholdt, R.B., Menovsky, A.A., Vandenberg, J., Nieuwenhuys, G.J., et al.: CePd2SI2 – a reduced-moment antiferromagnet. Solid State Commun. 66, 103– 107 (1988) 615. Movshovich, R., Graf, T., Mandrus, D., Thompson, J.D., Smith, J.L., Fisk, Z.: Superconductivity in heavy-fermion CeRh2Si2. Phys. Rev. B. 53, 8241–8244 (1996) 616. Walker, I.R., Grosche, F.M., Freye, D.M., Lonzarich, G.G.: The normal and superconducting states of CeIn3 near the border of antiferromagnetic order. Phys. C. 282, 303–306 (1997) 617. Hegger, H., Petrovic, C., Moshopoulou, E.G., Hundley, M.F., Sarrao, J.L., Fisk, Z., et al.: Pressure-induced superconductivity in quasi-2D CeRhIn5. Phys. Rev. Lett. 84, 4986–4989 (2000) 618. Kimura, N., Ito, K., Saitoh, K., Umeda, Y., Aoki, H., Terashima, T.: Pressure-induced superconductivity in noncentrosymmetric heavy-fermion CeRhSi3. Phys. Rev. Lett. 95 (2005) 619. Sugitani, I., Okuda, Y., Shishido, H., Yamada, T., Thamhavel, A., Yamamoto, E., et al.: Pressure-induced heavy-fermion superconductivity in antiferromagnet CeIrSi3 without inversion symmetry. J. Phys. Soc. Jpn. 75 (2006) 620. Saxena, S.S., Agarwal, P., Ahilan, K., Grosche, F.M., Haselwimmer, R.K.W., Steiner, M.J., et al.: Superconductivity on the border of itinerant-electron ferromagnetism in UGe2. Nature. 406, 587–592 (2000)

    15 Metallic Magnetic Materials

    805

    621. Aoki, D., Huxley, A., Ressouche, E., Braithwaite, D., Flouquet, J., Brison, J.P., et al.: Coexistence of superconductivity and ferromagnetism in URhGe. Nature. 413, 613–616 (2001) 622. Kobayashi, T.C., Fukushima, S., Hidaka, H., Kotegawa, H., Akazawa, T., Yamamoto, E., et al.: Pressure-induced superconductivity in ferromagnet UIr without inversion symmetry. Phys. B. 378–80, 355–358 (2006) 623. Huy, N.T., Gasparini, A., de Nijs, D.E., Huang, Y., Klaasse, J.C.P., Gortenmulder, T., et al.: Superconductivity on the border of weak itinerant ferromagnetism in UCoGe. Phys. Rev. Lett. 99 (2007)

    J. Ping Liu received his PhD in Physics from the University of Amsterdam, the Netherlands. He is currently a Distinguished Professor at the University of Texas at Arlington, United States and has worked in several institutions in China, Europe and the United States in research and development of permanent magnets, magnetic nanoparticles and nanocomposite materials.

    Matthew Willard received his PhD from Carnegie Mellon University in 2000. He worked at the U.S. Naval Research Laboratory in Washington, DC for 12 years prior to joining the Materials Science and Engineering faculty at Case Western Reserve University in Cleveland, OH. His expertise includes physical metallurgy, alloy design of soft magnetic alloys, and melt processing of alloys.

    Wei Tang received his PhD from the Northwestern Polytechnical University of China in 1996. He works now at Division of Materials Science and Technology, Ames Laboratory of USDOE as a Research Scientist. His research scope and interest focus on the studies of magnetic materials and development of permanent magnets with high performance.

    806

    J. P. Liu et al. Ekkes Brück received his PhD from University of Amsterdam in 1991. He worked at CMR Stanford University, University of Amsterdam and Federal University of Santa Catharina, Brazil. Currently he holds the chair of Fundamental Aspects of Materials and Energy at Delft University of Technology. Employing microscopic and macroscopic techniques, main research interests are materials for renewable energy and energy saving.

    Enke Liu received his PhD from Institute of Physics (IOP), Chinese Academy of Sciences (CAS) at Beijing in 2012. He works at State Key Lab. of Magnetism, IOP, CAS. He now works on the experimental realization and energy-band design of magnetic martensitic alloys and magnetic topological semimetals including the families of hexagonal MM’X, all-d-metal Heuslers and Shandites.

    Claudia Felser studied chemistry and physics at the University of Cologne and completed her doctorate in 1994. After postdoctoral fellowships at the Max Planck Institute (MPI) and the CNRS in Nantes, she joined the University of Mainz and became a Full Professor in 2003. She has been director and scientific member at the MPI for Chemical Physics of Solids since December 2011.

    Olivier Isnard received his PhD from Univ. Joseph Fourier and Institut Laue Langevin Grenoble France in 1993. Now Professor of Physics at Néel Institute, University Grenoble Alpes, he worked at the University of Missouri in USA, University of Amsterdam NL. Doctor Honoris Causa of both Babes Bolyai and Technical Universities in Cluj Romania, he works on magnetism of rare-earth transition metal intermetallics.

    15 Metallic Magnetic Materials

    807

    Sam Liu received his Ph.D. from the University of Dayton in 1989. He worked at the University of DaytonMagnetics Laboratory from 1989 to 2011. His research interests in the US were primarily focused on high-temperature metal-semiconductor contacts, high-temperature Sm-Co magnets, and nanocrystalline and nanocomposite rare earth permanent magnet materials.

    J. F. Herbst received his PhD in physics from Cornell University (Ithaca, NY) in 1974. During a career at the General Motors Research and Development Center he participated in a variety of new materials projects, including (1) the discovery of Nd2 Fe14 B and the development of permanent magnets based on that fascinating compound, and (2) hydrogen storage media for fuel cell vehicles.

    Fengxia Hu is a Professor at Institute of Physics, Chinese Academy of Science since 2008. She received her Ph.D in Physics at Institute of Physics, Chinese Academy of Science in 2002 with major of condensed matter physics. Her current interests include magnetocaloric effect, and magnetic refrigerant applications of various intermetallics, such as Fe-based NaZn13-type materials, NiMn-based Heusler alloys, and etc.

    808

    J. P. Liu et al. Anne de Visser is associate professor at the University of Amsterdam. His research is directed towards the understanding of collective phenomena in novel quantum materials, notably strongly correlated electron systems and topological materials. He is a pioneer in the field of heavy-fermion physics, quantum phase transitions and unconventional superconductivity, with an expertise in measuring transport, magnetic and thermal properties under extreme conditions.

    Metallic Magnetic Thin Films D. Wu

    16

    and X.-F. Jin

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . GaAs (001) Substrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structure and Magnetism of Magnetic 3d Transition Metals . . . . . . . . . . . . . . . . . . . . . . . . . . Cr/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mn/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fe/Cu (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Co/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ni/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structure and Magnetism of Magnetic 3d Transition Metal Alloys . . . . . . . . . . . . . . . . . . . . . Py/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cox Mn1−x /GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fex Cu1−x /GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fex Pd1−x /Cu (100) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    810 811 812 812 814 816 819 824 826 826 829 833 837 840

    Abstract

    Molecular beam epitaxy (MBE) is an advanced technique to grow singlecrystalline films. In this chapter we survey the epitaxial growth of 3d transition magnetic metal and their alloy films via MBE. Several metastable structure phases are obtained in the ultrathin film. For example, the body-centered cubic

    D. Wu National Laboratory of Solid State Microstructures and Department of Physics, Nanjing University, Nanjing, People’s Republic of China e-mail: [email protected] X.-F. Jin () Department of Physics and State Key Laboratory of Surface Physics, Fudan University, Shanghai, People’s Republic of China 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_19

    809

    810

    D. Wu and X.-F. Jin

    (bcc) phase of Ni, which does not exist in nature, can be epitaxially grown on GaAs (001). It is found that the interface structure and film growth temperature play a crucial role in determining the film structure and magnetism. With these metastable structure phase and precisely controlled composition of 3d transition magnetic metal and their alloy films, we explore several interesting physics such as the origin of the extremely small magnetocrystalline anisotropy in permalloy and the correlation between the structure and magnetism in Co1−x Mnx alloys.

    Introduction Significant progress has been made in the epitaxial growth of magnetic metallic films (MMFs) by molecular beam epitaxy (MBE) over the past four decades. Nowadays it is routinely growing single-crystalline MMFs on different substrates with submonolayer control. This allows us to understand the fascinating fundamental properties of the MMFs, which leads to important applications in current information technologies. A remarkable example is the discovery of the giant magnetoresistance (GMR) effect in metallic magnetic multilayers (e.g., Fe/Cr or Co/Cu superlattices) [1, 2] and the employment of the GMR in the read head of hard disks. Since then, a variety of spin transport phenomena were discovered and applied to spintronic devices. There are four ferromagnetic (FM) elementary metals, Fe, Co, Ni, and Gd. Fe, Co, and Ni and their alloys have been dominant materials in magnetism in both fundamental studies and applications, partially because their Curie temperature (Tc) is much higher than room temperature. Cr and Mn are the only two antiferromagnetic (AFM) metallic elements. According to the Slater-Pauling model [3, 4], the magnetic state is related to the occupation of d band around the Fermi level. These five magnetic elementary metals are all 3d transition metals with atomic numbers continuously changing from 24 to 28. The binary metal alloys with precisely controlled composition provide a model system to tune the Fermi level and manipulate the corresponding magnetic properties. Therefore, they and their alloys are the ideal platforms for a fundamental understanding of magnetism in the thin film state. Moreover, the magnetic properties of the films such as the magnetic anisotropy and the magnetic order are substantially different from these of bulk. A comprehensive and systematic investigation of the 3d MMFs is important. The recent technological applications of spintronics are mostly based on the 3d MMFs, and spintronics is also an important driving force for their investigation. For example, the spin-orbit torque phenomenon which can switch the magnetic moment by a dc current occurs in AlOx /Co/Pt with Co layers a few monolayers thick [5]. The 3d transition metals and alloys have a variety of crystallographic and magnetic phases in bulk, some of which only exist at high pressure or high temperature. In addition to these thermodynamically stable phases, metastable structures could be epitaxially grown on single-crystalline substrates. In this chapter we focus on the epitaxial growth and magnetic properties of the magnetic 3d transition metal and their alloy films.

    16 Metallic Magnetic Thin Films

    811

    GaAs (001) Substrates Among several film growth methods such as MBE, magnetron sputtering, and pulsed laser deposition (PLD), MBE is the most powerful technique to grow the highest-quality MMFs. GaAs (001) substrates are widely used for the epitaxial growth of 3d transition metals and their alloys. The cubic lattice constant is 0.565 nm. The top view of the unreconstructed GaAs (001) together with the corresponding Miller indices is schematically shown in Fig. 1a. If a cubic-structure film is grown on GaAs (001) surface, there are two kinds of epitaxial growth geometry: (i) the face-centered cubic (fcc) structure with (001)[110]film //(001)[100]GaAs , in which the epitaxial film rotates 45◦ around the surface normal, and (ii) the body-centered cubic (bcc) structure with (001)[100]film //(001)[100]GaAs , as schematically shown in Fig. 1b, c, respectively. The lattice constants are 0.399 nm for the fcc structure and 0.283 nm for the bcc structure without distortion. Lattice misfits are relatively small for 3d transition metals and alloys, meaning that GaAs is the ideal substrate for these metals. For example, the lattice misfits are 0.1% for bcc Co/GaAs film and − 8.6% for fcc Mn/GaAs film. Te-doped GaAs (001) single crystal wafers can be prepared by two different procedures after cleaning in acetone, water, and alcohol in an ultrasonic bath in

    Fig. 1 Schematic top view of the unreconstructed GaAs (001), together with rotated fcc structure film with (001)[110]film //(001)[100]GaAs and unrotated bcc structure film with (001)[100]film //(001)[100]GaAs [6]. LICENSE #: 4812820853366

    812

    D. Wu and X.-F. Jin

    sequence. One is a standard method for most substrates. The GaAs substrate is ◦ bombarded by 800 eV argon ions and then annealed at 500 C for several cycles in the ultrahigh vacuum (UHV). The other is a chemical etching method, using H2 SO4 :H2 O2 :H2 O = 5:1:1 to etch the substrate. Then it is loaded into the UHV ◦ system and flashed to 580 C. The observation of a clear (4 × 1) low-energy electron diffraction (LEED) or reflection high-energy electron diffraction (RHEED) pattern indicates a clean surface. Auger spectra show the surface is free of carbon and oxygen contamination. We find that the latter procedure gives a better substrate for epitaxial growth.

    Structure and Magnetism of Magnetic 3d Transition Metals Cr/GaAs (001) Chromium has a bcc structure in bulk with a lattice constant of 0.2885 nm. The lattice mismatch between GaAs (001) substrate and bcc structure Cr is less than 1.9%. The Cr film growth on GaAs (001) is expected to be bcc with the epitaxial relationship: (001)[100]Cr //(001)[100]GaAs and (001)[100]Cr //(001)[100]GaAs , as shown in Fig. 1c. The structure of Cr films grown on the GaAs (001) substrate is dependent on the growth temperature [7]. Figure 2a presents the typical RHEED pattern for a clean GaAs (001) surface, with an electron beam along the GaAs [110] direction. In the initial stage of the film growth at room temperature, the RHEED pattern shows some polycrystalline circles, seen in Fig. 2b. After the thickness is beyond ∼1.2 nm, a new diffraction pattern emerges and persists as long as Cr is being deposited. The typical pattern at this stage is shown in Fig. 2c. The previous study found that the orientation of the Cr/GaAs (001) film is (112)[110]Cr //(001)[110]GaAs and (112)[111]Cr //(001)[110]GaAs , instead of the geometry shown in Fig. 1c [8]. To determine the epitaxial orientation, we schematically draw these two growth geometries in Fig. 3a, b together with the corresponding Miller indices. There are two types of structures labeled as (I) and (II) for (112) plane growth geometry according to the structure symmetry. The RHEED patterns of type I and type II structures are calculated and shown in Fig. 2f, i, respectively. Comparing with the experimental results (Fig. 2c), we can immediately determine that only type I structure is formed. This result is further confirmed by X-ray diffraction (XRD) [7]. This growth geometry has not been found in other 3d transition metals and their alloys. It is still not understood why the (112) plane growth occurs since the lattice mismatch is as high as 25% along [110] direction or why only type I structure can be grown from the symmetry point of view. For growth temperatures between 130 and 160◦ C, the RHEED pattern is different from the room temperature growth case after the initial deposition of ∼1.4 nm, as shown in Fig. 2e. The growth geometry of Fig. 3a is calculated and shown in Fig. 2h, in good agreement with the experimental pattern of Fig. 3e. Therefore, the growth

    16 Metallic Magnetic Thin Films

    813

    Fig. 2 RHEED patterns of Cr grown on the GaAs (001) substrate at different growth temperatures and thickness, taken with the incident electron beam along the substrate [110] direction: (a) clean GaAs (001); (b) 1.0 nm, 25◦ C; (c) 5.0 nm, 25◦ C; (d) 1.6 nm, 100◦ C; (e) 4.0 nm, 160◦ C; (i–i) calculated diffraction patterns explained in the text [7]

    geometry of Fig. 3a is unambiguously determined to be an unrotated bcc structure. Again, this result is further confirmed by XRD [7]. For growth temperatures between 70 and 130◦ C, the RHEED pattern is a mixture of Fig. 2c, e, as shown in Fig. 2d. Figure 2g shows the mixture of the calculated pattern of Fig. 2f, h. It is clearly matched to the experimental results, indicating that the two growth geometries coexist in the Cr/GaAs (001) films.

    814

    D. Wu and X.-F. Jin

    Fig. 3 Epitaxial relationships between Cr and GaAs: (a) (001)[100]Cr //(001)[100]GaAs and (001)[100]Cr //(001)[100]GaAs ; (b) shadow I, (112)[110]Cr //(001)[110]GaAs and (112)[111]Cr // (001)[110]GaAs ; shadow II, (112)[110]Cr //(001)[110]GaAs and (112)[111]Cr //(001)[110]GaAs [7]. LICENSE #: 4812840298511

    Mn/GaAs (001) Several stable structure phases are available for bulk Mn [9]. The α-phase, which is a complex cubic structure with 58 atoms per unit cell, is stable below 1000 K. It is antiferromagnetic (AFM) below 95 K. The β-phase, which is also a complex cubic structure with 20 atoms per unit cell, is stable between 1000 and 1364 K and can be retained at room temperature by quenching. The β-phase is nonmagnetic. The γ -phase, which has the fcc structure, is stable between 1364 and 1410 K. The δ-phase, which is bcc structure, is stable between 1410 and 1450 K. Both γ - and δ-phases exist well beyond any magnetic ordering temperature. The metal-stable phases of Mn were successfully grown on Pd (001) and Ag (001) surfaces with distorted fcc structure [10, 11]. A nearly fcc structure with 3-monolayer (ML) Mn layers can be obtained in the (Mn/Ag)n superlattice on Ag (001) [12]. Epitaxial

    16 Metallic Magnetic Thin Films

    815

    growth on GaAs (001) provides another approach to stabilize metastable Mn at room temperature. The lattice constants of bcc and fcc structure Mn are estimated to be 0.289 nm and 0.365 nm, respectively, calculated by a simple constant volume approximation based on the Wigner-Seitz radius of α-phase Mn (0.143 nm) [11]. The lattice mismatches between GaAs (001) and bcc and fcc Mn are about 2% and 9%. Figure 4 presents the RHEED patterns of clean GaAs (001) surface and 8-nm-thick Mn/GaAs (001) film grown at 400 K with the incident electron beam along different directions [13]. The spotlike RHEED patterns reflect the rough surface and is the projection of the three-dimensional reciprocal lattice. The RHEED pattern remains the same when the sample is rotated 90◦ C, meaning that the singlecrystalline film is the cubic structure . In comparison with the RHEED pattern of GaAs (001) (Fig. 2a), the lattice constant of Mn film is determined to be 0.368 nm. The RHEED pattern of the sample rotated 45◦ C, i.e., electron beam along [100] direction, is shown in Fig. 4b [13]. The RHEED patterns of the rotated fcc Mn and unrotated bcc Mn are deduced from the crystal structure in Fig. 1 with electron beam along the [110] and the [100] directions. The results are shown in Fig. 5 [13]. The deduced RHEED patterns of the fcc structure is in good agreement with the experimental patterns (Fig. 4), suggesting that fcc structure Mn films are obtained. Figure 6 shows the XRD spectra of three Mn/GaAs (001) films grown at 300, 400, and 500 K, respectively [14]. The structure of the film grown at 300 K is polycrystalline α-phase, as shown in Fig. 6a. A (002) peak of the fcc structure is clearly present in XRD, as shown in Fig. 6b, confirming the RHEED result. The measured out-of-plane lattice constant of 0.362 nm is consistent with the value

    Fig. 4 RHEED patterns for (a) the clean GaAs (001) surface and (b), (c) fcc Mn thin films of 8 nm grown on the GaAs (001) substrate with different electron incident directions, respectively [13]. LICENSE #: 4812840822141

    816

    D. Wu and X.-F. Jin

    Fig. 5 RHEED diffraction patterns calculated for the rotated fcc Mn (001) and the unrotated bcc Mn (001) on the GaAs (001) surface [13]. LICENSE #: 4812840822141

    determined by the RHEED patterns. In addition to the fcc structure, we observe another structure with an out-of-plane lattice constant of 0.207 nm. We attribute it to the Mn2 As-type structure (003) peak with in-plane lattice constant a and b of 0.380 nm and out-of-plane lattice constant c of 0.627 nm [15]. This Mn2 As-type structure is a mixture of Mn-Ga-As and serves as a template for the fcc Mn to reduce the large lattice mismatch between the fcc Mn and GaAs (001). To confirm the structures, a cross-sectional high-resolution transmission electron microscope (HRTEM) is shown in Fig. 6d [14]. Obviously, the fcc Mn is grown on the interfacial Mn2 As-type phase. The lattice constants of these two structures are determined by HRTEM: a = b = 0.380 nm, c = 0.363 nm for the fcc-Mn film, and a = b = 0.380 nm, c = 0.208 nm for the fcc-Mn Mn2 As-type phase, consistent with the RHEED and XRD results. This finding unambiguously confirms the formation of the Mn2 As-type phase. For the film grown at 500 K, two crystallographic phases are observed in XRD similar to that of 400 K, as shown in Fig. 6c. We can identify these two structures as “fcc-Mn-type” and “Mn2 As-type” phases. The “fcc-Mn-type” is not pure Mn, but it is the mixture of Mn-Ga-As. The out-of-plane lattice constant of the “fcc-Mn-type” phase is 0.375 nm, slightly larger than that of the pure fcc Mn.

    Fe/Cu (001) Iron exhibits a rich variety of magnetic phases. Fe is bcc and FM in bulk. The fcc phase is stable above 910◦ C. It is predicted to be nonmagnetic, AFM, FM, or order as a spin density wave (SDW), depending on the lattice constant [16–22]. The fcc-Fe

    16 Metallic Magnetic Thin Films

    817

    Fig. 6 XRD spectra for Mn on GaAs (001), prepared at different growth temperature and film thickness: (a) 300 K, 20 nm Mn; (b) 400 K, 10 nm Mn; and (c) 500 K, 20 nm Mn. (d) HRTEM image for 10 nm Mn on GaAs (001) grown at 400 K [14]. LICENSE #: 4812841356913

    phase can be stabilized in small particles (d ∼ 50 nm) in a Cu matrix, and it is an AFM incommensurate SDW [23, 24]. Epitaxial growth of Fe on Cu (001) at 250–300 K can stabilize the bcc and fcc phases, depending on the film thickness [25–30]. For film thickness below ∼4 monolayers (ML) in region I, it grows in the FM face-centered tetragonal (fct) phase. For a film thickness between 5 and 11 ML (region II), it is the AFM fcc phase covered by an FM fct Fe surface layer. However, the energy of SDW is lower than that of the collinear type 1 AFM state. Its magnetic state is under debate. For the film thickness above 11 ML, it becomes the FM bcc phase. A clear RHEED oscillation can be observed for Fe grown on Cu (001) [31], indicating a layer-by-layer growth mode. It is noticed that quite different RHEED oscillations are found in the literature [Refs. 28, 30]. We found that the oscillation mode changes while the electron beam is measured along different azimuthal angles. Based on this fact, the previously reported oscillation modes can be well reproduced, as shown in Fig. 7 [31]. The film thickness is further accurately measured by in situ

    818

    D. Wu and X.-F. Jin

    Fig. 7 (a) and (b) Two types of RHEED oscillations for Fe/Cu (001) grown at 300 K [31]. LICENSE #: RNP/20/APR/024963

    scanning tunneling microscopy and a quartz crystal monitor. It is very critical to verify that the studied phase is fcc Fe. A martensitic phase transition from fcc to bcc occurs at about 10–12 ML, which can be observed by a scanning tunneling microscope (STM) [32]. Based on this fact, here, we use the observation of the martensitic phase transition by STM and low-energy electron diffraction (LEED) as a criterion to judge the realization of fcc Fe. Figure 8 shows the polar Kerr rotation angle at remanence as a function of film thickness measured at 70 K. A minimum is clearly seen at ∼7 ML and the signal does not drop down at 8.5 and 9 ML. These results cannot be explained by a collinear type 1 antiferromagnet, excluding the type 1 AFM phase of fcc Fe. Figure 9 shows three representative polar magneto-optic Kerr effect (MOKE) signals at remanence versus temperature for 6, 7, and 9 ML of Fe on Cu (001), respectively. The total moment of 7 ML (odd layer) is smaller than that of 6 and 8 ML (even layers) below ∼200 K, which is the effective ordering temperature Te of the AFM underlayers [30]. The coercivity Hc increases significantly when the temperature drops below Te, as seen in the inset of Fig. 9, and the hysteresis curves are shown in Fig. 9a–e. The enhancement of Hc is due to the exchange coupling between the top two FM surface layers and the bulk AFM underlayers below Te. No exchange bias loops are observed in these experiments, possibly due to the weak pinning effect of the thin AFM layer. We propose a SDW model to explain the data [31]. A SDW with Sz = Sz0 cos[q(z−z0 )] is assumed, where Sz is the z component of the magnetic moment, Sz0 is the normalization constant, q is the wave vector of SDW, z is the position in space along the z direction, and z0 is the initial phase term determined by the choice of the origin. Since the total magnetic moment increases below Te, the Fe magnetic moments at the FM/AFM interface must be ferromagnetically rather than antiferromagnetically coupled to each other. It means that z0 = 0 if choosing the origin at the topmost AFM plane, i.e., the third layer from the top. The inset

    16 Metallic Magnetic Thin Films

    819

    Fig. 8 Polar MOKE as a function of film thickness measured at 70 K together with the fitting curve using the incommensurate SDW with Sz = Sz cos(qz) and q = 2π/2.7d (solid line) [31]. LICENSE #: RNP/20/APR/024963

    Fig. 9 Temperaturedependent magnetization as a function of temperature for 6, 7, and 8 ML Fe on Cu (100), respectively; curves (a–e) are some representative hysteresis loops at different temperatures; (f) coercivity versus temperature for 8 ML Fe/Cu (001) [31]. LICENSE #: RNP/20/APR/024963

    of Fig. 10 shows the incommensurate SDW along z direction. The fitting of the experimental data to an SDW yields q = 2π/2.7d, where d is the interlayer distance of fcc Fe, as shown by the solid line in Fig. 8. The good fitting strongly suggests that the magnetic structure is SDW. The magnetic configurations of 6, 7, 8, and 9-ML Fe on Cu (001) can be constructed one by one as schematically shown in Fig. 10.

    Co/GaAs (001) The hexagonal close-packed (hcp) and fcc structures are, respectively, the stable and metastable phases of Co. Although the bcc structure is not available in bulk, it was realized by epitaxial growth on GaAs (110) substrates. A negative cubic magnetic anisotropy K1 and a strong in-plane uniaxial anisotropy Ku were proposed [33]. The negative K1 denotes the easy axes along bcc Co directions. However, it was pointed out that bcc Co phase is not a true metastable phase but a force-induced phase [34]. The bcc Co phase grown on GaAs (001) substrate was first reported by

    820

    D. Wu and X.-F. Jin

    Fig. 10 Magnetic structures proposed for 6, 7, 8, and 9-ML Fe on Cu (100); the inset gives the layer-dependent magnetic moments for fcc Fe along the z direction, z(d) = 0 corresponding to the first AFM layer. (Note: all the moments drawn here are lying in the planes parallel to the front plane of the structure section) [31]. LICENSE #: RNP/20/APR/024963

    Blundell et al. [35] They found a positive K1 for 4-nm Co/GaAs (001), contrary to a previous report. Another group reported that Co films grown on GaAs (001) were not bcc but two-domain hcp, resulting in positive K1 [36]. Theoretical calculation predicted that the surface layer contributes to a large positive K1 [37]. The structure of Co grown on GaAs (001) depends on film thickness and growth temperature. Figure 11a shows a typical RHEED pattern for a clean GaAs (001) surface, with an electron beam along the [110] direction. A new RHEED pattern appears at ∼0.2 nm for Co films grown at 150◦ C. This new pattern is clearly seen after ∼0.8 nm, shown in Fig. 11d [38]. The pattern remains the same by rotating the electron beam 90◦ to the [110] direction, indicating a fourfold in-plane geometry. This pattern is the same as the calculated bcc structure RHEED pattern (Fig. 11g). Moreover, we compare the RHEED pattern of the bcc Fe [39] (Fig. 11b) and fcc Mn (Fig. 11c) epitaxially grown on GaAs (001). From the similarity between Fig. 11b, d, again, we can immediately conclude that the structure is bcc. When the film thickness is beyond ∼2 nm, a new RHEED pattern develops together with the bcc pattern, shown in Fig. 11e. If we consider the new phase is hcp Co proposed earlier [36], the calculated pattern of the mixture of the bcc and hcp structure (Fig. 11h) is consistent with the experimental result (Fig. 11e). The growth geometry is determined to be (1210)[0001]Co //(001)[110]GaAs and (1210)[0001]Co //(001[110]GaAs . When the film thickness exceeds ∼6 nm, the RHEED pattern is dominated by the hcp structure, as shown in Fig. 11f. Interestingly, the pattern remains the same after rotating the electron beam 90◦ , contrary to the sixfold symmetry of the hcp structure. This implies the fourfold symmetry pattern might actually come from the two perpendicularly oriented hcp Co domains, with [0001]Co //(001)[110]GaAs and

    16 Metallic Magnetic Thin Films

    821

    Fig. 11 RHEED patterns for different thin films grown on the GaAs (001) substrate, taken with the incident electron beam along the substrate [110] direction: (a) clean GaAs (001), (b) 5.6-nm bcc Fe, (c) 8.0-nm fcc Mn, (d) 1.5-nm bcc Co, (e) 3.5-nm bcc and hcp Co, (f) 15-nm hcp Co, and (g–i) calculated diffraction patterns corresponding to (d–f), respectively. ◦ and • correspond to the two hcp domains described in the text [38]. LICENSE #: RNP/20/APR/024982

    822

    D. Wu and X.-F. Jin

    [0001]Co //(001[110]GaAs together [36]. The calculated pattern is shown in Fig. 11i according to the two-domain structure, in good agreement with the experimental result (Fig. 11f). In particular, the intensity of the second and fourth columns from right to left are much stronger than they ought to be, owing to the overlap of the two-domain diffraction. The structure of the Co/GaAs (001) film is directly studied by HRTEM, as shown in Fig. 12 [40]. The bilayer structure is clearly seen, with the bcc structure at the bottom and the hcp structure at the top. The diffraction pattern corresponding to this HRTEM picture does indicate that the top hcp Co film has a two-domain structure with c axes perpendicular to each other [38]. The thickness of the Co bcc structure is found to increase with increasing the growth temperature in the grown temperature range studied (60–160◦ C) [41]. A 1.4-nm bcc Co/GaAs (001) film, capped with 3.0-nm Au layer, measured by ex situ MOKE at room temperature, is shown in Fig. 13. The film exhibits strong uniaxial anisotropy with the easy axis along [110] direction. In order to extract the fourfold anisotropy field, rotating MOKE (ROTMOKE) measurements for the same sample [42] are shown in Fig. 13b. The ROTMOKE curve is fitted to the equation [43]: E/V = −μ0 MS H cos (θ − φ) + Ku sin2 φ + K1 /4sin2 (2φ + π/2) ,

    (1)

    where E is the free energy, V is the volume of the film, Ms is the saturation magnetization, H is the magnetic field, θ is the angle between H and Co [110] direction, and φ is the angle between magnetization and Co [110] direction. The

    Fig. 12 Cross-sectional HRTEM picture of the 29 nm Co/GaAs (001) film prepared ◦ at 150 C [40]. LICENSE #: 4813040909381

    16 Metallic Magnetic Thin Films

    823

    Fig. 13 (a) Longitudinal MOKE loops and (b) ROTMOKE data of 1.4-nm Co deposited at 120 ◦ C on GaAs (001). The solid line is the fit to Eq. (1) [41]. LICENSE #: RNP/20/APR/024983

    best fit yields the fourfold anisotropy field μ0 HK1 = 2 K1 /MS = −15.2 mT and the uniaxial anisotropy field μ0 HKu = 2Ku /MS = 22.0 mT. Since the magnetic moment of bcc Co is 1.53 μB [33], we deduce K1 = μ0 HK1 MS /2 = −6.5 kJ/m3 . The negative K1 means that the easy axes of the fourfold anisotropy are along Co directions. Furthermore, the anisotropy as a function of the growth temperature and thickness was systematically investigated by ROTMOKE [41]. Figure 14 shows HK1 fitted from the ROTMOKE curves as a function of the Co film thickness for different growth temperatures, 60, 120, and 160◦ C, and the thickness of the phase transition. HK1 for Co thinner than 1.2 nm deposited at 60 ◦ C is negative, and the absolute values increase monotonically with thickness, suggesting the bulk-like origin of the fourfold anisotropy. When the Co is thicker than 1.2 nm, |HK1 | first decreases to zero at 1.8 nm and then increases with increasing thickness (Fig. 14a). Above 1.2 nm, hcp Co, which contributes a positive K1 [36] and starts to develop [38, 44], leading to the increases of HK1 . For Co film grown at 120◦ C, K1 is negative for the bcc phase and starts to increase when the hcp phase starts to develop. The bcc structure Co films grown at 60–140◦ C always exhibit negative HK1 , as shown in Fig. 14d. Although the bcc structure film grown at 160◦ C is obtained below 2 nm, no Kerr signal is detected. It might be due to the strong reaction between Co and GaAs to form the nonmagnetic bcc CoGaAs compound. Above 2 nm, HK1 or K1 is positive for the mixture phases of the bcc and hcp structures (Fig. 14c).

    824

    D. Wu and X.-F. Jin

    Fig. 14 HK1 as a function of Co thickness measured on the Co wedge samples grown at (a) 60 ◦ C, (b) 120 ◦ C, and (c) 160 ◦ C. Each HK1 value is fitted from ROTMOKE experiments. The dashed line denotes the structure phase transition thickness. (d) HK1 versus growth temperature. HK1 is measured at the transition Co thickness. For Co grown at 160 ◦ C, there is no transition point [41]. The triangle data point is obtained from Refs. [38, 41]. LICENSE #: RNP/20/APR/024983

    Ni/GaAs (001) Unlike Fe and Co which have several structural phases in bulk, nickel only has one bulk structure—fcc. It was reported that a very thin bcc Ni phase less than 6 atomic layers (or 3 unit cells) could be epitaxially grown on a Fe(001) substrate [45, 46]. Above 6 ML, the structure is complicated [45]. The bcc structure Ni was measured to be FM with a magnetic moment of 0.55–0.80 μB /atom [47, 48] or 0.4 ± 0.45 μB /atom by different groups [49]. However, TC was found to lower than 77 K in Ref. [45] and higher than 300 K in Ref. [49]. In the Ni/Fe (001) system, the magnetic properties of bcc Ni can be strongly altered by Fe, considering the strong hybridization and magnetic coupling at the Ni/Fe interface [45, 48–51] For example, the enhanced Tc and induced magnetic moment of Ni could occur by the strong interface coupling [52, 53]. Therefore, it is desirable to obtain bcc Ni on a nonmagnetic substrate to investigate its intrinsic properties. Theoretically, it was predicted that bcc Ni with a lattice constant of 0.2773 nm at equilibrium is paramagnetic and a transition to an FM state occurs upon lattice expansion beyond 0.2815 nm [54]. However, Guo and Wang found that the magnetic phase transition occurs at 0.2730 nm [55]. We found that bcc Ni can be epitaxially grown on GaAs (001) [56, 57]. The growth temperature of 170 K is critical to achieve the thickest bcc Ni films of 3.5 nm. Figure 15a shows a cross-sectional HRTEM image of Ni (3.5 nm)/GaAs (001) with the electron beam along GaAs [110] direction. Ni is unambiguously identified to have a bcc structure [40] with a lattice constant of 0.28 nm. The orientation of the bcc Ni/GaAs (001) film is (001)[100]Ni //(001)[100]GaAs and (001)[110]Ni //(001)[110]GaAs . XRD experiments confirm the structure to be bcc, as shown in Fig. 15b. In addition to the GaAs (400) diffraction peak, a broad peak

    16 Metallic Magnetic Thin Films

    825

    Fig. 15 (a) Fourier-filtered HRTEM image of 3.5 nm bcc Ni/GaAs (001) with the electron beam along the [110] direction. (b) Grazing angle XRD spectrum with an ◦ incident angle at 0.2 [57]. LICENSE #: RNP/20/APR/024984

    superimposed on top of the narrow GaAs (400) peak is clearly observed (inset of Fig. 15b). This broad peak is identified to be the bcc Ni (200) diffraction from which an in-plane lattice constant of 0.282 nm is readily derived. The absence of the fcc Ni ◦ ◦ diffraction peaks at ∼25.9 for (200) and ∼ 38.1 proves that the bcc phase is pure. Figure 16a shows the MOKE signal at remanence measured from a Ni wedge sample at 120 K as a function of the Ni film thickness [57]. The extrapolated dashed line in Fig. 16a passes through the origin, indicating that there is no interfacial dead layer. The vanishing MOKE signal below ∼1.2 nm is attributed to the reduced Tc due to the finite-size scaling effect of Tc in ultrathin film [58]. In contrast, the disappearance of the magnetic signal for the fcc Ni/Cu (001) film is around 0.6 nm at 120 K [59]. To obtain Tc of bulk bcc Ni, Tc is fitted as a function of the bcc Ni thickness (Fig. 16b) by the finite-size scaling law [58]: T c (∞) − T c(l)/T c (∞) = (c/ l)λ ,

    (2)

    where Tc(∞) is Tc in the bulk limit, l is the film thickness, λ is a critical exponent, and c is a parameter. The fitting yields Tc(∞) = 456 K, which is much lower than Tc of bulk fcc Ni (627 K). The magnetic moment is measured to be 0.52 ± 0.08 μB /atom at 5 K by superconducting quantum interference device (SQUID) magnetometry, in good agreement with the theoretically calculated value of 0.54 μB /atom at the lattice constant of 0.282 nm [55]. The magnetocrystalline anisotropy K1 was determined by the ROTMOKE technique [42, 43]. Figure 16c presents the ROTMOKE data as a function of the in-plane magnetic field (μ0 H = 500 mT) direction for the 3.5-nm bcc Ni/GaAs (001) film. Similar to the procedure of determining the anisotropies of epitaxially grown Fex Cu1−x and Co films on GaAs (001) [41, 43], we fitted the experimental curve by considering Ku and K1 . The excellent fitting yields (μ0 HKu = 230 mT and HK1 = 180 mT (Fig. 16c). Using the foregoing obtained magnetic moment of bcc Ni, we obtained K1 = 40 kJm−3 , which is comparable with that of bcc Fe (K1 = 47 kJm−3 ). The positive sign of K1 indicates that bcc Ni has the same cubic magnetic easy axis as bcc Fe but opposite to that of fcc Ni (K1 = −57 kJm−3 ) [60]. We note that the presence of the uniaxial anisotropy in the thick bcc Ni films

    826

    D. Wu and X.-F. Jin

    Fig. 16 (a) MOKE signal at remanence as a function of the bcc Ni thickness measured at 120 K. The inset shows a representative hysteresis loop. (b) Tc as a function of film thickness. The solid line is a theoretical fit to Eq. (2) with Tc(∞) = 456 K, λ = 2.23, and c = 1.17 nm. (c) ROTMOKE data as a function of the magnetization directions. The lines are the theoretical fitting with different parameters [57]. LICENSE #: RNP/20/APR/024984

    could suggest that the bcc lattice has some degree of in-plane shear strain, similar to that of the thick bcc Fe layers on GaAs (001) [61]. In addition to the contrast in the magnetic anisotropy between fcc Ni and bcc Ni, their electronic band structures also show significant difference [57].

    Structure and Magnetism of Magnetic 3d Transition Metal Alloys Py/GaAs (001) Permalloy (Py) usually refers to Fex Ni1−x alloys with a stoichiometry of 0.18 ≤ x ≤ 0.25. Py has a vanishingly small magnetocrystalline anisotropy and magnetostriction but extremely large magnetic permeability. These unique properties make Py one of the most important soft magnets for a variety of applications, such as the cores of motors and the free layers of spin-valve magnetic reading heads. However, it is still unclear from the fundamental point of view why the cubic magnetic anisotropy energy of Py is so small, while both Fe and Ni have quite large magnetic anisotropy energy [62], albeit of opposite sign. In addition to the fundamental importance, a better understanding of the vanishing magnetic anisotropy in Py is crucial for the rational design of innovative intermetallic materials for various applications. It has long been believed that the vanishing of K1 in Py stems from the cancellation between the positive K1 of Fe and the negative K1 of Ni [62–64]. This cancellation picture works surprisingly well for the interpretation of the magnetic anisotropy phase diagram of Fex Ni1−x alloys [65, 66] and was even used to support findings of some first-principles calculations [67, 68]. However, this idea has not been experimentally verified.

    16 Metallic Magnetic Thin Films

    827

    Fig. 17 GIXRD spectrum with an incident angle at 0.3◦ for 5 nm Py, fitted by three identified peaks described in the text. The absence of peaks at 25.9◦ and 38.1◦ indicates the absence of fcc Py [70]. LICENSE #: RNP/20/APR/024985

    The structure of bulk Py is fcc. Since both Fe and Ni grow epitaxially on GaAs (001) in the bcc structure [57, 69], it is possible to grow the bcc structure Fex Ni1−x alloys on GaAs (001). The Fex Ni1−x alloys were grown at 200 K by co-deposition using MBE. In addition to the RHEED and LEED patterns suggesting the successful growth of the bcc Py film, similar to the bcc Fe and Ni films [56], the grazing incident XRD (GIXRD) proves the bcc structure of the Py film directly. Figure 17 presents a GIXRD spectrum of the Au (2 nm)/Fe0.25 Ni0.75 (5 nm)/GaAs (001) sample with a fixed photon energy at 8.052 keV and an incident angle at 0.3◦ [70]. The absence of peaks at 25.9◦ and 38.1◦ for fcc Py (200) and Py (220), respectively, excludes the existence of the fcc phase. The peak at 33.1◦ is composed of three parts: (1) an extremely narrow peak at 33.155◦ from the GaAs (400) diffraction (blue line in the inset); (2) a broad peak from the bcc Py (200) diffraction with an in-plane lattice constant of 0.2825 nm (red line); (3) a very broad and shallow peak from the fcc Au (220) diffraction (green line). We found similar results for samples with all compositions, meaning that bcc structure of Fex Ni1−x alloys is realized over the whole range of stoichiometry. To compare the magnetic properties of fcc and bcc Py, samples with two structures were prepared on one GaAs (001) substrate, as shown in the inset of Fig. 18. Half of the substrate was first covered by a 3-nm fcc Au seed layer by a shadow mask [71]. After removing the shadow mask, a thickness wedge of Py was then deposited on the whole sample. Therefore, we had fcc Py/Au/GaAs (001) [71] and bcc Py/GaAs (001) films with exactly the same stoichiometry on the same sample. Figure 18 shows the Kerr rotations at saturation for two structural Py as a function of the film thickness measured at 110 K. The different slope of Kerr rotations are attributed to the optical effect because the magnetic moment per atom is almost identical for bcc and fcc Py: mbcc = 1.03 ± 0.09μB and mfcc = 1.07 ± 0.09μB measured on the corresponding flat film by vibrating sample magnetometry (VSM).

    828

    D. Wu and X.-F. Jin

    Fig. 18 Kerr rotations of bcc and fcc Fe0.25 Ni0.75 as a function of the film thickness. The right inset shows the special sample prepared for this measurement. The left inset gives the finite-size scaling of Tc of the bcc Py film; dots for experimental data and line for the fitting [70]. LICENSE #: RNP/20/APR/024985

    Furthermore, Tc as a function of the film thickness in a bcc Py wedge was measured and shown in the inset of Fig. 18. Tc of the bulk bcc Py is obtained to be 553 K by fitting the curve to Eq. (2). It is significantly lower than Tc = 871 K for the bulk fcc Py [72]. The sample was covered by 2-nm-thick Au to perform ex situ longitudinal MOKE measurements. The top left inset of Fig. 19 shows the typical hysteresis loops of the 3-nm-thick bcc Py. Except a small uniaxial magnetic anisotropy caused by the substrate as generally found for 3d magnetic metals on III–V semiconductors [73], the in-plane cubic magnetic anisotropy is extremely small. To understand the anisotropy of Py, four compositional wedges of the bcc Fex Ni1−x /GaAs (001) samples were fabricated to cover the entire stoichiometry range [74]. K1 was obtained by the ROTMOKE measurements at room temperature for the whole range of compositions, as shown in Fig. 19 [42, 43]. K1 (x) curve shows a tubelike shape. In contrast, K1 (x) curve of fcc Fex Ni1−x is positive at Fe-rich side, passes through zero at x = 25%, and becomes negative at the Ni-rich side. The magnetization versus concentration of bcc Fex Ni1−x measured by VSM on the flat sample obeys very well the Slater-Pauling behavior, as shown in the top right inset of Fig. 19, in comparison to the high-spin to low-spin transition in the fcc Fex Ni1−x [75]. Moreover, it is observed in the same inset that the magnetization of bcc Fex Ni1−x measured by MOKE has nothing peculiar nearby the eminent “Invar point” (65% Fe), where the fcc-bcc structural phase transition happens in the conventional bulk Fex Ni1−x . First-principles calculations using the full potential linearized augmented plane wave method in conjunction with the generalized gradient approximation (GGA) clarify the experimental results [76, 77]. When x = 0.25, with the (2 × 2 × 2) supercells shown in the inset of Fig. 18b, the calculated spin magnetic moments for bcc and fcc Py are mbcc = 1.11μB /atom and mfcc = 1.14μB /atom, respectively. These results are in good accordance with the VSM data. Since the calculations were done in a chemically ordered supercell (e.g., the DO3 structure for the bcc lattice), the chemical disordering in experimental samples appears to be unimportant

    16 Metallic Magnetic Thin Films

    829

    Fig. 19 Cubic magnetic anisotropy constant (K1 ) versus iron concentration measured from the 3-nm-thick bcc Fex Ni1−x composition wedge (see the carton) by ROTMOKE at 300 K, together with the calculated K1 values are marked as red solid cycles. A set of representative hysteresis loops for bcc Fe0.25 Ni0.75 is shown in the inset. Averaged magnetic moment per atom in bcc Fex Ni1−x is also measured by VSM as a function of iron concentration together with a zoom around the “Invar point” obtained by MOKE [70]. LICENSE #: RNP/20/APR/024985

    for the magnetization [78]. The cubic magnetocrystalline anisotropy energies were calculated through the torque method for bcc Fex Ni1−x in several (2 × 2 × 2) supercells that preserve the cubic symmetry [79]. The results are presented as dots in Fig. 19 for direct comparison with the experiment. It is obvious that the theory reproduces well the general trend of the experimental data. The band-filling dependence of K1 obtained from a rigid band scheme is presented in Fig. 20 for several systems. It is clear to see that adding either a small amount of Fe to Ni or vice versa causes a reduction in K1 as guided by the arrows. To elucidate the element-specific contributions from Fe and Ni, we selectively switch off spin-orbit coupling (SOC) from different elements in the bcc Py. Intriguingly, the SOC of Fe only (denoted as Fe-SOC) gives a small negative contribution to K1 whereas Ni plays the dominant role. Both Ni and Fe behave very differently from their bulk forms. It seems that only the local Fe-Ni bonding is the most important factor for the quenching of the magnitude of K1 , whereas signs of K1 of individual constituents are not relevant.

    Cox Mn1−x /GaAs (001) Among the 3d transition metal alloys , the Mn-based alloys have attracted special attention because of their potential applications in the magnetic recording industry [80–82] The Cox Mn1−x alloy has shown interesting magnetic properties [83–85] Bulk Cox Mn1−x is FM with an hcp structure for 0 < x ≤ 0.32 but is AFM with an fcc structure for 0.32 < x ≤ 0.52. For x > 0.52; it is no longer possible to stabilize the fcc state [86]. The metastable bcc Co and fcc Mn can be epitaxially grown on GaAs (001) surfaces at room temperature [13, 38]. It provides us an ideal platform to

    830

    D. Wu and X.-F. Jin

    Fig. 20 Calculated K1 of Fe, Ni, and Fe3 Ni13 as a function of band filling. The vertical line at zero corresponds to the Fermi energy. “Fe-SOC” means only spin-orbit coupling of Fe is included and vice versa [70]. LICENSE #: RNP/20/APR/024985

    epitaxially grow the Cox Mn1−x alloys on GaAs (001) to investigate the correlation between their structure and magnetism. Figure 21 presents the typical RHEED patterns of the Cox Mn1−x /GaAs (001) films with different compositions with the incident electron beam along the [110] direction [87]. The spot-like patterns indicate island growth mode and allow us to determine the structure [14]. Figure 21a shows the RHEED pattern for a clean GaAs (001) surface, which serves as an internal scale to measure the film lattice constants. Figure 21b shows a typical RHEED pattern for the Cox Mn1−x films with composition 0 < x ≤ 0.44. The rectangular-shaped pattern corresponds to a bcc structure, which does not exist in the bulk and is similar to that of the Co/GaAs (001) film. The in-plane lattice constants of Co60 Mn40 are estimated to be a = b = 0.29 nm from the RHEED pattern (Fig. 21b). The out-of-plane lattice constant is determined to be 0.289 nm by XRD. For 0.78 ≤ x < 1, the RHEED pattern is square, as shown in Fig. 21c, suggesting an fcc structure, similar to that of the Mn/GaAs (001) film. In comparison, no stable bulk phases of Cox Mn1−x alloy can be obtained for x > 0.52. Similar to the above analyses, the lattice constants are estimated to be a = b = 0.36 nm and c = 0.360 nm by the RHEED pattern and XRD. The structure is more complicated and shows thickness dependence for 0.44 ≤ x ≤ 0.78. A typical case is given in Fig. 21d–f, in which the RHEED patterns of Co30 Mn70 thin films were recorded at different thicknesses. In the initial stage at film thickness thinner than 5 nm, the film is bcc (Fig. 21d). As the film gets thicker, a mixture of rectangular- and square-like diffraction patterns is observed (Fig. 21e), indicating the coexistence of the fcc and bcc phases on top of the initial bcc layer. When the film is thicker than ∼27 nm, the film becomes fcc (Fig. 21f). The coexistence of the fcc and bcc phases is further confirmed by XRD [87]. The thickness-dependent behavior in the intermediate range of composition is due to the competition between the fcc structure as Co/GaAs (001) film and bcc structure as Mn/GaAs (001) film. In the intermediate composition range, the bcc phase of the Cox Mn1−x alloys has a much smaller lattice misfit with the substrate (∼ 1%) than

    16 Metallic Magnetic Thin Films

    831

    Fig. 21 Representative RHEED patterns of Cox Mn1−x /GaAs (001) with different Mn composition. (a) Clean GaAs (001) surface; (b) bcc phase of Co60 Mn40 films; (c) fcc phase of Co10 Mn90 films; (d–f) thickness-dependent transition from bcc to fcc phase for the Co30 Mn70 film [87]. LICENSE #: RNP/20/APR/024986

    the fcc phase (∼ 9%), resulting in the initial growth of the bcc phase. However, the bcc phase does not exist in the bulk, indicating a relatively high-energy metastable state. When the film gets thicker, it tends to form the fcc phase. It is very unlikely that the bcc and fcc coexistence region corresponds to a phase separation of Mn-rich fcc and Co-rich bcc phases; otherwise the film structure in the final stage of growth would become much more complicated than the one we observed (Fig. 21f). Based on the foregoing results, a structural phase diagram as a function of composition is plotted in Fig. 22. Three regions are clearly seen, i.e., the Co-rich bcc phase region marked I, the Mn-rich fcc phase region marked III, and the intermediate region with the mixed bcc and fcc structures marked II.

    832

    D. Wu and X.-F. Jin

    Fig. 22 Structural phase diagram for Cox Mn1−x /GaAs (001) films [87]. LICENSE #: RNP/20/APR/024986

    After the structure characterization, the Cox Mn1−x /GaAs (001) films were carried out with the MOKE measurements. The films are in-plane anisotropy if they are FM. Figure 23 shows the typical longitudinal MOKE loops for the Cox Mn1−x /GaAs (001) films with different compositions after being normalized for the film thickness. The film shows strong ferromagnetism for x < 0.44. No hysteresis loop is observed for x > 0.78. The ferromagnetism is relatively weak for 0.44 ≤ x ≤ 0.78. Since the structure of the Cox Mn1−x /GaAs (001) films is thickness-dependent in the intermediate composition, a wedge-shaped sample was fabricated and measured by MOKE at room temperature to understand the correlation between the structure and magnetism. The results of the Kerr intensity versus thickness for a Co52 Mn48 film are shown in Fig. 24. The data can be divided into three regions: i) a magnetic dead layer (∼ 2 nm) is near the interface; ii) the linearly increasing Kerr intensity (∼ 2–6 nm) region means an FM layer; iii) the saturated Kerr intensity region (> 6 nm) means a non-FM layer on the top. The magnetic dead layer is caused by the interface reaction between film and substrate, as is quite common in such systems. The FM layer comes from the bcc-dominated Co52 Mn48 film. The non-FM layer has an fcc-dominated structure. Furthermore, the fcc-Cox Mn1−x /Fe bilayer exhibits an exchange bias effect, indicating the fcc structure Cox Mn1−x is AFM. Therefore, we conclude that Cox Mn1−x film is FM whenever its structure is bcc, and it is AFM whenever its structure is fcc. The magnetism is intimately correlated with a structure

    16 Metallic Magnetic Thin Films

    833

    Fig. 23 Longitudinal MOKE hysteresis curves for the Cox Mn1−x /GaAs (001) films at different Mn compositions [87]. LICENSE #: RNP/20/APR/024986

    for the Cox Mn1−x alloys. This strong correlation realized in experiments between the structure and magnetism of Cox Mn1−x alloys was further studied and confirmed by a first-principles linearized augmented plane-wave (LAPW) calculation with the local-spin-density approximation (LSDA) [87]. This correlation between structure and magnetism is also found in Fex Mn1−x /GaAs (001) films [88].

    Fex Cu1−x /GaAs (001) Since bulk Fe has the bcc structure, the magnetic properties of fcc Fe attract tremendous interest. Fcc Fe can only be prepared via precipitation in a Cu matrix or epitaxially grown on a Cu substrate [31]. Because of the extremely low bulk miscibility of either Cu in Fe or Fe in Cu, only ∼4% Fe dissolves into Cu (with fcc structure) and ∼ 10% Cu into Fe (with bcc structure) near their respective liquid points [89], and a long-standing challenge has been the goal of producing metastable Fex Cu1−x alloys with higher concentrations by different approaches, such as rapid quenching, vapor deposition, ion beam mixing, or mechanical alloying [90–94]. However, the samples prepared by these methods are polycrystalline. A

    834

    D. Wu and X.-F. Jin

    Fig. 24 Kerr intensity for a wedge-shaped Co52 Mn48 film at different thicknesses [87]. LICENSE #: RNP/20/APR/024986

    monocrystalline sample was achieved in a monolayer stacked Fe/Cu superlattice grown on the Cu (001) substrate by pulsed laser deposition [95]. This superlattice corresponds to only one composition of x = 0.5, Fe0.5 Cu0.5 . The magnetic properties of the Fex Cu1−x alloys with increasing x is controversial. It was reported that the Fe moment remains ∼2.2 μB /atom over a wide composition range at x > 0.50 and then quickly falls to zero as x decreases experimentally [96]. However, another group found that there were no noticeable changes in the magnetic moment over the entire composition range [91]. Firstprinciples calculations showed that the magnetic moment per Fe atom was almost independent of Fe-content over a wide range (x > 0.5) in both the bcc and fcc phases and then decreased and reached zero for x ≤ 0.25 [97]. However, it was found that the magnetic moment per Fe atom in the bcc phase increased from 2.2 μB to 2.62 μB as x decreased from 1 to 0.5 after taking into account the magneto-volume effect in the calculation [98]. Fe can be epitaxially grown on GaAs (001) with the bcc structure [39]. The RHEED pattern of bcc Fe/GaAs (001) film is shown in Fig. 25a [43]. The aspect ratio of a*/b*, shown in Fig. 25a, is close to 1.4. The RHEED patterns of the Cu/GaAs (001) films transform from rectangular to square shape as increasing the film thickness, as shown in Fig. 25b–d, meaning that the film is almost bcc-like for the first several layers, then body-centered tetragonal (bct), and finally in its thermodynamically stable fcc structure. The evolution of the corresponding aspect ratio of a*/b* extracted from the RHEED patterns as a function of the film thickness is plotted in Fig. 26.

    16 Metallic Magnetic Thin Films

    835

    Fig. 25 RHEED patterns for (a) Fe film, (b–d) Cu film, and (e–h) Fe0.25 Cu0.75 film on GaAs (001) with the electron beam along the [110] direction [43]. LICENSE #: RNP/20/APR/024987

    Fig. 26 The a*/b* aspect ratio extracted from RHEED pattern as a function of the film thickness for Fex Cu1−x at different x [43]. LICENSE #: RNP/20/APR/024987

    Similar to Cu/GaAs (001) film, the structure varies with the film thickness for the Fex Cu1−x films with any particular x value. Figure 25e–h shows a representative example of Fe0.25 Cu0.75 . The structure evolves from bcc to bct (closer to bcc), then to bct but closer to fcc, and finally to fcc. The structural evolution for different Fex Cu1−x films is summarized by the aspect ratio of a*/b* in Fig. 26. At the same thickness of 6 nm, a*/b* changes from 1 (fcc) for Cu to 1.4 (bcc) for Fe as increasing x. The Fex Cu1−x /GaAs (001) films were covered by 2 nm Au to conduct ex situ MOKE measurements. It is found that the ferromagnetic to non-ferromagnetic

    836

    D. Wu and X.-F. Jin

    Fig. 27 (a) The Kerr intensity at saturation is plotted as a function of the film thickness for Fe0.5 Cu0.5 . The inset shows a typical MOKE hysteresis loop, from which the Kerr intensity at saturation is obtained; (b) the effective magnetic moment per Fe concentration [43]. LICENSE #: RNP/20/APR/024987

    transition occurs at x ≈ 0.33. This result is quite different from the polycrystalline samples, where the transition is at x ≈ 0.5 [97]. Considering that the structure of the Fex Cu1−x films is dependent on the film thickness, the magnetization for different thickness was measured in the wedge samples at room temperature. The result of Fe0.5 Cu0.5 is shown in Fig. 27a. The straight line above 2 nm across origin indicates that there is no noticeable change of effective magnetic moment per Fe atom. The deviation from the straight line below 2 nm is due to the finite size effect of Tc [58]. The SQUID measurements were carried out at 5 K on the 6-nm Fex Cu1−x films to quantify the effective magnetic moment per Fe atom, assuming that Cu does not contribute to the magnetization. The results are shown in Fig. 27b. The effective magnetic moment per Fe atom decreases with increasing x, consistent with the previous theoretical calculation which is also plotted in the figure [98]. For the 6-nm-thick Fe0.33 Cu0.67 , the magnetization is too small to be detected. It is known that Fe/GaAs (001) has a fourfold anisotropy with the easy axis along the directions and a uniaxial anisotropy with the easy axis along the [110] direction [69]. The total energy density of the system is E/V = −μ0 MS H cos (α − θ ) + Ku sin2 θ     ◦ ◦ + K1 sin2 θ + 45 + K1 cos2 θ + 45 .

    (3)

    Here, the first term represents the Zeeman energy, α is the angle between the easiest axis (ea) and H (see the inset of Fig. 28b), and θ is the angle between the easiest axis and the magnetization; the second term is the uniaxial anisotropy energy, and the third term is the fourfold anisotropy energy. In order to extract the anisotropy field, we carried out the ROTMOKE measurements. The results are shown in Fig. 28. The ROTMOKE curve (Fig. 28a) is fitted to Eq. (3) (solid line in Fig. 28a) to yield the anisotropy field μ0 HKu = Ku /MS = 6.2 mT and μ0 HK1 = K1 /MS = 60.4 mT, respectively, for the Fe/GaAs (001) film. For the Fe0.50 Cu0.50 /GaAs (001) film,

    16 Metallic Magnetic Thin Films

    837

    Fig. 28 The torque momentl(θ) as a function of θ, (a) μ0 H = 80 mT, 4 nm Fe; (b) μ0 H = 20 mT, 8 nm Fe0.50 Cu0.50 . A schematic picture of the applied field H and the magnetization M is shown in the inset [43]. LICENSE #: RNP/20/APR/024987

    it is obvious that the ROTMOKE curve (Fig. 28b) cannot be fitted by considering only the uniaxial anisotropy (dashed line in Fig. 28b). The curve is well fitted with both uniaxial and fourfold anisotropies for μ0 HKu = 6.9 mT and μ0 HK1 = 4.7 mT as shown by the solid line in Fig. 28b. Although the fourfold anisotropy of Fe0.50 Cu0.50 /GaAs (001) is much weaker than that of pure Fe, it still retains.

    Fex Pd1−x /Cu (100) It is well-known that the magnetic long-range order does not exist in an isotropic two-dimensional (2D) Heisenberg system at any finite temperature [99]. However, it is hard to physically realize a 2D FM film without any anisotropy. Experimentally, it might be realized in the vicinity of the spin reorientation transition (SRT) in ultrathin magnetic films, because of the magnetocrystalline anisotropy, surface anisotropy, and dipolar anisotropy could compensate at the SRT point in some cases. Therefore, it is very interesting to study the correlation between SRT and the long-range magnetic order. Indeed, a suppression of magnetization or a reduction of Tc within the SRT region was reported in Fe/Cu (001), Fe/Ag (001), and Fe/Ni/Cu (001) [100– 102]. However, if it is a continuous SRT, the anisotropy never disappears within the SRT region, where the magnetization only rotates between in-plane and out-ofplane as Ni/Cu (001) [103, 104]. Interestingly, it was found that the SRT occurs at 87% Ni concentration in 9-ML Fex Pd1−x /Cu (100) and Tc is dramatically reduced [105]. Therefore, the Fex Pd1−x alloy provides us a good system to investigate the correlation between SRT and Tc [106]. A Fex Pd1−x composition wedge sample was epitaxially grown on Cu (001) to explore the composition-dependent SRT [74]. The magnetic properties of the sample were investigated through both MOKE and ac susceptibility measurements after capping 3 ML Cu on the sample [52, 107–109]. Two representative hysteresis loops of 9-ML sample measured by MOKE at room temperature were shown in

    838

    D. Wu and X.-F. Jin

    Fig. 29 (a) The longitudinal MOKE hysteresis loop of Ni0.57 Pd0.43 and polar MOKE hysteresis loop of Ni0.91 Pd0.09 measured on a 9-ML composition wedge sample. (b) The imaginary part of the ac susceptibility Im(χ) and remanent magnetization determined from polar MOKE of 9-ML composition wedge sample as a function of Ni composition [106]. LICENSE #: RNP/20/APR/024988

    Fig. 29a, one for Ni0.57 Pd0.43 with easy axis in-plane and the other for Ni0.91 Pd0.19 with a perpendicular easy axis. The remanent magnetization measured by the polar MOKE as a function of the Ni composition is plotted in Fig. 29b. The remanent magnetization increases gradually with increasing Ni concentration, suggesting that the magnetization rotates continuously from in-plane to out-of-plane. The imaginary part of the ac susceptibility Im(χ ) as a function of the Ni composition was also shown in Fig. 29b. Im(χ ) shows a pronounced peak at the SRT position, which is at 86% Ni concentration for 9-ML Fex Pd1−x /Cu (100), consistent with the previous report [105]. This SRT can be understood by the competition between the shape anisotropy and the magnetocrystalline anisotropy including magnetoelastic anisotropy originating from the lattice distortion and the surface anisotropy, similar to the Ni/Cu (001) system [104, 110, 111]. Tc of the 9-ML compositional Fex Pd1−x /Cu (100) film is determined by the ac susceptibility measurement, as shown in an example in the inset of Fig. 30a. The result of Tc and the susceptibility as a function of Ni composition is shown in Fig. 30a. The reduction of Tc is not found at the SRT position where the susceptibility exhibits a peak. Tc increases very slowly as a function of the Ni composition before SRT, while it increases more sharply after SRT. This behavior of the turning point of Tc at SRT indicates that the variation in anisotropy influences the increasing behavior of Tc. To confirm the correlation between SRT and Tc, a 7-ML compositional Fex Pd1−x /Cu (100) film was fabricated and measured using

    16 Metallic Magnetic Thin Films

    839

    Fig. 30 (a) Im(χ) and Curie temperatures versus Ni composition of 9-ML Nix Pd1−x composition wedge sample. An example of Curie temperature measurement by ac susceptibility is given in the inset, where the remanent magnetization determined by MOKE as a function of temperature is also shown. (b) Im(χ) and Curie temperatures versus Ni composition for 7 ML Nix Pd1−x composition wedge sample [106]. LICENSE #: RNP/20/APR/024988

    Fig. 31 Curie temperatures versus increasing perpendicular anisotropy and fourfold in-plane anisotropy, in which K is the anisotropy parameter and J is the FM exchange constant [106]. LICENSE #: RNP/20/APR/024988

    the same approach. The results are shown in Fig. 30b. Again, the turning point of Tc is at the SRT point, which is at a Ni concentration of 96% and no reduction in Tc is observed in the region of SRT, unlike another report (Ref. [105]). No reduction in Tc is observed around the SRT region in our experiments. One reason could be that its anisotropy never really disappears, but the easy direction of the magnetization rotates continuously within the transition [112]. The other possibility could be that at the SRT point there still exists some residual perpendicular anisotropy which is caused by variations in the perpendicular anisotropy across the film surface [113]. Another interesting phenomenon in our experiments is that increasing the perpendicular anisotropy may enhance Tc much more strongly than the fourfold in-plane anisotropy. A simple Monte Carlo simulation was performed

    840

    D. Wu and X.-F. Jin

    to calculate Tc of a 2D Heisenberg model with perpendicular anisotropy and fourfold in-plane anisotropy, respectively. The calculation shows that Tc increases much faster by increasing the perpendicular anisotropy than the fourfold anisotropy, shown in Fig. 31. This is because the fourfold in-plane anisotropy is a higherorder term compared to the perpendicular uniaxial anisotropy. The perpendicular anisotropy breaks the symmetry of the system more seriously than the fourfold one. As a consequence, uniaxial anisotropy should be much more effective at triggering ferromagnetism than fourfold anisotropy in a low-dimensional system. Acknowledgments One of the authors, XFJ, acknowledges the invaluable contribution of the past co-workers of Prof. G. S. Dong and graduated students.

    References 1. Baibich, M.N., Broto, J.M., Fert, A., Nguyen Van Dau, F., Petroff, F., Eitenne, P., Creuzet, G., Friederich, A., Chazelas, J.: Giant magnetoresistance of (001) Fe/(001) Cr magnetic superlattices. Phys. Rev. Lett. 61, 2472 (1988) 2. Binash, G., Grünberg, G.P., Saurenbach, F., Zinn, W.: Enhanced magnetoresistance in layered magnetic structures with antiferromagnetic interlayer exchange. Phys. Rev. B. 39, 4828 (1989) 3. Pauling, L.: The nature of the interatomic forces in metals. Phys. Rev. 54, 899–904 (1938) 4. Slater, J.C.: Electronic structure of alloys. J. Appl. Phys. 8, 385–390 (1937) 5. Miron, I.M., Garello, K., Gaudin, G., Zermatten, P.-J., Costache, M.V., Auffret, S., Bandiera, S., Rodmacq, B., Schuhl, A., Gambardella, P.: Perpendicular switching of a single ferromagnetic layer induced by in-plane current injection. Nature. 2011(476), 189 (2011) 6. Jin, X.F.: Interfaces between magnetic thin films and GaAs substrate. J. Electr. Spectr. Relat. Phen. 114–116, 771 (2001) 7. Qian, D., Liu, G.L., Loison, D., Dong, G.S., Jin, X.F.: Growth and structure of Cr thin films on GaAs (0 0 1). J. Cryst. Grow. 218, 197 (2000) 8. Wolf, J.A., Krebs, J.J., Prinz, G.A.: Growth and magnetic characterization of face centered cubic Co on (001) diamond. Appl. Phys. Lett. 65, 1057 (1994) 9. Weiss, R.J., Tauerr, K.J.: Thermodynamics and magnetic structures of the allotropic modifications of manganese. J. Phys. Chem. Solids. 4, 135 (1958) 10. Tian, D., Li, H., Wu, S.C., Jona, F., Marcus, P.M.: Epitaxy of Mn on Pd (001). Solid State Commun. 70, 199 (1989) 11. Jonker, B.T., Krebs, J.J., Prinze, G.A.: Growth and magnetic characterization of Mn films and superlattices on Ag (001). Phys. Rev. B. 39, 1399 (1989) 12. Idzerda, Y.U., Jonker, B.T., Elam, W.T., Prinze, G.A.: Identification of the Mn structure in Mn/Ag (001) superlattices. J. Appl. Phys. 67, 5385 (1990) 13. Jin, X., Zhang, M., Dong, G.S., Xu, M., Chen, Y., Xun Wang, X.G.Z., Shen, X.L.: Stabilization of face-centered-cubic Mn films via epitaxial growth on GaAs (001). Appl. Phys. Lett. 65, 3078 (1994) 14. Jin, X., Chen, Y., Lin, X.W., Dong, D.S., Chen, Y., Xu, M., Zhu, W.R., Wang, X., Shen, X.L., Li, L.: Interface structure of fcc Mn on GaAs (001). Appl. Phys. Lett. 70, 2455 (1997) 15. Dong, G.S., Xu, M., Chen, Y., Jin, X., Xun Wang: XPS study of Mn thin films grown on GaAs (001) surfaces. Surf. Interface Anal. 24, 653 (1996) 16. Wang, C.S., Klein, B.M., Krakauer, H.: Theory of magnetic and structural ordering in iron. Phys. Rev. Lett. 54, 1852 (1985) 17. Pinski, F.J., Staunton, J., Gyorffy, B.L., Johnson, D.D., Stocks, G.M.: Ferromagnetism versus antiferromagnetism in face-centered-cubic iron. Phys. Rev. Lett. 56, 2096 (1986)

    16 Metallic Magnetic Thin Films

    841

    18. Moruzzi, V.L., Marcus, P.M., Schwarz, K., Mohn, P.: Ferromagnetic phases of bcc and fcc Fe, Co, and Ni. Phys. Rev. B. 34, 1784 (1986) 19. Myrasov, O.N., Gubanov, V.A., Liechtenstein, A.I.: Spiral-spin-density-wave states in fcc iron: Linear-muffin-tin-orbitals band-structure approach. Phys. Rev. B. 45, 12330 (1992) 20. Uhl, M., Sandratskii, L.M., Kubler, J.: Spin fluctuations in γ-Fe and in Fe3 Pt Invar from local-density-functional calculations. Phys. Rev. B. 50, 291 (1994) 21. Körling, M., Ergon, J.: Gradient-corrected ab initio calculations of spin-spiral states in fcc-Fe and the effects of the atomic-spheres approximation. Phys. Rev. B. 54, R8293 (1996) 22. Zhou, Y.M., Wang, D.S., Kawazoe, Y.: Effective ab initio exchange integrals and magnetic phase transition in fcc Fe and Mn antiferromagnets. Phys. Rev. B. 59, 8387 (1999) 23. Tsunoda, Y.: Spin-density wave in cubic γ-Fe and γ-Fe100-x Cox precipitates in Cu. J. Physica. C. 1(10), 427 (1989) 24. Tsunoda, Y., Nishioka, Y., Nicklow, R.M.: Spin fluctuations in small γ-Fe precipitates. J. Magn. Magn. Mater. 128, 133 (1993) 25. Kalki, K., Chambliss, D.D., Johnson, K.E., Wilson, R.J., Chiang, S.: Evidence for martensitic fcc-bcc transition of thin Fe films on Cu (100). Phys. Rev. B. 48, 18344 (1993) 26. Wuttig, M., Feldmann, B., Thomassen, J., May, F., Zillgen, H., Brodde, A., Hannemann, H., Neddermeyer, H.: Structural transformations of fcc iron films on Cu (100). Surf. Sci. 291, 14 (1993) 27. Zillgen, H., Feldmann, B., Wuttig, M.: Structural and magnetic properties of ultrathin Fe films deposited at low temperature on Cu (100). Surf. Sci. 321, 32 (1994) 28. Thomassen, J., Feldmann, B., Wuttig, M., Ibach, H.: Magnetic live surface layers in Fe/Cu (100). Phys. Rev. Lett. 69, 3831 (1992) 29. Keune, W., Ezawa, T., Macedo, W.A.A., Glos, U., Schletz, K.P., Kirschbaum, U.: Magnetovolume effects in γ-Fe ultrathin films and small particles. Physica (Amsterdam). 161B, 269 (1989) 30. Li, D., Freitag, M., Pearson, J., Qiu, Z.Q., Bader, S.D.: Magnetic and structural instabilities of ferromagnetic and antiferromagnetic Fe/Cu (100). J. Appl. Phys. 76, 6425 (1994).; Phys. Rev. Lett. 72, 3112 (1994) 31. Qian, D., Jin, X.F., Barthel, J., Klaua, M., Kirschner, J.: Spin-density wave in ultrathin Fe films on Cu (100). Phys. Rev. Lett. 87, 227204 (2001) 32. Shen, J., Schmidthals, C., Woltersdorf, J., Kirschner, J.: Structural phase transformation under reversed strain: a comparative study of iron ultrathin film growth on nickel and copper (100). Surf. Sci. 407, 90 (1998) 33. Prinz, G.A.: Stabilization of bcc Co via epitaxial growth on GaAs. Phys. Rev. Lett. 54, 1051 (1985) 34. Liu, A.Y., Singh, D.J.: Elastic instability of bcc cobalt. Phys. Rev. B. 47, 8515 (1993) 35. Blundell, S.J., Gester, M., Bland, J.A.C., Daboo, C., Gu, E., Baird, M.J., Ives, A.J.R.: Structure induced magnetic anisotropy behavior in Co/GaAs (001) films. J. Appl. Phys. 73, 5948 (1993) 36. Gu, E., Gester, M., Hicken, R.J., Daboo, C., Tselepi, M., Gray, S.J., Bland, J.A.C., Brown, L.M.: Fourfold anisotropy and structural behavior of epitaxial hcp Co/GaAs (001) thin films. Phys. Rev. B. 52, 14704 (1995) 37. Klautau, A.B., Eriksson, O.: Ab initio calculation of the magnetocrystalline anisotropy and spin and orbital moments of a bcc Co (001) surface. Phys. Rev. B. 72, 014459 (2005) 38. Wu, Y.Z., Ding, H.F., Jing, C., Wu, D., Liu, G.L., Gordon, V., Dong, G.S., Jin, X.F., Zhu, S., Sun, K.: In-plane magnetic anisotropy of bcc Co on GaAs (001). Phys. Rev. B. 57, 11935 (1998) 39. Waldrop, J.R., Grant, R.W.: Interface chemistry of metal-GaAs Schottky-barrier contacts. Appl. Phys. Lett. 34, 630 (1979).; Prinz, G.A., Krebs, J.J.: Molecular beam epitaxial growth of single-crystal Fe films on GaAs. ibid. 39, 397 (1981); Dong, G.S., Chen, Y., Zhang, M., Jin, X.: Simple design of electron beam evaporators for 3d transition metals. J. Vac. Sci. Technol. A 13, 159 (1995)

    842

    D. Wu and X.-F. Jin

    40. Wu, Y.Z., Ding, H.F., Jing, C., Wu, D., Dong, G.S., Jin, X.F., Sun, K., Zhu, S.: Epitaxy and magnetism of Co on GaAs (001). J. Mag. Mag. Mat. 198-199, 297 (1999) 41. Xu, X.Y., Yin, L.F., Wei, D.H., Tian, C.S., Dong, G.S., Jin, X.F., Jia, Q.J.: Measurement of the thickness-dependent magnetic anisotropy of Co/GaAs (001). Phys. Rev. B. 77, 052403 (2008) 42. Mattheis, R., Quednau, G.: Determination of the anisotropy field strength in ultra-thin magnetic films using longitudinal MOKE and a rotating field: the ROTMOKE method. J. Magn. Magn. Mater. 205, 143 (1999) 43. Tian, Z., Tian, C., Yin, L., Wu, D., Dong, G., Jin, X., Qiu, Z.: Magnetic ordering and anisotropy of epitaxially grown Fex Cu1-x alloy on GaAs (001). Phys. Rev. B. 70, 012301 (2004) 44. Mangan, M.A., Spanos, G., Ambrose, T., Prinz, G.A.: Transmission electron microscopy investigation of Co thin films on GaAs (001). Appl. Phys. Lett. 75, 346 (1999) 45. Heinrich, B., Purcell, S.T., Dutcher, J.R., Urquhart, K.B., Cochran, J.F., Arrott, A.S.: Structural and magnetic properties of ultrathin Ni/Fe bilayers grown epitaxially on Ag (001). Phys. Rev. B: Cond. Matter. 38, 12879 (1988) 46. Wang, Z.Q., Li, Y.S., Jona, F., Marcus, P.M.: Epitaxial growth of body-centered-cubic nickel on iron. Solid State Commun. 61, 623 (1987) 47. Bland, J.A.C., Bateson, R.D., Johnson, A.D., Heinrich, B., Celinski, Z., Lauter, H.J.: Magnetic properties of ultrathin bcc Fe (001) films grown epitaxially on Ag (001) substrates. J. Magn. Magn. Mater. 93, 331 (1991) 48. Celinski, Z., Urquhart, K.B., Heinrich, B.: Using ferromagnetic resonance to measure the magnetic moments of ultrathin films. J. Magn. Magn. Mater. 166, 6 (1997) 49. Brookes, N.B., Clarke, A., Johnson, P.D.: Electronic and magnetic structure of bcc nickel. Phys. Rev. B. 46, 237 (1992) 50. Lee, J.I., Hong, S.C., Freeman, A.J., Fu, C.L.: Enhanced surface and interface magnetism of bcc Ni overlayers on Fe (001). Phys. Rev. B. 47, 810 (1993) 51. Lin, T., Schwickert, M.M., Tomaz, M.A., Chen, H., Harp, G.R.: X-ray magnetic-circulardichroism study of Ni/Fe (001) multilayers. Phys. Rev. B. 59, 13911 (1999) 52. Aspelmeier, A., Tischer, M., Farle, M., Russo, M., Baberschke, K., Arvanitis, D.: Ac susceptibility measurements of magnetic monolayers: MCXD, MOKE, and mutual inductance. J. Magn. Magn. Mater. 146, 256 (1995).; Wu, J., Jin, X.F.: Temperature-dependent magnetization in a ferromagnetic bilayer consisting of two materials with different Curie temperatures. Phys. Rev. B 70, 212406 (2004) 53. Rader, O., Vescovo, E., Redinger, J., Blügel, S., Carbone, C., Eberhardt, W., Gudat, W.: Feinduced magnetization of Pd: the role of modified Pd surface states. Phys. Rev. Lett. 72, 2247 (1994) 54. Moruzzi, V.L.: Singular volume dependence of transition-metal magnetism. Phys. Rev. Lett. 57, 2211 (1986).; Moruzzi, V.L., Marcus, P.M., Schwarz, K., Mohn, P.: Ferromagnetic phases of bcc and fcc Fe, Co, and Ni. Phys. Rev. B 34, 1784 (1986); Moruzzi, V.L., Marcus, P.M.: Magnetism in bcc 3d transition metals: onset and approach to the Hund’s-rule limit. Phys. Rev. B 38, 1613 (1988) 55. Guo, G.Y., Wang, H.H.: Gradient-corrected density functional calculation of elastic constants of Fe, Co and Ni in bcc, fcc and hcp structures. Chin. J. Phys. 38, 949 (2000) 56. Tang, W.X., Qian, D., Wu, D., Wu, Y.Z., Dong, G.S., Jin, X.F., Chen, S.M., Jiang, X.M., Zhang, X.X., Zhang, Z.: Growth and magnetism of Ni films on GaAs (001). J. Magn. Magn. Mater. 240, 404 (2002) 57. Tian, C.S., Qian, D., Wu, D., He, R.H., Wu, Y.Z., Tang, W.X., Yin, L.F., Shi, Y.S., Dong, G.S., Jin, X.F., Jiang, X.M., Liu, F.Q., Qian, H.J., Sun, K., Wang, L.M., Rossi, G., Qiu, Z.Q., Shi, J.: Body-centered-cubic Ni and its magnetic properties. Phys. Rev. Lett. 94, 137210 (2005) 58. Schulz, B., Schwarzwald, R., Baberschke, K.: Magnetic properties of ultrathin Ni/Cu (100) films determined by a UHV-FMR study. Surf. Sci. 307–309, 1102 (1994).; Zhang, R., Willis, R.F.: Thickness-dependent Curie temperatures of ultrathin magnetic films: effect of the range of spin-spin interactions. Phys. Rev. Lett. 86, 2665 (2001)

    16 Metallic Magnetic Thin Films

    843

    59. Tischer, M., Arvanitis, D., Yokoyama, T., Lederer, T., Tröger, L., Baberschke, K.: Temperature dependent MCXD measurements of thin Ni films on Cu (100). Surf. Sci. 307–309, 1096 (1994) 60. See Derek Craik: Magnetism. University of Nottingham, Nottingham (1995) 61. Thomas, O., Shen, Q., Schieffer, P., Tournerie, N., Lépine, B.: Interplay between anisotropic strain relaxation and uniaxial interface magnetic anisotropy in epitaxial Fe films on (001) GaAs. Phys. Rev. Lett. 90, 017205 (2003) 62. See, for example, Derek Craik: Magnetism: Principles and Applications, p. 392. Wiley, Chichester (1995) 63. Mckeehan, L.W.: Magnetic interaction and resultant anisotropy in unstrained ferromagnetic crystals. Phys. Rev. 52, 18 (1937) 64. Brooks, H.: Ferromagnetic anisotropy and the itinerant electron model. Phys. Rev. 58, 909 (1940) 65. Tarasov, L.P.: Ferromagnetic anisotropy of low nickel alloys of iron. Phys. Rev. 56, 1245 (1939) 66. Bozorth, R.M.: The permalloy problem. Rev. Mod. Phys. 25, 42 (1953) 67. James, P., Eriksson, O., Hjortstam, O., Johansson, B., Nordstrom, L.: Calculated trends of the magnetostriction coefficient of 3d alloys from first principles. Appl. Phys. Lett. 76, 915 (2000) 68. Weinberger, P., Szunyogh, L., Blaas, C., Sommers, C., Entel, P.: Magnetic properties of bulk Nic Fe1-c alloys, their free surfaces, and related spin-valve systems. Phys. Rev. B. 63, 094417 (2001) 69. Krebs, J.J., Jonker, B.T., Prinz, G.A.: Properties of Fe single-crystal films grown on (100) GaAs by molecular-beam epitaxy. J. Appl. Phys. 61, 2596 (1987) 70. Yin, L.F., Wei, D.H., Lei, N., Zhou, L.H., Tian, C.S., Dong, G.S., Jin, X.F., Guo, L.P., Jia, Q.J., Wu, R.Q.: Magnetocrystalline anisotropy in permalloy revisited. Phys. Rev. Lett. 97, 067203 (2006) 71. Noh, D.Y., Hwu, Y., Kim, H.K., Hong, M.: X-ray-scattering studies of the interfacial structure of Au/GaAs. Phys. Rev. B. 51, 4441 (1995).; Andersson, T.G., Le Lay, G., Kanski, J., Svensson, S.P.: Room-temperature growth of two-dimensional gold films on GaAs (001). Phys. Rev. B 36, 6231 (1987) 72. Wakelin, R.J., Yates, E.L.: A study of the order-disorder transformation in Iron-Nickel alloys in the region FeNi3. Proc. Phys. Soc. London, Sect. B. 66, 221 (1953) 73. Wastlbauer, G., Bland, J.A.C.: Structural and magnetic properties of ultrathin epitaxial Fe films on GaAs (001) and related semiconductor substrates. Adv. Phys. 54, 137 (2005) 74. Tian, C.S.: Ph. D. Thesis, Fudan University, Shanghai, China (2006) 75. Schumann, F.O., Willis, R.F., Goodman, K.G., Tobin, J.G.: Magnetic instability of ultrathin fcc Fex Ni1-x films. Phys. Rev. Lett. 79, 5166 (1997) 76. Wimmer, E., Krakauer, H., Weinert, M., Freeman, A.J.: Full-potential self-consistent linearized-augmented-plane-wave method for calculating the electronic structure of molecules and surfaces: O2 molecule. Phys. Rev. B. 24, 864 (1981).; Weinert, M., Wimmer, E., Freeman, A.J.: Total-energy all-electron density functional method for bulk solids and surfaces. Phys. Rev. B 26, 4571 (1982) 77. Perdew, J.P., Burke, K., Ernzerhof, M.: Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996) 78. Shull, C.G., Wilkinson, M.K.: Neutron diffraction studies of the magnetic structure of alloys of transition elements. Phys. Rev. 97, 304 (1955) 79. Wu, R.Q., Freeman, A.J.: Spin–orbit induced magnetic phenomena in bulk metals and their surfaces and interfaces. J. Magn. Magn. Mater. 200, 498 (1999) 80. Dieny, B., Speriosu, V.S., Gurney, B.A., Parkin, S.S.P., Whilhoit, D.R., Roche, K.P., Metin, S., Peterson, D.T., Nadimi, S.: Spin-valve effect in soft ferromagnetic sandwiches. J. Magn. Magn. Mater. 93, 101 (1991) 81. Sands, T., Harbison, J.P., Leadbeater, M.L., Allen Jr., S., Hull, G.W., Ramesh, R., Keramidas, V.G.: Epitaxial ferromagnetic τ-MnAl films on GaAs. Appl. Phys. Lett. 57, 2609 (1990)

    844

    D. Wu and X.-F. Jin

    82. Tanaka, M., Harbison, J.P., Deboeck, J., Sands, T., Philips, B., Cheeks, T.L., Keramidas, V.G.: Epitaxial growth of ferromagnetic ultrathin MnGa films with perpendicular magnetization on GaAs. Appl. Phys. Lett. 62, 1565 (1993) 83. Matsui, M., Ido, T., Sato, K., Adachi, K.: Ferromagnetism and antiferromagnetism in Co–Mn alloy. J. Phys. Soc. Jpn. 28, 791 (1970) 84. Rogers, D.J., Maeda, Y., Takei, K.: Compositional separation in Co-Mn magnetic thin films. J. Appl. Phys. 78, 5842 (1995) 85. Thomson, T., Reidi, P.C., Wang, Q., Zabel, H.: 59 Co and 55 Mn NMR of CoMn alloys and multilayers. J. Appl. Phys. 79, 6300 (1996) 86. Menshikov, A.Z., Takzei, G.A., Dorofeev, Y.A., Kazanstev, V.A., Kostyshin, A.K., Sych, I.I.: Magnetic phase diagram of cobalt – manganese alloys. Zh. Eksp. Teor. Fiz. 89, 1269 (1985) [Sov. Phys. JETP 62, 734 (1985)] 87. Wu, D., Liu, G.L., Jing, C., Wu, Y.Z., Loison, D., Dong, G.S., Jin, X.F., Wang, D.-S.: Magnetic structure of Co1-x Mnx alloys. Phys. Rev. B. 63, 214403 (2001) 88. Jing, C., Wu, Y.Z., Yang, Z.X., Dong, G.S., Jin, X.F.: Structure and magnetism of Fe1−x Mnx alloys on GaAs (001). J. Magn. Magn. Mater. 198–199, 270 (1999) 89. Hansen, M.: In: Hansen, M. (ed.) Constitution of Binary Alloys, p. 580. McGraw-Hill, New York (1958) 90. Sumiyama, K., Yoshitabe, T., Nakamura, Y.: XPS valence band and core level spectra of sputter-deposited Fe–Cu, Fe–Ag and Fe–Cu–Ag alloy films. J. Phys. Soc. Jpn. 58, 1725 (1989) 91. Chien, C.L., Liou, S.H., Kofalt, D., Wu, Y., Egami, T., McGuire, T.R.: Magnetic properties of Fex Cu100-x solid solutions. Phys. Rev. B. 33, 3247 (1986) 92. Crespo, P., Hernando, A., Yavari, R., Drbohlav, O., García Escorial, A., Barandiarán, J.M., Orúe, I.: Magnetic behavior of metastable fcc Fe-Cu after thermal treatments. Phys. Rev. B. 48, 7134 (1993) 93. Ambrose, T., Gavrin, A., Chien, C.L.: Magnetic properties of metastable fcc Fe-Cu alloys prepared by high energy ball milling. J. Magn. Magn. Mater. 124, 15 (1993) 94. Harris, V.G., Kemner, K.M., Das, B.N., Koon, N.C., Ehrlich, A.E., Kirkland, J.P., Woicik, J.C., Crespo, P., Hernando, A., Garcia Escorial, A.: Near-neighbor mixing and bond dilation in mechanically alloyed Cu-Fe. Phys. Rev. B. 54, 6929 (1996) 95. Manoharan, S.S., Klaua, M., Shen, J., Barthel, J., Jenniches, H., Kirschner, J.: Artificially ordered Fe-Cu alloy superlattices on Cu (001). I. Studies on the structural and magnetic properties. Phys. Rev. B. 58, 8549 (1998) 96. Uenishi, K., Kobayashi, K.F., Nasu, S., Hatano, H., Ishihara, K.N., Shingu, P.H.: Mechanical alloying in the Fe-Cu system. Z. Metallkd. 83, 132 (1992) 97. Serena, P.A., García, N.: Ferromagnetism in FeCu metastable alloys. Phys. Rev. B. 50, 944 (1994) 98. Wang, J.-T., Zhou, L., Kawazoe, Y., Wang, D.-S.: Ab initio studies on the structural and magnetic properties of FeCu superlattices. Phys. Rev. B. 60, 3025 (1999) 99. Mermin, M.D., Wagner, H.: Absence of ferromagnetism or antiferromagnetism in one-or twodimensional isotropic Heisenberg models. Phys. Rev. Lett. 17, 1133 (1966) 100. Pappas, D.P., Kamper, K.P., Hopster, H.: Reversible transition between perpendicular and in-plane magnetization in ultrathin films. Phys. Rev. Lett. 64, 3179 (1990).; Pappas, D.P., Brundle, C.R., Hopster, H.: Reduction of macroscopic moment in ultrathin Fe films as the magnetic orientation changes. Phys. Rev. B 45, 8169 (1992) 101. Qiu, Z.Q., Pearson, J., Bader, S.D.: Asymmetry of the spin reorientation transition in ultrathin Fe films and wedges grown on Ag (100). Phys. Rev. Lett. 70, 1006 (1993) 102. Wu, Y.Z., Won, C., Scholl, A., Doran, A., Zhao, H.W., Jin, X.F., Qiu, Z.Q.: Magnetic stripe domains in coupled magnetic sandwiches. Phys. Rev. Lett. 93, 117205 (2004).; Won, C., Wu, Y. Z., Choi, J., Kim, W., Scholl, A., Doran, A., Owens, T., Wu, J., Jin, X.F., Qiu, Z.Q.: Magnetic stripe melting at the spin reorientation transition in Fe/Ni/Cu (001). Phys. Rev. B 71, 224429 (2005)

    16 Metallic Magnetic Thin Films

    845

    103. Schulz, B., Baberschke, K.: Crossover from in-plane to perpendicular magnetization in ultrathin Ni/Cu (001) films. Phys. Rev. B. 50, 13467 (1994) 104. Farle, M., Mirwald-Schulz, B., Anisimov, A.N., Platow, W., Baberschke, K.: Higher-order magnetic anisotropies and the nature of the spin-reorientation transition in face-centeredtetragonal Ni (001)/Cu (001). Phys. Rev. B. 55, 3708 (1997) 105. Matthes, F., Seider, M., Schneider, C.M.: Strain-induced magnetic anisotropies in ultrathin epitaxial Nix Pd1-x alloy films. J. Appl. Phys. 91, 8144 (2002) 106. Yu, P., Yin, L.F., Wei, D.H., Tian, C.S., Dong, G.S., Jin, X.F.: Correlation between spin reorientation transition and Curie temperature of Nix Pd1-x alloy on Cu (001). Phys. Rev. B. 79, 212407 (2009) 107. Oepen, H.-P., Knappmann, S., Wulfhekel, W.: Ferro-and para-magnetic properties of ultrathin epitaxial Co/Cu films. ibid. 148, 90 (1995).; Garreau, G., Farle, M., Beaurepaire, E., Baberschke, K.: Curie temperature and morphology in ultrathin Co/W (110) films. Phys. Rev. B 55, 330 (1997) 108. Arnold, C.S., Venus, D.: Simple window-compensation method for improving the signal-tonoise ratio in measurements of the magneto-optic Kerr effect in ultrathin films. Rev. Sci. Instrum. 66, 3280 (1995) 109. Arnold, C.S., Johnston, H.L., Venus, D.: Magnetic susceptibility measurements near the multicritical point of the spin-reorientation transition in ultrathin fcc Fe (111)/2 ML Ni/W (110) films. Phys. Rev. B. 56, 8169 (1997) 110. O’Brien, W.L., Tonner, B.P.: Transition to the perpendicular easy axis of magnetization in Ni ultrathin films found by x-ray magnetic circular dichroism. Phys. Rev. B. 49, 15370 (1994) 111. Farle, M., Platow, W., Anisimov, A.N., Poulopoulos, P., Baberschke, K.: Anomalous reorientation phase transition of the magnetization in fct Ni/Cu (001). Phys. Rev. B. 56, 5100 (1997) 112. Fritzsche, H., Kohlhepp, J., Elmers, H.J., Gradmann, U.: Angular dependence of perpendicular magnetic surface anisotropy and the spin-reorientation transition. Phys. Rev. B. 49, 15665 (1994) 113. Heinrich, B., Monchesky, T., Urban, R.: Role of interfaces in higher order angular terms of magnetic anisotropies: ultrathin film structures. J. Magn. Magn. Mater. 236, 339 (2001)

    Di Wu is a professor in Department of Physics, Nanjing University, China. He received his B.S. (1997) and PhD (2001) at Fudan University, China. He had 6 years postdoctoral research experience in University of Utah and University of California, Riverside. His recent research focuses on spin current transport in magnetic insulators.

    846

    D. Wu and X.-F. Jin Xiaofeng Jin is a Professor in Department of Physics, Fudan University, China. He received his PhD at Fudan University in 1989. He served as the chair of the International Colloquium on Magnetic Films and Surfaces in 2012 and the International Union of Pure and Applied Physics Magnetism Commission C9 in 2014–2017. His research focuses on spin transport in thin film.

    Magnetic Oxides and Other Compounds

    17

    J. M. D. Coey

    Contents Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Rocks, Solid Solutions, and Percolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Principles of Oxide Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Iron Oxides and Hydroxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hematite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Maghemite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wüstite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Iron Hydroxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spinels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Garnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Orthoferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hexagonal Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Other Magnetic Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3d Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Perovskites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pyrochlores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4d and 5d Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4f Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5f Oxides and Related Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Related Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Halides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chalcogenides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pnictides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Silicates and Carbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Silicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Carbonates, Phosphates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    848 851 852 863 864 866 867 867 868 869 869 871 872 873 875 876 878 885 887 888 889 890 890 892 895 897 897 903

    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_17

    847

    848

    J. M. D. Coey

    Oxide Thin Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Substrates, Caps, and Buffers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thin Film Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Oxide Monolayers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxide Heterostructures and Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    903 903 904 909 914 917 917

    Abstract

    Magnetic oxides are important functional magnetic materials, both as permanent magnets and as soft magnetic materials for high-frequency and microwave applications. They have been central in developing our understanding of exchange in strongly correlated insulating and barely metallic systems, including the effects of magnetic frustration and disorder. Oxide thin films are playing an increasing role in spin electronics. This chapter begins with an outline of the natural distribution of magnetic elements, with an initial emphasis on naturally occurring oxide minerals, especially oxides, hydroxides, and silicates of iron and other magnetic cations. Principles of 3d oxide magnetism and crystal chemistry are introduced. Structure and intrinsic magnetic properties of the main structural families of binary and ternary oxides and related compounds, including halides, chalcogenides, and pnictides, are then presented in tabular form. Magnetism of 4d, 5d, and 5f oxides is also covered. Then methods of preparing oxide thin films are presented, and specific features of single oxide films, interfaces, and heterostructures are discussed.

    Background Terrestrial minerals are the source of the magnetic elements needed to make all the alloys and compounds of interest in this handbook. The ores are culled from the Earth’s surface or the topmost kilometer of the crust, where low-density compounds of oxygen and other electronegative elements segregate, while denser siderophilic materials tend to sink deep into the core, which is composed of 90% iron with a little nickel – a composition similar to that of Ni-Fe meteorites. In fact, iron is the most abundant element by weight in the whole Earth (35 wt%), but in the crust it occupies fourth position (6 wt%), after oxygen, silicon, and aluminum, or sixth position, after hydrogen and sodium if its concentration is expressed in atom percent (2.1 at%). Crustal abundances in at% are shown in Fig. 1a). Rock-forming minerals are therefore predominantly aluminosilicates, but two iron oxides, magnetite and hematite, feature among them. These were the first magnetic materials known to man, and they remain our primary source of iron, which provided the basis of technology and weaponry for over 2000 years.

    17 Magnetic Oxides and Other Compounds

    849

    a) Fe

    Log (abundance, ppm)

    5 4 Mn

    3 2

    Cr

    Ni Co

    1

    Ce

    Nd Pr

    Sm

    Gd Eu

    0

    Dy Tb

    Pm

    Er

    Yb

    Ho Tm

    -1

    b)

    3d elements

    4f elements

    Fig. 1 (a) Atomic abundances of the principal elements in the Earth’s crust and (b) abundances of magnetic 3d and 4f elements, plotted on a logarithmic scale

    Almost 1 atom in 40 in the crust is iron, and compared with other magnetic elements, the crustal abundance of iron is 40 times that of all the others put together. This 40/40 rule affords some perspective on the magnetism of natural oxides. Abundances of the magnetic elements are compared on a log scale in Fig. 1b), where it can be seen that there is about three orders of magnitude more Fe in the Earth’s crust than either of the other ferromagnetic 3d elements Co and Ni (Table 1). The magnetic elements, which have ions with unpaired electrons, are found predominantly among the 3d and 4f series. The light rare earths are not particularly rare – neodymium is as abundant as cobalt or zinc, but the abundances follow an odd/even variation with atomic number that reflects the stability of the atomic nuclei, superposed on a decline with increasing atomic number. The heavier 4f elements become increasingly uncommon. Crustal abundance is a rough proxy

    850

    J. M. D. Coey

    Table 1 Elements in the Earth’s crust (atomic %)

    Ion O2− Si4+ Al3+ H+ Na+ Fe2+/3+ Ca2+ Mg2+ K+

    1.0 0.8

    Tc/Tc (x= 1)

    0.4 0.2

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    0.0 0.0

    o o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    Antiferroo o magnetic pair o o

    0.2

    0.4

    x

    0.6

    0.8

    1.0

    o

    o

    o

    xp

    Configuration 2p6 2p6 2p6 – 2p6 3d6/5 3p6 2p6 3p6

    o

    o

    0.6

    Abundance 59.7 20.9 6.1 2.9 2.2 2.1 2.1 2.0 1.1

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    o

    Cluster

    Isolated o

    o

    Fig. 2 (a) Schematic variation of the magnetic ordering temperature as a function of the concentration x of magnetic ions in an oxide solid solution of magnetic and nonmagnetic cations M and N, (Mx N1-x )Oη , where η is a small integer or rational fraction. The magnetic interactions couple nearest-neighbor magnetic cations, and xp is the percolation threshold. The distribution of ions illustrated in (b) well below the percolation threshold

    for cost. The periodic table in Table 2 provides an indication of the costs of the elements in five broad bands. It is truly serendipitous for magnetism that the most abundant and cheapest magnetic element turns out to be the one that is ferromagnetic with the highest magnetization and the one that forms 3d cations with the greatest magnetic moments. The natural magnetic materials that feature in areas such as rock magnetism or biomineralization are iron oxides. Oxides exhibit a wide range of electric, magnetic, optical, and structural properties. They can be insulating, semiconducting, metallic, ferroelectric, piezoelectric, ferromagnetic, half-metallic, ferrimagnetic, antiferromagnetic, or superconducting. From a magnetic point of view, the dominant interaction in transition-metal oxides is usually antiferromagnetic superexchange, the coupling of the spins of two 3d cations separated by an oxygen anion, which leads to antiferromagnetic or ferrimagnetic order. Ferromagnetism is less common in oxides, where it is often associated with cation mixed-valence and orbital degeneracy, than it is in metals. See  Chap. 2, “Magnetic Exchange Interactions.”

    17 Magnetic Oxides and Other Compounds

    851

    Table 2 Periodic table of the elements, showing their approximate costs per kilogram, after [19]

    Rocks, Solid Solutions, and Percolation Oxide structures have great scope for accommodating cations of similar charge and ionic radius, and their properties can be considerably modified thereby. In some cases, there is an extended range or even complete solid solubility between chemically and structurally similar end-members. Pure, single-phase end-members are not the norm in nature. The relative abundance of Al and Fe in the crust (Fig. 1) means that iron (and other magnetic cations) are relatively sparse in aluminosilicate rocks, which are generally multiphase aggregates that reflect the temperature, pressure, and chemical environment in which the rock was formed. It may be possible to separate out one of the phases for examination, by physical or magnetic means. Otherwise, the whole multiphase rock or soil can be analyzed microscopically in order to characterize the natural mixture of minerals present. Iron is often an impurity in aluminosilicate minerals, replacing aluminum if trivalent or an alkaline earth cation if divalent. Iron-rich end-members are the exception. When the iron content in a solid solution is low, superexchange interactions among ferrous or ferric cations separated by a single oxygen anion are inhibited, and long-range magnetic order is impossible. The magnetically dilute minerals then exhibit Curie-Weiss paramagnetism. Percolation becomes possible when the iron concentration x on appropriate sites exceeds the percolation threshold

    852

    J. M. D. Coey

    xp , which depends on the crystal structure. A rough estimate of xp is 2/Z, where Z is the nearest-neighbor cation coordination number – 8 if the cations form a bcc lattice, 12 for an fcc lattice, 4 for a square lattice, etc. Magnetic order appears when x > xp , and the transition temperature increases with x, although the transition itself is broadened by the disorder (Fig. 2). Natural iron-rich, end-member oxides tend to be contaminated by aluminum, manganese, titanium, chromium, or other ions, which can modify the magnetic properties significantly; common examples are Al in hematite and Ti in magnetite. When discussing natural oxides here, we will often cite ideal chemical formulae, which differ from the more complex chemical composition of real natural samples. But many of these minerals are of interest in their own right; naturally occurring magnetic oxides have found important applications in paleomagnetism or archaeomagnetism, for example, and there are interesting natural examples of lowdimensional or mixed-valence magnetic materials. Nature, with its access to high pressures and geological timescales, provides us with a rich store of natural crystals that we are unable to replicate in the laboratory. Tetrataenite, the tetragonal, ordered L10 -structure FeNi alloy found in nickel-iron meteorites that have equilibriatesd over a huge timespan, is an example. Phase purity should not be a problem in synthetic oxides, but deliberate cation substitution can be a powerful tool to modify exchange interactions, conductivity, and sublattice magnetization in order to achieve optimal magnetic properties or to investigate how magnetism depends on composition.

    Principles of Oxide Magnetism Oxides are ionic compounds where small, highly charged metal cations are embedded in a lattice of larger oxygen anions that occupy 90% of the volume. Bonding is predominantly ionic. Oxygen accepts two electrons to form the divalent O2− anion with the particularly stable 2p6 closed shell, a configuration that is shared by 90% of all the ions in the Earth’s crust. The extra electrons are transferred to oxygen from the metals that form the positively charged cations; those listed in Table 2 are the ones most commonly encountered in oxides and other magnetic compounds. These cations are all smaller than the oxygen anion, for which the ionic radius is taken to be 140 pm. Other anions encountered in magnetic compounds include OH− (110 pm), F− (133 pm), Cl− (181 pm), S2− (184 pm), and N3− (146 pm). The ionic picture is, of course, oversimplified. The chemical bond between metal and oxygen has partionic and part-covalent character, governed by the electronegativity of the metal partner. Covalency is more pronounced in sites with a low coordination number and for cations like Si4+ or V4+ with a high formal charge state, as well as for anions such as S2− and N3− that are much less electronegative than oxygen or fluorine (Table 3). The crystal structures of most oxides may be regarded as a packing of a lattice of large oxygen anions held together by electrostatic interactions with the small metal cations in interstitial sites. The structures sometimes incorporate large cations that

    17 Magnetic Oxides and Other Compounds

    853

    Table 3 Ionic radii of cations in oxides. Values in brackets refer to a low-spin state Fourfold tetrahedral Mg2+ Zn2+ Al3+ Fe3+ 3d5 Si4+

    pm 57 60 39 49 26

    Sixfold octahedral V4+ 3d1 Cr4+ 3d2 Mn4+ 3d3 Mn2+ 3d5 Fe2+ 3d6 Co2+ 3d7 Ni2+ 3d8

    pm 58 55 53 83 78 (61) 75 (65) 69 (56)

    Sixfold octahedral Ti3+ 3d1 V3+ 3d2 Cr3+ 3d3 Mn3+ 3d4 Fe3+ 3d5 Co3+ 3d6 Ni3+ 3d7

    pm 67 64 62 65 65 61 (55) 60 (56)

    Eightfold cubic Ca2+ Sr2+ Ba2+ Pb2+ Y3+ La3+ Gd3+ 4f7

    pm 112 126 142 129 102 116 105

    substitute on oxygen sites. The packing may be body-centered cubic (perovskites), face-centered cubic (spinels, garnets), or hexagonal close-packed (hexaferrites, corundums). Interstices in these packings are normally of two types – tetrahedral and octahedral – which explains the emphasis on these two sites permeates any discussion of oxide magnetism. They are illustrated for the ideal face-centered cubic structure in Fig. 3, where both interstitial sites in the close-packed lattice have cubic symmetry. The radius of an ion that will just fit into a close-packed tetrahedral √ interstice is rtet = ( (3/2)−1)rO = 29 pm where rO = 140 pm is the ionic radius of O2− . The close-packed octahedral interstice is larger; it can just accommodate an √ ion of radius roct = ( 2 – 1)rO = 52 pm. Nevertheless, the magnetic cations listed in Table 2 are too big to fit into tetrahedral sites and even a bit too big to snugly occupy octahedral sites without distorting the oxygen lattice. The distortion may be uniform, retaining the cubic site symmetry, or uniaxial along a threefold or fourfold axis of the tetrahedron or octahedron. The 4f cations are larger, and they are eight- to twelve- fold coordinated in oxides. Rare-earth cations are usually trivalent, and their ionic radii in eightfold cubic coordination decrease from 116 pm for La3+ to 98 pm for Lu3+ . Eu2+ and Yb2+ are exceptions; they have a stable half-filled (4f7 ) or filled (4f14 ) 4f shells, with radii of 125 pm and 114 pm, respectively. The electrostatic coulomb interaction among the 3d electrons of a magnetic cation and between them and the negatively charged 2p6 shells of their oxygen anion neighbors are the fundamental interactions in oxide magnetism. A critical first step is to replace the spherically symmetric d wave function of the free ions with linear combinations that reflect the local crystal symmetry. A drastic simplification is the crystal field model, where we replace the oxygen anions by -2e point charges at the anion sites and consider the energy of the electrons in the unfilled 3d shell to see how the crystal field created by the O2− neighbors influences both the electronic structure and magnetic properties. The one-electron model [13], which ignores the on-site 3d-3d Coulomb interactions, is a fair approximation for d1 , d4 , d6 , and d9 ions when Hund’s first rule applies because these ions have a single electron or hole outside an empty (d0 ), half-filled (d5 ), or filled (d10 ) d-shell.

    854

    J. M. D. Coey

    Fig. 3 Tetrahedral and octahedral sites in a cubic-close-packed oxygen lattice

    When the tetrahedral or octahedral site is undistorted, the free ion eigenfunctions ψ0 , ψ±1 , and ψ±2 , with subscripts denoting the orbital magnetic quantum number ml , are replaced by suitable linear combinations that reflect the cubic site symmetry. They are:  √  ψxy = −i/ 2 (ψ2 − ψ−2 )  √  ψyz = i/ 2 (ψ1 + ψ−1 )  √  ψzx = −1/ 2 (ψ1 − ψ−1 )  √  ψx2−y2 = 1/ 2 (ψ2 + ψ−2 )

    (1)

    ψz2 = ψ0 Figure 4 compares the electron density for this basis set of d orbitals in an octahedral environment with that of the d electrons in a spherical potential. On inspecting the disposition of the oxygen electrons in the surrounding octahedron, it is obvious that the xy, yz, and zx orbitals are degenerate – they are labeled t2g orbitals. It is less obvious that the x2 – y2 and z2 orbitals, labeled eg , are degenerate, but they are clearly higher in energy because the 3d electron density is maximum near the negatively charged anions. The crystal field splitting in octahedral coordination (where the suffix g is included to indicate the presence of a center of symmetry) is illustrated in Fig. 5.

    17 Magnetic Oxides and Other Compounds

    855

    Fig. 4 The 3d orbitals for electrons in a free ion (top row) and the orbitals for the ion in a cubic crystal field (bottom row). Oxygen positions in a surrounding octahedron are marked by red dots that represent -2e point charges Fig. 5 Splitting of the t2g and eg orbitals in an octahedral environment. cf is the crystal field splitting calculated in a point charge model. The extra ligand field splitting due to the hybridization of 3d and sp orbitals gives the total splitting 

    The splitting of the one-electron levels is only partly explained by the electrostatic potential created by the oxygen anions; about half of it should be attributed to covalency associated with the different overlap of the t2g and eg orbitals with the lower-lying 2p oxygen orbitals, which introduces a greater bonding-antibonding splitting for the 2p-eg σ-bonds than the 2p-t2g π-bonds. This is the ligand field effect. The combined ionic and covalent effects result in a total splitting , which is about 1 eV in octahedral sites [5]. The crystal field splitting is reversed in pure cubic coordination, where the eg orbitals are lowest. The splitting is smaller and also reversed in tetrahedral sites, with the e orbitals lower and the t2 orbitals higher. The suffix g is dropped because a tetrahedron has no center of symmetry (Fig. 6). The influence of a distortion of the cubic site is to lift the degeneracy of the one-electron energy levels, as shown in Fig. 7. The splitting preserves the center of gravity of each set of orbitals. Whenever a set of orbitals is not fully occupied, the crystal field may lower the energy of the ion by an amount known as the crystal field stabilization energy cfse , which can dictate the site preference of ions in structures where there is a choice of

    856

    J. M. D. Coey

    Fig. 6 Energy splitting of the one-electron states in cation sites with octahedral, tetrahedral, and cubic coordination due to the crystalline electric field

    octahedral or tetrahedral site occupancy. For example, the energy of Cr3+ (3d3 ) in octahedral sites is 3 × (−2/5) O = −1.2 O , whereas for Fe3+ (3d5 ) it is zero. Cr3+ therefore has a strong octahedral site preference. Furthermore, a distortion of the environment can provide an additional stabilization. These considerations can outweigh purely steric effects. The Jahn-Teller effect is the tendency of some ions to spontaneously distort their local environment in order to gain some extra crystal field stabilization energy. A 3d3 ion in an octahedral site gains nothing from a distortion of the oxygen octahedron because of the center of gravity rule, but a 3d4 ion will tend to induce a local tetragonal deformation which splits the eg levels, lowering the energy of either the occupied dz2 or dx2–y2 orbital (Fig. 7). The Jahn-Teller effect is strongest for d4 and d9 ions in octahedral coordination and d1 and d6 ions in tetrahedral coordination. Two other important electronic effects encountered in 3d oxides are charge order and orbital order [62]. Here interionic coulomb interactions play a role. Charge

    17 Magnetic Oxides and Other Compounds z

    857

    z

    x

    x

    x

    y x

    x

    z

    x

    z

    z

    αt Δ cfse Δ0

    δt

    Fig. 7 Influence of a tetragonal elongation or flattening of an octahedral site on the splitting of the one-electron energy levels. The crystal field stabilization energy cfse is indicated for a 3d4 ion such as Mn3+ with a t2g 3 eg (x2 − y2 ) configuration

    order arises when a particular lattice site can be occupied by an ion in one of two different charge states. Examples are Mn3+ and Mn4+ or Fe2+ and Fe3+ . The average 3d occupancy is nonintegral, but the electrons may settle in an ordered pattern on the cation sites in a charge-ordered state with some lattice distortion. Otherwise, the surplus 3d electrons rapidly hop or tunnel among the d3 or d5 ion cores, a phenomenon known as mixed valence that occurs, for example, in mixed-valence manganites such as (La0.7 Sr0.3 ){Mn3+ 0.7 Mn4+ 0.3 }O3 or magnetite [Fe3+ ]{Fe2+ ,Fe3+ }O4 , where the brackets denote different crystallographic sites. The average 3d occupancy on octahedral sites {} in these examples is 3.7 or 5.5, respectively. Orbital order arises when the 3d occupancy is integral and single-valued, but the electron occupies a degenerate orbital. An example is Mn3+ (t2g 3 eg 1 ). A lattice distortion such as tetragonal compression and expansion on alternate sites can lead to alternate occupancy of dz2 and dx2-y2 orbitals, in a variant of the Jahn-Teller effect. The ordered electronic state generally reverts to a disordered state at a hightemperature phase transition, where the entropy of disorder Rln2 per mole that is released is comparable to the entropy Rln(2S + 1) released in the vicinity of a magnetic order-disorder phase transition.

    858

    J. M. D. Coey

    Coulomb interactions among electrons in the single electron states are critical, both within and between cations. Although the one-electron picture is rather easy to grasp and relate to energy-band calculations, it does no justice to these manyelectron interactions in 3d ions, especially for high-spin 3d ions with 2, 3, 7, or 8 electrons, which are in an F-state. Ions with five electrons are in an S state, which corresponds to a half-filled 3d shell with spherical symmetry, while those with one, four, six, or nine electrons are in an D-state, which maps onto a t2 or e one-electron state. Here S, D, and F refer to the values 0, 2, or 3 for the total orbital angular momentum L of the free ion allowed by Hund’s second rule, discussed below. The excited states of the 3dn shell which are probed in optical transitions have been calculated from crystal field theory, and they are represented on Tanabe-Sugano diagrams [101, 108]. In magnetism, it is usually sufficient to consider only the ground state. The intra-atomic interactions give rise to Hund’s rules for the spin and orbital moments of the many-electron ions. The first rule gives the net spin S, which is the maximum value, consistent with the Pauli principle that no two electrons can occupy the same quantum state with the same spin. It is simply obtained by populating the lowest orbitals with the available electrons. The interaction is represented by exchange coupling JH of a pair of electrons in different orbitals with the same spin, written in terms of an atomic Hamiltonian 

    HH = −JH α=β

    1 + 2s iα .s iβ 2

     (2)

    where typical values of the intra-atomic exchange JH are 0.8–0.9 eV for 3d electrons and 0.6–0.7 eV for 4d electrons. Values for the more compact 4f shell are higher. Here, s is the spin of an electron in a particular orbital at site i and the sum is over all electron pairs on the site. Two examples, for d4 and d6 , are shown in Fig. 8a and b. The energy contribution to (2) is JH for a ↑↑ pair and zero for a ↑↓ pair, so we have -6JH for d4 and -10JH for d6 . Hund’s first rule implies that the five ↓ orbitals lie above the five ↑ orbitals, which is normally the case unless the crystal/ligand field splitting is large or the on-site exchange is weak when it is possible to have a lowspin state, where in octahedral coordination, for example, the t2g ↓ orbitals are filled before the eg ↑ orbitals and Hund’s first rule no longer applies. This is illustrated in Fig. 8c and d, where the total spin moments are, respectively, S = 1 and S = 0. Low-spin states in oxides are sometimes found for Co3+ in octahedral coordination, which has S = 2 (t2g 3↑ eg 2↑ t2g 1↓ ) in the high-spin state and S = 0 (t2g 3↑ t2g 3↓ ) in the low-spin state. They are more common in more covalent compounds, where the ligand field is larger. Hund’s second rule maximizes the orbital angular momentum consistent with the value of S for the ion. The possibilities in Table 4 are a D-state with L = 2(d1 , d4 , d6 , d9 ), an F-state with L = 3(d2 , d3 , d7 , d8 ), or an S-state with L = 0 (d5 ). The

    17 Magnetic Oxides and Other Compounds

    859

    Fig. 8 Electronic configurations of an ion in octahedral coordination having four or six electrons. In (a) and (b), the usual high-spin states are illustrated, whereas (c) and (d) show the low-spin states that arise when the crystal field splitting cf exceeds the intra-ionic Hund’s rule exchange splitting H Table 4 Ground state terms for 3d and 4f ions, 2S + 1 LJ deduced from Hund’s rules 3d1 2D 3/2

    3d2 3F 2

    3d3 4F 3/2

    4f1

    4f2

    4f3

    2F 5/2

    3H 4

    4I 9/2

    4f4 5I 4

    3d4 5D 0

    3d5 6S 5/2

    4f5 4f6 6H 7F 5/2 0

    3d6 5D 4

    4f7 4f8 8S 7F 7/2 6

    3d7 4F 9/2

    4f9 4f10 6H 5 15/2 I8

    3d8 3F 4

    3d9 2D 5/2

    4f12 4f12 4I 3H 15/2 6

    4f13 2F 7/2

    nomenclature here is based on free ions, but it is retained for ions in solids, even though the orbital moment is almost completely quenched by the crystal field. The third rule invokes spin-orbit coupling, a relativistic interaction of great interest in magnetism, represented by the Hamiltonian:

    Hso = L.S

    (3)

    where  is the ionic spin-orbit coupling constant. In 3d ions with quenched orbital moments, the effect of the spin-orbit interaction is to restore a small orbital contribution, thereby creating magnetocrystalline anisotropy. Octahedral Co3+ is a good example. The crystal field in 4f ions is small in comparison with the spin-orbit interaction, so the orbital moment is unquenched and L and S couple together to form the total angular momentum J = L ± S, just as they do in free ions. We now consider briefly the intra-atomic coulomb correlations. Unlike the delocalized, nearly free electrons in broad energy bands in semiconductors and many metals, the 3d electrons in oxides are usually localized, with an integral number

    860

    J. M. D. Coey

    4s(T) conduction band 3dn±1 3dn

    U

    W

    3dn+1+ L 2p6(O) valence band

    3dn

    U

    W

    Fig. 9 Schematic density of states for a 3d oxide, showing the charge-transfer process that leads to a Mott insulator in an early transition-metal oxide and a charge-transfer insulator in a late transition-metal oxide. The oxide is insulating if the excited state cannot be accommodated within the bandwidth

    of them associated with each cation site. The 3d levels, illustrated schematically in Fig. 9, lie in the gap between a broad, full 2p(O) valence band and a broad, empty 4s(T) conduction band, where T is the 3d transition metal. The 3d levels lie near the top of the bandgap for cations like Ti3+ or V3+ at the beginning of the series and move toward the bottom of the gap for cations like Ni2+ or Cu2+ at the end. The 3d orbitals inevitably overlap to some extent, and in the tight-binding approximation with interatomic hopping integral t, this means that they form a narrow band of width W = 2Zt, where Z is the coordination number. The bandwidth decreases as the nuclear charge increases across the 3d series. Elementary band theory unphysically predicts that any partially filled band, however narrow, will always be conducting, because the theory neglects electronic correlations. Oxides are not usually metals, despite a 3d bandwidth of 2 eV or more, for reasons explained by Mott [75]. In a metal, the average 3dn configuration has to coexist with fluctuating 3dn ± 1 configurations in order for conduction to take place. The energy cost of transferring an electron from one site to the next is the coulomb energy Udd , which is the difference between the ionization energy and the electron affinity of a 3dn ion. When this exceeds W, Udd /W > 1, conduction cannot occur, and the oxide is known as a Mott insulator (or a Mott-Hubbard insulator). The d-d coulomb correlations transform what would otherwise be a metal into an insulator, because there is no place in the band to accommodate the correlation energy. For the heavier 3d ions, an easier charge-transfer process is from a filled oxygen 2p6 shell to an adjacent metal cation, creating a 2p5 configuration with a ligand hole and a 3dn + 1 cation configuration with an excess electron. This process costs an energy Upd , and when Udd > Upd > W, the oxide is known as a charge-transfer insulator.

    17 Magnetic Oxides and Other Compounds

    861

    The physics of Mott insulators is captured by the Hubbard’s deceptively simplelooking Hamiltonian for an s band. It is a favorite toy of condensed-matter theorists.

    H = −t,α c† iσ cj σ + U j ni↑ ni↓

    (4)

    Here the first term transfers an electron from site i to a neighboring site j, without flipping the spin. The second term is the on-site coulomb repulsion between two electrons of opposite spin on the same site. The nearest-neighbor interaction is antiferromagnetic because the wave function cannot spread out to neighboring sites, unless the spins are antiparallel. The coupling energy between the spins on adjacent sites is of order t2 /U. Modern electronic structure calculations, especially those based on density functional theory, have been good at predicting the magnetization and magnetic ordering temperature of metals, but they are less successful with oxides and other insulators, where the coulomb interaction parameter U is often has to be chosen to match the observed bandgap. Numerical calculations dispense with the simplified models that have helped us to develop an intuitive physical understanding of the many-electron ground state in terms of a handful of physical parameters that quantify the exchange, spin-orbit, and crystal field interactions that really call the shots in this area of magnetism. Oxides are rarely perfectly stoichiometric, so exactly integral d-band occupancy is unlikely. There will inevitably be spatial charge fluctuations associated with atomic defects, but they do not normally spoil the insulating state. Local lattice deformations arise because the 3d electrons or holes interact strongly with their ionic environment and create lattice polarons with a large effective mass that are effectively immobile. Furthermore, if an electron should hop to an unoccupied neighboring site in a 2D or 3D antiferromagnetic array, it leaves a hole behind and flips a spin with each subsequent hop, creating a ferromagnetic interaction while leaving a trail of magnetic defects behind it. The energy can be minimized by the electron returning to its starting point. A small excess of electrons or holes in the half-filled Hubbard band does not therefore make the model conducting. Direct overlap of the 3d orbitals of cations on neighboring sites to form the 3d band is relatively small, but the mixing of 3d states with the 2p states of neighboring oxygen anions is more significant. In these circumstances, the exchange interactions coupling the spins of next-neighbor cations are indirect, superexchange interactions. These were systematically investigated by Goodenough and Kanamori [50] in the 1960s. Their rules were simplified by Anderson [1] and the interactions are discussed in  Chap. 2, “Magnetic Exchange Interactions.” The coupling is usually antiferromagnetic, except if the M-O-M bond angle is close to 90◦ , when weak ferromagnetic interactions can occur. The antiferromagnetic interactions lead to either antiferromagnetic or ferrimagnetic order, depending on the lattice structure, but they may be frustrated by the lattice topology.

    862

    J. M. D. Coey

    In mixed-valence compounds, where the 3d electrons are delocalized, the electron hopping with spin memory leads to the ferromagnetic double exchange interaction, first described by Zener [116]. Both superexchange and double exchange are described by the Heisenberg Hamiltonian:

    H = −2JS i .S j

    (5)

    where a positive or negative sign of the exchange parameter J corresponds to a ferromagnetic or an antiferromagnetic intersite exchange interaction. Si is the total spin localized on site i. Exchange interactions with different symmetry are usually much weaker. The Dzyaloshinskii-Moriya (D-M) interaction, for example, which has the form 

    HDM = −D. S i × S j

    

    (6)

    where the vector D must lie along a uniaxial symmetry axis, was discovered in response to the problem of canted antiferromagnetism in hematite [74] and some other antiferromagnetic insulators. Finally, Fig. 10 provides an overview of the magnetic ordering temperatures in oxides. The histogram illustrates the distribution of Curie and Néel temperatures for

    Fig. 10 Magnetic transition temperatures for oxides. Ferromagnets are shown in blue, ferrimagnets in pink, and antiferromagnets in red. Data are based on [33]

    17 Magnetic Oxides and Other Compounds

    863

    a thousand magnetically ordered oxides. None of them exceeds 1000 K. Oxides of ions Fe3+ or Mn2+ with a 3d5 configuration and the maximum value S = 5/2 tend to order at the highest temperatures. Exchange between rare-earth ions in oxides is weak, on account of the limited spatial extent of the 4f wave functions and the large value of the on-site coulomb interaction U ∼ 6 eV. Hence t2 /U is small. The next five sections present the different families of magnetic oxides and related compounds. We use the structure types of natural minerals as a frame of reference to discuss a wide range of synthetic materials that are never found in the natural environment.

    Iron Oxides and Hydroxides Iron oxides and hydroxides constitute the most important group of magnetic minerals [35]. Most important are the famous magnetic rock-forming oxides magnetite and hematite. Then there is a series of crystalline or partly amorphous iron oxyhydroxides and hydroxides that are found in soils and sediments. A summary of the structural and intrinsic magnetic properties of the iron end-members of whole group is given in Table 5; substitution of other magnetic or nonmagnetic cations (especially Al3+ ) in the iron end-members will modify these values. Variability is ubiquitous in natural specimens; no two examples of a mineral with the same name are ever exactly the same. The crystal structures of the main binary iron oxides and hydroxides are illustrated in Fig. 11. We first present the structural and magnetic properties of each in turn and then discuss structurally related binary oxides with other magnetic cations. More complex, ternary oxides are treated in a following section.

    Table 5 Magnetic properties of iron oxides and hydroxides Mineral Wüstite Magnetite Hematite Maghemite Goethite Akaganénite

    Ideal formula FeO Fe3 O4 αFe2 O3 γFe2 O3 αFeOOH βFeO(OH,cl)

    Space group Fm3m Fd3m R3c P41 Pbnm I2/m

    Lepidocrocite Feroxhyte

    γFeOOH δFeOOH

    Cm21 P3ml

    Ferrihydrite

    5Fe2 O3. 9H2 O

    P63 mc

    Ferric gel Amakinite

    Fe(OH)3 nH2 O (FeMg)OH2

    – p3m1

    Lattice parameters 431 840 504; 1377 835; 2499 996; 302; 461 1056; 303; 1048 90.0◦ 308; 1250; 387 293; 460

    TC , TN (K) 198 853 980 950 360-405 295

    600; 910 poorly crystallized Amorphous 692; 1452

    70 455

    Order AF FI wF FI AF AF

    σ (Am2 kg−1 ) 300 >650 300 610 650 >350 >300 550 280–300 >300 >750 >600 900 >400 550

    to unreasonable moments per 3d ion (Table 33). The magnetism [62] (when it is not attributable to experimental artifacts or impurity phases such as embedded nanoclusters of metallic cobalt) is closely associated with lattice defects in the material. Furthermore, atomic-scale probes such as XMCD [102] and Mössbauer spectroscopy [28] failed to reveal any magnetic order of the 3d dopants at room temperature. In dilute Co-doped ZnO films, for example, no evidence could be found for ferromagnetic interactions among the dopants [76], and even in concentrated films with x = 0.6, the antiferromagnetic ordering temperature does not exceed 20 K [77]. What is remarkable about the magnetization curves is that they are essentially anhysteretic and exhibit little or no temperature dependence between 4 K and 400 K. Curves obtained at different temperatures often superpose after correcting the data for a diamagnetic or paramagnetic slope arising from the substrate or from the oxide film itself. The lack of temperature dependence excludes any possibility of superparamagnetism, for which the magnetization curves should superpose when M/Ms is plotted versus (H/T). The absence of hysteresis is a further argument against uniform magnetization, as would be expected for a DMS. The small net magnetization values Ms ≈ 10 kA m−1 inferred for thin films should lead to an appreciable anisotropy field, with a uniaxial anisotropy constant due to intrinsic crystal field or substrate-induced strain, and therefore produce coercivity at low temperature. Another serious problem is the high value of the Curie temperature. Relatively few insulating oxides are ferromagnetic (Fig. 10); they usually order antiferromagnetically or ferrimagnetically on account of the antiferromagnetic superexchange. Almost none order magnetically above 1000 K. It is unreasonable to expect oxide films with just 5% of magnetic ions to have a Curie temperature exceeding 100 K. The magnetic energy density varies as x

    914

    J. M. D. Coey

    √ or x, so high Curie temperatures could not be achieved with uniformly magnetized material. A defect-related origin rather than a magnetic dopant-related origin of the ferromagnetism is supported by its observation in some samples of undoped material, including films of HfO2 , ZnO, TiO2 , and In2 O3 . This phenomenon is known as d0 magnetism [20], and it reinforces the idea that while defects are necessary, 3d dopants are not essential. It is a challenge to associate the magnetism, which is often not quantitatively reproducible, with a specific surface, interface, or grain-boundary defects. Extended defects may be a more likely source of percolating magnetism than point defects. As regards the origin of d-zero magnetism two quite different ideas are in contention. One is to model the defects, probably oxygen vacancies, in terms of a narrow, spin-split defect-related impurity band. This can lead to a high Curie temperature since spin-wave excitations are suppressed in the half-metallic band. It is a Stoner model of the magnetism, involving a small fraction of the sample volume. The transition-metal dopants may nonetheless play a role when they exist in different charge states, providing a reservoir of charge for the impurity band to ensure filling to the point where the Stoner criterion is satisfied. This model is known as chargetransfer ferromagnetism [28]. An alternative view is that there is actually no spin ferromagnetism in these oxide films. The reversible nonlinear response to the magnetic field is really an unusual form of giant orbital paramagnetism, which has been related to the coherent response of quasi-two-dimensional electron systems to fluctuations of the vacuum electromagnetic field [32]. The definitive explanation remains an open question [20].

    Oxide Heterostructures and Interfaces Ultrathin nonmagnetic oxide barriers play an important role in spin electronics [7], which has become the most innovative area of applied magnetism in the twentyfirst century [113]. Early tunnel barriers were made of amorphous aluminum oxide, but the discovery that crystalline (100) MgO layers act as directional spin filters when grown on a (100) film of αFe, or a body-centered cubic Fe-Co alloy [82, 115], led to magnetic tunnel junctions (Fig. 28b) with a magnetoresistance in excess of 200%, which made them attractive both as magnetic sensors and memory or logic elements. Other nonmagnetic oxide tunnel barriers, such as spinel MgAl2 O4 , have been developed, but none rivals the efficiency of MgO. Thin films of magnetic oxides and sulfides such as ferromagnetic EuO or EuS and ferrimagnetic CoFe2 O4 or NiFe2 O4 can serve as tunnel spin filters with a nonmagnetic metal electrode, since the bandgap is different for minority- and majority-spin electrons [41]. Spin electronics [113] has progressed from its initial emphasis on magnetoresistive sensors for magnetic recording [119] to the development of a competitive nonvolatile magnetic memory and logic technology based on switching magnetic

    17 Magnetic Oxides and Other Compounds

    915

    tunnel junctions. At first the switching was based on current-generated magnetic fields, but that approach proved not to be scalable. Alternative approaches were then discovered using the angular momentum in a spin-polarized current to switch the bistable magnetic element directly (spin transfer torque) or indirectly by spin injection from a spin current in an adjacent heavy metal layer such as Ta, W, or Pt, making use of the spin Hall effect (spin-orbit torque). Magnetic oxides have so far played little role here, the best ferromagnetic layers being iron, cobalt, and their alloys. A long-term ambition has been to devise a form of oxide spin electronics, where memory or logic operations can be conducted in all-oxide thin film stacks grown on large-scale silicon wafers [8]. An attractive prospect is the use of electric fields (or strain) rather than electric currents to control magnetic switching in stacks with adjacent ferromagnetic (or ferrimagnetic) and ferroelectric (or piezoelectric) layers, thereby reducing the rapidly escalating energy cost of the big data revolution. Growth of oxide heterostructures is facilitated by the structure of the basic oxide scaffold, which is often close-packed, or else has a perovskite-type structure. The choice of conducting ferromagnetic oxides is quite limited: CrO2 , (La0.7 Sr0.3 )MnO3 , Fe3 O4 , and some double perovskites order magnetically well above room temperature, but a drawback of any oxide magnet is that its magnetization is less than a third that of iron. This is unavoidable because most of the unit cell volume is occupied by the large nonmagnetic O2− anions, with high-spin ferric iron Fe3+ or other magnetic ions in the interstices. Ferrimagnetic structures reduce the magnetization further. Interfacial dead layers tend to degrade the performance of the magnetic oxide layers, and the room-temperature candidate ferro- or ferrimagnetic oxides do not perform nearly as well as their metallic counterparts. However, SrRuO3 has played a part in low-temperature proof-of-principle perovskite-structure stacks when a good spin-polarized ferromagnetic metal was required. An advantage of all-oxide stacks is that ferroelectric layers such as BiTiO3 or BiFeO3 , a rare room-temperature multiferroic, can be readily incorporated into the stacks. The cycloidal character of the weak ferromagnetism is suppressed in thin films of BiFeO3 , and it then exhibits a large linear magnetoelectric effect, of order 2 nsm−1 . There is a prospect of using magnetoelectric switching combined with spin-orbit detection of state in new ultralow energy nonvolatile logic and memory beyond CMOS technology [69].

    Oxide Interfaces At interfaces between oxides, the three-dimensional symmetry of each oxide is broken, and proximity effects may emerge [30]. Interfacial exchange can modify the magnetic structure at the interface, if one or both of the oxides is magnetically ordered. Exchange bias, for example, may be a consequence. When the electronic structure of the two oxides differs, charge transfer may result in an interface with a different magnetic structure or different transport properties [30, 106]. The interfaces between polar and nonpolar oxides are especially interesting. Charged two-dimensional surfaces are unstable in nature unless there is nearby

    916

    J. M. D. Coey

    compensating charge, because of the energy density ½ε0 E2 associated with the uniform perpendicular electric field created in the surrounding space – an instability known as the polar catastrophe. Inevitably some electronic or ionic reconstruction must occur to resolve the problem. A celebrated and much-studied example is the (100) interface between formally nonpolar SrTiO3 and polar LaAlO3 [79]. Both are insulators, yet a conducting two-dimensional electron gas (2DEG) forms just below the interface in the SrTiO3 provided the thickness of the LaAlO3 overlayer exceeds four unit cells. The electron density and orbital character in the Ti dxy or dyz/zx bands can be modified by gating, and there is evidence of inhomogeneous ferromagnetic order in the 2DEG at liquid helium temperatures. Spin diffusion lengths can be on the order of a micron [106].

    Spin Pumping The technique of spin pumping has been developed to create spin currents perpendicular to the interface between ferro- or ferrimagnetic and nonmagnetic layers. It is related to the spin Hall effect, whereby a spin current is created perpendicular to a thin film of a heavy metal such as Ta, W, or Pt by passing an electric current through it. The spin current can modify the magnetic state of an adjacent ferromagnetic layer. In spin pumping, Larmor precession in the magnetic layer is excited by a radiofrequency magnetic field at a frequency determined by the ferromagnetic resonance that depends on an applied field or the anisotropy field of the layer. In that case, a rectified spin current is injected into the adjacent nonmagnetic layer, which may be detected by the inverse spin Hall effect. For this to work well, a magnetic material with very low damping is required, which makes yttrium-iron garnet the preferred option. The quality of the interface and the oxygen stoichiometry of the YIG are critical for efficient spin injection [114]. Magnonics Magnonics is the magnetic transport, storage, and processing of information using spin waves. See  Chap. 6, “Spin Waves”. The rich physics of the linear and nonlinear dynamics of multimode spin wave systems has been extensively investigated using yttrium-iron garnet strip lines [93]. The uniquely low damping in YIG allows spin wave propagation to be observed over distances of up to a centimeter, a thousand times longer than in a metal such as permalloy and far greater than the spin diffusion lengths encountered in spin electronics. Magnons, like diffusive spin currents, involve the transport of angular momentum, which in the case of magnons is correlated precession of the localized spin moments on neighboring sites. The key functions of propagation, storage, amplification, echo, and logic have all been demonstrated. Magnonics has the advantage of low power dissipation and may be combined with spin transfer torque to enable controlled generation of magnons along well-defined propagation directions [57]. Disadvantages are the large footprint of the devices (square microns) and the low charge ↔ magnon conversion efficiency at the input and output.

    17 Magnetic Oxides and Other Compounds

    917

    Conclusion Oxides continue to rival metals in practical importance as magnetic materials, and they have played a critical role in developing our understanding of collective magnetic order with localized electrons. Transition-metal oxide systems are especially useful for exploring the effects of frustration and spin-orbit coupling on magnetic order. Physical and chemical constraints of crystal structure and orbital overlap set limits on what is magnetically feasible and the associated parameters such as magnetization, anisotropy, Curie temperature, and magnetostriction. Traditionally, the space of achievable magnetism has been explored in bulk oxide solid solutions, leading to optimized functional materials – Fig. 19 showed a good example. In the future, strain and interface engineering in thin films may hold the key to the control of the magnetism of thin films or multilayers and potentially generate original oxidebased functionality for spin electronics. 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.

    References 1. Anderson, P.W.: Exchange in insulators: Superexchange, direct exchange and double exchange: Ch2. In: Rado, G.T., Suhl, H. (eds.) Magnetism I, pp. 25–83. Academic Press, New York (1963) 2. Artman, J.O., Murphy, J.C., Foner, C.S.: Magnetic anisotropy in antiferromagnetic corundumtype sesquioxides. Phys. Rev. 138, A912 (1965) 3. Ballet, O., Coey, J.M.D., Mangin, P., Townsend, M.G.: Ferrous talc - a planar antiferromagnet. Solid State Comm. 55, 787–790 (1985) 4. Ballet, O., Fuess, H., Wacker, K., Untersteller, E., Treutmann, W., et al.: Magnetic measurements of the synthetic olivine single crystals A2 SiO4 with a = Fe, co or Ni. J. Phys. Condens. Matter. 1, 4955–4979 (1989) 5. Ballhausen, C.J.: Introduction to Ligand Field Theory. McGraw Hill, New York (1962) 6. Bender, G., Koch, C., Oxborrow, C.A., Mørup, S., Madsen, M.B., Quinn, A.J., et al.: Magnetic properties of feroxhyte (δ-FeOOH). Phys. Chem. Miner. 22, 333–341 (1995) 7. Bibes, M., Villegas, J.E., Barthélémy, A.: Ultrathin oxide films and interfaces for electronics and spintronics. Adv. Phys. 60, 5 (2011) 8. Blank, D.H.A., Dekkers, M., Rijnders, G.: Pulsed laser deposition in Twente: from research tool towards industrial deposition. J. Phys. D. Appl. Phys. 47, 34006 (2014) 9. Bonnenberg, D., Boyd, E.L., Calhoun, B.A., Folen, V.J., Gräper, W., et al.: LandholdtBörnstein Numerical Data and Functional Relationships in Science and Technology New Series Group III, Vol 4 a and b, Magnetic and Other Properties of Oxides and Related Compounds. Springer, Berlin (1970) 10. Brabers, V.A.M.: Progress in spinel ferrite research, Ch 3. In: Buschow, K.H.J. (ed.) Handbook of Ferromagnetic Materials, vol. 8, pp. 189–324. Elsevier, Amsterdam (1995) 11. Bramwell, S.T., Gingras, M.J.P.: Spin ice state in frustrated magnetic pyrochlore materials. Science. 294, 1495–1501 (2001) 12. Brodsky, M.B.: Magnetic properties of the actinide elements and their metallic compounds. Rep. Prog. Phys. 41, 1548 (1978)

    918

    J. M. D. Coey

    13. Burns, R.G.: Mineralogical Applications of Crystal Field Theory, 2nd edn. Cambridge University Press, London (1993) 14. Cao, G., McCall, S., Shepard, M., Crow, J.E., Guertin, R.P.: Thermal, magnetic and transport properties of single-crystal Sr1−x Cax RuO3 (0 < x < 1.0). Phys. Rev. B. 56, 321–329 (1997) 15. Castelnovo, C., Moessner, R., Sondhi, S.L.: Magnetic monopoles in spin ice. Nature. 451, 42–45 (2009) 16. Catalano, S., Gibert, M., Fowlie, J., ´iñiguez, J., Triscone, J.-M., et al.: Rare-earth nickelates RNiO3 : thin films and heterostructures. Rep. Prog. Phys. 81, 046501 (2018) 17. Cho, A.Y., Arthur, J.R.: Molecular beam epitaxy. Prog. Solid State Chem. 10, 157 (1975) 18. Christen, H.M., Eres, G.: J. Phys. Conden. Mat. 20, 264005 (2008) 19. Coey, J.M.D.: Hard magnetic materials – a perspective. IEEE Trans. Magnetics. 47, 4671– 4681 (2011) 20. Coey, J.M.D.: Magnetism in d0 oxides. Nat. Mater. 18 (2019) 21. Coey, J.M.D.: Magnetism of dilute oxides, Ch 10. In: Tsymbal, E.Y., Zutic, I. (eds.) Handbook of Spin Transport and Magnetism, 2nd edn. CRC Press, Boca Raton (2019) 22. Coey, J.M.D., Chambers, S.A.: Oxide dilute magnetic semiconductors – fact or fiction? MRS Bull. 33, 1053 (2008) 23. Coey, J.M.D., Ghose, S.: Magnetic phase transitions in silicate minerals, Ch 9. In: Ghose, S., Coey, J.M.D., Salje, E. (eds.) Structural and Magnetic Phase Transitions in Minerals, pp. 162–184. Springer, New York (1988) 24. Coey, J.M.D., Readman, P.W.: New spin structure in an amorphous ferric gel. Nature. 246, 476–478 (1973) 25. Coey, J.M.D., Smith, P.A.I.: Magnetic nitrides. J. Magn. Magn. Mater. 200, 405–424 (1999) 26. Coey, J.M.D., Barry, A., Brotto, J.-M., Rakoto, H., Brennan, S., et al.: Spin flop in goethite. J. Phys. 7, 759–768 (1995) 27. Coey, J.M.D., Viret, M., von Molnar, S.: Mixed-valence manganites. Adv. Phys. 48, 167–293 (1999) 28. Coey, J.M.D., Stamenov, P., Gunning, R.D., Venkatesan, M., Paul, K.: Ferromagnetism in defect-ridden oxides and related materials. New J. Phys. 12, 053025 (2010) 29. Coey, J.M.D., Venkatesan, M., Xu, H.-J.: Introduction to magnetic oxides Ch 1. In: Ogale, S.B., Venkatesan, T.V., Blamire, M.G. (eds.) Functional Magnetic Oxides, pp. 1–30. WileyVCH, Weinheim (2013) 30. Coey, J.M.D., Ariando, Pickett, W.E.: Magnetism at the edge; new phenomena at oxide interfaces. Mater. Res. Bull. 38, 1040–1047 (2013) 31. Coey, J.M.D., Venkatesan, M., Stamenov, P.: Surface magnetism of strontium titanate. J. Phys. Condens. Matter. 28, 485001 (2016) 32. Coey, J.M.D., Ackland, K., Venkatesan, M., Sen, S.: Collective magnetic response of CeO2 nanoparticles. Nat. Phys. 12, 694–699 (2016) 33. Connolly, T.F., Copenhaver, E.D.: Bibliography of Magnetic Materials and Magnetic Transition Temperatures. Springer, New York (1972) 34. Corliss, L., Elliott, N., Hastings, J.: Magnetic structures of the polymorphic forms of manganous sulfide. Phys. Rev. 104, 924–928 (1956) 35. Cornell, R.M., Schwertmann, U.: The Iron Oxides; Structure, Properties, Reactions, Occurrence and Uses. VCH, Weinheim (2003) 36. Dietl, T., Ohno, H., Matsukura, F., Cibert, J., Ferrand, D.: Zener model description of ferromagnetism in zinc-blende magnetic semiconductors. Science. 287, 1019 (2000) 37. Dijkkamp, D., Venkatesan, T., Ogale, X.-D.W.S.B., Inam, A., et al.: Preparation of Y-Ba-Cu oxide superconductor thin films using pulsed laser evaporation from high Tc bulk material. Appl. Phys. Lett. 51, 619 (1987) 38. Dionne, G.F.: Magnetic Oxides. Springer, New York (2009) 39. Dunlop, D.J., Özdemir, Ö.: Rock Magnetism: Fundamentals and Frontiers. Cambridge University Press, Cambridge (1997) 40. Dzyaloshinsky, I.: A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. J. Phys. Chem. Solids. 4, 241–255 (1958)

    17 Magnetic Oxides and Other Compounds

    919

    41. Esaki, L., Stiles, P.J., von Molnar, S.: Magnetointernal field emission in junctions of magnetic insulators. Phys. Rev Lett. 19, 852 (1967) 42. Eschenfelder, A.H.: Magnetic Bubble Technology. Springer, Berlin (1980) 43. Frederichs, T., von Dobeneck, T., Bleil, U., Dekkers, M.J.: Towards the identification of siderite, rhodochrosite and vivianite in sediments by their low-temperature magnetic properties. Phys. Chem. Earth. 28, 669–679 (2003) 44. Fuess, H., Ballet, O., Lottermoser, W.: Magnetic phase transitions in olivines, Ch 10. In: Ghose, S., Coey, J.M.D., Salje, E. (eds.) Structural and Magnetic Phase Transitions in Minerals, pp. 185–201. Springer, New York (1988) 45. Gardner, J.S., Gingras, M.J.P., Greedan, J.E.: Magnetic pyrochlore oxides. Rev. Mod. Phys. 82, 53–107 (2010) 46. George, S.M.: Atomic layer deposition: an overview. Polymer. 1550, 125 (2010) 47. Gilleo, M.A.: Ferromagnetic insulators: garnets, Ch 1. In: Wohlfarth, E.P. (ed.) Handbook of Ferromagnetic Materials, vol. 2, pp. 1–54. Amsterdam, North Holland (1980) 48. Gong, C., Zhang, X.: Two dimensional magnetic crystals and emergent heterostructure devices. Science. 363 (2019) 49. Goniakowski, J., Finocchi, F., Noguera, C.: Polarity of oxide surfaces and nanostructures. Rep. Prog. Phys. 71, 016501 (2008) 50. Goodenough, J.B.: Magnetism and the Chemical Bond. Wiley-Interscience, New York (1963) 51. Greaves, C.: A powder neutron diffraction investigation of vacancy ordering and covalence in γFe2 O3 . J. Solid State Chem. 49, 325 (1983) 52. Greenwood, N.N., Gibb, T.C.: Mössbauer Spectroscopy. Chapman and Hall, London (1971) 53. Hazen, R.M., Jeanloz, R.: Wüstite (Fe1-x O): a review of its defect structure and physical properties. Rev. Geosci. Space Sci. 22, 37 (1984) 54. Hemberger, J., Krimmel, A., Kurz, T., Krug von Nidda, H.-A., Ivanov, V.Y., et al.: Structural, magnetic, and electrical properties of single-crystalline La1−x Srx MnO3 (0.4 < x < 0.85). Phys. Rev. B. 66, 094410 (2002) 55. Hill, R.J.: Physical vapour deposition, 2nd edn. Temescal, Berkley (1986) 56. Hill, A.H., Jiao, F., Bruce, P.G., Harrison, A., Kockelmann, W., Ritter, C.: Neutron diffraction study of mesoporous and bulk hematite, r-Fe2O3. Chem. Mater. 20, 4891–4899 (2008) 57. Hoffmann, A., Bader, S.D.: Opportunities at the frontiers of spintronics. Phys. Rev. Appl. 4, 047001 (2015) 58. Hurlbut, C.S., Klein, C.: Manual of Mineralogy, 19th edn. Wiley, New York (1977) 59. Jungwirth, T., Wang, K.Y., Mašek, J., Edmonds, K.W., König, J., et al.: Prospects for high temperature ferromagnetism in (Ga, Mn) as semiconductors. Phys. Rev. B. 72, 165204 (2005) 60. Kalid, M., Setzer, A., Ziese, M., Esquinazi, P., Spemann, D., et al.: Ubiquity of ferromagnetic signals in common diamagnetic oxides. Phys. Rev. B. 81, 214414 (2010) 61. Kanamori, J.: Anisotropy and magnetostriction of ferromagnetic and antiferromagnetic materials, Ch 4. In: Rado, G.T., Suhl, H. (eds.) Magnetism I, pp. 127–203. Academic Press, New York (1963) 62. Khomski, D.I.: Transition Metal Compounds. Cambridge University Press, Cambridge (2014) 63. Kojima, H.: Fundamental properties of hexagonal ferrites with the magnetoplumbite structure, Ch. 5. In: Wohlfarth, E.P. (ed.) Handbook of Ferromagnetic Materials, vol. 3, pp. 305–391. Amsterdam, North Holland (1982) 64. Krupicka, S., Novak, P.: Oxide spinels, Ch 4. In: Wohlfarth, E.P. (ed.) Handbook of Ferromagnetic Materials, vol. 3, pp. 189–303. North Holland, Amsterdam (1982) 65. Kuneš, J., Anisimov, V.I., Skornyakov, S.L., Lukoyanov, A.V., Vollhardt, D.: NiO: correlated band structure of a charge-transfer insulator. Phys. Rev. Lett. 99, 156404 (2007) 66. Lester, C., Ramos, S., Perry, R.S., Croft, T.P., Bewley, R.I., et al.: Field-tunable spin-densitywave phases in Sr3 Ru2 O7 . Nat. Mater. 14, 373–378 (2015) 67. Lu, H.-C., Chamorro, J.R., Wan, C., McQueen, T.M.: Universal single-ion physics in spinorbit-coupled d5 and d4 ions. Inorg. Chem. 57, 14443–14449 (2018) 68. Mackenzie, A.P., Maeno, Y.: The superconductivity of Sr2 RuO4 and the physics of spin-triplet pairing. Rev. Mod. Phys. 75, 657–712 (2003)

    920

    J. M. D. Coey

    69. Manipatruni, S., Nikonov, D.E., Lin, C.-C., Gosavi, T.A., Liu, H.-C., et al.: Scalable energyefficient magnetoelectric spin-orbit logic. Nature. 565, 35–43 (2019) 70. Matsumoto, Y., Murakami, M., Shono, T., Hasegawa, T., Fukumura, T., et al.: Roomtemperature ferromagnetism in transparent transition metal-doped titanium dioxide. Science. 291, 854 (2001) 71. McGuire, M.A.: Crystal and magnetic structures in layered transition metal dihalides. Crystals. 7, 120–145 (2017) 72. Mikhaylovskiy, R.V., Hendry, E., Secchi, A., Mentink, J.H., Eckstein, M., et al.: Ultrafast optical modification of exchange interactions in iron oxides. Nat. Commun. 6, 8190 (2015) 73. Moriya, T.: Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120, 91–98 (1960) 74. Morrish, A.H.: Canted Antiferromagnetism: Hematite. World Scientific, Singapore (1994) 75. Mott, N.F.: Metal-Insulator Transitions, 2nd edn. Taylor and Francis, London (1990) 76. Ney, A., Ollefs, K., Ye, S., Kammermeier, T., Ney, V., et al.: Absence of intrinsic ferromagnetic interactions of isolated and paired co dopant atoms in Zn1-x Cox O with high structural perfection. Phys. Rev. Lett. 100, 157201 (2008) 77. Ney, V., Henne, B., Lumetzberger, J., Wilhelm, F., Ollefs, K., et al.: Coalescence-driven magnetic order of the uncompensated antiferromagnetic Co doped ZnO. Phys. Rev. B. 94, 224405 (2016) 78. Obradors, X., Solans, X., Collomb, A., Samaras, D., et al.: Crystal structure of strontium hexaferrite. J. Solid State Chem. 72, 218–224 (1988) 79. Ohtomo, A., Hwang, H.Y.: A high-mobility electron gas at the LaAlO3 /SrTiO3 heterointerface. Nature. 427, 423–426 (2004) 80. Pan, F., Song, C., Liu, X.-J., Yang, Y.-C., Zeng, F.: Ferromagnetism and possible application in spintronics of transition-metal-doped ZnO films. Mater. Sci. Eng. Reports. 62, 1 (2008) 81. Park, J.H., Vescovo, E., Kim, H.J., Kwon, C., Ramesh, R., et al.: Direct evidence for a halfmetallic ferromagnet. Nature. 392, 794 (1998) 82. 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) 83. Parkinson, G.S.: Iron oxide surfaces. Surf. Sci. Reports. 71, 272–365 (2016) 84. Pearce, C.I., Pattrick, R.A.D., Vaughan, D.J.: Electric and magnetic properties of sulphides. Rev. Mineral. Geochem. 61, 127–180 (2006) 85. Pegg, J.T., Aparicio-Angles, X., Storr, M., de Leeuw, N.H.: DFT + U study of the structures and properties of the actinide dioxides. J. Nucle. Mater. 492, 269–278 (2017) 86. Penney, T., Shafer, M.W., Torrance, J.B.: Insulator-metal transition and long-range magnetic order in EuO. Phys. Rev. B. 5, 3669 (1972) 87. Philipp, J.B., Majewski, P., Alff, L., Erb, A., Gross, R., et al.: Structural and doping effects in half-metallic double perovskites A2 CrWO6 (A=Sr, Ba, and Ca) Phys. Rev. B. 68, 144431 (2003) 88. Porter, S.B., Venkatesan, M., Dunne, P., Doudin, B., Rode, K., et al.: Magnetic dead layers in La0.7 Sr0.3 MnO3 revisited. IEEE Trans. Magn. 56, 6000904 (2017) 89. Pullar, R.C.: Hexagonal ferrites a review of the synthesis, properties and applications of hexaferrite ceramics. Prog. Mater. Sci. 57, 1191–1334 (2012) 90. Schaefer, H.: Chemical Transport Reactions. Academic Press, New York (1964) 91. Schwertmann, U., Murad, E.: Effects of pH on the formation of goethite and hematite from ferrihydrite. Clays Clay Miner. 31, 277 (1983) 92. Sen, M.S., Wright, J.P., Attfield, J.P.: Charge order and three-site distortions in the structure of magnetite. Nature. 481, 173 (2011) 93. Serga, A.A., Chumak, A.V., Hildebrands, B.: YIG magnonics. J. Phys. D. Appl. Phys. 43, 264002 (2010) 94. Skomski, R., Coey, J.M.D.: Permanent Magnetism, pp. 257–264. IOP Publishing, Bristol (1999)

    17 Magnetic Oxides and Other Compounds

    921

    95. Skomski, R., Coey, J.M.D.: Magnetic anisotropy – how much is enough for a permanent magnet? Scripta Mater. 112, 3–8 (2016) 96. Smit, J., Wijn, H.P.J.: Ferrites. Philips Technical Library, Eindhoven (1959) 97. Stäblein, H.: Hard ferrites and plastoferrites, Ch. 7. In: Wohlfarth, E.P. (ed.) Handbook of Ferromagnetic Materials, vol. 3, pp. 441–601. Amsterdam, North Holland (1982) 98. Steinbeiss, E.: Ch 13. In: Ziese, M., Thornton, M.J. (eds.) Spin Electronics, pp. 296–315. Springer Verlag, Berlin (2001) 99. Sze, S.M., Lee, M.K.: Semiconductor Devices: Physics and Technology, Ch 12, 3rd edn. Wiley, New York (2012) 100. Takayama, T., Kato, A., Dinnebier, R., Nuss, J., Kono, H., et al.: Hyperhoneycomb iridate β-Li2 IrO3 as a platform for Kitaev magnetism. Phys. Rev. Lett. 114, 077202 (2015) 101. Tanabe, Y., Sugano, S.: On the absorption spectra of complex ions I – III. J. Phys. Soc. Jpn. 9, 753 (1954).; 9 766 (1954); 11 864 (1956) 102. Tietze, T., Gacic, M., Schütz, G., Jacob, G., Brück, S., et al.: XMCD studies on Co and Li doped ZnO magnetic semiconductors. New J. Phys. 10, 055009 (2008) 103. Treves, D.: Studies on orthoferrites at the Weizmann Institute of Science. J. Appl. Phys. 36, 1033–1039 (1965) 104. Valenzuela, R.: Magnetic Ceramics. Cambridge University Press, Cambridge (1994) 105. van Stapele, R.P.: Sulphospinels, Ch 8. In: Wohlfarth, E.P. (ed.) Ferromagnetic Materials, vol. 3, pp. 607–745. Amsterdam, North Holland (1982) 106. Varignon, J., Vila, L., Barthélemy, A., Bibes, M.: A new spin for oxide interfaces. Nat. Phys. 14, 322–325 (2018) 107. Vasala, S., Karppinen, M.: A2 B BO6 perovskites: a review. Prog. Solid State Chem. 43, 1–36 (2015) 108. Verwey, E.J.W.: Electronic conduction of magnetite (Fe3 O4 ) and its transition point at low temperature. Nature. 144, 327 (1937).; Verwey E.J.W., Haayman, P.W. Electronic conduction and transition point of magnetite (Fe3 O4 ). Physica 8, 979 (1941) 109. von Helmolt, R., Wecker, J., Holzapfel, B., Schultz, L., Samwer, K.: Giant negative magnetoresistance in perovskite-like La2/3 Ba1/3 MnOx ferromagnetic films. Phys. Rev. Lett. 71, 2331 (1993) 110. Walz, F.: The Verwey transition, − a topical review. J. Phys. Cond. Matter. 14, R285 (2002) 111. Wang, Z.-S., Oureshi, N., Yasin, S., Mukhin, A., Ressouche, E., et al.: Magnetoelectric effect and phase transitions in CuO in external magnetic fields. Nat. Comm. 7, 10295 (2016) 112. Wasey, A.H.M.A., Karmakar, D., Das, G.P.: Manifestation of long-range ordered state in VX2 (X = Cl, Br, I) systems. J. Phys. Condens. Matter. 25, 476001 (2013) 113. Xu, Y.-B., Awschalom, D.D., Nitta, J. (eds.): Handbook of Spintronics. Springer, Berlin (2015) 114. Yang, F.-Y., P. C.: Hammel: FMR-driven spin pumping in Y3 Fe5 O12 -based structures. J. Phys. D. Appl. Phys. 51, 253001 (2018) 115. 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) 116. Zener, C.: Interactions between the d-shells in the transition metals. Phys. Rev. 81, 40; 82 403 (1951) 117. Zhang, Z., Satpathy, S.: Electron states, magnetism and the Verwey transition in magnetite. Phys. Rev. B. 44, 13319 (1991) 118. Zhao, D.-P., Zhang, L.-G., Malik, I.A., Liao, M.-H., Cui, W.-Q., et al.: Finite temperature magnetism of CrSe and CrTe. Nano Res. 11, 3116–3121 (2018) 119. Ziese, M., Thornton, M.J.: Spin Electronics Lecture Notes on Physics, vol. 569. Springer, Berlin (2001)

    922

    J. M. D. Coey

    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, at 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.

    Dilute Magnetic Materials

    18

    Alberta Bonanni , Tomasz Dietl , and Hideo Ohno

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dilute Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dilute Magnetic Metals (DMMs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . p-Type Dilute Ferromagnetic Semiconductors (DFSs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Other Dilute Ferromagnetic Semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dilute Magnetic Semiconductors (DMSs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dilute Magnetic Topological Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heterogenous Magnetic Semiconductors and Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Energy States of Magnetic Dopants in Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Exchange Interactions Between Band and Localized Spins . . . . . . . . . . . . . . . . . . . . . . . . . . .

    926 928 928 928 929 929 930 930 931 936

    A. Bonanni Institut für Halbleiter- und Festkörperphysik, Johannes Kepler University, Linz, Austria e-mail: [email protected] T. Dietl International Research Centre MagTop, Institute of Physics, Polish Academy of Sciences, Warsaw, Poland WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan e-mail: [email protected] H. 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 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_21

    923

    924

    A. Bonanni et al.

    Effects of sp-d(f ) Exchange Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin-Splitting of Extended States: Weak Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin-Splitting of Extended States: Strong Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alloy and Spin-Disorder Scattering: Weak Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alloy and Spin-Disorder Scattering: Strong Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bound Magnetic Polarons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantum Localization and Mesoscopic Phenomena: Colossal Magnetoresistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Interplay of sp–d(f ) Exchange Interactions and Spin-Orbit Coupling . . . . . . . . . . . . . . . . Dominant Spin-Spin Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dipole-Dipole Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Direct Spin-Spin Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Superexchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Carrier-Mediated Spin-Spin Coupling: Intra- and Interband Contributions . . . . . . . . . . . . Magnetic Properties of Dilute Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin-Glass Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . p-Type Dilute Ferromagnetic Semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dilute Ferromagnetic Insulators and Topological Insulators . . . . . . . . . . . . . . . . . . . . . . . . Heterogenous Magnetic Semiconductors and Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    938 938 940 941 942 942 945 945 946 946 946 947 949 952 952 957 960 964 968

    Abstract

    This chapter reviews properties of dilute magnetic metals, semiconductors, and topological materials, with a focus on compound semiconductors. Comprehensive studies of these systems have bridged the physics and functionalities of semiconductors and magnetic materials as well as opened the door for novel concepts of spintronic devices by demonstrating the control of magnetization and spin currents by light, strain, electric fields, and currents in both topologically trivial and nontrivial systems (chapter “SSE”). Furthermore, dilute magnetic materials have gained a status of model systems for assessing the nature of thermodynamic phase transformations in the presence of spatial disorder, competing interactions, and frustration. In particular, studies of spin-glass freezing have challenged standard theoretical models of statistical physics and triggered the development of new conceptual and computational tools impacting presently various fields of science and computer hardware. At the same time, progress in controlling magnetic ion distribution at the nanoscale results in embedded nanostructures that have a potential for applications in electronics, spintronics, photovoltaics, plasmonics, and thermoelectrics. List of Symbols A aB bq F F

     N

    Spectral density of states Effective Bohr radius Fourier transform of electron density Free energy density Free energy density functional Mean free path Charge state

    18 Dilute Magnetic Materials N0 p P P

    Tf u Vkd V W x α β ηq φ(z), ψ(r)

    Cation density Hole carrier concentration Spin polarization Probability distribution Spin-glass freezing temperature Periodic part of the Bloch function Hybridization energy Matrix element of conduction band offset Matrix element of valence band offset Magnetic cation fractional content s-d exchange integral p-d exchange integral Fourier transform of magnetization Envelope functions

    List of Abbreviations 1D, 2D, 3D AF AP AFM AHE AMR BMP CMR CMS CPA DFT DMS DFS DMM DOS EPR FC fcc FM FMR GMR HR LDA LSDA LT MBE MIT MFA MOVPE MCD MR NC NN NNN RE

    One-, two-, three-dimensional Antiferromagnetic Atom probe Atomic force microscopy Anomalous Hall effect Anisotropic magnetoresistance Bound magnetic polaron Colossal magnetoresistance Condensed magnetic semiconductor Coherent potential approximation Density functional theory Dilute magnetic semiconductor Dilute ferromagnetic semiconductor Dilute magnetic material Density of states Electron paramagnetic resonance Field cooled Face-centered cubic Ferromagnetic Ferromagnetic resonance Giant magnetoresistance High resolution Local density approximation Local spin-density approximation Low temperature Molecular beam epitaxy Metal-insulator transition Mean-field approximation Metal organic vapor phase epitaxy (the same as MOCVD) Magnetic circular dichroism Magnetoresistance Nanocrystal Nearest neighbor Next nearest neighbor Rare earth

    925

    926 RKKY SQUID TEM TM TMR VCA wz XANES XAS XES XRD ZFC zb

    A. Bonanni et al. Ruderman-Kittel-Kasuya-Yosida Superconducting quantum interference device Transmission electron microscopy Transition metal Tunneling magnetoresistance Virtual-crystal approximation Wurtzite x-ray absorption near-edge structure x-ray absorption spectroscopy x-ray emission spectroscopy x-ray diffraction Zero-field cooled Zinc-blende

    Introduction Dilute localized magnetic moments are usually associated with open d or f shells of magnetic dopants but sometimes originate from nonmagnetic impurities and/or lattice defects. The dilute systems attract considerable attention for a number of reasons. Firstly, even a minute concentration of uncoupled magnetic impurities can change dramatically the transport and the optical characteristics. A relevant example is the Kondo effect in metals ( Chap. 2, “Magnetic Exchange Interactions”) used for the fabrication of efficient thermocouples at low temperatures. Another is represented by intra-impurity optical excitations in semiconductors and oxides, which dominate the optical properties in the spectral region below the fundamental gap ( Chap. 3, “Anisotropy and Crystal Field”), a property exploited successfully in stained glasses and broadband lasers [1]. Secondly, dilute systems bridge the physics and functionalities of semiconductors and magnetic materials. The existence of strong spin-dependent interactions between bands and open magnetic shells, coexisting with spin-orbit interactions, leads to giant magnetotransport and magnetooptical phenomena and opens the door for spintronic functionalities, the control of magnetization and spin current by light, strain, electric fields, and currents in both topologically trivial and nontrivial systems. Thirdly, the randomness in the distribution of interacting localized spins accounts for novel magnetic properties, not encountered in chemically ordered crystals. More specifically, dilute magnetic materials have gained the status of model systems for assessing the nature of thermodynamic phase transformations in the presence of spatial disorder, competing interactions, and frustration. Another appealing question concerns quantum phase transformations in disordered systems and how localized magnetism evolves into an itinerant or nonmagnetic ground state as a function of the magnetic ion concentration. Fourthly, materials science of dilute magnetic systems is surprisingly rich. It has been increasingly clear that information about the nominal chemical composition is rarely sufficient to account for the effective characteristics of these media. Actually, to a large extent, their properties, including the nature of the magnetic ground

    18 Dilute Magnetic Materials

    927

    state, depend on the way a given sample has been obtained. In particular, the distribution of magnetic ions over the specimen volume (random vs nonuniform), their incorporation into the lattice sites (substitutional vs interstitial), the resulting spin configuration, and spin-spin interactions depend on the substrate employed, growth conditions, co-doping, post-growth processing, surface reactions, and contamination. In this context the extensive use of nanocharacterization tools, such as synchrotron radiation, microscopy, particle beam, and scanning probe methods, is essential [2], as elaborated in  Chaps. 25, “Magnetic Scattering,” and  24, “Magnetic Imaging and Microscopy.” Together with more traditional techniques of magnetic resonance ( Chap. 26, “Electron Paramagnetic and Ferromagnetic Resonance”) and optical magnetospectroscopy ( Chap. 10, “Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films”), these tools not only visualize the micromagnetic characteristics but also reveal the distribution of magnetization and magnetic constituents at the nanoscale as well as determine properties of energy states relevant to magnetism with ever-improving spatial, chemical, spin, spectral, and more and more often time resolution. Results of such combined structural and spectroscopic investigations are now broadly employed to substantiate the interpretation of experimental findings and to benchmark theoretical modeling, particularly in newly developed systems [3]. Another specific feature of dilute magnetic materials is the small magnitude of the magnetization, which makes these systems sensitive to both contamination by ferromagnetic (FM) nanoparticles [4, 5, 6, 7] and experimental artifacts in SQUID measurements [8]. A related timely question concerns the origin of a temperature-independent anhysteretic magnetic signal often observed in nominally diamagnetic solids. The search for new physics, such as contribution from orbital currents, is now accompanied by an extensive use of spatially resolved analytic tools, examination of samples from a range of suppliers and processed in various ways [9]. Considering different hosts and magnetic ordering, dilute magnetic materials can be grouped into six classes, as outlined in section “Dilute Magnetic Materials”. A more detailed discussion of various aspects of dilute systems, together with examples of experimental and theoretical findings, is presented in sections “Energy States of Magnetic Dopants in Solids”, “Exchange Interactions Between Band and Localized Spins”, “Effects of sp-d(f ) Exchange Interactions”, “Dominant Spin-Spin Interactions”, and “Magnetic Properties of Dilute Magnetic Materials”. The key parameter characterizing dilute magnetic systems is the concentration of magnetic ions x. We warn the readers that various conventions exist in literature. Here, by x we understand the fractional cation content. For instance, x in the case of Mn-doped cadmium arsenide is defined by (Cd1−x Mnx )3 As2 . Within this definition, the concentration of magnetic √ dopants per√unit volume is xN0 , where the cation concentration N0 = 4/a03 , 4/( 3a 2 c), 12/( 3a 2 c), and 8/a03 for the zincblende and rock-salt, wurtzite, tetradymite, and elemental diamond-structure dilute magnetic semiconductors, respectively. Another convention of interest concerns the description of exchange coupling. Here, the exchange interaction between a pair of localized spins is expressed in the

    928

    A. Bonanni et al.

    form Hij = −2Jij S i · S j , whereas the exchange interaction between a band carrier and a localized spin is expressed as Hi = −Jsp−d(f ) s · S i .

    Dilute Magnetic Materials It is natural to distinguish the six distinct families of dilute magnetic materials as follows.

    Dilute Magnetic Metals (DMMs) In the limit of low concentrations of transition-metal impurities in nonmagnetic metals, some dopants may acquire giant magnetic moments, or, conversely, impurity moments can be screened by a cloud of band electrons (the Kondo effect), as described in  Chaps. 14, “Magnetism of the Elements” and  2, “Magnetic Exchange Interactions,” respectively. For higher concentrations, the energy scale becomes determined by long-range carrier-mediated RKKY interactions between localized spins ( Chap. 2, “Magnetic Exchange Interactions”). The randomness of impurity distribution (quenched disorder) and frustration ( Chap. 2, “Magnetic Exchange Interactions”) associated with alternations between FM and antiferromagnetic (AF) coupling as a function of the spin-spin distance results in a specific behavior, referred to as spin-glass freezing (section “Spin-Glass Systems”). Spinglass characteristics have been observed in numerous dilute metallic (crystalline or amorphous) but also in nonmetallic systems [10]. Unusual properties (such as a magnetic phase transition without anomalies in specific heat but with glassylike dynamics) have challenged standard theoretical models of statistical physics and triggered the development of new conceptual and computational tools [11, 12] impacting various fields of science [13] and computational hardware [14]. The prevailing view is that in the case of long-range RKKY interactions specific to DMMs, glassy dynamics reflect the wandering of spin configurations between local free energy minima (hierarchical dynamics) [15], whereas the spontaneous formation and annihilation of compact droplet excitations [16] describes better nonmetallic materials such as DMSs, in which spin-spin interactions are shortranged. As in the case of other materials, “intrinsic” properties of DMMs are affected by materials issues such as oxidation [17] and a nonuniform distribution of the magnetic constituent [18].

    p-Type Dilute Ferromagnetic Semiconductors (DFSs) In these systems, a high density of delocalized or weakly localized holes leads to long-range FM interactions between TM cations, which dominates over the superexchange ( Chap. 2, “Magnetic Exchange Interactions”), and is well described by the p–d Zener model [19] (section “p–d Zener Model”). The flagship example here is (Ga,Mn)As [20], but to this category belong also other families of compounds doped with transition metals, in which holes originate from point defects, e.g.,

    18 Dilute Magnetic Materials

    929

    (Pb,Sn,Mn)Te [21], or acceptor impurities, e.g., (Cd,Mn)Te/(Cd,Mg)Te:N [22], rather than from Mn. Groundbreaking spintronic functionalities demonstrated for these systems [23,24], rely on the strong p–d coupling between localized spins and hole carriers as well as on a sizable Luttinger spin-orbit interaction in p-like orbitals forming the valence band or originating from inversion asymmetry of the structure. The reported magnitudes of Curie temperature TC reach 200 K in (Ga,Mn)As [25, 26, 27], (Ge,Mn)Te [28, 29], and (K,Ba)(Zn,Mn)2 As2 [30] with less than 10% of Mn cations, xeff < 0.1, measured by saturation magnetization in moderate fields, μ0 H  5 T (section “p-Type Dilute Ferromagnetic Semiconductors”). A numerous functionalities evident in (Ga,Mn)As and related DFSs (magnetization control by an electric field, current-induced domain-wall motion, anisotropic tunneling magnetoresistance) are now transferred to magnetic metals supporting ferromagnetism above room temperature. A specific feature of p-type DFSs is the interplay of carrier-mediated ferromagnetism with carrier localization, which results in spatial fluctuations of magnetization and superparamagnetic signatures generated by nonuniformities in carrier density that appear in the vicinity of the metal-toinsulator transition [31, 32].

    Other Dilute Ferromagnetic Semiconductors The Goodenough-Kanamori-Anderson rules ( Chap. 2, “Magnetic Exchange Interactions”) indicate in which cases a superexchange can lead to FM shortrange coupling between localized spins. According to experimental and theoretical studies, such a mechanism operates for Mn3+ ions in GaN and accounts for TC values reaching about 13 K at x ≈ 10% in semi-insulating wurtzite Ga1−x Mnx N [33] (section “Dilute Ferromagnetic Insulators and Topological Insulators”). The dependence of TC on x corroborates the percolation theory [34] that implies the coexistence of percolating and non-percolating FM clusters at T > 0. Ferromagnetic superexchange and its variants dominate in magnetic semiconductors [35] such as Cr-spinels and CdCr2 Se4 [36] (TC = 130 K) as well as in EuO (TC = 68 K) and EuS (TC = 16 K) [37], known for showing colossal magnetoresistance effects in the vicinity of TC [38]. By electron doping (oxygen vacancies, Eu substituted by Gd), the TC of Eu compounds can be enhanced by about 50 K in agreement with the RKKY theory that explains also TC ≈ 160 mK in n-Zn1−x Mnx O:Al with x = 0.03 [39].

    Dilute Magnetic Semiconductors (DMSs) This material family was initially named semimagnetic semiconductors [40], as it comprises standard semiconductors doped with magnetic impurities that are randomly distributed, electrically inactive, and do not undergo any long-range spin ordering [41,42]. Typical representatives are (Cd,Mn)Te, (Zn,Co)O, (Hg,Fe)Se, and (Pb,Eu)S. Short-range AF superexchange between transition metal ions or longrange dipolar interactions between rare-earth magnetic moments lead to spin-glass

    930

    A. Bonanni et al.

    freezing at Tf < 1 K for x < 0.1 (section “Spin-Glass Systems”). The use of magnetooptical and quantum transport techniques has revealed and quantified the influence of sp–d coupling upon the exciton and Landau-level spectra, quantum Hall effect, quantum localization, universal conductance fluctuations, and spin dynamics in DMS systems of various dimensionality [43] (section “Effects of sp-d(f ) Exchange Interactions”) as well as made it possible to demonstrate electrical spin injection [44]. Conversely, the spin subsystem has been probed by the Faraday effect [45] or quantum noise [46] (section “Spin-Glass Systems”). The physics of bound magnetic polarons, a single electron interacting with spins localized within the confining potential of impurities or quantum dots, has been advanced in DMSs [47] (section “Bound Magnetic Polarons”). Most of the end-member compounds, for instance, MnSe and EuTe, are antiferromagnets, which together with related materials, like CuMnAs, constitute the building blocks for the emerging field of AF spintronics [48]. Another ultimate limit of DMSs is represented by qubit systems consisting of single magnetic ions in single quantum dots [49].

    Dilute Magnetic Topological Materials In these systems magnetization-induced giant p–d exchange spin-splitting of Dirac cones turns helical states into chiral states. Striking consequences of this transformation include (i) the precise quantization of Hall conductance σxy = e2 / h demonstrated for thin layers of FM (Bi,Sb,Cr)2 Te3 at millikelvin temperatures [50], as predicted theoretically [51]; (ii) the efficient magnetization switching by spinlocked electric currents in these ferromagnets [52]; and (iii) the formation of Weyl semimetals from 3D Dirac spin-glass materials, such as (Cd,Mn)3 As2 or strained (Hg,Mn)Te [53] (section “Dilute Ferromagnetic Insulators and Topological Insulators”). Furthermore, inverted band ordering specific to topological matter enhances the role of the long-range interband Bloembergen-Rowland contribution to spin-spin interactions, resulting in higher spin-glass freezing temperatures Tf in topological semimetals, such as (Cd,Mn)3 As2 and (Hg,Mn)Te, compared to topologically trivial II-VI Mn-based DMSs [54]. This exchange, taken into consideration within the p–d Zener model [55] (section “p–d Zener Model”) and named Van Vleck’s mechanism, was proposed to lead to ferromagnetism in topological insulators of V-, Cr-, or Fecontaining bismuth/antimony chalcogenides [51]. However, according to ab initio studies and corresponding experimental data, the superexchange in the case of V and Cr and the RKKY coupling in Mn-doped films appear to contribute significantly to ferromagnetism in tetradymite V2 -VI3 compounds [56]. Ferromagnetism mediated by carriers in surface or edge Dirac cones is rather weak for realistic values of spatial extension of these states into the bulk [23].

    Heterogenous Magnetic Semiconductors and Oxides These materials are characterized by a highly nonrandom distribution of magnetic elements brought about by binodal or spinodal nanodecomposition (chemical phase separation) or nanoprecipitation (crystallographic phase separation) [3]

    18 Dilute Magnetic Materials

    931

    (section “Heterogenous Magnetic Semiconductors and Oxides”). Here, nanoregions with high concentrations of magnetic cations, denoted as condensed magnetic semiconductors – CMS [57] – account for superparamagnetic, superferromagnetic ( Chap. 20, “Magnetic Nanoparticles”), or FM-like properties persisting typically above room temperature. The CMSs appear in the shapes of nanodots (the dairiseki phase) or nanocolumns (the konbu phase) buried in semiconducting or insulating hosts [58]. Such nanocomposites, fabricated in a self-organized fashion and containing metallic magnetic nanostructures, have a potential for applications in electronics, spintronics, photovoltaics, plasmonics, and thermoelectrics [3].

    Energy States of Magnetic Dopants in Solids A good starting point for the description of dilute magnetic materials is the Vonsovskii model, according to which the electron states can be divided into two categories: (i) localized magnetic states, typically originating from open d or f atomic shells, and (ii) extended band states built up of s, p, and also d atomic orbitals. The understanding of how particular hosts perturb open shells of magnetic impurities or defects is built on two pillars: • Crystal-field theory: This is essentially a first-order perturbation theory that shows how electrostatic potentials generated by nonmagnetic ligands split open multi-electron atomic shells of magnetic dopants ( Chap. 3, “Anisotropy and Crystal Field”) or point defects depending on their location in the lattice. These splittings, and thus the resulting charge and spin configurations of the impurity ground and excited states, are described for various crystal structures by the Tanabe-Sugano diagrams as a function of the crystal-field strength ( Chap. 4, “Electronic Structure: Metals and Insulators”). In more general terms, characteristics of possible states, including the forms of corresponding spin Hamiltonians [59], are dictated by the spin-containing group theory, whose parameters are provided experimentally, and reproduced increasingly successfully by relativistic ab initio computations ( Chap. 3, “Anisotropy and Crystal Field”). In general, electronic energies are lowered by coupling to the lattice, by a local deformation. This is particularly relevant for orbitally degenerate states, the mechanism described in terms of a static or dynamic Jahn-Teller effect ( Chap. 3, “Anisotropy and Crystal Field”). In extreme cases coupling to the lattice can lead to a negative value of the intra-site correlation energy Ueff . • Anderson Hamiltonian: The Anderson model of magnetic impurities ( Chap. 4, “Electronic Structure: Metals and Insulators”) describes a competition between the intra-atomic Coulomb energy U (that accounts for the presence of open d and f shells) and quantum hopping of electrons between these localized shells and extended band states (quantified by the hybridization matrix element Vkd ). Thus, hybridization enlarges the localization radius of magnetic electrons, which lowers their intra-site correlation energy. This effect explains (i) the existence of numerous levels corresponding to subsequent charge states of impurities or defects within the semiconductor energy gap (the Haldane-Anderson mecha-

    932

    A. Bonanni et al.

    nism) [60] and (ii) the disappearance of localized magnetic moments for Vkd > U, the case of many combinations of magnetic impurities and metallic hosts. However, as long as localized magnetic moments survive in a given host, so that the corresponding d states do not contribute to the Fermi volume, the presence of hybridization does not invalidate the expectations of the crystal-field theory but only modifies the magnitudes and microscopic origins of particular crystal-field parameters. Figure 1 depicts ground-state energy levels corresponding to two charge states of cation-substitutional TM ions in zinc-blende or wurtzite II-VI and III-V compounds. These two states have the donor and acceptor character (lower and upper Hubbard bands, respectively), and the energy difference between them is the effective correlation energy Ueff . For either of these material families, the energies of TM levels are to a good approximation universal, in the sense that they can be presented in one plot if the relative positions of the band edges in particular compounds are shifted according to the band offsets known from heterostructure studies. For particular charge states, the spin localized on the TM ion assumes in these compounds the highest possible value, according to Hund’s rule. In the case of ions with the orbital momentum L = 0, the Jahn-Teller effect and spin-orbit coupling affect positions and splittings of the levels. If there is no net magnetic moment associated with the ion, the case of Fe2+ in II-VI compounds, magnetization is of the Van Vleck type at low temperatures. The diagrams in Fig. 1 make it possible to assess the electrical activity of a given TM impurity in a given host and the variation of its charge state with the extrinsic doping. In particular, the levels residing within the bandgap act as carrier traps and, thus, may serve for obtaining a semi-insulating material [62]. Alternatively, TM impurities donate electrons if the corresponding donor state is above the bottom of the conduction band (e.g., Sc in CdSe) or holes if the acceptor state is below the top of the valence band (e.g., Mn in GaAs). It is important to note that these resonant states are accompanied by shallow donor or acceptor levels, respectively, created by the Coulomb potential of the charged TM impurities, i.e., Sc3+ in CdSe and Mn2+ in GaAs. Another type of level that is not depicted in Fig. 1 is the so-called charge transfer state or Zhang-Rice polaron [63]. It appears if the magnitude of p–d hybridization is sufficiently large (the strong coupling limit) to result in bound states of holes on anions surrounding isoelectronic (neutral) TM impurities, the case of, e.g., Fe3+ in GaN and Mn2+ in ZnO [64]. The same mechanism accounts for a gradual increase of the Mn2+ acceptor binding energy in III–V compounds on going from the antimonides, through arsenides and phosphides, to the nitrides, as a contribution of p-d hybridization to the binding energy increases when the cation-anion distance diminishes [23]. A similar set of data has been collected for rare-earth impurities in various families of compounds (see Ref. [42] for the case of II–VI semiconductors). On the other hand, no such data are so far available for newly developed topological materials, such as thallium/nobium pnictides and bismuth/antimony chalcogenides in which TM doping plays an increasingly important role.

    Energy (eV)

    CdTe HgTe

    2

    CdSe

    ZnO

    A(0/-)

    CdS

    ZnS

    4

    ZnTe

    933

    ZnSe

    18 Dilute Magnetic Materials

    D(0/+)

    0

    Sc Ti V Cr Mn Fe Co Ni Cu 1

    2

    3

    4

    d d d d

    5

    d

    6

    7

    8

    9

    d d d d

    3

    4

    5

    GaN

    InSb

    InP InAs

    A(0/-)

    AlAs

    GaP

    6

    GaSb

    2

    InN

    Energy (eV)

    1

    d d d d d d

    GaAs

    Ti V Cr Mn Fe Co

    2

    AlP

    AlN

    4

    AlSb

    -2

    0 D(0/+)

    -2

    Fig. 1 Approximate positions of ground-state levels derived from d shells of cation-substitutional transition-metal impurities in two different charge states relative to the conduction and valence band edges of II-VI (upper panel) and III-V (lower panel) compounds. Triangles and squares denote d N /d N −1 donor (D) state (donating the electron, N → N − 1, becomes positively charged) and d N /d N +1 acceptor (A) states (accepting the electron, N → N + 1, becomes negatively charged), respectively. Typically, the wave function of these states contains a substantial contribution from p orbitals of neighboring anions. These diagrams include neither shallow donor or acceptor levels accompanying resonant states in the conduction and valence band, respectively, nor mid-gap hole traps appearing for strong p–d hybridization, the case of oxides and nirides. (Adapted from Langer et al. [61] and Zunger [60] updating the values of band offsets)

    934

    A. Bonanni et al.

    Detailed information on energy levels corresponding to the ground and excited states at a given charge/spin configuration and coupling to the lattice is essential for designing light emitters, particularly broadband optically pump lasers of TM-doped oxides or chalcogenides [1]. If an impurity level participating in optical transitions overlaps with a band continuum, the absorption spectrum may assume a form of the Fano resonance. In metals, in turn, resonant scattering of carriers is expected if the Fermi energy εF coincides with the ground-state level of impurities, the case presented in Fig. 2a. However, if the Fermi level is located in-between impurity Hubbards bands, electron scattering rate is enlarged, too – this time by the presence of the Abrikosov-Shul resonance at the Fermi level (Fig. 2b), which is formed by electrons dynamically bound to localized spins, according to theory of the Kondo effect in the strong coupling limit ( Chap. 2, “Magnetic Exchange Interactions”). A number of phenomena appear when increasing the content of the magnetic constituent. Below, we consider first insulator-to-metal transitions transforming localized centers into extended states, either merging with the existing bands or forming additional bands. Next, the case of resonant levels is described. Finally, the question is addressed what happens if the impurity concentration surpasses the solubility limit. The origin and effects of exchange coupling between carriers and localized spins as well as the mechanisms and consequences of interactions among localized spins are discussed in subsequent sections. Anderson-Mott transition: Within this model, localization of carriers in their parent band results from quantum interference of many-body and single-carrier scattering amplitudes [65]. For disorder-modified hole-hole interactions and hole

    E

    (a)

    E

    (b)

    HF

    E

    DOS (c)

    HF

    E

    DOS (d)

    HF HF DOS

    DOS

    Fig. 2 Schematic forms of density of states (DOS) if magnetic impurities introduce states overlapping with a continuum of band states; (a) resonant state; (b) Abrikosov-Shul resonance at the Fermi level εF appearing below the Kondo temperature; (c) Efros-Shklovskii Coulomb gap at εF in the impurity DOS appearing if inhomogeneous broadening of the impurity band is larger than lifetime broadening; (d) hybridization gap for Kondo insulators

    18 Dilute Magnetic Materials

    935

    scattering by ionized cation-substitutional Mn2+ acceptors in III-V semiconductors, the critical concentration pc is about 1016 cm−3 for Insb, 1020 cm−3 for GaAs, and above 4 × 1021 cm−3 for GaN. These values of pc are consistent with the empirical Mott criterion pc ≈ (0.26/aB )3 , where aB is an effective Bohr radius of the dopant in question, here the hole localization radius in a single neutral Mn3+ = Mn2+ + h acceptor center [23]. Obviously, Anderson-Mott physics is relevant to magnetic semiconductors also in the case when carriers originate from extrinsic dopants, such as oxygen vacancies in EuO or In in (Cd,Mn)Se. It should be noted that no available ab initio method can describe the Anderson-Mott localization. Mott-Hubbard transition: Zinc-blende MnTe appears to be a localized antiferromagnet meaning that 3d levels do not undergo a Mott-Hubbard insulator-to-metal transition (see  Chap. 17, “Magnetic Oxides and other Compounds”). In contrast, zinc-blende MnAs is presumably an itinerant ferromagnet, which indicates that at a certain value of Mn density in (Ga,Mn)As delocalization of Mn 3d levels occurs. This issue, despite its relevance to a number of dilute magnetic semiconductors, has not yet been experimentally explored, presumably because materials with low concentrations of magnetic impurities are usually investigated. Theoretical studies on disorder effects in the Mott-Hubbard transition are under way [66]. Coulomb gap and charge ordering: Inter-site interactions within the impurity band give rise to the Efros-Shklovskii Coulomb gap in the one-carrier density of states at the Fermi level [67], as shown in Fig. 2c. Let’s consider a concentration of magnetic impurities low enough that no Mott-Hubbard transition occurs but sufficiently large for the band carriers to be delocalized and the Fermi level pinned by the resonant state. Under these conditions, the Coulomb gap in the impurity band decouples it from the continuum and precludes resonant scattering of band carriers. At the same time, the momentum relaxation rate for scattering by ionized impurities is diminished, as the inter-site Coulomb interactions impose spatial ordering of impurity charges. Accordingly, record-high electron mobilities up to 2×107 cm2 /Vs were observed in zero-gap semiconductors when the Fermi level was pinned by either resonant donors (HgSe:Fe [68]) or acceptors [(Hg,Mn)Te:Cu [69]]. A similar effect was found in (Sn,Gd)Te [70]. Hybridization gap and Kondo insulators: The Kondo effect vanishes if the magnitude of RKKY coupling between magnetic impurities surpasses the Kondo temperature, which for TM impurities in metals occurs at densities typically below 1%. However, there is a class of compounds including SmB6 , Ce3 Bi4 Pt3 , FeSi [71] for which the Anderson Hamiltonian remains valid. For the periodic case in question, there exists a solution that points to the presence of a gap in the density of states at the crossing of the magnetic level with a continuum of band states, the case illustrated in Fig. 2d. If the Fermi level is located in this gap, the bulk is insulating at low temperature, the case for SmB6 that possesses additionally topologically protected surface states conducting in the limit of zero temperature [71]. According to one of the scenarios, in heavy-fermion metals ( Chap. 15, “Metallic Magnetic Materials”), the Fermi level is located just above or below the hybridization gap, where the density of states is largely enhanced by a sizable component coming from magnetic shells that are periodically arranged but highly localized around parent atoms. How dilution of the magnetic constituent by nonmagnetic atoms affects

    936

    A. Bonanni et al.

    fragile coherence leading to the hybridization gap and in some cases to topological states remains largely an open question. Beyond the solubility limit: The stability of a particular alloy A1−x Bx against the decomposition is described by the mixing energy, i.e., the alloy energy in respect to the weighted energies of the end compounds:

    E(x) = EAB − (1 − x)EA − xEB .

    (1)

    Due to the aforementioned hybridization, d shells perturb bonding, which often makes E(x) convex, meaning that the alloy is unstable against binodal decomposition into the end compounds. In reality the degree of nanoscale nonuniformities in the alloy composition and in the local crystal structure, which set at a given growth or annealing temperature, are determined by complex interplay between energy, entropy, and diffusion barriers in the material volume and at the growth surface. These effects are known in the DMS literature as spinodal nanodecomposition, and their dynamics is described by the Cahn-Hilliard equation [3]. As a result of the spinodal nanodecomposition, electronic structure of magnetic dopants may vary on the nanoscale and depend more on growth and processing procedures than on an average chemical composition, as elaborated in section “Heterogenous Magnetic Semiconductors and Oxides”.

    Exchange Interactions Between Band and Localized Spins The key feature of dilute magnetic materials is the existence of strong spindependent coupling between magnetic ions and band carriers, which accounts for the unique low-energy properties – transport, optical, and thermodynamic characteristics of these systems. Neglecting non-scalar corrections that can appear for magnetic ions with nonzero orbital momentum, L = 0, a one-electron Hamiltonian that perturbs dynamics of the Bloch electrons assumes the Kondo form, HK =  K i Hi , where HiK = [V (r − R i ) − J (r − R i )s · S i ].

    (2)

    Here V (r − R i ) is a spin-independent potential change introduced by the magnetic impurity, and J (r − R i ) is a short-range exchange energy operator between the carrier spin s and the localized spin S residing at R i . In general, V (r − R i ) contains a long-range Coulomb part, present if the magnetic impurity is not isoelectronic, and a short-range alloy potential Vs (r − R i ), giving rise to a central-cell correction or a band offset in the impurity and alloy nomenclature, respectively. There are two mechanisms contributing to the spin-dependent part of the Kondo coupling J : (i) the exchange part of the Coulomb interaction between the effective mass and localized electrons, potential exchange which is FM, J > 0, and (ii) quantum hopping (hybridization) between band and local states resulting in a spindependent partial delocalization of electrons, which lowers their kinetic energy to

    18 Dilute Magnetic Materials Table 1 Magnitudes of the intra-atomic exchange energies J for singly ionized free atoms Mn+ and Eu+ , as collected by Dietl et al. [72]

    937 Mn+ 4s-3d 4p-3d 4d-3d

    meV 392 196 107

    Eu+ 6s-4f 6p-4f 5d-4f

    meV 52 33 215

    a degree dependent on the mutual spin orientation. This mechanism, referred to as the kinetic exchange, can be of either sign. It also contributes significantly to the magnitude of Vs . Potential exchange: Since the exchange part of the Coulomb interaction is shortranged, the potential exchange for the cation-substitutional magnetic impurities is sizable only if the carrier wave is appreciable at the cation sites. Table 1 contains exchange energies deduced from the energy differences ES+1/2 − ES−1/2 for an electron occupying various outer shells in free standing Mn+ (S = 5/2) and Eu+ (S = 7/2) ions. In the picture of ionic bonding, the value J4s−3d = 0.39 eV would determine entirely the exchange energy Js−d in the s-type band, as there is no hybridization between the Γ6 states and the levels derived from the d shell (eg and t2 states). Actually, the covalency (s–s hybridization) leads to spreading of the electron wave function between cation and anion sites. In accord with this insight, the values of Js−d (named N0 α in the DMS literature), as determined from magnetooptics experiments, are reduced to the range 190  N0 α  320 meV for a series of paramagnetic II-VI compounds doped with Mn as well as with Cr, Fe, and Co [73], in agreement with ab initio computations [74]. This reduction, as seen by observations of the spin-splitting, is even stronger in the case of Mn-based III-V DMSs, in which Mn ions are negatively charged [75]. In the case of Eu chalcogenides and mono-oxide, the magnitude of spin-splitting of the conduction band [76] points to the exchange energy J = 0.17 eV. This value is consistent with the data for Eu+ displayed in Table 1, as the wave function of conduction band electrons is primarily built of Eu 5d atomic orbitals. Kinetic exchange: By employing a canonical transformation, Schrieffer and Wolff demonstrated how to derive the kinetic part of the Kondo Hamiltonian from the Anderson Hamiltonian for impurities with a single open orbital embedded in a nonmagnetic metal. Since then a number of generalizations have been proposed, in particular taking into account realistic level structures of particular magnetic impurities but assuming that the carrier wave functions are in the form of plane waves [77]. The kinetic term overweighs the potential exchange and accounts for the AF sign of J leading to the Kondo effect in diluted magnetic metals ( Chap. 2, “Magnetic Exchange Interactions”). In the case of DMSs, hybridization between periodic parts uk of the Bloch wave functions and magnetic shells is considered. In particular, there is a strong hybridization between Γ8 and t2 states, which affects their relative position and leads to large values of |Jp−d ≡ N0 β| ≈ 1 eV collected for a number of tetrahedrally coordinated Mn-based DMSs in Fig. 3. If the relevant effective mass state is above the t2g level (the case of Mn-based DMSs), then β < 0, but otherwise β can be

    938

    A. Bonanni et al.

    Exchange energy, –N0β (eV)

    4 3

    β = – 0.054 eV nm3

    2

    ZnO

    CdS ZnSe CdSe ZnS ZnTe GaAs CdTe InAs

    1

    GaN

    0 ZnO

    -1

    GaN

    solid symbols: photoemission, XAS open symbols: magnetooptics

    -2 0

    10

    20

    30

    40

    50

    Cation concentration, N0 (nm-3) Fig. 3 Compilation of experimentally determined energies of the p-d exchange interaction −N0 β for various Mn-based DMSs as a function of the cation concentration N0 . Solid symbols denote the values evaluated from photoemission and x-ray absorption spectra. The values shown by open symbols were determined within the virtual-crystal and mean-field approximations (i.e., neglecting effects of strong coupling) from excitonic or band splittings in a magnetic field. Solid line corresponds to a constant value of β across the DMS series. (The data collected by Dietl and Ohno [23])

    positive [the case of Zn1−x Crx Se [78]]. These findings are supported by theoretical considerations [73, 74]. It should be emphasized that the role of hybridization depends crucially on the symmetry of band states and the parity of the localized state. The description of carrier-impurity interactions had to be reconsidered entirely in rock-salt DMSs, such as (Pb,Mn)Te [72], as well as when rare-earth atoms are involved, as in (Pb,Eu)Se [72]. Finally, we note that appropriately far away from the band extremum, the kp method [79] ceases to work and, additionally, the term exp(ikr) in the carrier wave function generates additional hybridization with the open magnetic shells.

    Effects of sp-d(f ) Exchange Interactions Spin-Splitting of Extended States: Weak Coupling Within the time-honored virtual-crystal approximation (VCA) and mean-field approximation (MFA), the host crystal symmetry is restored by replacing HK (Eq. 2) by its translationally invariant form. Taking into account only the short-range part in the spin-independent term,

    18 Dilute Magnetic Materials

    Hav =

    

    939

    x[Vs (r − R j ) − J (r − R j )s · S],

    (3)

    j

    where now j runs over all atoms in the sublattice containing magnetic ions (typically the cation sublattice) and S is a thermodynamic average value of spin operators Si . In terms of S, macroscopic spin magnetization M o (T , H ) = −g0 μB xN0 S, where g0 = 2.0. In the case of impurities with nonzero orbital momentum L (e.g., the rare-earth case [80]), M o (T , H ) = −gJ μB xN0 J , and S = (gJ − 1)J  ( Chap. 3, “Anisotropy and Crystal Field”). The above approach is suitable for describing the influence of magnetic impurities onto the spin-resolved band structure under two conditions: (i) there is no correlation in the distribution of magnetic ions, as the presence of any superstructure would lower point symmetry; (ii) the total potential introduced by a single impurity is too weak to give rise to bound states, so that the effect of HK on extended states can be treated perturbatively (weak coupling limit). Within a first-order perturbation theory and in the spirit of the multiband effective mass approximation [79, 41], the effect of Hav is described by matrix elements ui |Vs |ui  and ui |J |ui , where ui are periodic Kohn-Luttinger amplitudes corresponding to bands at the relevant band extremum. In the case of substitutional magnetic impurities in zinc-blende DMSs, for which typically bands with Γ6 , Γ7 , and Γ8 symmetries at k = 0 should be taken into account, there are four relevant matrix elements: V = uc |Vs |uc  and α = uc |J |uc ; W = uv |Vs |uv  and β = uv |J |uv  involve s-type (uc ) and p-type (uv ) periodic parts of the Bloch wave functions at k = 0, respectively. This formalism can be extended into lowdimensional nanostructures and topological materials, particularly by mapping the multiband effective mass approach onto a multiorbital tight-binding approximation [81, 82]. In general, due to kp coupling between bands and the spin-orbit interaction, the impurity-induced shift and spin-splitting of bands depend in a complex way on k, M 0 , and the angle between them. However, at k = 0 this shift is xN0 V and xN0 W for the Γ6 and Γ8 band, respectively, whereas the corresponding splitting is hω ¯ s = (h) αMo /gμB for the Γ6 band and h¯ ωs = βMo /gμB for the heavy-hole Γ8 band at k = 0. For x = 0.05, N0 β = −1 eV (see Fig. 3), and saturated spins S = 5/2, (h) |hω ¯ s | = 0.125 eV. A combination of the MFA and the effective mass formalism has been successfully applied to describe spin-splittings of bands and excitonic spectra in bulk, films, and various dimensionality quantum structures of DMSs as well as spinsplittings of Landau levels in these systems as a function of temperature and a magnetic field [41, 42, 43, 83]. Such studies have provided quantitative information on the magnitudes of sp–d(f ) exchange integrals in various families of DMSs; some examples are collected in Fig. 3. The VCA and MFA formalism can readily be employed to assess spin-splittings of boundary states specific to topological materials [84, 51]. When the relevant physics involves length scales greater than an average distance between magnetic ions (see  Chap. 7, “Micromagnetism”), it is convenient to work

    940

    A. Bonanni et al.

    within the continuous medium approximation. In the case under consideration, it assumes a form of the molecular-field approximation, in which the spin-independent part leads to a position-independent energy shift, whereas the spin-dependent part is given by Hsp−d(f ) = Jsp−d(f ) s · M(r)/N0 gμB ,

    (4)

    where Jsp−d(f ) is the relevant exchange energy (Js−d ≡ N0 α and Jp−d ≡ N0 β in tetrahedrally coordinated semiconductors) and M(r) is the local spin magnetization that shows spatial variations due to the presence of, for instance, magnetic domains or thermodynamic fluctuations of magnetization. According to Eq. 4, band carriers experience a molecular field H ∗ = Jsp−d(f ) M/N0 gg ∗ μ2B , where g ∗ is the carrier Landé factor. Similarly, the spins are in a molecular field generated by band carriers, H s = Jsp−d(f ) nP /N0 gμB , where n and P are carrier density and spin polarization, respectively. The field H s displaces the EPR frequency of localized spins (the Knight shift), which constitutes another way of determining the magnitude and sign of sp–d(f ) exchange energies [70]. This method is particularly useful if spin-orbit interactions make g ∗ and g different, ensuring the absence of a spin bottleneck [85]. Exchange spin-splitting of bands and associated spin polarization of carriers allows one to generate spin currents essential for spin injection, spin-transfer torque, GMR, TMR, and other spintronic capabilities of magnetic and dilute magnetic systems [23, 24].

    Spin-Splitting of Extended States: Strong Coupling If the local perturbation of the crystal potential introduced by a single impurity HiK is strong enough to bind a carrier even in the absence of a Coulomb potential, VCA and MFA do not describe adequately the influence of magnetic doping on the band states. This is particularly the case for nitrides and oxides, where a relatively short bond length results in a sizable p–d hybridization. In this strong coupling case, the spectral density of states Asz (ω) was determined by a non-perturbative generalized alloy theory for noninteracting Heisenberg spins. A maximum of Asz (ω) provides the position of the valence band top (if holes can be trapped) for particular spin orientations in respect to magnetization Mz (T , H ) of localized spins at given values of the integrals W and β as well as of the width of the potential well, which controls the binding ability of the neutral magnetic impurity [86]. This approach demonstrates that both the band shift and splitting can even assume an opposite sign to that expected within the VCA and MFA, as extended states are pushed away from the impurity, if the spectrum contains a bound state for a given spin orientation. As shown in Fig. 3, the values of β determined from excitonic splittings neglecting this effect appear as anomalous and do not follow chemical trends in the case of nitrides and oxides.

    18 Dilute Magnetic Materials

    941

    Another case is represented by magnetic impurities that are electrically active, acting as donors or acceptors, so that HiK contains also a long-range Coulomb potential. The effect of such impurities on spin-splitting of band states depends on whether the dopants are occupied or ionized. In the former case, a band carrier is a subject of exchange coupling not only to the spin of the magnetic ion but also to the spin of the carrier bound to it [87]. If magnetic donors or acceptors are ionized, the Coulomb potentials renormalize the probability of finding a band carrier in the vicinity of magnetic ions [87, 75]. In either case, the band spin-splitting and the apparent sp–d exchange integral can be substantially modified compared to cases when magnetic ions are not electrically active.

    Alloy and Spin-Disorder Scattering: Weak Coupling In general, the Hamiltonian describing alloy and spin-disorder scattering Hsc is determined by a difference between the full Kondo Hamiltonian (Eq. 2) and its VCA and MFA version (Eq. 3) serving to determine the band structure of magnetic alloys. In order to obtain the relevant scattering cross section, two principal approaches have been put forward. The first is the phase-shift theory introduced by Friedel as the d resonance model and then successfully applied to describe the residual resistivity [88], electron spin-resonance linewidth [89], and the electronic structure [90] of noble metals with transition-metal impurities. Another approach makes use of a perturbation theory whose weak coupling version corresponds to the Born approximation. A particularly transparent formula for the momentum relaxation time τ is obtained assuming that kB T is greater than energies of magnetic excitations as well as that the spin-splitting hω ¯ s is much smaller than εF ( Chap. 9, “Magnetotransport”). For cubic materials and s-type wave functions and in terms of the band shift V and exchange integral α one obtains for cubic materials and assuming that the z direction is along the magnetization Mo (T , H ), a derivative of the de Gennes-Friedel and Béal-Monod-Weiner formulae [91], 1/τ = 1/τal + 2/τsx + 1/τsz .

    (5)

    The alloy and spin-disorder scattering rates are given by 1/τal = π xN0 V 2 ρ(ε)/h; ¯ π α 2 ρ(ε) k B T χ⊥ ; 4hg ¯ 02 μ2B     π α 2 ρ(ε) x(1 − x) dMo (T , H ) 2 = k B T χ + , N0 dx 4hg ¯ 02 μ2B

    1/τsx =

    1/τsz

    where magnetic susceptibilities of localized spins

    (6)

    942

    A. Bonanni et al.

    χ⊥ = Mo (T , H )/H and χ = ∂Mo (T , H )/∂H,

    (7)

    enter the formulae for τsi via the fluctuation-dissipation theorem. The term in square brackets in Eq. 6 describes the effect of chemical disorder (inhomogeneous broadening). For spherical bands and in the 3D case, the density of states (DOS) ρ(ε) = m∗ k(ε)/π 2 h¯ 2 ; for a single electric subband in 2D systems, k(ε) should be replaced by π/L˜ W , where the participation length L˜ W = 1/

     dz|φ(z)|4 .

    (8)

    Here, φ(z) is the envelope function describing the carrier confinement, and the integration extends over the region where the magnetic ions reside. In the vicinity of the critical temperature for magnetic ordering, where χ tends to diverge, several additional effects have to be considered in evaluating τsi (T ), including the q dependence of χ , screening of magnetization fluctuations by carrier redistribution, and finite quantum coherence length [92]. A careful description of the pertinent effects will be essential in order to explain the behavior of resistance and its temperature derivative at criticality [93]. Since the exchange part of Hsc breaks spin rotation symmetry, spin-disorder scattering, together with spin-orbit effects (section “Interplay of sp–d(f ) Exchange Interactions and Spin-Orbit Coupling”), accounts for carrier spin relaxation in magnetic systems as well as controls the magnitude of quantum interference effects underlying Anderson-Mott localization and mesoscopic transport phenomena (section “Quantum Localization and Mesoscopic Phenomena: Colossal Magnetoresistance”).

    Alloy and Spin-Disorder Scattering: Strong Coupling A manifestation of strong coupling in spin-disorder scattering is the Kondo enhancement of the resistivity below a certain temperature, which appears for AF coupling between a single localized impurity spin and spins of the carrier sea (Fig. 2b and  Chap. 2, “Magnetic Exchange Interactions”). The effect has been widely observed in diluted magnetic metals but not yet in degenerate p-type diluted magnetic semiconductors, presumably because a relative small magnitude of DOS compared to metals shifts the Kondo phase to rather low temperatures in DMSs, where it competes with localization phenomena and spin-orbit effects.

    Bound Magnetic Polarons The concept of bound magnetic polarons (BMPs), spin complexes consisting of a donor electron and a bath of localized impurity spins residing within electron’s confinement region, was introduced by von Molnar and Kasuya in the context

    18 Dilute Magnetic Materials

    943

    of europium chalcogenides. Thus, BMPs exist in the strongly localized regime of the Anderson-Mott transition in magnetic and dilute magnetic semiconductors containing carriers or excitons trapped in singly occupied localized states [94]. Since a typical number of localized spins contributing to a BMP is greater than 10, they can be treated classically and described within the molecular-field approximation, i.e., characterized by the order parameter M(r). The starting point of the Dietl-Spałek model of BMPs [47] is the electron spin Hamiltonian s ·  with eigenvalues describing spin-split electron energies ± 12 . In this approach, the exchange part of

    s−d

    α = | g0 μB

     drM(r)|ψ(r)|2 |,

    (9)

    where α is the s–d exchange integral and ψ(r) is the electron envelope function. Hence, in the presence of a collinear magnetic field B and an average magnetization M 0 of bath spins S in the absence of donor electrons,  assumes the form  = g ∗ μB B +

     α (M 0 + bq ηq ). g0 μB q

    (10)

    Here g ∗ is the electron effective Landé factor; bq and ηq are Fourier components of |ψ(r)|2 and of bath magnetization fluctuations M(r) − M 0 , respectively. If the electron localization length is much longer than an average distance between the bath spins, the summation over q can be extended to infinity, but otherwise an appropriate cutoff qmax should be implemented. Except in the immediate vicinity of the spin ordering temperature, the probability distribution of ηq is, to a good approximation, Gaussian for any mixed state that can be described by spin temperature with variance, according to the fluctuationdissipation theorem, determined by an appropriate integral over ω of the imaginary part of χq (ω). Since  is linear in ηq , the central limit theorem implies that the probability distribution of  in the absence of the electron, PS (), is also Gaussian: PS () = ZS−1 exp[−( z − 0 )2 /2σ 2 − 2⊥ /2σ⊥2 ],

    (11)

    where ZS = (2π )3/2 σ σ⊥2 is the probability-normalizing constant insuring that dPS () = 1 and 2 = 2z + 2⊥ with the z-axis taken along the magnetic field. The three parameters characterizing cubic systems are then given by

    0 = g ∗ μB B +  2 σ (⊥)

    =

    α g0 μB

    2

    α Mo , g0 μB kB T

     q

    χ¯ (⊥)q |bq |2 /μ0 .

    (12) (13)

    944

    A. Bonanni et al.

    If energies of bath magnetic excitations are smaller than the thermal energy kB T and their q-dependencies are irrelevant, then χ¯ (⊥)q (T , H ) are given in Eq. 7. Far from magnetization saturation, χ⊥ = χ = χ . By adding the free energy of the electron, so that P() = Z −1 2 cosh( /2kB T )PS (),

    (14)

    the BMP effects in thermal equilibrium are taken into account with the partition function Z being determined by the probability normalization condition. Accordingly, the BMP free energy becomes F = −kB T ln(Z/ZS ), from which the BMP contribution to magnetization and specific heat is obtained. Furthermore, by minimizing F in respect to the BMP localization radius, self-trapping effects can be assessed [95]. Integrating P() over  directions, one obtains the probability distribution of the electron spin-splitting that can be directly probed by a variety of optical and magnetic resonance experiments. In particular, theoretically determined P( ) describes quantitatively the shape, width, and position of the spin-flip Raman scattering line in a number of n-type DMSs [96]. Additional inhomogeneous spin-splitting broadening caused by statistical fluctuations of composition x (chemical disorder), leading to different 0 values in particular BMPs, can also be incorporated into the theory [42]. In a paramagnetic phase and in the absence of an external magnetic field, P( ) attains a maximum at = 2εp and (8kB T εp )1/2 for εp kB T and εp kB T , respectively, where the polaron energy εp (T ) =

    α 2 χ (T ) 32πg02 μ2B aB3

    ,

    (15)

    where aB is an effective Bohr radius of the relevant carrier localized state. As seen, spontaneous spin-splitting appears even in the paramagnetic phase. It is dominated by the polaron effect, i.e., by bath spin polarization produced by the carrier spin at low temperatures, whereas thermodynamic magnetization fluctuations lead to a molecular field whose contribution to takes over in the high-temperature regime. Since non-scalar spin-spin interactions, such as Dzyaloshinskii-Moriya coupling, break spin momentum conservation, the BMP formation time is of the order of T2(S) of the bath spins. Hence, up to a time scale of T1(S) , adiabatic rather than isothermal magnetic susceptibilities describe energetics of the system [97]. Furthermore, by solving the quantum Liouville equation with the Hamiltonian Hs = sˆ ·  and averaging the resulting spin-density matrix with P(), one obtains information on dephasing of the electron spin for t < T2 at arbitrary values of M0 (T , H ) and magnetic field B [47].

    18 Dilute Magnetic Materials

    945

    Quantum Localization and Mesoscopic Phenomena: Colossal Magnetoresistance In addition to contributing to the momentum relaxation rate, spin-disorder scattering plays an essential role in controlling the magnitude of quantum coherence effects in charge transport ( Chap. 5, “Quantum Magnetism”), particularly in metals, where the DOS and, thus, 1/τsi are relatively large. Even residual magnetic impurities or spin-carrying defects in nominally ultra-pure noble metals perturb quantum interference phenomena below 100 mK [98]. The presence of giant spin-splitting affects in a dramatic way quantum localization [31] and mesoscopic phenomena [99] in DMSs. Also spin-disorder scattering in the paramagnetic phase, presumably involving preformed FM clusters, enhances localization and leads to colossal negative magnetoresistance (CMR) in the vicinity of TC [38, 31] ( Chap. 9, “Magnetotransport”). The presence of CMR is a trademark of magnetic and dilute magnetic semiconductors. It allowed testing of dynamic renormalization group equations at the magnetic field-induced AndersonMott insulator-to-metal transition [100]. At the same time, quantitative description of the mutual relationship between magnetism and Anderson-Mott localization is one of open issues in the physics of dilute magnetic materials.

    Interplay of sp–d(f ) Exchange Interactions and Spin-Orbit Coupling This interplay is a core resource for spintronic applications of magnetic materials, accounting for a range of magnetotransport [AMR, AHE ( [101],  Chap. 9, “Magnetotransport,” 6)] and magnetooptic ( Chap. 10, “Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films”) phenomena as well as the magnitude of critical currents for spin-transfer [102], spin-orbit torque [103], and spin-locking in topological materials [104,82]. In the spin-orbit Hamiltonian Hso = μB s·(E×v)/c, the relevant electric field E originates from atomic and impurity potentials, lack of inversion symmetry in some crystal structures or at interfaces, space-charge layers, or external sources, whereas the required velocity v from drift motion, pseudomomentum k, atomic orbital motion, and a real velocity in a crystal, often described in terms of the Berry curvature dependent on the topology of a given Brillouin zone ( Chap. 5, “Quantum Magnetism”). A number of these functionalities, particularly those insensitive to disorder, can be examined numerically by exploiting progress in the development of fully relativistic codes ( Chap. 10, “Magnetooptics and Laser-Induced Dynamics of/in Thin Metalic Films”). In the case of DMSs, particularly rewarding and numerically efficient has been the application of semiempirical multiband kp and multiorbital tight-binding approaches taking into account the presence of magnetic impurities within VCA and MFA [23, 24], the methodology generalized more recently

    946

    A. Bonanni et al.

    to predict the anomalous quantum Hall effect in topological materials [84, 51] (sections “p–d Zener Model” and “Dilute Ferromagnetic Insulators and Topological Insulators”). The incorporation of spin-orbit effects into theories of spin-disorder scattering, magnetic polarons, strong coupling, and quantum localization, particularly in the case of overlapping p-type subbands, is technically more demanding.

    Dominant Spin-Spin Interactions Dipole-Dipole Interactions Dipolar interactions are ubiquitous in nature and due to their long-range may dominate other types of interactions in the limit of low magnetic impurity concentrations, particularly in magnetic insulators with rare earth impurities, where competing exchange coupling decays rapidly with spin-spin distance. Due to randomness, the resulting ground state is a spin glass, as discussed in section “Spin-Glass Systems”.

    Direct Spin-Spin Interactions We consider pairs of interacting spins. In contrast to the direct Coulomb interaction (the Hartree term) that decays as inverse distance, the exchange coupling between spins (the Fock term) is nonzero (at least neglecting spin-orbit effects) only if there is an overlap between the relevant wave functions. This is the case of molecular dimers, such as Mn2 , for which the AF exchange integral comes from the kinetic exchange (quantum hopping between d orbitals of the two magnetic atoms, i.e., the direct interaction). The determined value of J for the molecular dimer Mn2 [105] is shown in Fig. 4 in comparison to AF exchange energies JNN of the nearest-neighbor (NN) Mn, Co, and Eu cation dimers in II-VI and IV-VI families of DMSs. The data for these impurities in various hosts point to a relatively slow decay of AF JN N with distance rN N . Furthermore, −JN N at given rNN is substantially larger for tetrahedral environment (II-VI DMSs) compared to the octahedral case (IV-VI DMSs). This suggests the presence of indirect, i.e., anion-mediated, exchange mechanisms in the case of TM impurities in compound semiconductors. In agreement with the expectation, the smallest magnitudes of AF exchange energies are observed in the case of Eu cation dimers, for which spins reside on the highly localized 4f shells. Surprisingly, however, in the end compounds, i.e., in Eu monochalcogenides, the nearest-neighbor interaction is FM. According to Kasuya’s model, in the presence of FM spin arrangement, d–f splitting of the conduction band lowers the energy denominator for quantum hopping between f and d states on the NN Eu ions, JNN ≈ xJdf |Vdf |2 /(Ed − Ef )2 . In the case of f orbitals and for x = 1 (i.e., giant splitting of the conduction band), this interaction is expected to be stronger than the competing x-independent AF coupling.

    18 Dilute Magnetic Materials

    947

    100

    Exchange energy - JNN (K)

    ZnS 10

    ZnSe ZnTe CdSe CdTe CdS HgSe HgTe PbSe

    1 PbTe

    0.1 10

    PbS

    20

    Mn2

    ZnO

    triangles: Co squares: Mn circles: Eu

    30

    40

    Cation concentration N0 (nm ) -3

    Fig. 4 Compilation of experimentally determined antiferromagnetic exchange integrals −JNN for the nearest-neighbor Mn, Co, and Eu cation dimers in various DMSs as a function of the cation concentration N0 , i.e., the inverse cation dimer distance for fcc lattices rNN = (4/N0 )1/3 /21/2 , as collected by Shapira [106], Savoyant et al. [107], and Story [108]. In the case of the wurtzite compounds, values corresponding to two nonequivalent dimers are shown. The convention, Hij = −2Jij S i ·S j , is employed. The magnitude of −J for Mn2 molecular dimer is also presented [105]

    Superexchange This indirect coupling between localized spins exists in virtually all compounds with localized spins and typically dominates in nonmetallic materials containing transition metals (TMs). The superexchange ( Chap, 2, “Magnetic Exchange Interactions”) proceeds via hybridization of magnetic levels with orbitals of atoms located between the spins in question, usually, but not always, p orbitals of relevant anions. For this indirect interaction, within the lowest-order perturbation theory in the hybridization, the interaction Hamiltonian is bilinear, Hij = −2S i Jˆij S j . Components of the exchange tensor Jˆij are proportional to |Vkd(f ) |4 and decay fast with the distance rij corresponding to NN, NNN, NNNN, ... spin pairs, typically on average exponentially, though not necessarily monotonically, as the direction of r ij in respect to the orientation of molecular orbitals is relevant, too. The scalar (isotropic) part of Jˆij [i.e., Jij I , where I is the identity (unit) matrix] usually dominates though the remaining components brought about by spin-orbit coupling, DM = −2D DM ·(S × particularly the Dzyaloshinskii-Moriya antisymmetric term, Hij i S j ), play often an important role in media with inversion asymmetry, also associated

    948

    A. Bonanni et al.

    with proximity to an interface. In general, group theory determines the form, if not the magnitude, of Jˆij for particular spin pairs and hosts.

    DMSs with Transition Metals Antiferromagnetic superexchange: A vast majority of nonmetallic TM compounds are antiferromagnets or ferrimagnets. Figure 4 presents the nearest-neighbor (NN) AF exchange energies JNN for Mn and Co dimers in II-VI and IV-VI DMSs, as determined from inelastic neutron scattering and magnetization steps in high magnetic fields in samples with a few percent of TM ions, as well as by modeling dependencies of magnetization and specific heat on a magnetic field and temperature. These experiments, particularly magnetization steps [106], provided also some information on Jij for more distant neighbors as well as, together with the EPR linewidth studies, on the Dzyaloshinskii-Moriya term whose magnitude increases on going from sulphides to selenides and tellurides [109]. In general, the employed ab initio and tight-binding approaches [73, 110, 107] describe satisfactorily the observed sign and magnitude of JNN . Ferromagnetic superexchange: According to the Goodenough-KanamoriAnderson rules ( Chap, 2, “Magnetic Exchange Interactions”), superexchange is FM for certain charge states of TM ions and bond arrangements. It was found employing a tight-binding approximation that Jij > 0 in the case of Cr2+ [111] and Mn3+ [33] ions in a tetrahedral environment. These expectations are corroborated by experimental results for (Ga,Mn)N, as discussed in section “Dilute Ferromagnetic Insulators and Topological Insulators” (Fig. 11). According to ab initio studies of Bi and Sb tetradymite chalcogenides [56], the coupling is FM for pairs of Ti, V, Cr, and Mn spins, whereas it assumes an AF character for Fe and Co pairs. If the computational approach properly describes the localized nature of magnetism in these systems, the determined chemical trend indicates that superexchange is the relevant interaction mechanism. IV-VI DMSs with Rare-Earth Metals In the most-studied Eu monochalcogenides, where spin S = 7/2 resides on the highly localized 4f shell, magnetic properties can be described in terms of FM NN and AF next NN (NNN) exchange integrals, J1 and J2 , respectively. Pressure studies revealed a strong dependence of Ji values on the lattice parameter and, −20 −10 thus, on the Eu pair distance rij , J1 (rNN ) ∝ RNN and J2 (rNNN ) ∝ rNNN [112]. As mentioned in section “Direct Spin-Spin Interactions”, according to Kasuya’s model, the direct Eu-Eu interaction involving quantum hopping between f and d states is relevant. At larger distances indirect AF coupling via anions takes over, which determines the magnitude of J2 , in agreement with the presence of p–f hybridization revealed by ab initio computations [113]. As shown in Fig. 4, in Eudoped lead chalcogenides, only antiferromagnetic coupling is observed [108], as the magnitude of Kasuya’s mechanism diminishes with decreasing x (section “Direct Spin-Spin Interactions”). A resonant enhancement of AF f –f coupling occurs when the Fermi level is pinned by donor 5d states degenerate with the valence band in p+ -(Sn,Gd)Te [70].

    18 Dilute Magnetic Materials

    949

    Carrier-Mediated Spin-Spin Coupling: Intra- and Interband Contributions RKKY Interaction As shown by Ruderman, Kittel, Kasuya, and Yosida, under the presence of a sp– d(f ) exchange interaction, a localized spin S i generates a spin polarization of the carrier sea that affects another localized spin S j ( Chap, 2, “Magnetic Exchange Interactions”). Since spin polarization shows Friedel oscillations (associated with the Kohn anomaly due to the Fermi energy cutoff), the sign of resulting bilinear coupling energy Jij , as given by the lowest-order perturbation theory in Jsp−d(f ) , changes between FM and AF as a function of rij with period λF /2, where λF is the de Broglie wavelength of carriers at εF ; the period is short in metals but substantially longer than an average distance between spins in doped semiconductors, where carrier density typically n < xN0 . The interaction amplitude is proportional to (Jsp−d(f ) /N0 )2 and the density of carrier states at εF and decays as rij−d , where d is the DOS dimensionality. Similarly to superexchange, spin-orbit interactions give rise to the presence of non-scalar terms in spin-spin coupling. Bloembergen-Rowland Mechanism This mechanism takes into account the presence of spin polarization of occupied bands generated by localized spins. Since here the perturbation theory involves interband matrix elements, this coupling is expected to be particularly strong in the inverted band structure cases of topological materials, such as (Hg,Mn)Te [114] and (Bi,Cr)2 Se3 [51]. In the case of (Hg,Mn)Te and related alloys and to the lowest order in perturbation theory that leads to bilinear coupling, Jij is ferromagnetic for the NN γ and antiferromagnetic for more distant pairs for which it decays as 1/rij , where γ varies between 4 and 5 in the HgTe type of systems [114], a much slower decrease with rij compared to superexchange. Furthermore, this mechanism is relevant for ptype tetrahedrally coordinated DMSs, in which quantum hopping between various valence subbands is allowed [55]. p–d Zener Model In the frame of the superexchange, RKKY, and Bloembergen-Rowland approaches, the thermodynamic properties of the system are inferred from the eigenvalues of the Hamiltonian which contains the relevant bilinear interactions S i Jˆij S j for all pairs of localized spins at a given value of the Fermi energy of the spin-unpolarized host. According to the p–d Zener model, implemented with mean-field and molecularfield approximations, one looks for the spin magnetization of localized magnetic ions M(r) that minimizes – for given T , H , and carrier density p – the total free energy functional F[M(r)]. This functional includes (i) a term related to the localized spin in the absence of carriers FS [M(r)] and (ii) the carrier contribution Fc [M(r)] in the presence of magnetization produced by localized spins [23]. If the energetics is dominated by spatially uniform magnetization M and the magnetic ions carry no orbital momentum (L = 0), the free energy density of localized spins in the magnetic field H is expressed as

    950

    A. Bonanni et al.

    

    M

    FS [M](T , H ) =

    dM o · h(M o ) − M · H .

    (16)

    0

    Here, h(M o ) is the inverse function of M o (h), where M o is the experimental macroscopic magnetization of the localized spins in the absence of carriers in the magnetic field h and at temperature T . In this way, the contribution due to intrinsic spin-spin interactions, like superexchange, is taken into account. The carrier-related term Fc [M(r)] is assessed within the multiband kp [79] or multiorbital tight-binding theory applicable for a given host including the sp–d(f ) coupling and the spin-orbit effects. The incorporation of spin-orbit interactions is imperative [19], as – together with single-ion anisotropy – it accounts for the magnetic anisotropy and for other specific properties of the magnetic systems. Within this formulation, intra-band and interband contributions (involving the RKKY and Bloembergen-Rowland mechanisms, respectively) are inherently accounted for. The model is valid for arbitrary M and allows taking account of strain, confinement, the presence of boundary states in topological materials, and the Landau quantization of the carrier spectrum. The carrier correlation effects are tied in, by introducing a Fermi-liquid-like parameter AF , which dilates the Pauli spin susceptibility of the hole liquid. Such effects would be arduous to treat within the RKKY and the BloembergenRowland approaches for three reasons. First, the spin-orbit coupling leads to non-scalar terms in the spin-spin Hamiltonian. Second, the host is assumed to be spin unpolarized, the approximation breaking down in a ferromagnetic phase. Third, there is an increasing amount of evidence that a description of magnetism in terms of a bilinear exchange Hamiltonian is inadequate if the range of spin-spin interactions is greater than the average distance between spins, as in the case of DFSs. The implementation of the p–d Zener model for the determination of various DFS properties and functionalities is now well established, at least for collinear magnetism and for systems not too close to the metal-insulator transition [23, 24]. Here below, we summarize the p–d Zener model of TC [19]; it is extensively reviewed elsewhere [23].

    Theory of Curie Temperature Near the Curie temperature TC and at H = 0, where M is small and the free energy is an even function of M, one expects Fc [M] − Fc [0] ∼ −M 2 . It is convenient to parameterize this dependence by a generalized carrier spin susceptibility χ˜ c , which is related to the magnetic susceptibility of the carrier liquid according to χc = AF (g ∗ μB )2 χ˜ c . In terms of χ˜ c , Fc [M] = Fc [0] − AF χ˜c β 2 M 2 /2(gμB )2 .

    (17)

    By expanding BS (M) for small M and introducing the spin susceptibility of the magnetic ions in the absence of carriers, χ˜ S = χ /(gμB )2 , one arrives at the meanfield formula allowing to determine TC :

    18 Dilute Magnetic Materials

    951

    AF β 2 χ˜ S (TC , q)χ˜ c (TC , q) = 1,

    (18)

    where β should be replaced by the s–d exchange integral α in the case of electrons (section “Exchange Interactions Between Band and Localized Spins”) and q denotes the Fourier component of the magnetization texture, for which TC attains the highest value. It is convenient to parametrize experimental values of χ˜ S by χ˜ S = xeff N0 /S(S + 1)/(T − Θp ), where xeff = x and Θp = 0 take into account coexisting superexchange interactions. For the spatially uniform magnetization (q → 0), one then obtains TC = TF + Θp ,

    (19)

    TF = xeff N0 S(S + 1)AF χ˜ c (TC )β 2 /3kB .

    (20)

    where TF is given by

    The magnitude of χ˜ c appearing in Eq. 18 can be determined from the linear response theory. Assuming that Jsp−d /N0 ≡ β in Eq. 4, so that only anion p bands are relevant [55]: χ˜ c = lim 2 q→0

     |ui,k |sM |uj,k+q |2 fi (k)[1 − fj (k + q)] Ej (k + q) − Ei (k)

    ij k

    ,

    (21)

    where the term i = j describes the intraband (RKKY-like) coupling, whereas the interband (Bloembergen-Rowland-like) interactions correspond to i = j terms; sM is the component of the spin operator along the direction of magnetization; ui,k is the periodic part of the Bloch wave function; and fi (k) is the Fermi-Dirac distribution function for the i-th band. In thin films, heterostructures, and superlattices, owing to the formation of interfacial space charge layers, the hole density and corresponding Curie temperatures TC [p(z)] are nonuniform even for a uniform distribution of acceptors and donors. The role of nonuniformity in the carrier distribution grows on reducing the thickness t of magnetic layers and is particularly relevant in those structures where p(z) can be tuned electrostatically, for instance, by the gate voltage. When t is larger than the phase coherence length Lφ , the region with the highest TC value determines TC of the whole structure. If, however, t < Lφ , the value of local magnetization M(z) and TC are determined by the distribution p(z) across the whole channel thickness. In this regime two situations are relevant. If disorder is strong  < t, the scattering broadening makes dimensional quantization irrelevant though quantum mechanical non-locality remains important. Under these conditions the magnitude of a layer’s TF can be expressed as [115]  TF =

     dzTF [p(z)]

     dzp(z)2 /[

    dzp(z)]2 ,

    (22)

    952

    A. Bonanni et al.

    where TF (p) is to be determined from the relevant 3D model and p(z) is to be evaluated from the Poisson equation taking into account the pinning of the Fermi energy by surface states. The opposite limit of weak disorder,  t, was also considered [116,22]. Owing to a typically large confinement-induced splitting between heavy- and light-hole subbands, only one ground-state heavy-hole subband is occupied, for which the p–d exchange is of the Ising type, Hpd = −N0 βsz Sz , so that TF = N0 xeff S(S + 1)AF β 2 m∗ /(12π h¯ 2 kB L˜ W ),

    (23)

    Here m∗ denotes the in-plane effective mass, and L˜ W is an effective width of the region occupied by carriers and spins relevant to ferromagnetism given in Eq. 8 [22, 117]. As seen, owing to a steplike form of DOS in the 2D case, TF does not depend on the hole density in this case. The expression for TF was generalized further to the case of arbitrary degeneracy of the hole liquid and by including effects of disorder via scattering broadening of DOS [118]. Theoretical approaches were developed to allow evaluation of the Curie temperature for ferromagnetic ordering of magnetic impurities mediated by Dirac carriers at the surface of 3D topological insulators [119, 120]. The Ising type of exchange was assumed, Hex = −N0 Jz sz Sz , leading to a gapped dispersion given by 2 1/2 ε(k) = ±[(Jz M/2gμB )2 + (hv ¯ f k) ] ,

    (24)

    where vf is the Fermi velocity. For such a case [119], in our notation, TF = N0 xeff S(S + 1)AF rJz2 (Ec − |εF |)/(24π h¯ 2 kB vf2 L˜ W ),

    (25)

    where r is the number of Dirac cones at a given surface, Ec is a cutoff energy associated with the termination of the Dirac surface band, and L˜ W is the penetration depth of Dirac carriers related to their envelop function according to Eq. 8. The formation of spin-density waves is expected in the case of carrier-mediated ferromagnetism in 1D systems [117].

    Magnetic Properties of Dilute Magnetic Materials Spin-Glass Systems The coexistence of randomness and frustration in spin-spin interactions ( Chap, 2, “Magnetic Exchange Interactions”) results in a distinct magnetic behavior known as spin-glass freezing. The randomness is typically brought about by dilution (frozen disorder), and therefore the spin glass is a low-temperature magnetic phase of a number of dilute magnetic materials [10].

    18 Dilute Magnetic Materials

    953

    A well-known example of a frustrated system is a triad of antiferromagnetically interacting spins located at corners of a triangle. Accordingly, spin-glass freezing is observed in, for instance, zinc-blende or wurtzite II-VI Mn-based DMSs such as Cd1−x Mnx Te, x  0.65 in which AF superexchange (section “DMSs with Transition Metals”) is the dominant exchange mechanism for all Mn-Mn distances, and the end compounds (x = 1) are AF [54, 106]. To the spin-glass family belong also other types of DMSs such as rock salt Sr1−x Eux S with x  0.5 [121] and spinel Cd(In1−x Crx )2 S4 with x  0.85 [122], in which the frustration is enforced by a competition between different interactions: the NN and the NNN exchange coupling is FM and AF, respectively, and the end compounds (x = 1) are FM (section “IV-VI DMSs with Rare-Earth Metals”). Another example is tetragonal LiHox Y1−x F4 with x  0.20 [123], characterized by (i) strong single-ion magnetic anisotropy that turns Ho ions into Ising spins, (ii) weak exchange coupling between them meaning that magnetism is driven by dipolar interactions that are FM or AM depending on the position of a given Ho pair in respect to the c-axis, and (iii) FM ordering at x = 1. Actually, spin-glass freezing was discovered and most thoroughly studied, not in semiconductors, but in dilute magnetic metals [10], such as CuMn, in which the RKKY exchange interaction operates (section “RKKY Interaction”). Despite their diversity, these materials exhibit a similar set of spin-related properties [10, 122, 12], including: (1) A characteristic cusp in the temperature dependence of magnetic susceptibility χ (T ), whose maximum in the static and H → 0 limits defines the freezing temperature Tf (Fig. 5); (2) Divergence of a coefficient of nonlinear magnetic susceptibility anl = limH →0 ∂ 2 χ (H )/∂H 2 at T → Tf+ ; (3) Broadening and shift of the cusp position to higher temperatures with the magnitude and frequency f of the magnetic field H , respectively; (4) A thermal equilibrium behavior of field-cooled magnetization MFC (T , H ) that shows no thermal hysteresis when quenching below Tf and annealing above Tf in a constant magnetic field H ; (5) Out-of-equilibrium magnetization dynamics below Tf in response to field changes, which fulfils the relation MFC (T , H = 0, t) + MZFC (T , H = 0, t) = MFC (T , H ), where MFC (T , H = 0, t) and MZFC (T , H = 0, t) denote timedependent magnetizations after removing and applying the magnetic field H , respectively, and the zero-field-cooled magnetization MZFC (T , 0, t) ≡ 0 in a spin-glass case (Fig. 5); (6) Other slow-dynamic glassy phenomena at T < Tf including: (i) aging, a dependence of the magnetization evolution rate on the waiting time, when the system was kept at given T < Tf and H , before H was changed; (ii) rejuvenation, by partial erasing of aging by abrupt lowering of temperature; and (iii) memory, the partial erasing of rejuvenation by returning to initial temperature; (7) No changes in the spin structure factor S(q, T ) or resistivity ρ(T ) on crossing Tf ;

    954

    A. Bonanni et al.

    Fig. 5 Magnetization in the vicinity of spin-glass freezing temperature Tf in CdCr1.7 In0.3 S4 . Zero-field-cooled (ZFC), field-cooled (FC), and thermoremanent (TMR) magnetizations in 10 Oe (10 kA/m/4π ) are shown. (From Vincent et al. [122])

    (8) No anomaly in the temperature dependence of spin-specific heat at Tf but rather a broad maximum at T > Tf . According to the present understanding [12] and in contrast to dielectric window glasses, spin glasses undergo at Tf a continuous phase transformation to a distinct thermal equilibrium spin state. This new phase is characterized by a nonzero value of the Edwards-Anderson order parameter: qEA = lim S i · S i (t), t→∞

    (26)

    where the bar denotes the configuration average with respect to random spin positions and the brackets denote the thermal or dynamical average. The existence of a phase transition at nonzero temperatures is supported by a careful experimental examination of static [124] and dynamic [125, 45] critical scaling behavior of magnetization as function of T , H , and f in the immediate vicinity of Tf for AgMn, Fe0.5 Mn0 .5 TiO3 , and (Cd,Mn)Te, respectively. Another relevant question is whether the lower critical dimensionality for Heisenberg spins is dl < 3, so that the observed phase transition in 3D cases is driven by magnetic anisotropy that effectively makes the spins Ising-like. Figure 6 shows Tf of CuMnPt and AgMnAu as a function of a magnetic anisotropy parameter A that increases with the concentration of Pt and Au impurities bringing a sizable spin-orbit interaction to this Heisenberg spin system [126]. As seen, the data point to nonzero values of Tf in the limit A → 0. A similar conclusion emerges from Fig. 7 showing comparable magnitudes of Tf in Mn-doped II-VI selenides and tellurides, despite the strength of the Dzyaloshinskii-Moriya interaction differing

    18 Dilute Magnetic Materials

    955

    Freezing temperature Tf (K)

    35 CuMnPt AgMnAu

    30

    25

    20

    15 0.0

    0.1

    0.2

    0.3

    0.4

    Magnetic anisotropy parameter A0.8 Fig. 6 Spin-glass freezing temperature Tf in CuMn4% and AgMn5.5% as a function of a magnetic anisotropy parameter A to power 0.8, changed by co-doping with Pt (concentration between 0 and 1.96%) and with Au (concentration between 0% and 5%), respectively. Straight lines are guides for the eye showing that Tf > 0 for A → 0. (Adapted from Fert et al. [126])

    by a factor of about two for these two families of Heisenberg DMSs [127]. Altogether these findings suggest that the spin-glass phase transition occurs also for the Heisenberg case. However, somewhat surprisingly critical exponents appear similar for Heisenberg [45] and Ising [125] spins. A great deal of information on spin glasses comes from Monte Carlo numerical simulations [11, 12]. Within the Edwards-Anderson model, classical spins residing in a cubic lattice are coupled by NN exchange interactions Jij , whose magnitudes are randomly selected from a Gaussian distribution with a zero mean value. There is a converging view that for such a model Tf > 0 for both Ising and Heisenberg spins in the 3D case [12, 128]. There is also a growing amount of evidence that spin-glass freezing of Heisenberg spins is accompanied by a chiral order, i.e., by a nonzero value of the chiral order parameter: (μ)

    qchiral = S i−μˆ · (S i × S i+μˆ ),

    (27)

    where μˆ is the unit lattice vector along x, y, or z [12, 128]. The presence of chiral ordering is thought to account for the lack of proportionality between the anomalous Hall resistivity and magnetization observed in spin glasses, such as AuMn [129]. Since a number of optimization problems can be mapped onto the determination (z) of thermal equilibrium spin orientations {Si } for a given set of interaction magnitudes {Jij }, progress in spin-glass computation tools has had a significant impact on other fields [13]. A particularly timely question concerns the effectiveness

    956

    A. Bonanni et al.

    x 1.0

    0.0156

    0.064

    Zn1-xMnxSe Zn1-xMnxTe Cd1-xMnxTe Cd1-xMnxSe Hg1-xMnxTe Hg1-xMnxSe (Cd1-xMnx)3As2

    Tf(K)

    10

    1

    0.1 1.0

    1.5

    2.0

    2.5

    3.0

    3.5

    4.0

    x-1/3 Fig. 7 Spin-glass freezing temperature Tf in various bulk II1−x Mnx -VI DMSs and (Cd1−x Mnx )3 As2 as a function of x −1/3 (points), as implied by the percolation theory for spin-spin couplings decaying exponentially with the pair distance [34]. Straight line is a guide for the eye. Relatively high values of Tf for inverted band structure semiconductors Hg1−x Mnx Te(Se) at x  0.1 and (Cd1−x Mnx )3 As2 point to the presence of an additional exchange interaction, presumably ferromagnetic interband Bloembergen-Rowland (Van Vleck) contribution (sections “Bloembergen-Rowland Mechanism” and “p–d Zener Model”). (The data points collected by Gała¸zka et al. [54])

    (z)

    of quantum annealing procedures [14] that aims at accelerating the search for Si  by adding a transverse field Hx , which means that i S (z) zi ceases to be a constant of motion and, thus, the ground state {Si(z) } for Hx → 0 can be achieved by quantum tunneling. On the theoretical side, two approaches have generated a considerable interest. One is the Sherrington-Kirkpatrick model, in which the probability distribution P(Jij ) is Gaussian but the range of interactions between classical Ising spins is infinite, so that a mean-field approach should be valid. Its implementation proceeds by employing a replica trick allowing, in the configurational averaging of the free energy over P(Jij ), the replacement of ln Z[Jij Si(z) Sj(z) ] by (Z[Jij Si(z) Sj(z) ])n , i.e., by n replicas of the system. In the time-honored and exact solution proposed by Parisi, order parameters involving different pairs of replicas (actually their overlap) are not identical (replica symmetry breaking – RSB). Within the RSB model, wandering of the coupled spin system between closely lying local free energy minima accounts for glass-like dynamics (hierarchical dynamics) [15]. An

    18 Dilute Magnetic Materials

    957

    Almeida-Thouless line separates the paramagnetic and spin-glass phase in the H –T plane. Another proposal, the droplet theory [16], is developed for short-range interactions. In this picture, there exists a unique spin ground state (and its time-reversal counterpart), whereas local deviations from it, in the form of compact and spatially uncorrelated droplet-like regions, constitute low-energy excitations accounting for the spin-glass dynamics. Exchange stiffness of spins adjacent to the droplet surface determines the droplet energy E that, because of randomness and frustration, increases rather slowly with the droplet radius R, E ∼ R θ , where θ < d/2. As the droplet magnetic moment m ∼ R d/2 , an arbitrarily small magnetic field H magnetizes droplets with appropriately large radii R ∼ H −2/(d−2θ) and, thus, destabilizes the spin-glass state. One of the signatures of glassy dynamics is the 1/f noise characterized by a power spectrum S(f ) ∼ 1/f γ , where 1  γ  2 is found experimentally in a variety of systems and expected theoretically within a number of disordered media models. While magnetization noise was detected directly in spin glasses [130], particularly accurate information was obtained by exploiting an outstanding sensitivity of mesoscopic electronic systems to spin configurations [131, 46]. As expected, studies of resistance as a function of time reveal 1/f noise over a wide frequency range in nanostructures of CuMn [131] and (Cd,Mn)Te [46] below Tf . However, more insight into the nature of spin glasses is provided by examining the Fourier transform of the time dependence of S(f ) determined over several consecutive time intervals (say a series of 70 S(f )’s, each extracted from resistance noise collected over consecutive 1-h intervals). Such a second spectrum s (2) (f ) is white for independent fluctuators (Gaussian noise) but assumes a nonwhite character if events of spin reconfigurations are correlated. As shown in Fig. 8, the fluctuators appear independent for Cd0.93 Mn0.07 Te [46]. However, some correlations are found for Cd0.8 Mn0.2 Te [46], and they are significantly stronger for Cu0.91 Mn0.09 [131]. These results lead to the conclusion that the hierarchical model of the spin-glass phase applies to dilute magnetic metals, in which spin-spin interactions are longrange, whereas the droplet picture is more appropriate for DMSs with short-range antiferromagnetic coupling.

    p-Type Dilute Ferromagnetic Semiconductors In virtually all semiconductor families, the coexistence of hole carriers with dilute Mn2+ ions results in ferromagnetic ordering at sufficiently low temperatures. This ferromagnetism is mediated by itinerant or weakly localized holes that are strongly coupled to Mn spins via p–d hybridization. In DFSs of III-V [(Ga,Mn)As, (In,Mn)As [20]] and V-VI [(Bi,Mn)2 Te3 [133]] compounds and also of elemental semiconductors [(Ge,Mn) [134]], the holes are provided by Mn ions themselves, which act as acceptors. Deviation from stoichiometry, such as cation vacancies, is another source of holes, the case of IV-VI DFSs [(Ge,Mn)Te [28,29], (Pb,Sn,Mn)Te

    958

    A. Bonanni et al.

    1.5

    Log second spectral density (Hz-1)

    1.0

    0.5 Cd0.8 Mn0.2Te 0.0

    Cd0.93 Mn0.07Te Gaussian background

    -0.5 10-3 2.0

    10-2

    10-1

    Cu0.9 Mn0.1 Cd0.8 Mn0.2 Te

    1.5 1.0 0.5 0.0 10-3

    10-2

    10-1

    Frequency f (Hz) (2,φ)

    Fig. 8 Upper panel: Second noise spectra in the phase domain sf of (Cd,Mn)Te quantum wires at 50 mK in the bandwidth from 0.1 to 0.6 Hz [46] (points) are compared to expectations for (2) the Gaussian 1/f noise (solid line). Lower panel: Frequency dependence of sf (normalizing to and subtracting the Gaussian expectation) for Cd0.8 Mn0.2 Te at 50 mK [46] and Cu0.91 Mn0.09 at 11 K [132] (points). Lines, except for the Gaussian background, are guides for the eye. (From Jaroszy´nski et al. [46])

    [21]] and also of I-II-V systems [Li(Zn,Mn)As [135]]. In certain compounds holes can be introduced by co-doping with anion- or cation-substitutional acceptors [e.g., (Cd,Mn)Te/(Cd,Mg,Zn)Te:N [22] and (K,Ba)(Zn,Mn)2 As2 [30], respectively]. In the case of (Ge,Mn)Te, ferromagnetism coexists with ferroelectricity [136]. These systems share a number of common features affecting their magnetic properties. First, the hole-mediated interaction competes with a short-range antiferromagnetic superexchange, so that there is an optimum Mn content for achieving the highest TC (so far reaching ∼200 K in (Ga,Mn)As, (Ge,Mn)Te [28, 29], and (K,Ba)(Zn,Mn)2 As2 [30]) which according to magnetization saturation corresponds to xeff  0.1.

    18 Dilute Magnetic Materials

    959

    Second, these materials show a strong tendency to phase separation (see section “Heterogenous Magnetic Semiconductors and Oxides”) – dilute Mn spins can coexist with nanocrystals of, for instance, ferromagnetic Mn-rich (Mn,Ga)As in (Ga,Mn)As [137] or antiferromagnetic MnTe in (Ga,Mn)Te [29]. Third, in addition to the solubility limit of Mn ions, there exists a selfcompensation mechanism with generation of donor defects (e.g., Mn interstitials or As antisites) that limit the achievable hole density (section “Heterogenous Magnetic Semiconductors and Oxides”). These and other defects may alter magnetic properties. Fourth, as in any doped semiconductor, carriers in DFSs undergo the AndersonMott metal-to-insulator transition (MIT) at pc ≈ (0.26/aB )3 , where the effective Bohr radius aB is diminished by coupling to localized spins, rather substantially in the strong coupling limit (section “Spin-Splitting of Extended States: Strong Coupling”). Nanoscale critical fluctuations in the carrier density on approaching and crossing the MIT result in the appearance of spin-glass (section “Spin-Glass Systems”) and superparamagnetic regions [31]. However, despite these materials issues, films of (Ga,Mn)As and related compounds, optimized by appropriate epitaxy and annealing procedures, show excellent micromagnetic properties, which has opened the door to demonstrating novel functionalities and to verifying the understanding of carrier-mediated ferromagnetism and spintronic phenomena in DFS-based devices [23, 24]. In particular, it has been found (by employing the independently determined exchange energy N0 β as well as kp or tight-binding parameters) that the p–d Zener model (sections “p–d Zener Model” and “p-Type Dilute Ferromagnetic Semiconductors”) even within MFA and neglecting disorder allows one to describe semiquantitatively and often quantitatively pertinent phenomena as a function of temperature, magnetic field, pressure, strain, and dimensionality [23, 24]. The studied properties include (i) thermodynamic characteristics such as Curie temperature, spin magnetization, specific heat, carrier spin polarization, and orbital magnetization; (ii) micromagnetic parameters including magnetic anisotropy, exchange stiffness, spin wave dispersion, domain-wall width and intrinsic resistance; and (iii) spintronic functionalities, e.g., electric-field control of magnetization, spin-transfer and spin-orbit torques, intrinsic component of the anomalous Hall effect, interlayer coupling, spin injection efficiency, current-induced domain-wall velocity and intrinsic pinning, MCD, TMR, and TAMR. The progress in semiconductor spintronics of p-type DFSs has triggered search for similar effects in metals that by supporting ferromagnetic ordering above room temperature are more suitable for applications. As an example, Fig. 9 presents the predicted and observed values of TC in p-type (III,Mn)V compounds. As seen, there is a good agreement in TC values and chemical trends. An analysis of theoretical results indicates that a contribution of interband terms involving valence band subbands (the Bloembergen-Rowland mechanism) to the magnitudes of TC is about 30%. For verification of how TC varies with the hole concentration p, particularly useful are studies of TC (p) in a single sample, as such a dependence is virtually independent of poorly known values of Θp in Eq. 19 and background concentrations of compensating donors. According to numerical results for the p–d Zener model

    960

    A. Bonanni et al.

    Fig. 9 Magnitudes of the Curie temperature TC predicted within the p–d Zener model for various p-type III-V semiconductors containing 5% of Mn and 3.5 × 1020 holes per cm3 (Eq. 20) compared to the highest experimentally determined values (upper and lower panel, respectively). (The data collected by Dietl [138])

    [55], γ = d ln TC /d ln p = 0.6–0.8 in the relevant region of hole densities. This prediction was confirmed experimentally by tracing the dependence TC (p) in (Ga,Mn)As [139] and (Ga,Mn)P [140] films irradiated by ions that produce holecompensating donor defects. However, detailed studies of changes in TC induced by the gate voltage Vg in metal-insulator-semiconductor (MIS) structures of (Ga,Mn)As [115, 141] led to an entirely different value, γ = d ln TC /d ln(−Vg ) = 0.19 ± 0.02 [141]. As shown in Fig. 10, this finding was elucidated by the p–d Zener model generalized to the case of a nonuniform hole distribution obtained by solving the Poisson equation in thin (Ga,Mn)As layers (Eq. 23), in which the Fermi level at the surface is pinned in the gap region by surface states [115, 141].

    Dilute Ferromagnetic Insulators and Topological Insulators In the course of the years, Ga1−x Mnx N has gained the status of a model dilute magnetic insulator, in which strong p–d coupling leads to tight binding of the hole at a Mn acceptor (the Zhang-Rice polaron) precluding the insulator-to-metal transition,

    18 Dilute Magnetic Materials

    insulator (Ga,Mn)As GaAs

    TC (K)

    82 80

    -3

    1.5 -7 MV/cm

    20

    84

    p(z) (10 cm )

    88 86

    961

    1.0

    78 76 74 72 0.5

    0.5 +7 MV/cm

    0.0

    z

    2 nm

    3D p-d Zener (uniform distribution)

    p-d Zener + non-uniform hole distribution 0.23 TC v p

    Experiment (Device B2) 0.21

    TC v p

    1

    1.5 20

    2

    2.5

    -3

    pS / t (10 cm ) Fig. 10 Curie temperature TC in a metal-oxide-semiconductor structure, in which the gate electric field EG changes the hole distribution (inset) and density (the areal hole concentration ps normalized to the thickness t of the (Ga,Mn)As channel). The solid line represents the generalized p-d Zener model for thin layers (Eq. 23), whereas the dotted line shows the dependence predicted by the p–d Zener model for the 3D case (Eq. 20). (From Nishitani et al. [141])

    at least, up to x  0.1. Magnetic properties of Ga1−x Mnx N depend strongly on the growth method and growth conditions [3]. In particular, spin-spin interactions are AF in samples containing a high concentration of Mn2+ ions, presumably due to the presence of compensating donor defects or impurities, such as Si or O. The strength of this AF interaction is similar to that observed in Mn-based II-VI DMSs , and the coupling is to be assigned to AF superexchange discussed in section “DMSs with Transition Metals”. In weakly compensated samples, where – according to intra-center optical transitions, EPR, XANES, and XES – Mn3+ ions prevail, low-temperature ferromagnetism was found with the TC magnitudes reaching about 13 K at x ≈ 10%, as shown in Fig. 11. The tight-binding model of the superexchange, which described successfully AF interaction specific to Mn2+ ions in II-VI DMSs [73], was used to compute exchange integrals Jij for Mn3+ pairs occupying 20 subsequent cation neighbor positions in GaN [33]. These values of Jij incorporated to Monte Carlo simulations provide the magnitudes of TC in agreement with experimental results (Fig. 11). The dependence TC (x) can also be obtained from the percolation theory, according to which overlapping sheers start to form a percolation cluster if their diameter rij ≥ 0.87(N0 x)−1/3 and spins are coupled if |Jij |  kB T [34]. Since the magnitudes of computed Jij decay approximately exponentially, Jij =

    962

    A. Bonanni et al.

    1000

    1

    0.125

    x

    0.0156

    0.008

    Curie temperature (K)

    experiment theory - superexchange TBA + Monte Carlo

    100

    Ga1-xMnxN 10

    1

    0.1

    1

    2

    3

    4

    5

    x-1/3 Fig. 11 Experimental Curie temperatures TC as a function of x −1/3 in Ga1 − xMnx N (symbols) as collected by [33] compared to theory within the tight-binding approximation and Monte Carlo simulations, serving to determine TC (stars [33]). The slope of the straight solid line agrees quantitatively with the percolation theory [34] for the magnitude of the exponential decay of spinspin coupling with the pair distance evaluated within the tight-binding approximation

    J0 exp(−rij /b), one obtains [34] TC (x) = T0 exp[−0.87(N0 x)−1/3 /b], where b = 0.11 nm in Ga1−x Mnx N, the expectations corroborated by experimental and Monte Carlo data (Fig. 11). This model implies also that all spins join the percolation cluster only in the limit T → 0, whereas ferromagnetic and superparamagnetic regions coexist at T > 0 in dilute ferromagnets with a finite range of spin-spin interactions, | drJ (r)| < ∞. An additional mechanism of FM coupling operates in topological materials. One of the consequences of the inverted band structure specific to these systems is a large contribution of anion p-type wave functions to Kohn-Luttinger amplitudes uik of both the valence and conduction bands. Accordingly, appreciable magnitudes of TF , described in Eqs. 20 and 21, can be expected from interband polarization (the Bloembergen-Rowland mechanism) even if the Fermi energy resides in the bandgap [51]. This FM interaction competes with AF superexchange in the case of Mn2+ ions and, as shown in Fig. 7, enhances the spin-glass freezing temperature of Hg1−x Mnx Te, Hg1−x Mnx Se, and (Cd1−x Mnx )3 As2 in the topologically nontrivial regime. Without strain, Hg1−x Mnx Te and Hg1−x Mnx Se are topological semimetals for x  0.06, whereas under tensile and compressive strain, they become topological insulators [142] and Weyl semimetals [53], respectively, in this low x limit.

    18 Dilute Magnetic Materials

    963

    [Cr0.11(BixSb1-x)0.89]2Te3

    Fig. 12 Experimental Curie temperatures TC as a function of the Bi concentration x as well as electron and hole carrier density in a thin film of [Cr0.11 (Bix Sb1−x )0.89 ]2 Te3 . (From Chang et al. [143])

    Non-cubic (Cd1−x Mnx )3 As2 , whose band structure is similar to compressively strained Hg1−x Mnx Te, is a Weyl semimetal over a wide Mn concentration range. The FM Bloembergen-Rowland mechanism works together with the superexchange in the case of early TM ions, for which the superexchange is theoretically expected to be FM for II-VI DMSs [73] as well as for V2 -VI3 DMSs, such as Bi2 Se3 , Bi2 Te3 , and Sb2 Se3 doped with Ti, V, or Cr [56]. Relatively high TC values observed in (Sb,Cr)2 Te3 [144] and particularly in (Sb,V)2 Te3 [145] confirm this view. A negligible role of bulk and boundary carriers in mediating FM coupling is further confirmed by results for [Cr0.11 (Bix Sb1−x )0.89 ]2 Te3 presented in Fig. 12, which demonstrate that TC is virtually independent of the Fermi-level position in the bands and across the bandgap filled up by topological boundary states [143]. The most striking development is the demonstration of accurate quantization of the Hall conductance |σxy | = e2 / h in thin films ferromagnetic V2 -VI3 topological insulators in the absence of an external magnetic field under conditions that charge transport proceeds only via the chiral edge state [50]. The precision of quantization improves with lowering temperature, reaching the level of 10−4 in the tens of mK range. The existence of superparamagnetic clusters in dilute ferromagnets, as mentioned above, accounts presumably for the presence of residual backscattering at T > 0.

    964

    A. Bonanni et al.

    Heterogenous Magnetic Semiconductors and Oxides The magnetic properties discussed above have concerned systems with a random distribution of cation-substitutional magnetic impurities, except for possible JahnTeller distortion for ions with a nonzero value of the orbital momentum, like the case of Mn3+ in GaN. A challenge but also a resource of dilute magnetic materials is a crucial dependence of their properties on the spatial distribution of magnetic ions and their position in the crystal lattice. These nanoscale structural characteristics depend, in turn, on the growth and processing protocols, as in the extensively studied field of magnetic metallic alloys and nanocomposites [146] ( Chap. 15, “Metallic Magnetic Materials”). By employing a range of photon, electron, and particle beam methods, with structural, chemical, and spin resolution down to the nanoscale, it has become possible to correlate surprising magnetic properties with the spatial arrangement and the electronic configuration of the magnetic constituent [3]. A number of striking properties of nonmetallic dilute magnetic materials discussed in this section stem from a contribution of the p–d hybridization to the binding energy, which results in attractive chemical forces between TM impurities. Numerous ab initio studies of TM-doped semiconductors and oxides, by determining the pairing energy of TM cation dimers [147, 3], have revealed that such forces exist in virtually all studied combinations of TM ions and nonmetallic hosts, except for Mn in II-VI chalcogenides in which Mn-induced states lie far from the Fermi level (see Fig. 1). These attractive forces can be overcompensated by Coulomb repulsion if co-doping with electrically active dopants changes the charge state of the magnetic impurities [148] though, in general, the effect of co-doping can be more intricate leading, for instance, to the formation of impurity complexes or affecting TM diffusion coefficients.

    Phase Separation Effects in (Ga,Mn)As It is instructive to take the canonical dilute ferromagnetic semiconductor Ga1−x Mnx As with a nominal Mn content x ≈ 5% as an example and to consider a palette of issues associated with the incorporation of Mn in GaAs [23, 3]: Cation-substitutional randomly distributed Mn ions: A 3D atomic probe analysis implies that at length scales larger than ∼1 nm, Mn ions are distributed randomly if growth is performed below 300 ◦ C by employing low-temperature molecular beam epitaxy (LT-MBE). These ions account for the ferromagnetic properties of (Ga,Mn)As. However, the magnitude of the saturation magnetization points systematically to xeff < x. An appropriate Ga/As ratio during the growth reduces substantially the concentration of As antisite defects that act as donors compensating the holes introduced by Mn. Interstitial Mn ions: According to combined Rutherford backscattering (RBS) and particle-induced x-ray emission (PIXE) measurements [149, 150], in samples obtained by LT-MBE, a significant fraction of Mn ions occupies interstitial positions in which they are antiferromagnetically coupled to substitutional Mn ions and form double donors compensating the holes mediating ferromagnetic coupling. Actually,

    18 Dilute Magnetic Materials

    965

    Fig. 13 Temperature dependence of remanent magnetization and resistivity (inset) for three Ga0.76 Al0.24 As/Ga1−x Mnx As/Ga0.76 Al0.24 As quantum well (QW) structures. The width of the QW is 5.6 nm, x = 0.06. Beryllium acceptors were introduced either into the first barrier (grown before the ferromagnetic QW, I-MDH) or into the second barrier (N-MDH), or the sample was undoped, as marked. The presence of holes originating from Be either enhances TC (N-MDH) or reduces it by generation of Mn interstitials during the epitaxy (I-MDH). (From Wojtowicz et al. [150])

    this self-compensation effect makes the formation of the interstitials energetically favorable, as holes in the bonding states (valence band) enlarge the system energy. For this reason, an increase of TC by modulation doping is only effective if the p-type barrier is deposited after the (Ga,Mn)As channel, as shown in Fig. 13. The interstitials can be driven toward the surface and neutralized by oxidation under low-temperature annealing Tan  200 ◦ C, which improves the ferromagnetic characteristics. Mn dimers: In the bulk zinc-blende structure, the crystallographic directions ¯ [110] and [110] are equivalent, whereas on the (001) surface, they are not. One of the consequences is a different energy of NN Mn dimers oriented along these ¯ two nonequivalent surface axes, the lower value corresponding to the [110] case, as there is an As atom below the surface connecting the two Mn ions, increasing their binding energy due to p–d hybridization. This means that epitaxial growth results ¯ in a surplus of [110]-oriented Mn dimers. This growth-related lowering of crystal symmetry is thought to account for the in-plane uniaxial magnetic anisotropy of (Ga,Mn)As [151]. The existence of this anisotropy is essential for a number of (Ga,Mn)As functionalities. Spinodal nanodecomposition: Epitaxial growth above 300 ◦ C or annealing above 400 ◦ C of films prepared by LT MBE results, according to TEM studies, in the

    966

    A. Bonanni et al.

    aggregation of Mn-rich (Mn,Ga)As zinc-blende nanocrystals embedded in Mnpoor GaAs, indicating that surface and volume diffusion barriers can be overcome at these temperatures, respectively. Though it is unknown whether (Mn,Ga)As is on the metal or insulator side of the Mott-Hubbard transition, the magnetic nanocrystals formed in this way are referred to as condensed magnetic semiconductors (CMSs). Under specific growth conditions (low Mn concentrations), (Ga,Mn)As decomposed by the chemical phase separation shows superparamagnetic characteristics. More often, however, superferromagnetic properties ( Chap. 20, “Magnetic Nanoparticles”) are observed. In particular, despite rather small nanocrystal diameters, typically below 6 nm, weakly temperature-dependent open magnetization loops persist up to 360 K [152]. This points to a rather high TC value of the individual nanocrystals, and also to the presence of coupling between them, perhaps mediated by a combination of strain and spin-orbit interactions [153]. In contrast to (Ga,Mn)As grown by LT MBE, this segregated system shows many appealing functionalities, such as large magnetooptical effects, persisting up to room temperature [154]. MnAs precipitation: At high growth or annealing temperatures Tan  600 ◦ C, CMSs in the form of hexagonal MnAs nanocrystals embedded coherently in zincblende GaAs (crystallographic phase separation) with diameters up to 500 nm are formed, showing magnetic properties and TC  318 K, specific to free-standing MnAs.

    Phase Separation Effects Beyond (Ga,Mn)As There are other relevant properties associated with the low solubility of TM ions in semiconductors and oxides. Precipitation of TM-rich compounds or TM grains: Examples here are metallic precipitates of Fe3 N in (Ga,Fe)N [155] and of Co in (Zn,Co)O [156], as revealed by TEM and x-ray diffraction (XRD). They account for ferromagnetic-like signatures in magnetic measurements persisting above room temperature and coexisting at low temperatures with a paramagnetic component originating from diluted TM ions. Effect of co-doping by electrically active impurities: In the above two cases, codoping by Si and Al donors, respectively, inhibits the precipitation, reducing the high-temperature ferromagnetic-like response. By contrast, according to energydispersive x-ray spectroscopy (EDS), co-doping with nitrogen acceptors stops the phase separation in (Zn,Cr)Te, whereas co-doping with iodine donors, presumably by compensating native acceptor defects, leads to an enhanced aggregation of Cr cations, which results in room-temperature ferromagnetic-like magnetic, magnetooptical, and magnetotransport properties [157]. Impurity complexes: The application of extended x-ray absorption fine structure (EXAFS) and x-ray emission spectroscopy (XES) demonstrated that co-doping of (Ga,Mn)N with Mg acceptors results in the formation of cation complexes MnMgk , where the number of Mg ions 0 ≤ k ≤ 3 bound to Mn determines the Mn charge and spin state, 3+ ≤ n ≤ 5+ and 2 ≥ S ≥ 1, respectively [158]. These changes in the Mn electronic configuration alter magnetic properties and intra-impurity optical transitions. In particular, Mn-Mg3 (Mn5+ ) complexes show strong broadband photoluminescence in the infrared, which persists to room

    967

    Pairing energy (eV)

    18 Dilute Magnetic Materials

    wz-GaN:TM

    Number of TMs in the cluster (n)

    Fig. 14 Computed pairing energies Ed as a function of the number n of TM cations in Cr, Mn, and Fe clusters in bulk wz-GaN (circles) and on Ga(0001) surface (triangles). According to these data, Mn ions will not, whereas Fe ions will aggregate during the epitaxy. (Adapted from Gonzalez Szwacki et al. [159])

    temperature, opening the door for laser applications in spectral windows relevant for telecommunication [158]. Surface aggregation: Epitaxial growth proceeds under conditions allowing for surface migration, which leads to aggregation of TM cations if the dimer pairing energy Ed is negative at the growth surface rather than in the bulk. According to ab initio data collected in Fig. 14, the values of Ed for Cr, Mn, and Fe cations in the bulk GaN are negative, and |Ed | is much larger than kB T under conditions of epitaxial growth of GaN. However, Ed becomes positive for Mn surface cation dimers. This may explain why the solubility limit of Mn compared to that of Fe is about one order of magnitude greater under the same growth conditions [159]. Formation of nanocolumns: There is a steady progress in controlling spinodal nanodecomposition in DMSs, particularly the nanocrystal size distribution as well as lateral and vertical ordering [47]. Remarkably, under specific growth conditions, the surface aggregation and associated mismatch strain lead in some systems to the self-organized growth of TM-rich nanocolumns embedded in the TM-poor host, an effect observed in (Al,Cr)N [161], (Ge,Mn), and (Zn,Cr)Te [3]. The resulting TM distribution is illustrated in Fig. 15 for the case of (Ge,Mn). These nanocolumns are typically metallic and thus can serve as nanocontacts (in, e.g., energy harvesting systems) or as active devices [3]. For instance, by interrupting the TM flow for a certain time during the epitaxy, tunnel barriers can be formed resulting in a dense array of nanoscale TMR junctions or single-electron transistors [162].

    968

    A. Bonanni et al.

    33x33x10 nm3

    28x28x50 nm3 Fig. 15 Visualization of Mn-rich nanocolumns in a Ge0.94 Mn0.06 film by 3D atomic probe. In the nanocolumns, Ge1−x Mnx with x  0.5 is formed. (Reproduced from Mouton et al. [160], with the permission of AIP Publishing)

    Acknowledgments The work of A. B. was supported by the Austrian Science Foundation, FWF (P31423 and P26830), and by the Austrian Exchange Service (ÖAD) project PL-01/2017. T. D. acknowledges a support by the Foundation for Polish Science through the IRA Programme financed by EU within SG OP Programme.

    References 1. Sorokin, E., Naumov, S., Sorokina, I.T.: Ultrabroadband infrared solid-state lasers. IEEE J. Sel. Top. Quant. Electr. 11, 690–712 (2005). https://doi.org/10.1109/JSTQE.2003.850255 2. Bonanni, A.: (Nano)characterization of semiconductor materials and structures. Semicon. Sci. Technol. 26, 060301 (2011). https://doi.org/10.1088/0268-1242/26/6/060301 3. Dietl, T., Sato, K., Fukushima, T., Bonanni, A., Jamet, M., Barski, A., Kuroda, S., Tanaka, M., Hai, P.N., Katayama-Yoshida, H.: Spinodal nanodecomposition in semiconductors doped with transition metals. Rev. Mod. Phys. 87, 1311–1377 (2015). https://doi.org/10.1103/ RevModPhys.87.1311 4. Abraham, D.W., Frank, M.M., Guha, S.: Absence of magnetism in hafnium oxide films. Appl. Phys. Lett. 87, 252502 (2005). https://doi.org/10.1063/1.2146057 5. Grace, P.J., Venkatesan, M., Alaria, J., Coey, J.M.D., Kopnov, G., Naaman, R.: The origin of the magnetism of etched Silicon. Adv. Mater. 21, 71 (2009). https://doi.org/10.1002/adma. 200801098 6. Matsubayashi, K., Maki, M., Tsuzuki, T., Nishioka, T., Sato, N.K.: Parasitic ferromagnetism in a hexaboride? Nature 420, 143 (2002). https://doi.org/10.1038/420143b 7. Makarova, T.L., Sundqvist, B., Höhne, R., Esquinazi, P., Kopelevich, Y., Scharff, P., Davydov, V., Kashevarova, L.S., Rakhmanina, A.V.: Retraction: magnetic carbon. Nature 440, 707 (2006). https://doi.org/10.1038/nature04622

    18 Dilute Magnetic Materials

    969

    8. Sawicki, M., Stefanowicz, W., Ney, A.: Sensitive SQUID magnetometry for studying nanomagnetism. Semicon. Sci. Technol. 26, 064006 (2011). https://doi.org/10.1088/02681242/26/6/064006 9. Coey, J.M.D., Venkatesan, M., Stamenov, P.: Surface magnetism of strontium titanate. J. Phys.: Condens. Matter 28, 485001 (2016). https://doi.org/10.1088/0953-8984/28/48/485001 10. Mydosh, J.A.: Spin glasses: redux: an updated experimental/materials survey. Rep. Progr. Phys. 78, 052501 (2015). https://doi.org/10.1088/0034-4885/78/5/052501 11. Binder, K., Young, A.P.: Spin glasses: experimental facts, theoretical concepts, and open questions. Rev. Mod. Phys. 58, 801–976 (1986). https://doi.org/10.1103/RevModPhys.58.801 12. Kawamura, H., Taniguchi, T.: Spin glasses. In: Buschow, K.H.J. (ed.) Handbook of Magnetic Materials, vol. 24, pp. 1–138. Elsevier, Amsterdam (2015). https://doi.org/10.0116/bs.hmm. 2015.08001 13. Stein, D.L., Newman, C.M.: Spin Glasses and Complexity. Princeton University Press, Princeton (2013). http://press.princeton.edu/titles/9917.html 14. Rønnow, T.F., Wang, Z., Job, J., Boixo, S., Isakov, S.V., Wecker, D., Martinis, J.M., Lidar, D.A., Troyer, M.: Defining and detecting quantum speedup. Science 345, 420–424 (2014). https://doi.org/10.1126/science.1252319 15. Mezard, M., Parisi, G., Virasoro, M.A.: Spin glass theory and beyond. World Scientific, Singapore (1987). https://doi.org/10.1142/0271 16. Fisher, D.S., Huse, D.A.: Equilibrium behavior of the spin-glass ordered phase. Phys. Rev. B 38, 386–411 (1988). https://doi.org/10.1103/PhysRevB.38.386 17. de Vegvar, P., Lévy, L.: Reply to comment on Conductance fluctuations of mesoscopic spin glasses. Phys. Rev. Lett. 68, 3485–3485 (1992). https://doi.org/10.1103/PhysRevLett.68.3485 18. Bouchiat, H., Dartyge, E., Monod, P., Lambert, M.: Diffuse x-ray determination of local atomic order in a spin-glass alloy system. Phys. Rev. B 23, 1375–1383 (1981). https://doi. org/10.1103/PhysRevB.23.1375 19. Dietl, T., Ohno, H., Matsukura, F., Cibert, J., Ferrand, D.: Zener model description of ferromagnetism in zinc-blende magnetic semiconductors. Science 287, 1019 (2000). https:// doi.org/10.1126/science.287.5455.1019 20. Ohno, H.: Making nonmagnetic semiconductors ferromagnetic. Science 281, 951 (1998). https://doi.org/10.1126/science.281.5379.951 21. Story, T., Gała¸zka, R.R., Frankel, R.B., Wolff, P.A.: Carrier-concentration–induced ferromagnetism in PbSnMnTe. Phys. Rev. Lett. 56, 777 (1986). https://doi.org/10.1103/PhysRevLett. 56.777 22. Haury, A., Wasiela, A., Arnoult, A., Cibert, J., Tatarenko, S., Dietl, T., d’Aubigne, Y.M.: Observation of a ferromagnetic transition induced by two-dimensional hole gas in modulation-doped CdMnTe quantum wells. Phys. Rev. Lett. 79, 511 (1997). https://doi.org/ 10.1103/PhysRevLett.79.511 23. Dietl, T., Ohno, H.: Dilute ferromagnetic semiconductors: physics and spintronic structures. Rev. Mod. Phys. 86, 187–251 (2014). https://doi.org/10.1103/RevModPhys.86.187 24. Jungwirth, T., Wunderlich, J., Novák, V., Olejník, K., Gallagher, B.L., Campion, R.P., Edmonds, K.W., Rushforth, A.W., Ferguson, A.J., Nˇemec, P.: Spin-dependent phenomena and device concepts explored in (Ga,Mn)As. Rev. Mod. Phys. 86, 855–896 (2014). https:// doi.org/10.1103/RevModPhys.86.855 25. Olejník, K., Owen, M.H.S., Novák, V., Mašek, J., Irvine, A.C., Wunderlich, J., Jungwirth, T.: Enhanced annealing, high Curie temperature and low-voltage gating in (Ga,Mn)As: a surface oxide control study. Phys. Rev. B 78, 054403 (2008). https://doi.org/10.1103/PhysRevB.78. 054403 26. Wang, M., Campion, R.P., Rushforth, A.W., Edmonds, K.W., Foxon, C.T., Gallagher, B.L.: Achieving high Curie temperature in (Ga,Mn)As. Appl. Phys. Lett. 93, 132103 (2008). https:// doi.org/10.1063/1.2992200 27. Chen, L., Yang, X., Yang, F., Zhao, J., Misuraca, J., Xiong, P., von Molna`r, S.: Enhancing the Curie temperature of ferromagnetic semiconductor (Ga,Mn)As to 200 K via nanostructure engineering. Nano Lett. 11, 2584 (2011). https://doi.org/10.1021/nl201187m

    970

    A. Bonanni et al.

    28. Fukuma, Y., Asada, H., Yamamoto, J., Odawara, F., Koyanagi, T.: Large magnetic circular dichroism of Co clusters in Co-doped ZnO. Appl. Phys. Lett. 93, 142510 (2008) 29. Hassan, M., Springholz, G., Lechner, R.T., Groiss, H., Kirchschlager, R., Bauer, G.: Molecular beam epitaxy of single phase GeMnTe with high ferromagnetic transition temperature. J. Cryst. Growth 323, 363 (2011). https://doi.org/10.1016/j.jcrysgro.2010.10.135 30. Zhao, K., Chen, B., Zhao, G., Yuan, Z., Liu, Q., Deng, Z., Zhu, J., Jin, C.: Ferromagnetism at 230 K in (Ba0.7 K0.3 )(Zn0.85 Mn0.15 )2 As2 diluted magnetic semiconductor. Chin. Sci. Bull. 59, 2524–2527 (2014). https://doi.org/10.1007/s11434-014-0398-z 31. Dietl, T.: Interplay between carrier localization and magnetism in diluted magnetic and ferromagnetic semiconductors. J. Phys. Soc. Jpn. 77, 031005 (2008). https://doi.org/10.1143/ JPSJ.77.031005 32. Richardella, A., Roushan, P., Mack, S., Zhou, B., Huse, D.A., Awschalom, D.D., Yazdani, A.: Visualizing critical correlations near the metal-insulator transition in Ga1−x Mnx As. Science 327, 665 (2010). https://doi.org/10.1126/science.1183640 33. Stefanowicz, S., Kunert, G., Simserides, C., Majewski, J.A., Stefanowicz, W., Kruse, C., Figge, S., Li, T., Jakieła, R., Trohidou, K.N., Bonanni, A., Hommel, D., Sawicki, M., Dietl, T.: Phase diagram and critical behavior of a random ferromagnet Ga1−x Mnx N. Phys. Rev. B 88, 081201(R) (2013). https://doi.org/10.1103/PhysRevB.88.081201 34. Korenblit, I.Y., Shender, E.F., Shklovsky, B.I.: Percolation approach to the phase transition in the very dilute ferromagnetic alloys. Phys. Lett. A 46, 275 (1973). https://doi.org/10.1016/ 0375-9601(73)90219-3 35. Methfessel, S., Mattis, D.C.: Magnetic semiconductors. In: Flugge, S. (eds.) Handbuch der Physik, vol. 18, p. Pt. 1. Springer, Berlin (1968) 36. Kioseoglou, G., Hanbicki, A.T., Sullivan, J.M., van ’t Erve, O.M.J., Li, C.H., Erwin, S.C., Mallory, R., Yasar, M., Petrou, A., Jonker, B.T.: Electrical spin injection from an n-type ferromagnetic semiconductor into a III-V device heterostructure. Nat. Mater. 3, 799–803 ( 2004). https://doi.org/10.1038/nmat1239 37. Wachter, P.: Europium chalcogenides: EuO, EuS, EuSe and EuTe. In: Gschneidner, K.A. Jr., Eyring, L. (eds.) Handbook on the Physics and Chemistry of Rare Earth, vol. 2, p. 507. NorthHolland, Amsterdam (1979). https://doi.org/10.1016/S0168-1273(79)02010-9 38. Nagaev, E.L.: Colossal-magnetoresistance materials: manganites and conventional ferromagnetic semiconductors. Phys. Rep. 346, 387 (2001). https://doi.org/10.1016/S03701573(00)00111-3 39. Andrearczyk, T., Jaroszy´nski, J., Sawicki, M., Van Khoi, L., Dietl, T., Ferrand, D., Bourgognon, C., Cibert, J., Tatarenko, S., Fukumura, T., Jin, Z., Koinuma, H., Kawasaki, M.: Ferromagnetic interactions in p- and n-type II-VI diluted magnetic semiconductors. In: Miura, N., Ando, T. (eds.) Proceedings 25th International Conference on Physics of Semiconductors, Osaka, 2000, p. 235. Springer, Berlin, (2001) 40. Gała¸zka, R.R.: Semimagnetic semiconductors. In: Wilson, B.L.H. (ed.) Proceedings 14th International Conference on the Physics of Semiconductors, Edinburgh 1978, p. 133. IoP, Bristol (1978) 41. Furdyna, J.K., Kossut, J.: Diluted Magnetic Semiconductors. Semiconductors and Semimetals, vol. 25. Academic Press, New York (1988) 42. Dietl, T.: (Diluted) magnetic semiconductors. In: Mahajan, S. (ed.) Handbook of Semiconductors, vol. 3B, p. 1251. North Holland, Amsterdam (1994) 43. Gaj, J.A., Kossut, J. (eds.): Introduction to the Physics of Diluted Magnetic Semiconductors. Springer, Berlin (2010). https://doi.org/10.1007/978-3-642-15856-8 44. Fiederling, R., Keim, M., Reuscher, G., Ossau, W., Schmidt, G., Waag, A., Molenkamp, L.W.: Injection and detection of a spin-polarized current in a light-emitting diode. Nature 402, 787 (1999). https://doi.org/10.1038/45502 45. Leclercq, B., Rigaux, C., Mycielski, A., Menant, M.: Critical dynamics in Cd1−x Mnx Te spin glasses. Phys. Rev. B 47, 6169–6172 (1993). https://doi.org/10.1103/PhysRevB.47.6169 46. Jaroszy´nski, J., Wróbel, J., Karczewski, G., Wojtowicz, T., Dietl, T.: Magnetoconductance noise and irreversibilities in submicron wires of spin-glass n+ -Cd1−x Mnx Te. Phys. Rev. Lett. 80, 5635 (1998). https://doi.org/10.1103/PhysRevLett.80.5635

    18 Dilute Magnetic Materials

    971

    47. Dietl, T.: Spin dynamics of a confined electron interacting with magnetic or nuclear spins: a semiclassical approach. Phys. Rev. B 91, 125204 (2015). https://doi.org/10.1103/PhysRevB. 91.125204 48. Jungwirth, T., Marti, X., Wadley, P., Wunderlich, J.: Antiferromagnetic spintronics. Nat. Nanotech. 11, 231–241 (2016). https://doi.org/10.1038/nnano.2016.18 49. Kobak, J., Smole´nski, T., Goryca, M., Papaj, M., Bogucki, A., Gietka, K., Koperski, M., Rousset, J.-G., Suffczy´nski, J., Janik, E., Nawrocki, M., Golnik, A., Kossacki, P., Pacuski, W.: Designing quantum dots for solotronics. Nat. Commun. 5, 3191 (2014). https://doi.org/ 10.1038/ncomms4191 50. Chang, C.-Z., Zhang, J., Feng, X., Shen, J., Zhang, Z., Guo, M., Li, K., Ou, Y., Wei, P., Wang, L.-L., Ji, Z.-Q., Feng, Y., Ji, S., Chen, X., Jia, J., Dai, X., Fang, Z., Zhang, S.-C., He, K., Wang, Y., Lu, L., Ma, X.-C., Xue, Q.-K.: Experimental observation of the quantum anomalous Hall effect in a magnetic topological insulator. Science 340, 167–170 (2013). https://doi.org/10. 1126/science.1234414 51. Yu, R., Zhang, W., Zhang, H.-J., Zhang, S.-C., Dai, X., Fang, Z.: Quantized anomalous Hall effect in magnetic topological insulators. Science 329, 61–64 (2010). https://doi.org/10.1126/ science.1187485 52. Fan, Y., Kou, X., Upadhyaya, P., Shao, Q., Pan, L., Lang, M., Che, X., Tang, J., Montazeri, M., Murata, K., et al.: Electric-field control of spin–orbit torque in a magnetically doped topological insulator. Nat. Nanotechnol. 352 (2016). https://doi.org/10.1038/nnano.2015.294 53. Bulmash, D., Liu, C.-X., Qi, X.-L.: Prediction of a Weyl semimetal in Hg1−x−y Cdx Mny Te. Phys. Rev. B 89, 081106 (2014). https://doi.org/10.1103/PhysRevB.89.081106 54. Gała¸zka, R.R.: II–VI compounds – Polish perspective. Phys. Stat. Sol. (B) 243, 759 (2006). https://doi.org/10.1002/pssb.200564601 55. Dietl, T., Ohno, H., Matsukura, F.: Hole-mediated ferromagnetism in tetrahedrally coordinated semiconductors. Phys. Rev. B 63, 195205 (2001). https://doi.org/10.1103/PhysRevB. 63.195205 56. Vergniory, M.G., Otrokov, M.M., Thonig, D., Hoffmann, M., Maznichenko, I.V., Geilhufe, M., Zubizarreta, X., Ostanin, S., Marmodoro, A., Henk, J., Hergert, W., Mertig, I., Chulkov, E.V., Ernst, A.: Exchange interaction and its tuning in magnetic binary chalcogenides. Phys. Rev. B 89, 165202 (2014). https://doi.org/10.1103/PhysRevB.89.165202 57. Bonanni, A., Dietl, T.: A story of high-temperature ferromagnetism in semiconductors. Chem. Soc. Rev. 39, 528 (2010). https://doi.org/10.1039/b905352m 58. Fukushima, T., Sato, K., Katayama-Yoshida, H., Dederichs, P.H.: Spinodal decomposition under layer by layer growth condition and high Curie temperature quasi-one-dimensional nano-structure in dilute magnetic semiconductors. Jpn. J. Appl. Phys. 45, L416 (2006). https:// doi.org/10.1143/JJAP.45.L416 59. Ludwig, G.W., Woodbury, H.H.: Electron spin resonance in semiconductors, pp. 223–304. Academic Press, New York (1962) 60. Zunger, A.: Electronic structure of 3d transition-atom impurities in semiconductors. In: Seitz, F., Turnbull, D. (eds.) Solid State Physics, vol. 39, p. 275. Academic Press, New York (1986) 61. Langer, J.M., Delerue, C., Lannoo, M., Heinrich, H.: Transition-metal impurities in semiconductors and heterojunction band lineups. Phys. Rev. B 38, 7723 (1988). https://doi.org/10. 1103/PhysRevB.38.7723 62. Nolte, D.D.: Semi-insulating semiconductor heterostructures: optoelectronic properties and applications. J. Appl. Phys. 85, 6259 (1999). https://doi.org/10.1063/1.370284 63. Khomskii, D.I.: Transition Metal Compounds. Cambridge University Press, Cambridge (2014). https://doi.org/10.1017/CBO9781139096782 64. Pacuski, W., Suffczy´nski, J., Osewski, P., Kossacki, P., Golnik, A., Gaj, J.A., Deparis, C., Morhain, C., Chikoidze, E., Dumont, Y., Ferrand, D., Cibert, J., Dietl, T.: Influence of s,p-d and s-p exchange couplings on exciton splitting in Zn1−x Mnx O. Phys. Rev. B 84, 035214 (2011). https://doi.org/10.1103/PhysRevB.84.035214 65. Belitz, D., Kirkpatrick, T.R.: The Anderson-Mott transition. Rev. Mod. Phys. 66, 261 (1994). https://doi.org/10.1103/RevModPhys.66.261

    972

    A. Bonanni et al.

    66. Byczuk, K., Hofstetter, W., Vollhardt, D.: Anderson localization vs. Mott-Hubbard metalinsulator transition in disordered, interacting lattice fermion systems. J. Mod. Phys. B 24, 1727 (2010). https://doi.org/10.1142/S0217979210064575 67. Shklovskii, B.I., Efros, A.L.: Electronic Properties of Doped Semiconductors, Springer, Berlin (1984). https://doi.org/10.1007/978-3-662-02403-4 ´ ¸ tek, K., Dietl, T., Kossut, J.: Resonant states in semiconductors: a 68. Wilamowski, Z., Swia quantitative study of HgSe:Fe. Solid State Commun. 74, 833 (1990). https://doi.org/10.1016/ 0038-1098(90)90945-8 69. Sawicki, M., Dietl, T., Plesiewicz, W., Sekowski, P., Sniadower, L., Baj, M., Dmowski, L.: Influence of an acceptor state on transport in zero-gap Hg1−x Mnx Te. In: Landwehr, G. (ed.) Application of High Magnetic Fields in Physics of Semiconductors, p. 382. Springer, Berlin (1983). https://doi.org/10.1007/3-540-11996-5_55 70. Story, T., Górska, M., Łusakowski, A., Arciszewska, M., Dobrowolski, W., Grodzicka, E., Gołacki, Z., Gała¸zka, R.R.: New mechanism of f –f exchange interactions controlled by Fermi level position. Phys. Rev. Lett. 77, 3447–3450 (1996). https://doi.org/10.1103/ PhysRevLett.77.3447 71. Dzero, M., Xia, J., Galitski, V., Coleman, P.: Topological Kondo insulators. Ann. Rev. Conden. Matter Phys. 7, 249–280 (2016). https://doi.org/10.1146/annurev-conmatphys-031214014749 ´ 72. Dietl, T., Sliwa, C., Bauer, G., Pascher, H.: Mechanisms of exchange interactions between carriers and Mn or Eu spins in lead chalcogenides. Phys. Rev. B 49, 2230 (1994). https://doi. org/10.1103/PhysRevB.49.2230 73. Kacman, P.: Spin interactions in diluted magnetic semiconductors and magnetic semiconductor structures. Semicond. Sci. Technol. 16, R25 (2001). https://doi.org/10.1088/0268-1242/ 16/4/201 74. Chanier, T., Virot, F., Hayn, R.: Chemical trend of exchange coupling in diluted magnetic II-VI semiconductors: ab initio calculations. Phys. Rev. B 79, 205204 (2009). https://doi.org/ 10.1103/PhysRevB.79.205204 75. Adhikari, R., Stefanowicz, W., Faina, B., Capuzzo, G., Sawicki, M., Dietl, T., Bonanni, A.: Upper bound for the s-d exchange integral in n-(Ga,Mn)N:Si from magnetotransport studies. Phys. Rev. B 91, 205204 (2015). https://doi.org/10.1103/PhysRevB.91.205204 76. Steeneken, P.G., Tjeng, L.H., Elfimov, I., Sawatzky, G.A., Ghiringhelli, G., Brookes, N.B., Huang, D.-J.: Exchange splitting and charge carrier spin polarization in EuO. Phys. Rev. Lett. 88, 047201 (2002). https://doi.org/10.1103/PhysRevLett.88.047201 77. Irkhin, V.Y., Irkhin, Y.P.: Hybridization and Kondo effect in systems with degenerate d and f shells. Zh. Eksp. Teor. Fiz. 107, 616–627 (1995). www.jetp.ac.ru/cgi-bin/dn/e_080_02_0334. pdf 78. Mac, W., Khoi, N.T., Twardowski, A., Gaj, J.A., Demianiuk, M.: Ferromagnetic p–d exchange in Zn1-x Crx Se diluted magnetic semiconductor. Phys. Rev. Lett. 71, 2327–2330 (1993). https://doi.org/10.1103/PhysRevLett.71.2327 79. Winkler, R.: Spin-Orbit Coupling Effects in Two-Dimensional Electron and Hole Systems. Springer Tracts in Modern Physics, vol. 191 Springer, Berlin (2003). Note in this book a convention is adopted according to which a sign of the sp-d Hamiltonian is reversed. http:// www.springer.com/gp/book/9783540011873 80. Jensen, J., Mackintosh, A.R.: Rare Earth Magnetism. Structures and Excitations. The International Series of Monographs on Physics Clarendon Press, Oxford (1991). www.fys. ku.dk/~jjensen/Book/Ebook.pdf 81. Sankowski, P., Kacman, P., Majewski, J.A., Dietl, T.: Spin-dependent tunneling in modulated structures of (Ga,Mn)As. Phys. Rev. B 75, 045306 (2007). https://doi.org/10.1103/PhysRevB. 75.045306 82. Qi, X.-L., Zhang, S.-C.: Topological insulators and superconductors. Rev. Mod. Phys. 83, 1057–1110 (2011). https://doi.org/10.1103/RevModPhys.83.1057 83. Bauer, G., Pascher, H., Zawadzki, W.: Magneto-optical properties of semimagnetic lead chalcogenides. Semicond. Sci. Technol. 7, 703 (1992). https://doi.org/10.1088/0268-1242/ 7/6/001

    18 Dilute Magnetic Materials

    973

    84. Liu, C.-X., Qi, X.-L., Dai, X., Fang, Z., Zhang, S.-C.: Quantum anomalous Hall effect in Hg1−y Mny Te quantum wells. Phys. Rev. Lett. 101, 146802 (2008). https://doi.org/10.1103/ PhysRevLett.101.146802 85. Barnes, S.E.: Theory of electron spin resonance of magnetic ions in metals. Adv. Phys. 30, 801–938 (1981). https://doi.org/10.1080/00018738100101447 86. Dietl, T.: Hole states in wide band-gap diluted magnetic semiconductors and oxides. Phys. Rev. B 77, 085208 (2008). https://doi.org/10.1103/PhysRevB.77.085208 ´ 87. Sliwa, C., Dietl, T.: Electron-hole contribution to the apparent s–d exchange interaction in IIIV dilute magnetic semiconductors. Phys. Rev. B 78, 165205 (2008). https://doi.org/10.1103/ PhysRevB.78.165205 88. Toyoda, T., Kume, K.: Residual resistivities of Au- or Cu-based 4d transition metal dilute alloys. Solid State Commun. 15, 1889–1890 (1974). https://doi.org/10.1016/00381098(74)90110-0 89. Monod, P., Schultz, S.: Conduction electron spin-flip scattering by impurities in copper. J. Phys. (Paris) 43, 393–401 (1982). https://doi.org/10.1051/jphys:01982004302039300 90. Braspenning, P.J., Zeller, R., Dederichs, P.H., Lodder, A.: Electronic structure of nonmagnetic impurities in Cu. J. Phys. F: Metal Phys. 12, 105 (1982). https://doi.org/10.1088/ 0305-4608/12/1/011 91. Dietl, T., Sawicki, M., Dahl, M., Heiman, D., Isaacs, E.D., Graf, M.J., Gubarev, S.I., Alov, D.L.: Spin-flip scattering near the metal-to-insulator transition in Cd0.95 Mn0.05 Se:In. Phys. Rev. B 43, 3154–3163 (1991). https://doi.org/10.1103/PhysRevB.43.3154 92. Fisher, M.E., Langer, J.S.: Resistive anomalies at magnetic critical points. Phys. Rev. Lett. 20, 665 (1968). https://doi.org/10.1103/PhysRevLett.20.665 93. Novák, V., Olejník, K., Wunderlich, J., Cukr, M., Výborný, K., Rushforth, A.W., Edmonds, K.W., Campion, R.P., Gallagher, B.L., Sinova, J., Jungwirth, T.: Curie point singularity in the temperature derivative of resistivity in (Ga,Mn)As. Phys. Rev. Lett. 101, 077201 (2008). https://doi.org/10.1103/PhysRevLett.101.077201 94. Yakovlev, D.R., Ossau, W.: Magnetic polarons. In: Gaj, J.A., Kossut, J. (eds.) Introduction to the Physics of Diluted Magnetic Semiconductors. Springer Series in Materials Science, vol. 144. Spinger, Berlin (2010). https://doi.org/10.1007/978-3-642-15856-8_7 95. Benoit á La Guillaume, C.: Free magnetic polarons in three, quasi-two, and quasi-one dimensions. phys. stat. sol. (B) 175, 369–380 (1993). https://doi.org/10.1002/pssb.2221750208 96. Peterson, D.L., Bartholomew, D.U., Ramdas, A.K., Rodriguez, S.: Magnetic field and temperature tuning of resonant Raman scattering in diluted magnetic semiconductors. Phys. Rev. B 31, 7932 (1985). https://doi.org/10.1103/PhysRevB.31.7932 97. Dietl, T., Peyla, P., Grieshaber, W., Merle d’Aubigné, Y.: Dynamics of spin organization in diluted magnetic semiconductors. Phys. Rev. Lett. 74, 474 (1995). https://doi.org/10.1103/ PhysRevLett.74.474 98. Pierre, F., Birge, N.O.: Dephasing by extremely dilute magnetic impurities revealed by Aharonov-Bohm oscillations. Phys. Rev. Lett. 89, 206804 (2002). https://doi.org/10.1103/ PhysRevLett.89.206804 99. Jaroszy´nski, J., Wróbel, J., Sawicki, M., Kami´nska, E., Sko´skiewicz, T., Karczewski, G., Wojtowicz, T., Piotrowska, A., Kossut, J., Dietl, T.: Influence of s-d exchange interaction on universal conductance fluctuations in Cd1−x Mnx Te:In. Phys. Rev. Lett. 75, 3170 (1995). https://doi.org/10.1103/PhysRevLett.75.3170 100. Wojtowicz, T., Dietl, T., Sawicki, M., Plesiewicz, W., Jaroszy´nski, J.: Metal-insulator transition in semimagnetic semiconductors. Phys. Rev. Lett. 56, 2419–2422 (1986). https:// doi.org/10.1103/PhysRevLett.56.2419 101. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A.H., Ong, N.P.: Anomalous Hall effect. Rev. Mod. Phys. 82, 1539 (2010). https://doi.org/10.1103/RevModPhys.82.1539 102. Ralph, D.C., Stiles, M.D.: Spin transfer torques. J. Magn. Magn. Mater. 320, 1190–1216 (2008). https://doi.org/10.1016/j.jmmm.2007.12.019 103. Manchon, A., Koo, H.C., Nitta, J., Frolov, S.M., Duine, R.A.: New perspectives for Rashba spin-orbit coupling. Nat. Mater. 14, 871–882 (2015). https://doi.org/10.1038/nmat4360

    974

    A. Bonanni et al.

    104. Hasan, M.Z., Kane, C.L.: Colloquium: topological insulators. Rev. Mod. Phys. 82, 3045–3067 (2010). https://doi.org/10.1103/RevModPhys.82.3045 105. Barborini, M.: Neutral, anionic, and cationic manganese dimers through density functional theory. J. Phys. Chem. A 120, 1716–1726 (2016). https://doi.org/10.1021/acs.jpca.5b12169 106. Shapira, Y., Bindilatti, V.: Magnetization-step studies of antiferromagnetic clusters and single ions: exchange, anisotropy, and statistics. J. Appl. Phys. 92, 4155 (2002). https://doi.org/10. 1063/1.1507808 107. Savoyant, A., D’Ambrosio, S., Kuzian, R.O., Daré, A.M., Stepanov, A.: Exchange integrals in Mn- and Co-doped II-VI semiconductors. Phys. Rev. B 90, 075205 (2014). https://doi.org/ 10.1103/PhysRevB.90.075205 108. Story, T.: Semimagnetic semiconductors based on lead chalcogenides,” In: Khokhlov, D. (ed.) Lead Chalcogenides, Physics and Applications. Optoelectronic Properties of Semiconductors and Supelattices, vol. 18. Taylor and Francis, New York, (2002) 109. Samarth, N., Furdyna, J.K.: A proposed interpretation of EPR linewidth in diluted magnetic semiconductors. Solid State Commun. 65, 801–804 (1988). https://doi.org/10.1016/00381098(88)90508-X 110. Larson, B.E., Ehrenreich, H.: Anisotropic superexchange and spin-resonance linewidth in diluted magnetic semiconductors. Phys. Rev. B 39, 1747–1759 (1989). https://doi.org/10. 1103/PhysRevB.39.1747 111. Blinowski, J., Kacman, P., Majewski, J.A.: Ferromagnetic superexchange in Cr-based diluted magnetic semiconductors. Phys. Rev. B 53, 9524 (1996). https://doi.org/10.1103/PhysRevB. 53.9524 112. Söllinger, W., Heiss, W., Lechner, R.T., Rumpf, K., Granitzer, P., Krenn, H., Springholz, G.: Exchange interactions in europium monochalcogenide magnetic semiconductors and their dependence on hydrostatic strain. Phys. Rev. B 81, 155213 (2010). https://doi.org/10.1103/ PhysRevB.81.155213 113. Kuneš, J., Ku, W., Pickett, W.E.: Exchange coupling in Eu monochalcogenides from first principles. J. Phys. Soc. Japan 74, 1408–1411 (2005). https://doi.org/10.1143/JPSJ.74.1408 114. Lewiner, C., Bastard, G.: Indirect exchange interaction in extremely non-parabolic zero-gap semiconductors. J. Phys. C: Solid State Phys. 13, 2347 (1980). https://doi.org/10.1088/00223719/13/12/016 115. Sawicki, M., Chiba, D., Korbecka, A., Nishitani, Y., Majewski, J.A., Matsukura, F., Dietl, T., Ohno, H. Experimental probing of the interplay between ferromagnetism and localization in (Ga, Mn)As. Nat. Phys. 6, 22 (2010). https://doi.org/10.1038/nphys1455 116. Dietl, T., Haury, A., d’Aubigne, Y.M.: Free carrier-induced ferromagnetism in structures of diluted magnetic semiconductors. Phys. Rev. B 55, R3347 (1997). https://doi.org/10.1103/ PhysRevB.55.R3347 117. Dietl, T., Cibert, J., Ferrand, D., Merle d’Aubigné, Y.: Carrier-mediated ferromagnetic interactions in structures of magnetic semiconductors. Mater. Sci. Eng. B 63, 103–110 (1999). https://doi.org/10.1016/S0921-5107(99)00059-8 118. Boukari, H., Kossacki, P., Bertolini, M., Ferrand, D., Cibert, J., Tatarenko, S., Wasiela, A., Gaj, J.A., Dietl, T.: Light and electric field control of ferromagnetism in magnetic quantum structures. Phys. Rev. Lett. 88, 207204 (2002). https://doi.org/10.1103/PhysRevLett. 88.207204 119. Liu, Q., Liu, C.-X., Xu, C., Qi, X.-L., Zhang, S.-C.: Magnetic impurities on the surface of a topological insulator. Phys. Rev. Lett. 102, 156603 (2009). https://doi.org/10.1103/ PhysRevLett.102.156603 120. Abanin, D.A., Pesin, D.A.: Ordering of magnetic impurities and tunable electronic properties of topological insulators. Phys. Rev. Lett. 106, 136802 (2011). https://doi.org/10.1103/ PhysRevLett.106.136802 121. Maletta, H., Felsch, W.: Insulating spin-glass system Eux Sr1−x S. Phys. Rev. B 20, 1245–1260 (1979). https://doi.org/10.1103/PhysRevB.20.1245 122. Vincent, E.: Ageing, rejuvenation and memory: the example of spin-glasses, pp. 7–60. Springer, Berlin (2007)

    18 Dilute Magnetic Materials

    975

    123. Andresen, J.C., Katzgraber, H.G., Oganesyan, V., Schechter, M.: Existence of a thermodynamic spin-glass phase in the zero-concentration limit of anisotropic dipolar systems. Phys. Rev. X 4, 041016 (2014). https://doi.org/10.1103/PhysRevX.4.041016 124. Bouchiat, H.: Determination of the critical exponents in the AgMn spin glass. J. de Phys. (Paris) 47, 71–88 (1986). https://doi.org/10.1051/jphys:0198600470107100 125. Gunnarsson, K., Svedlindh, P., Nordblad, P., Lundgren, L., Aruga, H., Ito, A.: Dynamics of an Ising spin-glass in the vicinity of the spin-glass temperature. Phys. Rev. Lett. 61, 754–757 (1988). https://doi.org/10.1103/PhysRevLett.61.754 126. Fert, A., De Courtenay, N., Bouchiat, H.: Experimental arguments for a non-zero transition temperature in RKKY Heisenberg spin glasses without anisotropy. J. de Physique 49, 1173– 1178 (1988). https://doi.org/10.1051/jphys:019880049070117300 127. Larson, B.E., Ehrenreich, H.: Exchange in II-VI-based magnetic semiconductors (invited). J. Appl. Phys. 67, 5084–5089 (1990). https://doi.org/10.1063/1.344681 128. Campos, I., Cotallo-Aban, M., Martin-Mayor, V., Perez-Gaviro, S., Taranco, A.: Spin-glass transition of the three-dimensional Heisenberg spin glass. Phys. Rev. Lett. 97, 217204 (2006). https://doi.org/10.1103/PhysRevLett.97.217204 129. Fabris, F.W., Pureur, P., Schaf, J., Vieira, V.N., Campbell, I.A.: Chiral anomalous Hall effect in reentrant AuFe alloys. Phys. Rev. B 74, 214201 (2006). https://doi.org/10.1103/PhysRevB. 74.214201 130. Bouchiat, H., de Courtenay, N., Monod, P., Ocio, M., Refregier, P.: Critical behavior and magnetic fluctuations in spin glasses. Japan. J. Appl. Phys. 26, S3-3, 1951 (1987). https://doi. org/10.7567/JJAPS.26S3.1951 131. Weissman, M.B.: What is a spin glass? A glimpse via mesoscopic noise. Rev. Mod. Phys. 65, 829–839 (1993). https://doi.org/10.1103/RevModPhys.65.829 132. Weissman, M.B., Israeloff, N.E., Alers, G.B.: Spin-glass fluctuation statistics: mesoscopic experiments in CuMn. J. Magn. Magn. Mater. 114, 87–130 (1992). https://doi.org/10.1016/ 0304-8853(92)90336-M 133. Hor, Y.S., Roushan, P., Beidenkopf, H., Seo, J., Qu, D., Checkelsky, J.G., Wray, L.A., Hsieh, D., Xia, Y., Xu, S.-Y., Qian, D., Hasan, M.Z., Ong, N.P., Yazdani, A., Cava, R.J.: Development of ferromagnetism in the doped topological insulator Bi2−x Mnx Te3 . Phys. Rev. B 81, 195203 (2010). https://doi.org/10.1103/PhysRevB.81.195203 134. Park, Y.D., Hanbicki, A.T., Erwin, S.C., Hellberg, C.S., Sullivan, J.M., Mattson, J.E., Ambrose, T.F., Wilson, A., Spanos, G., Jonker, B.T.: A group-IV ferromagnetic semiconductor: Mnx Ge1−x . Science 295, 651 (2002). https://doi.org/10.1126/science.1066348 135. Deng, Z., Jin, C.Q., Liu, Q.Q., Wang, X.C., Zhu, J.L., Feng, S.M., Chen, L.C., Yu, R.C., Arguello, C., Goko, T., Ning, F., Zhang, J., Wang, Y., Aczel, A.A., Munsie, T., Williams, T.J., Luke, G.M., Kakeshita, T., Uchida, S., Higemoto, W., Ito, T.U., Gu, B., Maekawa, S., Morris, G.D., Uemura, Y.J.: Li(Zn,Mn)As as a new generation ferromagnet based on a I-II-V semiconductor. Nat. Commun. 2, 422 (2011). https://doi.org/10.1038/ncomms1425 136. Przybyli´nska, H., Springholz, G., Lechner, R.T., Hassan, M., Wegscheider, M., Jantsch, W., Bauer, G.: Magnetic-field-induced ferroelectric polarization reversal in the multiferroic Ge1−x Mnx Te semiconductor. Phys. Rev. Lett. 112, 047202 (2014). https://doi.org/10.1103/ PhysRevLett.112.047202 137. Wang, K.Y., Sawicki, M., Edmonds, K.W., Campion, R.P., Rushforth, A.W., Freeman, A.A., Foxon, C.T., Gallagher, B.L., Dietl, T.: Control of coercivities in (Ga,Mn)As thin films by small concentrations of mnas nanoclusters. Appl. Phys. Lett. 88, 022510 (2006). https://doi. org/10.1063/1.2162856 138. Dietl, T.: A ten-year perspective on dilute magnetic semiconductors and oxides. Nat. Mater. 9, 965 (2010). https://doi.org/10.1038/nmat2898 139. Mayer, M.A., Stone, P.R., Miller, N., Smith, H.M., Dubon, O.D., Haller, E.E., Yu, K.M., Walukiewicz, W., Liu, X., Furdyna, J.K.: Electronic structure of Ga1−x Mnx As analyzed according to hole-concentration-dependent measurements. Phys. Rev. B 81, 045205 (2010). https://doi.org/10.1103/PhysRevB.81.045205

    976

    A. Bonanni et al.

    140. Winkler, T.E., Stone, P.R., Li, T., Yu, K.M., Bonanni, A., Dubon, O.D.: Compensationdependence of magnetic and electrical properties in Ga1−x Mnx P. Appl. Phys. Lett. 98, 012103 (2011). https://doi.org/10.1063/1.3535957 141. Nishitani, Y., Chiba, D., Endo, M., Sawicki, M., Matsukura, F., Dietl, T., Ohno, H.: Curie temperature versus hole concentration in field-effect structures of Ga1−x Mnx As. Phys. Rev. B 81, 045208 (2010). https://doi.org/10.1103/PhysRevB.81.045208 142. Brüne, C., Liu, C.X., Novik, E.G., Hankiewicz, E.M., Buhmann, H., Chen, Y.L., Qi, X.L., Shen, Z.X., Zhang, S.C., Molenkamp, L.W.: Quantum Hall effect from the topological surface states of strained bulk HgTe. Phys. Rev. Lett. 106, 126803 (2011). https://doi.org/10.1103/ PhysRevLett.106.126803 143. Chang, C.-Z., Zhang, J., Liu, M., Zhang, Z., Feng, X., Li, K., Wang, L.-L., Chen, X., Dai, X., Fang, Z., Qi, X.-L., Zhang, S.-C., Wang, Y., He, K., Ma, X.-C., Xue, Q.-K.: Thin films of magnetically doped topological insulator with carrier-independent long-range ferromagnetic order. Adv. Mater. 25, 1065–1070 (2013). https://doi.org/10.1002/adma.201203493 144. Zhou, Z., Chien, Y.-J., Uher, C.: Thin film dilute ferromagnetic semiconductors Sb2−x Crx Te3 with a Curie temperature up to 190 K. Phys. Rev. B 74, 224418 (2006). https://doi.org/10. 1103/PhysRevB.74.224418 145. Zhou, Z., Chien, Y.-J., Ctirad, U.: Thin-film ferromagnetic semiconductors based on Sb2−x Vx Te3 with TC of 177 K. Appl. Rev. Lett. 87, 112503 (2005). https://doi.org/10.1063/ 1.2045561 146. Enders, A., Skomski, R., Sellmyer, D.J.: Designed magnetic nanostructures. In: Liu, J.P., Fullerton, E., Gutfleisch, O., Sellmyer, D.J. (eds.) Nanoscale Magnetic Materials and Applications. Springer (2009). https://doi.org/10.1007/978-0-387-85600-1_3 147. Sato, K., Bergqvist, L., Kudrnovský, 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). https://doi.org/10.1103/RevModPhys.82.1633 148. Dietl, T.: Self-organised growth controlled by charge states of magnetic impurities. Nat. Mater. 5, 673 (2006). https://doi.org/10.1038/nmat1721 149. Yu, K.M., Walukiewicz, W., Wojtowicz, T., Kuryliszyn, I., Liu, X., Sasaki, Y., Furdyna, J.K.: Effect of the location of Mn sites in ferromagnetic Ga1−x Mnx As on its Curie temperature. Phys. Rev. B 65, 201303 (2002). https://doi.org/10.1103/PhysRevB.65.201303 150. Wojtowicz, T., Lim, W.L., Liu, X., Dobrowolska, M., Furdyna, J.K., Yu, K.M., Walukiewicz, W., Vurgaftman, I., Meyer, J.R.: Enhancement of Curie temperature in Ga1−x Mnx As/ Ga1−y Aly As ferromagnetic heterostructures by Be modulation doping. Appl. Phys. Lett. 83, 4220 (2003). https://doi.org/10.1063/1.1628815 ´ 151. Birowska, M., Sliwa, C., Majewski, J.A., Dietl, T.: Origin of bulk uniaxial anisotropy in zincblende dilute magnetic semiconductors. Phys. Rev. Lett. 108, 237203 (2012). https://doi.org/ 10.1103/PhysRevLett.108.237203 152. Moreno, M., Trampert, A., Jenichen, B., Däweritz, L., Ploog, K.H.: Correlation of structure and magnetism in GaAs with embedded Mn(Ga)As magnetic nanoclusters. J. Appl. Phys. 92, 4672 (2002). https://doi.org/10.1063/1.1506402 153. Korenev, V.L., Salewski, M., Akimov, I.A., Sapega, V.F., Langer, L., Kalitukha, I.V., Debus, J., Dzhioev, R.I., Yakovlev, D.R., Müller, D., Schröder, C., Hövel, H., Karczewski, G., Wiater, M., Wojtowicz, T., Kusrayev, Y.G., Bayer, M.: Long-range p–d exchange interaction in a ferromagnet-semiconductor hybrid structure. Nat. Phys. 12, 85–91 (2016). https://doi.org/10. 1038/nphys3497

    18 Dilute Magnetic Materials

    977

    154. Tanaka, M.: Ferromagnet (MnAs)/III-V semiconductor hybrid structures. Semicond. Sci. Technol. 17, 327 (2002). https://doi.org/10.1088/0268-1242/17/4/306 155. Bonanni, A., Navarro-Quezada, A., Li, T., Wegscheider, M., Mat˘ej, Z., Holý, V., Lechner, R.T., Bauer, G., Rovezzi, M., D’Acapito, F., Kiecana, M., Sawicki, M., Dietl, T.: Controlled aggregation of magnetic ions in a semiconductor: an experimental demonstration. Phys. Rev. Lett. 101, 135502 (2008). https://doi.org/10.1103/PhysRevLett.101.135502 156. Venkatesan, M., Stamenov, P., Dorneles, L.S., Gunning, R.D., Bernoux, B., Coey, J.M.D.: Magnetic, magnetotransport, and optical properties of Al-doped Zn0.95 Co0.05 O thin films. Appl. Phys. Lett. 90, 242508 (2007). https://doi.org/10.1063/1.2748343 157. Kuroda, S., Nishizawa, N., Takita, K., Mitome, M., Bando, Y., Osuch, K., Dietl, T.: Origin and control of high temperature ferromagnetism in semiconductors. Nat. Mater. 6, 440 (2007). https://doi.org/10.1038/nmat1910 158. Devillers, T., Rovezzi, M., Gonzalez Szwacki, N., Dobkowska, S., Stefanowicz, W., Sztenkiel, D., Grois, A., Suffczy´nski, J., Navarro-Quezada, A., Faina, B., Li, T., Glatzel, P., d’Acapito, F., Jakieła, R., Sawicki, M., Majewski, J.A., Dietl, T., Bonanni, A.: Manipulating Mn–Mgk cation complexes to control the charge- and spin-state of Mn in GaN. Sci. Rep. 2, 722 (2012). https://doi.org/10.1038/srep00722 159. Gonzalez Szwacki, N., Majewski, J.A., Dietl, T.: Aggregation and magnetism of Cr, Mn, and Fe cations in GaN. Phys. Rev. B 83, 184417 (2011). https://doi.org/10.1103/PhysRevB.83. 184417 160. Mouton, I., Lardé, R., Talbot, E., Cadel, E., Genevois, C., Blavette, D., Baltz, V., Prestat, E., Guillemaud, P.B., Barski, A., Jamet, M.: Composition and morphology of self-organized Mnrich nanocolumns embedded in Ge: correlation with the magnetic properties. J. Appl. Phys. 112, 113918 (2012). https://doi.org/10.1063/1.4768723 161. Gu, L., Wu, S., Liu, H., Singh, R., Newman, N., Smith, D.: Characterization of Al(Cr)N and Ga(Cr)N dilute magnetic semiconductors. J. Magn. Magn. Mater. 290, 1395 (2005). https:// doi.org/10.1016/j.jmmm.2004.11.446 162. Fukushima, T., Sato, K., Katayama-Yoshida, H.: Ab initio design of fabrication process and shape control of self-organized tera-bit-density nano-magnets in dilute magnetic semiconductors by two-dimensional spinodal decomposition. Phys. Stat. Sol. (C) 3, 4139 (2007). https:// doi.org/10.1002/pssc.200672830 Alberta Bonanni graduated in physics at the University of Trieste – Italy. She worked at the University of Minnesota and at the Synchrotron Center of the University of Madison-Wisconsin in the USA. She is now at the Johannes Kepler University in Linz, Austria, and her scientific interests are focused on the physics of nitride compounds and quantum materials.

    978

    A. Bonanni et al. Tomasz Dietl obtained his PhD from the Institute of Physics, Polish Academy of Sciences in Warsaw, where he is presently a head of the International Centre for Interfacing Magnetism and Superconductivity with Topological Matter “MagTop.” He is also a PI at the Advanced Institute for Materials Research at Tohoku University in Sendai, Japan.

    Hideo Ohno received his PhD from the University of Tokyo in 1982. He joined Hokkaido University from 1982 and he was a visiting scientist at the IBM T. J. Watson Research Center for 1.5 years. He was appointed professor at Tohoku University in 1994 and is president since 2018. His research interests include spintronics and semiconductor science and technology.

    Single-Molecule Magnets and Molecular Quantum Spintronics

    19

    Gheorghe Taran, Edgar Bonet, and Wolfgang Wernsdorfer

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Single Ion Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Multi-spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Giant Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantum Tunneling of Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Landau–Zener–Stückelberg (LZS) Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Spin Parity and Quantum Phase Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantum Coherence in Molecular Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Resonant Photon Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Rabi Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Molecular Quantum Spintronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Direct Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Indirect Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantum Algorithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    980 982 982 987 988 988 989 991 994 994 994 995 996 999 1001 1002 1002

    Abstract

    This chapter gives an overview of the main phenomenologies related to the magnetism of single-molecule magnets (SMMs) and covers some important achievements in the field of molecular spintronics. We start by discussing the dominant interactions at sub-Kelvin temperatures in the framework of spin G. Taran · W. Wernsdorfer () Physikalisches Institute, KIT, Karlsruhe, Germany e-mail: [email protected]; [email protected] E. Bonet Néel Institute, CNRS, Grenoble, France 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_18

    979

    980

    G. Taran et al.

    Hamiltonian models. The application of the general formalism to mononuclear and polynuclear complexes allows us to illustrate the power of the spin models in explaining both static properties (e.g., magnetic bistability) and dynamic ones (e.g., quantum tunneling of magnetization). We show how SMMs were used as a vehicle to explore quantum phenomenologies like nonadiabatic spin transitions, spin parity effect, the Berry phase interference, and quantum coherence while covering milestone results that brought the field closer to providing basic components of quantum devices. The last section is devoted to recent achievements in the field of molecular spintronics with emphasis on basic experimental designs that allowed the implementation of Grover’s quantum algorithms at the singlemolecule level. The successful transposition of the properties of the molecular magnets into functional devices is a proof of the deep understanding acquired in the two decades of scientific effort since the birth of this research field.

    List of Abbreviations and Symbols

    SMM QTM ZFS SQUID LZS EPR EM SD RD STM CNT

    Single-molecule magnets Quantum tunneling of magnetization Zero-field splitting Superconducting quantum interference device Landau–Zener–Stückelberg Electron paramagnetic resonance Electromagnetic Spin dot Read-out dot Scanning tunnel microscope Carbon nanotube

    Introduction Traditionally, the word magnet evokes a material in which a long-range magnetic order arises from local exchange interactions. The observed magnetic properties, like remanence and hysteresis, result from a collective behavior. However, developments in the last two decades showed that isolated molecules can bear large magnetic moments that exhibit bistability like traditional magnets (Fig. 1). They have therefore been called single-molecule magnets (SMMs). The field of SMMs started with the discovery of the large magnetic moment of the Mn12 O12 (CH3 COO)16 (H2 O)4 (here denoted as Mn12 -ac, Fig. 2) molecular cluster [1] and the observation of its magnetic bistability [2]. Later, quantum tunneling of magnetization (QTM) [3, 4] was evidenced in Mn12 -ac, as well as ground state quantum tunneling [5] and topological quantum phase interference effects [6] in [Fe8 O2 (OH)12 (tacn)6 ]8+ (or simply Fe8 , Fig. 2). These first discovered

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    981

    Fig. 1 Size scale spanning atomic to nanoscale dimensions. On the far right is shown a highresolution transmission electron micrography of a typical 3 nm diameter cobalt nanoparticle containing about 1000 Co atoms. The Mn84 molecule is a 4.2 nm diameter particle. Also shown for comparison are the indicated smaller Mn nanomagnets which are drawn to scale. An alternative means of comparison is the Néel vector (N) which is the scale shown. The arrows indicate the magnitude of the Néel vectors for the indicated SMMs which are 7.5, 22, 61, and 168 for Mn4 , Mn12 , Mn30 , and Mn84 , respectively. (Reprinted from [10])

    Fig. 2 Left: Structure of the Mn12 SMM with the formula [Mn12 O12 (CH3 COO)16 (H2 O)4 ] · 2CH3 COOH·4H2 O. Middle: Structure of the Fe8 SMM with the formula [Fe8 O2 (OH)12 (tacn)6 ]8+ where tacn is a macrocyclic ligand. Right: Mononuclear terbium complex TbPc2 . The upper Pc ligand is a mirror image of the lower Pc ligand but rotated by 45◦

    SMMs lead to breakthrough observations and to this day remain among the most investigated systems [7]. SMMs are constituted by an inner core of magnetic ions which is surrounded by a shell of organic ligands. They come in a variety of shapes and sizes and permit selective substitution of the ligands in order to alter the coupling to the environment. It is also possible to replace the magnetic ions, thus changing the magnetic properties without greatly modifying the structure and the interaction with its surroundings.

    982

    G. Taran et al.

    The technological interest in the field of SMMs is fueled by the desire to use molecular structures as memory units and quantum bits (qubits). There are currently two different approaches that try to satisfy the observed miniaturization tendencies in the current state-of-the-art technology. The first one is a top-down approach in which the nanometer-sized objects are obtained by reducing the dimensions of a bulk material (a common way for obtaining magnetic nanoparticles). The second method is a bottom-up approach [8], which for the field of SMMs means enhancing the magnetic moment of the molecule by ion substitution or through adding new magnetic centers to the molecule. Figure 1 shows a dimensional comparison between the above presented methods. One big advantage of SMMs over magnetic nanoparticles is their monodisperse properties, as chemical synthesis yields a large number of molecules with identical characteristics. In many cases, the SMMs can be made to form insulating monocrystals; thus, the environment of each molecule is very similar. This gives access to the properties of a single molecule from the measurements performed on the ensemble. The found alignment between the magnetic principal axis of the molecule and the crystallographic axes [9] is especially important when determining the intrinsic characteristics of the magnetic centers. SMMs belong to the mesoscopic realm (we refer to the vector space associated with the spin degree of freedom) where properties from both the quantum and the classical worlds transpire. The interest of working in this dimensional range comes from the desire of using quantum properties characterizing microscopic systems while addressing and manipulating them with simple means. We start by presenting the spin Hamiltonian formalism, alongside introducing the concepts essential for understanding the magnetism of molecular magnets. Then, we show how SMMs were used as a vehicle to explore quantum phenomenologies like quantum tunneling of magnetization (Sect. “Quantum Tunneling of Magnetization”), spin parity effect and the Berry phase interference (Sect. “Spin Parity and Quantum Phase Interference”), and quantum coherence (Sect. “Quantum Coherence in Molecular Magnets”). The last section is devoted to recent achievements in the field of molecular spintronics.

    Spin Hamiltonian The first step in describing the static and dynamic properties of SMMs is writing the characteristic Hamiltonian. The application of the general formalism to mononuclear and polynuclear complexes will allow us to highlight both the strengths and the range of validity of the different spin models.

    Single Ion Spin Hamiltonian A molecular complex with a single magnetic center is characterized by a total angular momentum (J ) which results from the coupling between the spin (S) and orbital angular momentum (L).

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    983

    The coupling between J and the external magnetic field (H ) is modeled through the Zeeman term: HZ = μB μ0 H gJ ˆ where gˆ is a tensor that links the total angular momentum to the magnetic moment and μB is the Bohr magneton. The effect of the ligand field can be accounted for through the use of the equivalent Stevens operators [11, 12]. Thus, the total Hamiltonian is: H=

    q 2S  

    Bqk Oqk + HZ

    (1)

    q=2 k=0

    where Oqk are equivalent Stevens operators and Bqk are parameters related to the ligand field and electronic structure of the ion. The Stevens operators are functions of Jz , J− , J+ and are listed in [11, 12]. The index q represents the order of the operators. The sum includes only even terms as only even terms generate nonzero matrix elements. The subscript k denotes the operator’s rotational symmetry.

    Transition Metal Ions As a first example, we consider a transition metal ion compound characterized by a weak coupling between the spin and orbital angular momentum. As a consequence, the Hamiltonian, in a first approximation, can be written using only spin degrees of freedom. The spin–orbit interaction, treated as a perturbation, is nonetheless very important because it couples the magnetic ions to the surrounding organic ligands, leading to critical contributions to the anisotropy of the molecule. By considering only the quadratic terms from (1) in zero external magnetic field, the Hamiltonian of the system is given by: H = DSz2 + E(Sx2 − Sy2 )

    (2)

    where D and E are commonly used notations that stand for the Stevens coefficients 3B20 and B22 , respectively. D is negative when z-axis is chosen to coincide with the easy axis of the molecule, which will be assumed in the following, and |E|  |D| in most cases. Considering only the first term in the above Hamiltonian, we see that the states pertaining to the S multiplet and labeled by the quantum number m = −S...S, are only doubly degenerate: E(±m) = −|D|m2 . For this reason, the magnetic anisotropy constant D is often denoted as zero-field splitting (ZFS). The discrete energy levels follow a parabola, illustrated in Fig. 3. U = |D|S 2 is called the anisotropy barrier. The two sides of the barrier correspond to opposite orientations of the magnetic moment; thus, the ZFS term gives the approximate energy barrier the spin has to overcome to flip its orientation.

    984

    G. Taran et al.

    Fig. 3 Left: Energy levels of a spin S = 10 subject to the ZFS term. Right: A zoom of the energy level diagram near an avoided level crossing where m and m are the quantum numbers of the energy levels. PLZ is the Landau–Zener tunnel probability when sweeping the applied field from the left to the right over the anticrossing. The greater the gap Δm,m and the slower the sweeping rate, the higher is the tunnel probability PLZ (equation (9))

    The height of the barrier and the characteristic time of the experiment lead to the definition of the blocking temperature (TB ), as the temperature under which the phonons do not have sufficient energy to promote reversal processes (e.g., for Mn12 -ac, TB ∼ 4 K for a characteristic time of the experiment of 1 s). Increasing the blocking temperature remains a central research goal in the field of SMMs [13, 14, 15] with the current record anisotropy barrier being 1760 K and TB ∼ 60 K [16,17]. The last term in (2) breaks the axial symmetry of the system and strongly affects the low-temperature relaxation of the magnetization (Sect. “Quantum Tunneling of Magnetization”).

    Lanthanide(III) Ions As a second example, let’s look at the lanthanide (III) ion complexes whose magnetic properties are mainly determined by the strongly anisotropic and partially filled 4f orbitals. Opposite to the case of transition metal ions, the angular orbital momentum of the rare earth atoms is not quenched by the ligand field, and the spin– orbit coupling is much stronger due to their larger nuclear charge. Thence, neither L or S are good quantum numbers, the magnetic moment being described by a total angular momentum, J = |L − S| . . . (L + S), where L + S is the ground state for more than half-shell filling (Hund’s rule). For instance, the Tb3+ ion exhibits the [Xe]4f 8 electronic structure, which leads to a spin S = 3 and an orbital angular momentum L = 3. The ground state of the Tb3+ ion is thus J = L+S = 6. The strong spin–orbit coupling leads to a separation of about 2900 K between the ground state and the excited state (J = 5). We can therefore restrict the following discussion to the ground state multiplet, J = 6, comprised of 2J + 1 (degenerate) substates |J, mJ . The Tb3+ ion can be embedded between two parallel phthalocyaninato (Pc) ligand planes. The crystal field interaction generates a quantization axis oriented perpendicular to the Pc planes. The Tb3+ ion is coordinated by four nitrogen atoms from each ligand plane, and the upper Pc ligand (blue in Fig. 2) is a mirror reflection of the lower one (with respect to the x-y plane) rotated by 45◦ around the z-axis. The

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    985

    Tb3+ is therefore exposed to an antiprismatic ligand field that was modeled by the following Hamiltonian [18]: Hlf = B20 O20 + B40 O40 + B44 O44 + B60 O60 + B64 O64

    (3)

    where Bqk coefficients are written as a product of a constant depending on the electronic shell of the ion [11] and a Akq factor characterizing the ligand field interaction. The ligand field parameters, Akq , were determined experimentally by fitting NMR and magnetic susceptibility measurements and can be found in [18]. The diagonalization of Hlf in the |J, mJ  eigenbasis reveals that the ligand field partially lifts the degeneracy of the 2J + 1 substate in the ground state multiplet J = 6. The ground state doublet, mJ = ±6, is separated from the first excited doublet, mJ = ±5, by approximately 600 K (see [18] and Fig. 4). Then, each mJ doublet splits in an external magnetic field, the resulting Zeeman diagram being depicted in Fig. 4 [19]. Therefore, at cryogenic temperatures, the TbPc2 singlemolecule magnet behaves with a good approximation as an Ising spin system. The Tb3+ ion also carries a nuclear spin with I = 3/2 and a natural abundance of 100%. The electric and magnetic interactions between the nuclear spin and the electronic shell of the Tb3+ ion are represented by a hyperfine and a quadrupolar term [20]: ˆ + Ahyp I · J + IPˆquad I HTbPc2 = Hlf + μB μ0 H gJ

    (4)

    where Ahyp = 26.7 mK is the hyperfine constant and Pˆquad is the quadrupole tensor. The hyperfine interaction, Ahyp I · J , splits the electronic states from the ground state doublet mJ = ±6 into (2I + 1) hyperfine states, labeled by the additional nuclear quantum number mI = −3/2 . . . 3/2.

    Fig. 4 Left: Ligand field splitting of the ground state multiplet in different mononuclear lanthanide complexes LnPc2 . (Modified from [18]). Right: Calculated Zeeman diagram for the ground state multiplet, J = 6, in a TbPc2 SMM. (Extracted from [19])

    986

    G. Taran et al.

    Because I > 1/2, the 159 Tb nucleus has a quadrupolar moment that couples to the electric field gradient through IPˆquad I with the dominant term being the axial component with magnitude Pquad = 17 mK. The tensorial nature of Pˆquad reflects the non-axial character (with respect to the easy axis of the ligand field) of the quadrupolar interaction which together with transverse terms in Hlf was shown to be responsible for the observed tunneling dynamics of TbPc2 in diluted single crystals [20] (Fig. 5). 1

    a)

    0.5 0 M/M s

    M/M s

    Δ

    -0.5 -1

    Energy (K)

    0.6

    b)

    0.3 0 -0.3 -0.04

    -0.02

    0.00

    0.02

    0.04

    Bz (T) Fig. 5 (a) Zoom on the magnetic hysteresis loop of an assembly of TbPc2 SMMs in a diluted single crystal (1% TbPc2 in an YPc2 matrix) measured with the microSQUID setup at T = 40 mK. The insets show: (left) a zoom of a level anticrossing between two hyperfine states and (right) the entire hysteresis loop. The red markers (right inset) indicate the continuous change of magnetization at high fields attributed to phonon-assisted relaxation. (b) Zeeman diagram of a TbPc2 SMM. The characteristic steps in the magnetization curve around zero field correspond to QTM events in the single crystal and are highlighted by dashed lines. (Changed from [21])

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    987

    Multi-spin Hamiltonian Although a number of molecules having a single magnetic ion have been synthesized [22], most SMMs contain multiple magnetic centers within their organic shell, and their description thus requires a multi-spin Hamiltonian. For a pair of spins coupled through exchange interaction, the Hamiltonian can be written as [23]: Hexc = S1 Jˆ12 S2

    (5)

    In general, the interaction tensor Jˆ12 is not symmetric, so it’s customary to put the above equation in a form that highlights an isotropic term (J12 S1 · S2 ) that favors a parallel alignment of the spins, an anisotropic, symmetric component (S1 Dˆ12 S2 ) that tends to align the spins along a certain direction in space, and Dzyaloshinskii– Moriya term (d12 · (S1 × S2 )) that works toward orienting the spins perpendicular to each other. Usually, the isotropic term is significantly larger than the others, and its sign will dictate the ferromagnetic or antiferromagnetic character of the ground state. The generalization to a system containing an arbitrary number of spins is done by putting together the single ion anisotropies with the pairwise exchange interactions, leading to the following multi-spin Hamiltonian: H=

     i

    Si Dˆ i Si +

    

    ˆ Sj Si Ji,j

    (6)

    ij

    where we considered only the quadratic term in the ligand field Hamiltonian. Even with this simplification, the number of parameters involved can be quite large, especially for clusters with high nuclearity. This difficulty cannot be completely eliminated, even when one considers a very symmetric molecule. As an example, let’s consider the archetypal Mn12 -ac molecule (Fig. 2). It contains an inner ring of four Mn(IV) ions (S4 = 3/2) and an outer ring of eight Mn(III) (S3 = 2) coupled together through superexchange interactions involving oxygen bridges. The number of states that characterizes the system is (2S4 + 1)4 (2S3 + 1)8 , which yields an overwhelming 108 eigenstates. Because the Mn12 -ac molecules crystallize in a body-centered tetragonal crystal, the unit cell contains two equivalent Mn ions so one independent zero-field splitting tensor is needed. When choosing a coupling scheme with only isotropic interactions [2], another four exchange parameters are required. Adjusting this model to experimental results, such as electron paramagnetic resonance (EPR) measurements, is far from trivial. Nevertheless, this model is valuable for low nuclearity systems [24]. A common way to shrink the dimension of the Hilbert space in which we have to study the system is done by replacing pairs of strongly interacting spins with a single equivalent spin.

    988

    G. Taran et al.

    Giant Spin Hamiltonian Applying the above procedure recursively to all spin pairs leads to the giant spin approximation, which describes in an effective way the ground state multiplet. This simplification works especially well when the temperature is low enough so that the dominant energy in the system is the exchange interaction. In this case, the system is always found in its spin ground state (e.g., S = 10 for Fe8 and Mn12 -ac), allowing the use of the single ion model presented at the beginning of this section. Therefore, we write the giant spin model: HGS = DSz2 + gμB μ0 Hz Sz + HT

    (7)

    where D and g are now effective parameters and HT describes non-axial interactions and can be written as a function of rising and lowering spin operators (S+ , S− ). In order to relate the effective parameters of the giant spin model and the parameters characterizing each individual magnetic center, one can use projection techniques [25], where single ion anisotropies are projected onto the vector space corresponding to the molecular ground state. This method is less accurate when the Hilbert space is very large, like in the case of Mn12 -ac, but the model retains its use because of its simplicity and intuitive form. The energy range in which the giant spin model is valid can be found in the multi-spin Hamiltonian. Indeed, the weakest exchange link gives with a good approximation the upper temperature limit to which the predictions are expected to be accurate. Experimentally one can observe the crossover to higher spin subspace by determining the magnetic moment associated with the molecular cluster [26,27].

    Quantum Tunneling of Magnetization For a ligand field Hamiltonian described only by the ZFS term, the magnetic field lifts the twofold degeneracy of the ±m energy levels. However, two different levels  (denoted by m and m ), align for specific field values Hrm,m , applied along the magnetic principal axis: 

    Hrm,m =

    |D|(m + m ) gμB μ0

    (8)

    Due to transverse terms of the spin Hamiltonian, the degeneracy can be transformed into an avoided level crossing; see Fig. 3. The energy separation is called the tunnel splitting, Δm,m , and is a central parameter that characterizes the quantum tunneling dynamics of spins. There are several theoretical tools that can be used to determine the tunnel splitting, including path integral formalism [28, 29], perturbation theory [30, 31], and numerical methods. The latter allows the possibility to consider arbitrarily

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    989

    complex spin Hamiltonians but leaves little intuition on the system’s behavior. Thus, a combination of the above techniques is to be preferred when performing both quantitative and qualitative analyses of the magnetic properties of SMMs. At an avoided level crossing, the eigenvectors of the Hamiltonian are a linear combination of the base vectors that correspond to negative and positive spin projections. This means that in these states, the magnetization of the spin has a nonzero probability to be found on either side of the barrier. This behavior is called spin tunneling, that is, the spin is in quantum resonance between the opposite orientation states. At a constant field tuned at an avoided level crossing, a spin initially situated on one side of the well will oscillate coherently between the mixed states with a  characteristic angular frequency, ωTm,m , related to the tunnel splitting through the m,m relation: Δm,m = 2hω ¯ T . The field interval where this behavior is predicted to Δ

    

    happen is given by [23]: δH0 = gμB μ0m,m |m−m | . This quantity is called the bare tunnel −9 width and can be as small as 10 T. However, environmental interactions broaden the observed width of the resonance and hinder coherent oscillations [32,33,34,35].

    Landau–Zener–Stückelberg (LZS) Model The nonadiabatic transition between the states of a two-level system was first discussed by Landau, Zener, and Stückelberg [36, 37, 38]. The original work by Zener concentrates on the electronic states of a bi-atomic molecule, while Landau and Stückelberg considered two atoms that undergo a scattering process. Their solution to the time-dependent Schrödinger equation of a two-level system driven through resonance has been applied to many physical systems, and it became an important tool for studying tunneling transitions. The LZS model was used to analyze spin tunneling in nanoparticles and clusters [39, 40, 41, 42]. The tunneling probability, PLZ , between the states m and m , after sweeping the applied field at a constant rate, α, through the resonance (Fig. 3), is given by:   δH0 PLZ = 1 − exp −π ωT α

    (9)

    With the LZS model in mind, we can now start to understand qualitatively the hysteresis loops of SMMs (Fig. 6), which exhibit steps of fast relaxation separated by regions where the magnetization is almost constant. The steps happen at specific fields where the levels from both sides of the well are mixed by transverse terms in the spin Hamiltonian, as discussed in the previous section. Let us start at a large negative magnetic field Hz . At very low temperatures, all molecules are in the m = 10 ground state (Fig. 6). As the applied field Hz is ramped down to zero, all molecules will stay in the m = 10 ground state. When ramping the field over the Δ10,−10 region, at Hz ≈ 0, there is a Landau–Zener probability, P10,−10 , for the spins to tunnel from the m = 10 to the m = −10

    990

    G. Taran et al.

    Fig. 6 Top: Hysteresis loops of a single crystal of Fe8 molecular clusters at 0.04 K and different field sweep rates. The loops display a series of steps, separated by plateaus. Arrows indicate the resonance transitions highlighted in the figure below. Bottom: Zeeman diagram of the S = 10 manifold of Fe8 as a function of the field applied along the easy axis. From bottom to top, the levels are labeled with quantum numbers m = ±10, ±9, . . .. The levels cross at equidistant field values given by μ0 Hz ≈ n × 0.22 T with n = 1, 2, 3 . . .. The arrows are explained in the text

    state. P10,−10 depends on the sweeping rate (9); the slower the sweeping rate, the larger is the tunneling probability. This is clearly demonstrated in the hysteresis loop measurements showing larger steps for slower sweeping rates [43,6]. When the field Hz is now further increased, there is a remaining fraction of molecules in the now metastable m = 10 state. The next chance to escape from this state is when the field reaches the Δ10,−9 region. There is a Landau–Zener tunnel probability P10,−9 to tunnel from the m = 10 to the m = −9 state. As m = −9 is an excited state, the molecules in this state relax quickly to the m = −10 state by emitting a phonon. An analogous procedure happens when the applied field reaches the Δ10,−10+n regions (n = 2, 3, . . . ) until all molecules are in the m = −10 ground state, that is, the magnetic moment of all molecules is reversed. As phonon emission can only change the molecule state by Δm = ±1 or ±2, there is a phonon cascade for higher applied fields. Figure 5 shows the magnetic hysteresis loop, measured by microSQUID technique [44], of a crystal containing TbPc2 SMMs randomly distributed in a diamagnetic, isostructural matrix formed by YPc2 molecules, with [TbPc2 ]/[YPc2 ] ratio of 1%. Upon sweeping the magnetic field from −1 T up to positive fields as small as 0.05 T, approximately 75% of the TbPc2 SMMs undergo quantum tunneling

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    991

    transitions, resulting in sharp steps in the magnetization curve. The QTM transitions take place between mixed states of nuclear and electronic origin; thus, both spin projections can change [20]. The remaining SMM reverse their magnetic moment at larger magnetic fields by a direct relaxation process [45]. When using the above formalism for a quantitative analysis, one should keep in mind that the LZS model is exact only for an isolated spin. Deviations from the ideal coherent dynamics are due to environmental interactions of both elastic (dephasing) and inelastic (relaxation and excitation) nature [21]. Thus, in order to try to search an agreement with LZS theory, one should use large sweeping rates, so that in the time the resonance is swept, the local environmental field does not change significantly. The model also doesn’t include relaxation through phonons so one should work at very low temperatures and focus on ground state tunneling. LZS theory was successfully used to determine the tunnel splitting in molecular systems such as Fe8 [6] and Mn4 [43]. As mentioned, the agreement with the experiment must be searched in the fast sweeping rate regime, which has the drawback of a small sensitivity. To overcome this difficulty, the resonance can be swept repeatedly. This way, the probability for the spin to remain in the original state, for small variations of the magnetization, after n back-and-forth sweeps, is proportional to (1 − 2nPLZ ), with PLZ given by (9).

    Spin Parity and Quantum Phase Interference When discussing quantum tunneling of magnetization, it was emphasized that transverse terms remove the degeneracy of the eigenstates at a level crossing (Hz = Hr ) and thus promote relaxation. Actually, the general problem of eigenvalue degeneracy is discussed by the von Neumann–Wigner theorem [46,47] which states the need of at least two parameters (e.g., the components of the magnetic field), to control the degeneracy property of a Hermitian matrix. In this section, we discuss the cases when the degeneracy is predicted theoretically and in some cases observed experimentally through the absence of tunneling at level intersections.

    Spin Parity n ; thus, a transition between The general form of a transverse term of order n is Bn S± n  the levels m and m is made possible by applying S± operator an integer number of times. Consequently, the degeneracy is removed only when the change in the spin projection is a multiple of the perturbation’s order: n|(m − m ). The above quantum tunneling selection rule is called spin parity effect and has its origin in the symmetry of the non-axial Hamiltonian term. For example, in the case of the Mn3 cluster [48], with a C3 rotational symmetry, level splitting should occur only if the selection rule m − m = 3k is satisfied. Another interesting example is the observation of Kramer’s degeneracy in halfinteger spin molecular clusters [49], which tells us that in zero applied magnetic field, the ground state is at least doubly degenerate. This degeneracy can be lifted

    992

    G. Taran et al.

    by a transverse field (a first-order perturbation) of external or internal origin (e.g., dipolar and hyperfine field). In general, the selection rules are rarely observed. Possible explanations involve asymmetries introduced by crystal defects, solvent disorder [50], abovementioned dipolar and hyperfine interactions, or Dzyaloshinskii–Moriya interaction [51] fuelled by low molecular symmetry and strong exchange coupling. In the case when one gets to control the above factors, the spin parity effect can be clearly evidenced, as seen for Mn3 [48].

    Quantum Phase Interference Suppression of tunneling can also occur for certain values of the transverse field, without involving the above-presented selection rule. The interference effect in spin systems was predicted theoretically [28] and then observed experimentally [6] using LZS approach and is generally called the Berry phase interference. The mechanics behind the explanation of this phenomena makes it similar to the observation of critical current oscillations in the Josephson junctions [52, 53] and Bohm–Aharonov oscillations [54]. A semiclassical description, which has a rather intuitive picture associated with it, describes the tunneling motion of the spin with the help of the path integral formalism [28]. The initial and final states are represented by two points on the Bloch sphere (Fig. 7). The interference is due to a nonzero phase difference

    Fig. 7 Unit sphere showing degenerate minima A and B which are joined by two tunnel paths (heavy lines). The hard, medium, and easy axes are taken in the x-, y-, and z-direction, respectively. The constant transverse field Htrans for tunnel splitting measurements is applied in the xy-plane at an azimuth angle ϕ. At zero applied field H = 0, the giant spin reversal results from the interference of two quantum spin paths of opposite directions in the easy anisotropy yz-plane. For transverse fields along an axis perpendicular to the easy axis, by using Stokes’ theorem, it has been shown that the path integrals can be converted in an area integral, yielding that destructive interference – that is, a quench of the tunneling rate – occurs whenever the area enclosed by the two paths is kπ/S, where k is an odd integer. (Reprinted from [6])

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    993

    Fig. 8 Measured tunnel splitting Δ as a function of transverse field for (a) several azimuth angles ϕ at m = ±10 and (b) ϕ ≈ 0◦ , as well as for quantum transition between m = −10 and (10 − n). Note the parity effect that is analogous to the suppression of tunneling predicted for half-integer spins. It should also be mentioned that internal dipolar and hyperfine fields hinder a complete quench of Δ which is predicted for an isolated spin. (Reprinted from [6])

    acquired by moving in the parametric space along a closed path containing these two points. The magnetic field has the ability to modulate the phase by modifying the paths, so an interference pattern is observed as the transverse field is increased. Figure 8 depicts tunnel splitting oscillations of a Fe8 complex showing equally spaced minima as a function of the transverse field. The difference between the maxima and the minima of the tunnel splitting can be as large as one order of magnitude. The anisotropy of the Fe8 molecule, similar to many low symmetry molecules, is modeled by adding a biaxial crystal field (HT = E(Sx2 − Sy2 )) to the uniaxial term. Therefore, when the tunnel splitting between the levels S and S − n is measured, the spin parity effect is observed (the second-order transverse anisotropy term forbids odd resonances at zero field). Also, a monotonic increase of Δ with n is observed: the lower is the energy barrier, the higher are the tunneling rates. The predicted field separation between two consecutive minima of the tunnel splitting, when we consider only the biaxial term, is given by: ΔH =

    2  2E(E + D) gμB

    (10)

    Using the anisotropy parameters determined in [55], one obtains a field separation of 0.26 T, which is smaller than the experimental value of 0.4 T. In order to account for this difference, higher order terms must be considered. This is easily done in the numerical approach that involves the diagonalization of the system’s Hamiltonian. Other systems in which it was possible to observe the Berry phase interference include Mn12 complexes [56,57,58], Mn4 [59], and dimer molecular magnets [60].

    994

    G. Taran et al.

    Quantum Coherence in Molecular Magnets The study of coherence in molecular magnets is very important for the potential applications in the field of quantum information processing. Here we briefly account on the quest for the observation of coherent time evolution of molecular spins under electromagnetic (EM) radiation.

    Resonant Photon Absorption In EPR experiments involving molecular magnets [55, 61, 62], transitions between the m and m states of the S multiplet happen when the energy of the incoming photons is equal to the energy difference between the states: hν = |E(m) − E(m )|, obeying the selection rule Δm = ±1. An important question that had to be answered concerns the effect of the photon-induced excitations on tunneling. The first important results [63] were obtained when a Fe8 complex was investigated using micro-Hall bars in a dilution refrigerator. The study proved that the lifetime of the excited states is large enough (relative to the tunneling time) for an enhanced relaxation to be observed. In order to clearly show the photonassisted quantum tunneling regime, circular polarized microwave radiation was used because it allows to select the Δm = +1 or Δm = −1 transitions. Thus, an asymmetric hysteresis loop is observed. From the dependence of the transition rates on the power of the microwave source used, it was shown that the effective spin temperature (the parameter which describes the occupation number of the excited states) depends linearly on the power of the EM field. Subsequent experimental work further proved the photon-assisted tunneling regime [64,65,66] leading toward eventual observation of Rabi oscillations [62].

    Rabi Oscillations The coherent evolution of the system between two eigenstates coupled by the EM field is described by the Rabi model [68,69]. If the system’s initial state is |m, then the probability to find it in the state |m , at time t, is proportional to sin2 (ΩR t). The Rabi frequency, ΩR , is proportional to the amplitude of the perturbative field. This property is used in the spin-echo measurement protocols [70] to determine the characteristic relaxation times of the system. The longitudinal relaxation time (T1 ) is obtained from the recovery of the equilibrium magnetization after an inversion pulse, π − T − π2 − τ − π − τ − echo, is applied, where the variable is T and τ is being kept constant. T1 is directly connected to the coupling of the spins to the phonon bath and thus can be significantly long at mK temperatures. The second important characteristic time is the phase coherence time (T2 ) and is determined by spin–spin interactions. As the name suggests, it tells us the time over which the memory of the quantum phase is preserved, so the quantum properties of the spin can be exploited. Employing a similar Hahn-echo sequence, π2 − τ − π −

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    995

    Fig. 9 Comparison of state-of-the-art T2 values of notable molecular and solid-state electronic spin qubits. (Extracted from [67])

    τ − echo, with τ variable, T2 is obtained from the echo signals that are fitted to a stretched exponential, I (τ ) = I (0) exp(−(2τ/T2 )x ). The interaction with environmental spins of nuclear and electronic origin, phonons, and photons represents the main sources of decoherence [71] in SMMs [72, 73] and thus limits T2 . In order to reduce the dipolar interactions, one usually chooses a system with a small collective spin [65, 74, 62] because the magnitude of the interactions scales as the square of the spin magnitude. Then, SMMs can be diluted without greatly effecting their individual properties [74, 62, 75]. Also, the initial choice of a molecular complex with S = 1/2 [74] avoided the problems associated with the distribution of the anisotropy axis. The above measurements led to the first observations of a long phase coherence time in Cr7 Ni (S = 1/2) and Cr7 Mn (S = 1) clusters [74]. Afterward, Rabi oscillations were observed in V15 [62, 76], and T2 on μs scale was measured in Fe8 [77] through the application of a high transverse magnetic field. Recent developments showed that a strict control over lattice rigidity and hyperfine interactions can lead to significantly large phase coherence times, even when compared to other qubit candidates ( [75] and Fig. 9). The coherence also has been shown to be preserved at room temperature [78, 79]. These are valuable observations that encourage possible spintronics applications.

    Molecular Quantum Spintronics Molecular quantum spintronics is a relatively new research field that combines spintronics, molecular electronics, and quantum computing with the aim of creating new spintronics molecular devices that exploit the quantum properties seen at the microscopic level [80,81,82]. These devices facilitate the read-out and manipulation of the spin states pertaining to a molecular magnet, leading to structures that can perform basic quantum operations. The idea of using SMMs as magnetic centers in spintronics devices is supported by their unique characteristics, namely, weak spin–orbit coupling in transition metal

    996

    G. Taran et al.

    ion compounds, tunable environmental interactions that can lead to a long coherence time (Sect. “Quantum Coherence in Molecular Magnets”), and chemically controlled functionalities like switchability with light or electric field [83]. The coupling to external structures can be facilitated by choosing an appropriate ligand. Delocalized bonds, which mediate the interaction between the magnetic ions, often imply great conduction properties. The above-outlined properties allowed the realization of some essential circuit elements, like [81] molecular spin transistors [84, 85], molecular spin valves [86] and spin filters, and molecular double-dot devices.

    Direct Coupling The first configuration that we consider involves a direct coupling mechanism in which the molecules are connected to metallic electrodes through chemical bonds (Fig. 10). This facilitates a strong coupling between the conduction electrons and the molecular spin but also implies a strong back action on the spin state. The setup was tested experimentally when a low-temperature scanning tunnel microscope (STM) was used to probe molecular magnets deposited on a conducting surface, leading to the observation of reversible chiral switching [87] and electrical

    Fig. 10 Schematic representation of different device geometries for molecular spintronics devices. (a) Three-terminal spin-dot device. This is a direct coupling scheme, in which the current flows through the spin dot (SD). (b) Three-terminal double-dot device. This is an indirect coupling scheme, in which the current flows through a second nonmagnetic quantum dot, the read-out dot (RD), which is coupled with the spin dot (SD) via exchange interaction (J). (c) Supramolecular spin valve device, with two SD’s coupled to the RD. The current in the RD is therefore sensitive to relative spin orientation in the two SD’s. (d) Multiterminal multi-dot device. (Reprinted from [45])

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    997

    current control over the spin state [88]. The spin-polarized STM allows direct space resolved observations of spin-split orbitals [89]. Another configuration is represented by a three terminal device where a molecular magnet bridges the gap of a break junction. A gate electrode is used to tune the transport properties of the conduction electrons [90]. First implementation of this design involved a Mn12 compound bound covalently to gold electrodes. To explore different conduction regimes, weak binding ligands were also used. In both cases the peripheral group acted as a tunnel barrier, helping to conserve the magnetic properties of the cluster. By knowing the magnetic properties of the redox species that are formed when the electrons pass through the molecule, one can compare the information obtained from spectroscopic transport measurements and established experimental methods like: magnetization measurements, EPR or neutron scattering. First studies involving Mn12 in a magnetic field showed that it exhibits transistor like properties [90, 91]. Both degeneracy in zero field and non-linear behavior of the excitations as a function of field are typical of tunneling via a magnetic molecule. However, follow up experiments indicated an alteration of the Mn12 magnetic properties during the deposition on gold electrodes [92]. In contrast to Mn12 , both Fe4 [93] and TbPc2 [94] are known to preserve their magnetic properties upon deposition. In the following we show how such a three terminal device, containing TbPc2 deposited on electromigrated gold electrodes, was used to read-out not only the electronic spin state but also the nuclear spin state of Tb3+ [94].

    Read-Out of a Single Nuclear Spin The Tb3+ center was shown to have a stable redox state which suggests that the current is not flowing through the ion. Instead, it is more likely that the Pc ligands, that contain a conjugated π system, are the conducting medium. Transport measurements revealed a single charge degeneracy with a weak spin-1/2 Kondo effect. When studying the position of the Kondo peak as a function of the applied field (supplemental information of [94]), an exchange interaction of about 0.35 T between a quantum dot with S = 1/2 and the Tb3+ ion was found. This relatively strong interaction indicates that the quantum dot and the Tb3+ ion are fairly close together, suggesting a binding geometry between the molecule and the electrodes as shown in Fig. 11. The spin states of the Tb3+ ion are not influenced by the charge state of the ligands, thus indicating that the Pc ligands behave as read-out dots. The response to a varying magnetic field was studied, using the charge degeneracy point as the working point. The experiment showed single jumps of the conductance as the field was swept over the zero-field region. These sudden changes in conductance are connected to the relaxation of the spin through quantum tunneling. In order to construct the statistics of the magnetization reversal, the resonance region was swept 22,000 times. The resulting histogram, showed in Fig. 12, illustrates that the magnetization reversal occurs at four distinct field values. The results are in perfect agreement with the discussion on quantum tunneling behavior of TbPc2 presented in

    998

    G. Taran et al.

    Fig. 11 Artist view of (left) a nuclear spin qubit transistor based on a single TbPc2 molecular magnet, coupled to source, drain, and gate (not shown) electrodes, and (right) a three-spin-dot device. (Reprinted from [85]) Fig. 12 Zeeman diagram and nuclear spin detection procedure. Top: Zeeman diagram of the TbPc2 molecular magnet, showing the hyperfine split electronic spin ground state doublet as a function of the external magnetic field H|| parallel to the easy-axis of magnetization. The ligand field-induced avoided level crossings (colored rectangles) allow for tunneling of the electron spin. Middle: The jumps of the conductance of the read-out quantum dot during magnetic field sweeps are nuclear spin dependent. Bottom: Histogram of the positions of about 75,000 conductance jumps, showing four non-overlapping Gaussian-like peaks. Each conductance jump can be assigned to a specific nuclear spin state

    Sect. “Landau–Zener–Stückelberg (LZS) Model”. Thus, the position of a particular transition tells us not only the state of the electronic spin but also the nuclear spin state. Because there is no overlap between the distribution of the switching fields, it is possible to determine the state of the nuclear spin through a single measurement.

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    999

    The lifetime of the nuclear states can be determined by doing subsequent measurements separated by a waiting time [85]. A T1 of order of tens of seconds was observed, which shows that this is a non-invasive method. The hyperfine Stark effect was used to manipulate the nuclear spin within a single molecule with an individual relaxation time (T1 ) exceeding 20 s [85]. Rabi, Ramsey and Hahn echo experiments (Fig. 13), performed on the 159 Tb nuclear spin [95], compare well with similar experiments involving P or Bi defects in Si. Ramsey T2 -times are 300 microseconds with π/2 pulse-lengths of 100 ns, i.e., about 1000 spin manipulations are possible before decoherence sets in. This allowed the implementation of Grover’s algorithm using the four states of the nuclear spin [96].

    Indirect Coupling In order to avoid the strong back action characterizing the above presented devices, the molecule can be coupled to the electrodes by indirect means, that is, by establishing a link between the SMM (a spin dot) and a non-magnetic molecular conductor (read-out dot) connected to the electrodes (Fig. 10). The coupling between the spin dot and read-out dot is usually realized through an exchange interaction and can be modulated by a gate voltage. Among possible candidates for read-out dots (e.g., nanowires, carbon nanotubes, nano-SQUIDs and ligands), carbon nanotubes are special due to their outstanding structural, mechanical and electrical properties. Because a carbon nanotube is essentially a 1D molecular conductor, with a diameter of the same order as a molecular magnet, a good coupling is easily achieved. At low temperature, it presents electronic properties like Coulomb blockade [97] and Kondo effect [98]. Thus, carbon nanotube conductance is sensitive to any charge fluctuations including the spin reversal processes. A number of devices with a carbon nanotube as the detector and TbPc2 as the spin dot showed that the indirect coupling mechanism allows the determination of both the electronic and nuclear spin of Tb3+ ion. Urdampilleta et al. [86] showed that a carbon nanotube functionalized with two SMMs has a supramolecular valve behavior (Fig. 14). Indeed, the two TbPc2 molecules behave as a spin polarizeranalyzer system exhibiting a strong magnetoresistive effect. Mediated by exchange interaction, the magnetic moment of each molecule induces a localized spin-polarized dot in the carbon nanotube quantum dot that can be controlled by a magnetic field (Fig. 14). Depending on the relative orientation of the molecular spins, a high- or low-conductance regime can be observed. A butterfly hysteresis loop, characteristic of spin valve devices, with a magnetoresistance up to 300% has been measured for temperatures smaller than 1 K. For a more detailed description, the reader may refer to Urdampilleta et al. [86]. Increasing the number of SMMs coupled to the carbon nanotube (Fig. 10) and addressing each individual molecule through a local gate may allow the implementation of complex quantum computation protocols.

    Fig. 13 Qubit coherent manipulations. Each color is a transition between consecutive levels: red being the transition between the ground state and the first excited level, and the next two in green and blue. (a) Rabi oscillations, the frequency of these oscillations can be tuned from 1.5 to 9 MHz. The maximum visibility is respectively 0.6, 0.9 and 0.6. (b) By recording the maximum of the visibility of the Rabi oscillations as function of the detuning, the shape of the three transitions can be measured. (c) Ramsey fringes give a coherence time of the order of 0.3 ms for the three transitions. (d) The Results of the Hahn-echo measurement show that we are far from the limit T1 = 2T2 . This indicates that the decoherence mechanism is not due to the relaxation process. (Changed from [96])

    1000 G. Taran et al.

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    1001

    Fig. 14 Spin valve behavior in a supramolecular spintronics device based on a carbon nanotube quantum dot functionalized with TbPc2 SMMs. Figures from [86,99]: (a) Butterfly hysteresis loop at T = 40 mK. (b) Antiparallel spin configuration: the spin state in dot A is reversed with respect to that of dot B. The energy mismatch between levels with different spins results in a current blockade. (c) Parallel spin configuration for both molecules A and B. Energy levels with same spin are aligned allowing electron transport through the carbon nanotube

    Quantum Algorithms The research aimed at using molecular magnets for qubit encoding started mainly after the theoretical proposal of Leuenberger and Loss [100, 101], followed by others [102,103,104,105,106,107]. We shall follow DiVincenzo [108] to summarize the steps taken toward successful implementation of a quantum computer using molecular magnets. To use molecular magnets as qubits, external constrains (e.g., temperature, electric or magnetic field) should be applied in order to confine the system to a subset of two levels (|0 and |1). The scalability of the system is facilitated by the synthetic bottom-up fabrication process that guarantees cheap production of identical molecular units. The implementation of proposed quantum algorithms [109, 110, 111, 112] involves the use of one-qubit and two-qubit gates. The former represents a rotation on the Bloch sphere. The output of the gate, after realizing an electric or magnetic coupling to the spin and applying an EM pulse sequence (Sect. “Quantum Coherence in Molecular Magnets”), is the state: α|0 + β|1. A strong-coupling regime, at high temperatures, between a molecular spin ensemble and microwave resonators, has been achieved. The possibility of coupling strongly with single molecules has been put forward, and experiments are in progress. This opens a way to develop scalable architectures using molecular spins coupled to quantum circuits [113, 114, 115, 116, 117]. Moreover, by molecular engineering of the crystal field, molecular spins can also be manipulated by electric fields [118, 119, 120, 84, 121]. A two-qubit gate can be implemented by controlling the exchange interaction between the spins. Schemes and compounds for switchable effective qubit–qubit interactions in the presence of permanent exchange couplings are now available

    1002

    G. Taran et al.

    [122, 123]. Spin entanglement between and within molecules was shown by different techniques. Molecular spin clusters/arrays offer an incredible variety of spin topologies to test entanglement at the molecular scale [119]. It is necessary that the system remains in a coherent state (preserve α and β), for a time considerably larger than the computational clock time. This is an important requirement for the system to be amenable to quantum error corrections. As noted in Sect. “Quantum Coherence in Molecular Magnets”, long coherence time (T2 ) at low temperature has been reported for many ensembles of molecular spin systems with T2 approaching 1 ms in nuclear spin-free environments [74, 124, 72]. Recent reports have shown microsecond coherence times and Rabi oscillations at room temperature [78, 125, 75, 126, 127, 79]. Strategies to protect spin states from decoherence (e.g., via atomic clock transitions) have been experimentally tested by fine engineering of molecular states and levels [128]. The development of theoretical schemes to implement quantum error correction codes in molecules with multiple spin degrees of freedom has also been started [129, 130, 131, 132, 133]. Another important requirement is the possibility to initialize and read out the qubit state. As seen in previous sections, the read-out of a single molecular spin located at a tunnel junction or on a CNT/graphene quantum dot has been demonstrated [134, 86, 135]. All these achievements are important milestones in molecular magnetism. They bring this research field closer to being able to provide the future basic components of quantum devices [136].

    Conclusion This chapter gives an overview of the main phenomenologies related to magnetic properties of molecular magnets and presents some important achievements in the field of molecular spintronics, aimed toward the implementation of a quantum computer. We highlight the symbiotic relationship between fundamental research and technological application as a central feature of this research field, the successful transposition of the properties of the molecular magnets into functional devices being a proof of the deep understanding acquired in the last decades.

    References 1. Caneschi, A., Gatteschi, D., Sessoli, R., Barra, A.L., Brunel, L.C., Guillot, M.: Alternating current susceptibility, high field magnetization, and millimeter band epr evidence for a ground s = 10 state in [Mn12 O12 (CH3 COO)16 (H2 O)4 ].2CH3 COOH.4H2 O. J. Am. Chem. Soc. 113(15), 5873–5874 (1991) 2. Sessoli, R., Tsai, H.-L., Schake, A.R., Sheyi, W., Vincent, J.B., Folting, K., Gatteschi, D., Christou, G., Hendrickson, D.N.: High-spin molecules:[mn12o12 (o2cr) 16 (h2o) 4]. J. Am. Chem. Soc. 115(5), 1804–1816 (1993) 3. Friedman, J.R., Sarachik, M., Tejada, J., Ziolo, R.: Macroscopic measurement of resonant magnetization tunneling in high-spin molecules. Phys. Rev. Lett. 76(20), 3830 (1996)

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    1003

    4. Thomas, L., Lionti, F., Ballou, R., Gatteschi, D., et al.: Macroscopic quantum tunnelling of magnetization in a single crystal of nanomagnets. Nature 383(6596), 145 (1996) 5. Sangregorio, C., Ohm, T., Paulsen, C., Sessoli, R., Gatteschi, D.: Quantum tunneling of the magnetization in an iron cluster nanomagnet. Phys. Rev. Lett. 78(24), 4645 (1997) 6. Wernsdorfer, W., Sessoli, R.: Quantum phase interference and parity effects in magnetic molecular clusters. Science 284(5411), 133–135 (1999) 7. Friedman, J.R., Sarachik, M.P.: Single-molecule nanomagnets. Annu. Rev. Condens. Matter Phys. 1(1), 109–128 (2010) 8. Lu, W., Lieber, C.M.: Nanoelectronics from the bottom up. Nature Mater. 6(11), 841–850 (2007) 9. Wernsdorfer, W., Chakov, N., Christou, G.: Determination of the magnetic anisotropy axes of single-molecule magnets. Phys. Rev. B 70(13), 132413 (2004) 10. Tasiopoulos, A.J., Vinslava, A., Wernsdorfer, W., Abboud, K.A., Christou, G.: Giant singlemolecule magnets: a {Mn84} torus and its supramolecular nanotubes. Angewandte Chemie 116(16), 2169–2173 (2004) 11. Stevens, K.: Matrix elements and operator equivalents connected with the magnetic properties of rare earth ions. Proc. Phys. Soc. Sect. A 65(3), 209 (1952) 12. Abragam, A., Bleaney, B.: Electron paramagnetic resonance of transition ions. OUP, Oxford (2012) 13. Waldmann, O.: A criterion for the anisotropy barrier in single-molecule magnets. Inorg. Chem. 46(24), 10035–10037 (2007) 14. Ishikawa, N., Sugita, M., Ishikawa, T., Koshihara, S.-Y., Kaizu, Y.: Mononuclear lanthanide complexes with a long magnetization relaxation time at high temperatures: a new category of magnets at the single-molecular level. J. Phys. Chem. B 108(31), 11265–11271 (2004) 15. Langley, S.K., Wielechowski, D.P., Moubaraki, B., Murray, K.S.: Enhancing the magnetic blocking temperature and magnetic coercivity of {Cr III2 Ln III2} single-molecule magnets via bridging ligand modification. Chem. Commun. 52(73), 10976–10979 (2016) 16. Goodwin, C.A., Ortu, F., Reta, D., Chilton, N.F., Mills, D.P.: Molecular magnetic hysteresis at 60 kelvin in dysprosocenium. Nature 548(7668), 439 (2017) 17. Layfield, R., Guo, F.-S., Day, B., Chen, Y.-C., Tong, M.-L., Mansikamäkki, A.: A dysprosium metallocene single-molecule magnet functioning at the axial limit. Angew. Chem. Int. Ed. 56(38), 11445–11449 (2017) 18. Ishikawa, N., Sugita, M., Okubo, T., Tanaka, N., Iino, T., Kaizu, Y.: Determination of ligand-field parameters and f-electronic structures of double-decker bis (phthalocyaninato) lanthanide complexes. Inorg. Chem. 42(7), 2440–2446 (2003) 19. Ishikawa, N., Sugita, M., Wernsdorfer, W.: Quantum tunneling of magnetization in lanthanide single-molecule magnets: bis (phthalocyaninato) terbium and bis (phthalocyaninato) dysprosium anions. Angew. Chem. Int. Ed. 44(19), 2931–2935 (2005) 20. Taran, G., Bonet, E., Wernsdorfer, W.: The role of the quadrupolar interaction in the tunneling dynamics of lanthanide molecular magnets. J. Appl. Phys. 125(14), 142903 (2019) 21. Taran, G., Bonet, E., Wernsdorfer, W.: Decoherence measurements in crystals of molecular magnets. Phys. Rev. B 99(18), 180408 (2019) 22. Zheng, Y.-Z., Zheng, Z., Chen, X.-M.: A symbol approach for classification of moleculebased magnetic materials exemplified by coordination polymers of metal carboxylates. Coord. Chem. Rev. 258, 1–15 (2014) 23. Gatteschi, D., Sessoli, R., Villain, J.: Molecular Nanomagnets, vol. 5. Oxford University Press on Demand, Oxford (2006) 24. Hill, S., Datta, S., Liu, J., Inglis, R., Milios, C.J., Feng, P.L., Henderson, J.J., del Barco, E., Brechin, E.K., Hendrickson, D.N.: Magnetic quantum tunneling: insights from simple molecule-based magnets. Dalton Trans. 39(20), 4693–4707 (2010) 25. Gatteschi, D., Bencini, A.: Electron Paramagnetic Resonance of Exchange Coupled Systems. Springer, Berlin/Heidelberg (1990) 26. Tupitsyn, I., Barbara, B.: Quantum tunneling of magnetization in molecular complexes with large spins–effect of the environment. Magnetism: molecules to materials: 5 Volumes Set, pp. 109–168 (2001)

    1004

    G. Taran et al.

    27. Barbara, B.: Quantum tunneling of the collective spins of single-molecule magnets: from early studies to quantum coherence. In: Molecular Magnets, pp. 17–60. Springer, Berlin/Heidelberg (2014) 28. Garg, A.: Topologically quenched tunnel splitting in spin systems without kramers’ degeneracy. EPL (Europhys. Lett.) 22(3), 205 (1993) 29. Schilling, R.: Quantum spin-tunneling: a path integral approach. In: Quantum Tunneling of Magnetization: QTM’94, pp. 59–76. Springer, Dordrecht (1995) 30. Garanin, D.: Spin tunnelling: a perturbative approach. J. Phys. A: Math. Gen. 24(2), L61 (1991) 31. Hartmann-Boutron, F., Politi, P., Villain, J.: Tunneling and magnetic relaxation in mesoscopic molecules. Int. J. Modern Phys. B 10(21), 2577–2637 (1996) 32. 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 magnetization. Phys. Rev. Lett. 82(19), 3903 (1999) 33. Wernsdorfer, W., Caneschi, A., Sessoli, R., Gatteschi, D., Cornia, A., Villar, V., Paulsen, C.: Effects of nuclear spins on the quantum relaxation of the magnetization for the molecular nanomagnet fe 8. Phys. Rev. Lett. 84(13), 2965 (2000) 34. Alonso, J.J., Fernández, J.F.: Tunnel window’s imprint on dipolar field distributions. Phys. Rev. Lett. 87(9), 097205 (2001) 35. Tupitsyn, I., Stamp, P., Prokof’ev, N.: Hole digging in ensembles of tunneling molecular magnets. Phys. Rev. B 69(13), 132406 (2004) 36. Landau, L.: Zur theorie der energieubertragung. II. Phys. Z. Sowjetunion 2(46), 1–13 (1932) 37. Zener, C.: Non-adiabatic crossing of energy levels. In: Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, vol. 137, pp. 696–702. The Royal Society (1932) 38. Stueckelberg, E.: Theory of inelastic collisions between atoms(theory of inelastic collisions between atoms, using two simultaneous differential equations). Helv. Phys. Acta (Basel) 5, 369–422 (1932) 39. Miyashita, S.: Dynamics of the magnetization with an inversion of the magnetic field. J. Phys. Soc. Jpn. 64(9), 3207–3214 (1995) 40. Miyashita, S.: Observation of the energy gap due to the quantum Tunneling making use of the landau-zener mechanism. J. Phys. Soc. Jpn. 65(8), 2734–2735 (1996) 41. Thorwart, M., Grifoni, M., Hänggi, P.: Tunneling and vibrational relaxation in driven multilevel systems. Phys. Rev. Lett. 85, quant-ph/9912024, 860 (2000) 42. Leuenberger, M.N., Loss, D.: Incoherent zener tunneling and its application to molecular magnets. Phys. Rev. B 61(18), 12200 (2000) 43. Wernsdorfer, W., Bhaduri, S., Vinslava, A., Christou, G.: Landau-zener tunneling in the presence of weak intermolecular interactions in a crystal of mn 4 single-molecule magnets. Phys. Rev. B 72(21), 214429 (2005) 44. Wernsdorfer, W.: From micro-to nano-squids: applications to nanomagnetism. Superconductor Sci. Technol. 22(6), 064013 (2009) 45. Ganzhorn, M., Wernsdorfer, W.: Molecular quantum spintronics using single-molecule magnets. In: Molecular Magnets, pp. 319–364. Springer, Berlin/Heidelberg (2014) 46. Bruno, P.: Berry phase, topology, and degeneracies in quantum nanomagnets. Phys. Rev. Lett. 96(11), 117208 (2006) 47. Demkov, Y.N., Kurasov, P.B.: Von neumann-wigner theorem: Level repulsion and degenerate eigenvalues. Theor. Math. Phys. 153(1), 1407–1422 (2007) 48. Henderson, J., Koo, C., Feng, P., Del Barco, E., Hill, S., Tupitsyn, I., Stamp, P., Hendrickson, D.: Manifestation of spin selection rules on the quantum tunneling of magnetization in a single-molecule magnet. Phys. Rev. Lett. 103(1), 017202 (2009) 49. Wernsdorfer, W., Bhaduri, S., Boskovic, C., Christou, G., Hendrickson, D.: Spin-parity dependent tunneling of magnetization in single-molecule magnets. Phys. Rev. B 65(18), 180403 (2002) 50. Del Barco, E., Kent, A.D., Hill, S., North, J., Dalal, N., Rumberger, E., Hendrickson, D., Chakov, N., Christou, G.: Magnetic quantum tunneling in the single-molecule magnet mn12acetate,” J. Low Temp. Phys. 140(1–2), 119–174 (2005)

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    1005

    51. De Raedt, H., Miyashita, S., Michielsen, K., Machida, M.: Dzyaloshinskii-moriya interactions and adiabatic magnetization dynamics in molecular magnets. Phys. Rev. B 70(6), 064401 (2004) 52. Josephson, B.D.: Possible new effects in superconductive tunnelling. Phys. Lett. 1(7), 251– 253 (1962) 53. Koelle, D., Kleiner, R., Ludwig, F., Dantsker, E., Clarke, J.: High-transition-temperature superconducting quantum interference devices. Rev. Modern Phys. 71(3), 631 (1999) 54. Aharonov, Y., Bohm, D.: Significance of electromagnetic potentials in the quantum theory. Phys. Rev. 115(3), 485 (1959) 55. Barra, A.-L., Debrunner, P., Gatteschi, D., Schulz, C.E., Sessoli, R.: Superparamagneticlike behavior in an octanuclear iron cluster. EPL (Europhys. Lett.) 35s(2), 133 (1996) 56. Wernsdorfer, W., Soler, M., Christou, G., Hendrickson, D.: Quantum phase interference (berry phase) in single-molecule magnets of [mn 12] 2-. J. Appl. Phys. 91(10), 7164–7166 (2002) 57. Wernsdorfer, W., Chakov, N., Christou, G.: Quantum phase interference and spin-parity in mn 12 single-molecule magnets. Phys. Rev. Lett. 95(3), 037203 (2005) 58. Adams, S., da Silva Neto, E.H., Datta, S., Ware, J., Lampropoulos, C., Christou, G., Myaesoedov, Y., Zeldov, E., Friedman, J.R.: Geometric-phase interference in a mn 12 singlemolecule magnet with fourfold rotational symmetry. Phys. Rev. Lett. 110(8), 087205 (2013) 59. Quddusi, H.M., Liu, J., Singh, S., Heroux, K., Del Barco, E., Hill, S., Hendrickson, D.: Asymmetric berry-phase interference patterns in a single-molecule magnet. Phys. Rev. Lett. 106(22), 227201 (2011) 60. Ramsey, C.M., Del Barco, E., Hill, S., Shah, S.J., Beedle, C.C., Hendrickson, D.N.: Quantum interference of tunnel trajectories between states of different spin length in a dimeric molecular nanomagnet. Nat. Phys. 4(4), 277–281 (2008) 61. Mossin, S., Stefan, M., ter Heerdt, P., Bouwen, A., Goovaerts, E., Weihe, H.: Fourth-order zero-field splitting parameters of [mn (cyclam) br2] br determined by single-crystal w-band epr. Appl. Magn. Reson. 21(3–4), 587–596 (2001) 62. Bertaina, S., Gambarelli, S., Mitra, T., Tsukerblat, B., Müller, A., Barbara, B.: Quantum oscillations in a molecular magnet. Nature 453(7192), 203–206 (2008) 63. Sorace, L., Wernsdorfer, W., Thirion, C., Barra, A.-L., Pacchioni, M., Mailly, D., Barbara, B.: Photon-assisted tunneling in a fe 8 single-molecule magnet. Phys. Rev. B 68(22), 220407 (2003) 64. Bal, M., Friedman, J.R., Suzuki, Y., Mertes, K., Rumberger, E., Hendrickson, D., Myasoedov, Y., Shtrikman, H., Avraham, N., Zeldov, E.: Photon-induced magnetization reversal in the fe 8 single-molecule magnet. Phys. Rev. B 70(10), 100408 (2004) 65. Wernsdorfer, W., Müller, A., Mailly, D., Barbara, B.: Resonant photon absorption in the lowspin molecule v15. EPL (Europhys. Lett.) 66(6), 861 (2004) 66. Petukhov, K., Wernsdorfer, W., Barra, A.-L., Mosser, V.: Resonant photon absorption in fe 8 single-molecule magnets detected via magnetization measurements. Phys. Rev. B 72(5), 052401 (2005) 67. Moreno-Pineda, E., Godfrin, C., Balestro, F., Wernsdorfer, W., Ruben, M.: Molecular spin qudits for quantum algorithms. Chem. Soc. Rev. 47(2), 501–513 (2018) 68. Rabi, I.I.: Space quantization in a gyrating magnetic field. Phys. Rev. 51(8), 652 (1937) 69. Grifoni, M., Hänggi, P.: Driven quantum tunneling. Phys. Rep. 304(5), 229–354 (1998) 70. Schweiger, A., Jeschke, G.: Principles of Pulse Electron Paramagnetic Resonance. Oxford University Press on Demand, Oxford (2001) 71. Zurek, W.H.: Decoherence, einselection, and the quantum origins of the classical. Rev. Modern Phys. 75(3), 715 (2003) 72. Takahashi, S., Tupitsyn, I., Van Tol, J., Beedle, C., Hendrickson, D., Stamp, P.: Decoherence in crystals of quantum molecular magnets. Nature 476(7358), 76–79 (2011) 73. Shim, J., Bertaina, S., Gambarelli, S., Mitra, T., Müller, A., Baibekov, E., Malkin, B., Tsukerblat, B., Barbara, B.: Decoherence window and electron-nuclear cross relaxation in the molecular magnet v 15. Phys. Rev. Lett. 109(5), 050401 (2012)

    1006

    G. Taran et al.

    74. Ardavan, A., Rival, O., Morton, J.J., Blundell, S.J., Tyryshkin, A.M., Timco, G.A., Winpenny, R.E.: Will spin-relaxation times in molecular magnets permit quantum information processing? Phys. Rev. Lett. 98(5), 057201 (2007) 75. Zadrozny, J.M., Niklas, J., Poluektov, O.G., Freedman, D.E.: Millisecond coherence time in a tunable molecular electronic spin qubit. ACS Cent. Sci. 1(9), 488–492 (2015) 76. Yang, J., Wang, Y., Wang, Z., Rong, X., Duan, C.-K., Su, J.-H., Du, J.: Observing quantum oscillation of ground states in single molecular magnet. Phys. Rev. Lett. 108(23), 230501 (2012) 77. Takahashi, S., van Tol, J., Beedle, C.C., Hendrickson, D.N., Brunel, L.-C., Sherwin, M.S.: Coherent manipulation and decoherence of s = 10 single-molecule magnets. Phys. Rev. Lett. 102(8), 087603 (2009) 78. Bader, K., Dengler, D., Lenz, S., Endeward, B., Jiang, S.-D., Neugebauer, P., van Slageren, J.: Room temperature quantum coherence in a potential molecular qubit. Nat. Commun. 5, 5304 (2014) 79. Atzori, M., Tesi, L., Morra, E., Chiesa, M., Sorace, L., Sessoli, R.: Room-temperature quantum coherence and rabi oscillations in vanadyl phthalocyanine: toward multifunctional molecular spin qubits. J. Am. Chem. Soc. 138(7), 2154–2157 (2016) 80. Cleuziou, J.-P., Wernsdorfer, W., Bouchiat, V., Ondarçuhu, T., Monthioux, M.: Carbon nanotube superconducting quantum interference device. Nat. Nanotechnol. 1(1), 53–59 (2006) 81. Bogani, L., Wernsdorfer, W.: Molecular spintronics using single-molecule magnets. Nat. Mater. 7(3), 179–186 (2008) 82. Roch, N., Florens, S., Bouchiat, V., Wernsdorfer, W., Balestro, F.: Quantum phase transition in a single-molecule quantum dot. Nature 453(7195), 633–637 (2008) 83. Sanvito, S., Rocha, A.R.: Molecular-spintronics: the art of driving spin through molecules. J. Comput. Theor. Nanosci. 3(5), 624–642 (2006) 84. Zyazin, A.S., van den Berg, J.W., Osorio, E.A., van der Zant, H.S., Konstantinidis, N.P., Leijnse, M., Wegewijs, M.R., May, F., Hofstetter, W., Danieli, C., et al.: Electric field controlled magnetic anisotropy in a single molecule. Nano Lett. 10(9), 3307–3311 (2010) 85. Thiele, S., Vincent, R., Holzmann, M., Klyatskaya, S., Ruben, M., Balestro, F., Wernsdorfer, W.: Electrical readout of individual nuclear spin trajectories in a single-molecule magnet spin transistor. Phys. Rev. Lett. 111(3), 037203 (2013) 86. Urdampilleta, M., Klyatskaya, S., Cleuziou, J.-P., Ruben, M., Wernsdorfer, W.: Supramolecular spin valves. Nat. Mater. 10(7), 502–506 (2011) 87. Fu, Y.-S., Schwöbel, J., Hla, S.-W., Dilullo, A., Hoffmann, G., Klyatskaya, S., Ruben, M., Wiesendanger, R.: Reversible chiral switching of bis (phthalocyaninato) terbium (iii) on a metal surface. Nano Lett. 12(8), 3931–3935 (2012) 88. Komeda, T., Isshiki, H., Liu, J., Zhang, Y.-F., Lorente, N., Katoh, K., Breedlove, B.K., Yamashita, M.: Observation and electric current control of a local spin in a single-molecule magnet. Nat. Commun. 2, 217 (2011) 89. Schwöbel, J., Fu, Y., Brede, J., Dilullo, A., Hoffmann, G., Klyatskaya, S., Ruben, M., Wiesendanger, R.: Real-space observation of spin-split molecular orbitals of adsorbed singlemolecule magnets. Nat. Commun. 3, 953 (2012) 90. Heersche, H., De Groot, Z., Folk, J., Van Der Zant, H., Romeike, C., Wegewijs, M., Zobbi, L., Barreca, D., Tondello, E., Cornia, A.: Electron transport through single mn 12 molecular magnets. Phys. Rev. Lett. 96(20), 206801 (2006) 91. Jo, M.-H., Grose, J.E., Baheti, K., Deshmukh, M.M., Sokol, J.J., Rumberger, E.M., Hendrickson, D.N., Long, J.R., Park, H., Ralph, D.: Signatures of molecular magnetism in single-molecule transport spectroscopy. Nano Lett. 6(9), 2014–2020 (2006) 92. Mannini, M., Sainctavit, P., Sessoli, R., Cartier dit Moulin, C., Pineider, F., Arrio, M.-A., Cornia, A., Gatteschi, D.: Xas and xmcd investigation of mn12 monolayers on gold. Chem.– A Eur. J. 14(25), 7530–7535 (2008) 93. Burzurí, E., Zyazin, A., Cornia, A., Van der Zant, H.: Direct observation of magnetic anisotropy in an individual fe 4 single-molecule magnet. Phys. Rev. Lett. 109(14), 147203 (2012)

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    1007

    94. Vincent, R., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Balestro, F.: Electronic read-out of a single nuclear spin using a molecular spin transistor. Nature 488(7411), 357–360 (2012) 95. Thiele, S., Balestro, F., Ballou, R., Klyatskaya, S., Ruben, M., Wernsdorfer, W.: Electrically driven nuclear spin resonance in single-molecule magnets. Science 344(6188), 1135–1138 (2014) 96. Godfrin, C., Ferhat, A., Ballou, R., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Balestro, F.: Operating quantum states in single magnetic molecules: implementation of grover’s quantum algorithm. Phys. Rev. Lett. 119(18), 187702 (2017) 97. Charlier, J.-C., Blase, X., Roche, S.: Electronic and transport properties of nanotubes. Rev. Modern Phys. 79(2), 677 (2007) 98. Nygård, J., Cobden, D.H., Lindelof, P.E.: Kondo physics in carbon nanotubes. Nature 408(6810), 342–346 (2000) 99. Urdampilleta, M., Klyatskaya, S., Ruben, M., Wernsdorfer, W.: Landau-zener tunneling of a single tb 3+ magnetic moment allowing the electronic read-out of a nuclear spin. Phys. Rev. B 87(19), 195412 (2013) 100. Leuenberger, M.N., Loss, D.: Quantum computing in molecular magnets. Nature 410(6830), 789–793 (2001) 101. Leuenberger, M.N., Loss, D.: Grover algorithm for large nuclear spins in semiconductors. Phys. Rev. B 68(16), 165317 (2003) 102. Meier, F., Levy, J., Loss, D.: Quantum computing with spin cluster qubits. Phys. Rev. Lett. 90(4), 047901 (2003) 103. Troiani, F., Ghirri, A., Affronte, M., Carretta, S., Santini, P., Amoretti, G., Piligkos, S., Timco, G., Winpenny, R.: Molecular engineering of antiferromagnetic rings for quantum computation. Phys. Rev. Lett. 94(20), 207208 (2005) 104. Carretta, S., Santini, P., Amoretti, G., Troiani, F., Affronte, M.: Spin triangles as optimal units for molecule-based quantum gates. Phys. Rev. B 76(2), 024408 (2007) 105. Troiani, F., Affronte, M.: Molecular spins for quantum information technologies. Chem. Soc. Rev. 40(6), 3119–3129 (2011) 106. Lehmann, J., Gaita-Arino, A., Coronado, E., Loss, D.: Spin qubits with electrically gated polyoxometalate molecules. Nat. Nanotechnol. 2(5), 312–317 (2007) 107. Bartolomé, J., Luis, F., Fernández, J.F.: Molecular magnets. Phys. Appl. Springer (2014) 108. DiVincenzo, D.P., et al.: The physical implementation of quantum computation. arXiv preprint quant-ph/0002077 (2000) 109. Deutsch, D., Jozsa, R.: Rapid solution of problems by quantum computation. In: Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, vol. 439, pp. 553–558. The Royal Society (1992) 110. Steane, A.: Quantum computing. Rep. Progress Phys. 61(2), 117 (1998) 111. Grover, L.K.: A fast quantum mechanical algorithm for database search. In: Proceedings of the Twenty-Eighth Annual ACM Symposium on Theory of Computing, pp. 212–219. ACM (1996) 112. Shor, P.W.: Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer. SIAM Rev. 41(2), 303–332 (1999) 113. Abe, E., Wu, H., Ardavan, A., Morton, J.J.: Electron spin ensemble strongly coupled to a three-dimensional microwave cavity. Appl. Phys. Lett. 98(25), 251108 (2011) 114. Chiorescu, I., Groll, N., Bertaina, S., Mori, T., Miyashita, S.: Magnetic strong coupling in a spin-photon system and transition to classical regime. Phys. Rev. B 82(2), 024413 (2010) 115. Clauss, C., Bothner, D., Koelle, D., Kleiner, R., Bogani, L., Scheffler, M., Dressel, M.: Broadband electron spin resonance from 500 mhz to 40 ghz using superconducting coplanar waveguides. Appl. Phys. Lett. 102(16), 162601 (2013) 116. Jenkins, M., Hümmer, T., Martínez-Pérez, M.J., García-Ripoll, J., Zueco, D., Luis, F.: Coupling single-molecule magnets to quantum circuits. New J. Phys. 15(9), 095007 (2013) 117. Carretta, S., Chiesa, A., Troiani, F., Gerace, D., Amoretti, G., Santini, P.: Quantum information processing with hybrid spin-photon qubit encoding. Phys. Rev. Lett. 111(11), 110501 (2013)

    1008

    G. Taran et al.

    118. Trif, M., Troiani, F., Stepanenko, D., Loss, D.: Spin electric effects in molecular antiferromagnets. Phys. Rev. B 82(4), 045429 (2010) 119. Troiani, F., Bellini, V., Candini, A., Lorusso, G., Affronte, M.: Spin entanglement in supramolecular structures. Nanotechnology 21(27), 274009 (2010) 120. Nossa, J., Islam, M., Canali, C.M., Pederson, M.: First-principles studies of spin-orbit and dzyaloshinskii-moriya interactions in the {Cu 3} single-molecule magnet. Phys. Rev. B 85(8), 085427 (2012) 121. Graham, M.J., Zadrozny, J.M., Shiddiq, M., Anderson, J.S., Fataftah, M.S., Hill, S., Freedman, D.E.: Influence of electronic spin and spin–orbit coupling on decoherence in mononuclear transition metal complexes. J. Am. Chem. Soc. 136(21), 7623–7626 (2014) 122. Timco, G.A., Carretta, S., Troiani, F., Tuna, F., Pritchard, R.J., Muryn, C.A., McInnes, E.J., Ghirri, A., Candini, A., Santini, P., et al.: Engineering the coupling between molecular spin qubits by coordination chemistry. Nat. Nanotechnol. 4(3), 173–178 (2009) 123. Ferrando-Soria, J., Pineda, E.M., Chiesa, A., Fernandez, A., Magee, S.A., Carretta, S., Santini, P., Vitorica-Yrezabal, I.J., Tuna, F., Timco, G.A., et al.: A modular design of molecular qubits to implement universal quantum gates. Nat. Commun. 7, 11377 (2016) 124. Ardavan, A., Bowen, A.M., Fernandez, A., Fielding, A.J., Kaminski, D., Moro, F., Muryn, C.A., Wise, M.D., Ruggi, A., McInnes, E.J., et al.: Engineering coherent interactions in molecular nanomagnet dimers. arXiv preprint arXiv:1510.01694 (2015) 125. Zadrozny, J.M., Niklas, J., Poluektov, O.G., Freedman, D.E.: Multiple quantum coherences from hyperfine transitions in a vanadium (iv) complex. J. Am. Chem. Soc. 136(45), 15841– 15844 (2014) 126. Tesi, L., Lucaccini, E., Cimatti, I., Perfetti, M., Mannini, M., Atzori, M., Morra, E., Chiesa, M., Caneschi, A., Sorace, L., et al.: Quantum coherence in a processable vanadyl complex: new tools for the search of molecular spin qubits. Chem. Sci. 7(3), 2074–2083 (2016) 127. Gómez-Coca, S., Urtizberea, A., Cremades, E., Alonso, P.J., Camón, A., Ruiz, E., Luis, F.: Origin of slow magnetic relaxation in kramers ions with non-uniaxial anisotropy. Nat. Commun. 5 4300 (2014) 128. Shiddiq, M., Komijani, D., Duan, Y., Gaita-Ariño, A., Coronado, E., Hill, S.: Enhancing coherence in molecular spin qubits via atomic clock transitions. Nature 531(7594), 348–351 (2016) 129. Hodges, J.S., Yang, J.C., Ramanathan, C., Cory, D.G.: Universal control of nuclear spins via anisotropic hyperfine interactions. Phys. Rev. A 78(1), 010303 (2008) 130. Santini, P., Carretta, S., Troiani, F., Amoretti, G.: Molecular nanomagnets as quantum simulators. Phys. Rev. Lett. 107(23), 230502 (2011) 131. Zhang, Y., Ryan, C.A., Laflamme, R., Baugh, J.: Coherent control of two nuclear spins using the anisotropic hyperfine interaction. Phys. Rev. Lett. 107(17), 170503 (2011) 132. Whitehead, G.F., Cross, B., Carthy, L., Milway, V.A., Rath, H., Fernandez, A., Heath, S.L., Muryn, C.A., Pritchard, R.G., Teat, S.J., et al.: Rings and threads as linkers in metal–organic frameworks and poly-rotaxanes. Chem. Commun. 49(65), 7195–7197 (2013) 133. Ueda, A., Suzuki, S., Yoshida, K., Fukui, K., Sato, K., Takui, T., Nakasuji, K., Morita, Y.: Hexamethoxyphenalenyl as a possible quantum spin simulator: an electronically stabilized neutral π radical with novel quantum coherence owing to extremely high nuclear spin degeneracy. Angew. Chem. Int. Ed. 52(18), 4795–4799 (2013) 134. Candini, A., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Affronte, M.: Graphene spintronic devices with molecular nanomagnets. Nano Lett. 11(7), 2634–2639 (2011) 135. Urdampilleta, M., Klayatskaya, S., Ruben, M., Wernsdorfer, W.: Magnetic interaction between a radical spin and a single-molecule magnet in a molecular spin-valve. ACS Nano 9(4), 4458–4464 (2015) 136. Ghirri, A., Troiani, F., Affronte, M.: Quantum computation with molecular nanomagnets: achievements, challenges, and new trends. In: Molecular Nanomagnets and Related Phenomena, pp. 383–430. Springer (2014)

    19 Single-Molecule Magnets and Molecular Quantum Spintronics

    1009

    Taran Gheorghe born in R. Moldova in 1990. He finished his bachelor and master degree in physics at University “Alexandru Ioan Cuza,” Romania. Currently, he is pursuing a doctoral degree in molecular magnetism at the Karlsruhe Institute of Technology in Germany.

    Edgar Bonet studied physics and computer science at the “École Normale Supérieure” in Lyon. He completed his PhD at the Louis Néel laboratory, in Grenoble, France, in 1999. After a 2-year postdoctoral position at Cornell University, he came back to Grenoble as a permanent researcher. He now works at the Institut Néel on nanomagnetism and molecular magnetism.

    Wolfgang Wernsdorfer born in Germany in 1966, studied physics at the University of Würzburg and École Normale Supérieure in Lyon. In 1993, he became a doctoral researcher at the Low Temperature Laboratory and the Laboratoire de Magnetism in Grenoble, France, where he then stayed as a researcher. In 2004, he became directeur de recherche at the Institut NEEL in Grenoble. In 2016, he accepted a position as Humboldt Professor at the Institute of Physics and the Institute of Quantum Materials and Technologies at the KIT.

    Magnetic Nanoparticles

    20

    Sara A. Majetich

    Contents Ideal Single-Domain Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . How Real Magnetic Nanoparticles Can Be Different . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Nanoparticle Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . General Nanoparticle Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Liquid Phase Syntheses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Applications of Magnetic Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Important Magnetic Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Separation and Manipulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Hyperthermia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Particle Imaging (MPI) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Contrast Agents for Magnetic Resonance Imaging (MRI) . . . . . . . . . . . . . . . . . . . . . . . . . Ferrofluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Recording Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nonbiomedical Topics of Recent Interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . L10 FePt Nanoparticles for Magnetic Recording Media . . . . . . . . . . . . . . . . . . . . . . . . . . . Core-Shell NPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Surface Effects and Spin Canting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Frontiers and Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    1012 1015 1017 1017 1018 1021 1021 1023 1024 1026 1026 1028 1029 1030 1030 1032 1033 1035 1037

    Abstract

    The physics of ideal single-domain nanoparticles is reviewed, followed by a discussion of how the behavior of real particles can differ. The main synthetic approaches are surveyed, and the advantages of different methods are identified. The magnetic properties of the nanoparticles are related to their use in applica-

    S. A. Majetich () Physics Department, Carnegie Mellon University, Pittsburgh, PA, 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_20

    1011

    1012

    S. A. Majetich

    tions including magnetic separation, ferrofluids, magnetic recording media, and biomedicine. Topics of recent interest are discussed, along with future directions.

    Monodomain magnetic nanoparticles (NPs) may be treated to first order as macrospins, but their detailed behavior can be quite complex. They have been extensively studied because they serve as models for magnetic recording media, and they are used in magnetic fluids and biomedical applications. Here we will focus on small magnetic nanoparticles prepared by chemical means that can be dispersed in liquids and moved by external forces. We start with a short description of the physics that unifies the field of magnetic nanoparticles and enables comparison of different materials. Next we discuss the differences between real and idealized NPs. Different approaches to NP synthesis are presented, along with explanations of constraints imposed by their intended use. The roles of magnetic NPs in applications in biomedicine, ferrofluids, and magnetic recording are discussed, along with the potential for future applications. Finally there is a survey of some recent trends in magnetic NP research. There have been numerous reviews on magnetic nanoparticles, from the perspective of theoretical models [1], collective behavior [2], and biomedical applications [3–5]. For reasons of space, some subfields will be limited or omitted, including nanoparticle synthesis, lithographically patterned nanoparticles, and atomistic simulations. This review is intended as a guide to the magnetism that unites the field of magnetic nanoparticles, together with a survey of their applications and current research trends.

    Ideal Single-Domain Nanoparticles In 1930 Frenkel and Dorfman first suggested the existence of ferromagnetic domains [6]. Bitter confirmed them by dusting fine magnetic particles over the surface of a ferromagnetic material [7]. We now know that in bulk ferromagnetic materials, the spins are parallel due to short-range quantum mechanical exchange interactions. There are also long-range magnetostatic interactions, and the combination of exchange and magnetostatic interactions leads to the spontaneous formation of magnetic domains, where locally the spins are parallel but over the sample the domain magnetization directions form magnetic flux closure paths to minimize the overall energy. Below a critical size, it is not energetically favorable to form a domain wall, and the particle is said to be monodomain. An ellipsoidal particle of a ferromagnetic or ferromagnetic material will be uniformly magnetized. The magnetic moment of the particle μ = μ0 Ms V, where Ms is the saturation magnetization and V is the particle volume. The maximum monodomain diameter can be estimated from the parameters of the bulk ferromagnetic (FM) or ferromagnetic (FiM) material [8]:

    20 Magnetic Nanoparticles

    1013

    Fig. 1 A monodomain magnetic nanoparticle. Here the direction of the particle moment μ depends on the effective anisotropy, which determines the easy axis direction, and the external magnetic field H

    Easy Axis

    µ H

    dcr =

    72(Ak)1/2 μ0 Ms2

    (1)

    Here A is the exchange stiffness and K is the anisotropy. A and Ms depend on the material, while the anisotropy has contributions from the crystallographic structure and direction and particle shape. Figure 1 shows the case for spherical particle of a uniaxial material, where the energy is minimized when the magnetization lies along an easy axis. For more complex cases with biaxial and cubic anisotropy, see Ref. [1]. Maximum single-domain sizes for different materials based on bulk parameters have been tabulated elsewhere [9]. While Eq. 1 is useful as a guide, there are many reasons it can be inaccurate. Surface spins have fewer near neighbors for exchange coupling. Symmetry breaking at the surface leads to an additional anisotropy contribution that can distort the NP spin configuration [10]. If the oxidation state of surface atoms differs from that of the interior, as with metal NPs, the effective Ms will be reduced compared with that for the bulk metal, even if the spins are not canted. For materials with very large magnetocrystalline anisotropy, the theoretical dcr could be hundreds of nanometers, but structural defects such as grain boundaries or antiphase boundaries limit the experimentally observed domain size. At the opposite extreme, with very low magnetocrystalline anisotropy, there is little energetic cost if they are not perfectly collinear. Here the domain wall width is a better estimate of the single-domain threshold. In bulk ferromagnets magnetization reversal occurs through domain wall motion, but in single-domain particles, the exchange-coupled spins remain parallel to one another as they rotate coherently, as a macrospin [11, 12]. The switching field or coercivity, Hc , is a maximum for particles at the largest monodomain size [13] and can be predicted from an energy barrier model [14], where the barrier to reversal is given by   Ms H 2 E = KV 1 − . 2K Hc drops until it reaches zero at the superparamagnetic particle size dSP :

    (2)

    1014

    S. A. Majetich

     dSP =

    6kT B π K ln (τmeas /τ0 )

    1/3 ,

    (3)

    where τ 0 is the inverse of the Larmor precession frequency, ∼10−9 s, and τ meas is the measurement time. The threshold temperature for superparamagnetism is called the blocking temperature, TB , defined as TB = KV ln (τmeas /τ0 ) /k,

    (4)

    but it too will depend on the measurement time. For a superconducting quantum interference effect device (SQUID) magnetometer, the time is ∼100 s, but Mössbauer spectroscopy and neutron scattering have measurement times of ∼10−7 s or shorter. In magnetic hyperthermia, the frequency ranges from 100 kHz to 1 MHz, and particles that look superparamagnetic based on SQUID magnetometry may be blocked over the shorter time scales [2]. By definition, superparamagnetic particles do not interact with each other, and the magnetization of an ensemble of superparamagnets, relative to its value at zero temperature, is a Langevin function: M 1 = L(x) = coth(x) − , M(0) x

    (5)

    where x = μH/kT. Figure 2 shows an example of a Langevin function fit. It is important that the sample is dilute (∼0.1 vol.%), or magnetostatic interactions of the particles will distort the curve and will typically lead to underestimation of the particle moment μ.

    Fig. 2 Langevin function fit of a dilute sample of 8.5 nm Fe3 O4 nanoparticles. Below the blocking temperature of ∼65 K, the two branches of the loop have been averaged. (unpublished results)

    20 Magnetic Nanoparticles

    1015

    The dynamics of relaxation for superparamagnets are predicted by the Néel model [12, 15], an ensemble of identical, rigid, noninteracting particles has a magnetization that decays exponentially with time. The Néel-Brown model assumes an Arrhenius law for the escape rate from a metastable state. Rearranging Eq. 4, the rate for a given temperature is τ 1 = τ0 1 exp (−KV /kB T ) .

    (6)

    This is a reasonable approximation if the particle spends the majority of its time in the local or global ground state, rather than in a transition state. To understand dynamics on shorter time scales (100 ns or less), the Landau-Lifshitz-Gilbert equation provides a more accurate picture of precessional switching [16].

    How Real Magnetic Nanoparticles Can Be Different The view of magnetic NPs as macrospins is a helpful starting point, but real particles can be more complex in many ways. Here we describe a few of the most common differences. The nanoparticles need not be chemically uniform. Most nanoparticles have a surfactant or polymer coating that forms a diffusion barrier to oxygen, but does not prevent oxidation of sensitive materials. Sudden exposure to air can cause magnetic metal NPs of Fe or Co to catch fire, but gradual exposure leads to a core-shell structure. Metals oxidize by O2 adsorption and dissociation on the surface, followed by metal ion migration to react and nucleate the metal oxide. The oxide shells surrounding magnetic metal cores are usually polycrystalline, with ∼1–2 nm grains having different crystallographic orientations, relative to the interior. In extreme cases, the Kirkendall effect is observed, in which the particles become hollow, and all that remains is a porous metal oxide shell [17]. Since Fe and Co oxides can be magnetically ordered. Bulk Fe3 O4 and γ-Fe2 O3 are ferromagnetic, while FeO is antiferromagnetic. Bulk CoO and Co3 O4 are antiferromagnetic. When these phases form a thin shell around a metallic core, there can be exchange bias between metal core and oxide shell spins [18], just as in bilayer thin films [19, 20]. However, the behavior is much more complex due to the polycrystalline nature of the shell, which leads to a wide variety of superexchange interactions within a single particle. Many NPs show reduced Ms compared with that of the bulk material. In some cases thin magnetic films and NPs are said to have a “dead layer” without ascribing an origin, but there are many hypotheses about the underlying reasons. Some are inherent in the symmetry breaking at the surface. This introduces a new contribution to the anisotropy that can distort the spins from the completely parallel configuration of the ideal monodomain particle. Negative surface anisotropy favors spins that lie in the plane of the surface [10, 21]. Positive surface anisotropy favors perpendicular orientation [10]; at moderate values this leads to surface distortion akin to a point dipole pattern. If the NP cores are amorphous, the surface anisotropy will dominate [22]. Many groups have assumed that the surface anisotropy contribution

    1016

    S. A. Majetich

    is inversely proportional to the NP diameter, but in reality many different factors contribute to an effective anisotropy [23]. Not all NPs show significant reduction in Ms [24–26], but in real NPs there are two additional factors that can reduce the surface magnetization: structural disorder and chemical variations. A spin glass model was proposed to explain the behavior of ball-milled ferrite NPs [27, 28]. Materials that have a combination of ferro- and antiferro-magnetic exchange interactions, such as spinel ferrites and L10 FePt [29], are most susceptible to disorder-induced magnetization reductions. However, this disorder could occur either at the surface or in the interior, in the form of antiphase boundaries [25]. NP surfaces have a unique form of disorder due to their curved surfaces; a broader range of near neighbor exchange interactions, relative to the interior, can lead to a magnetic frustration that would reduce the net magnetization. Most NPs have something bonded to their surface to prevent agglomeration, and unless the species is physisorbed, it will alter the chemical composition of the surface. An example of this would be oleic acid bonding to iron oxide NPs, where the two oxygen ions of the carboxylic acid bind with a surface Fe ion. Surfactant coverage varies with the preparation method, and also with the type of surface. Oleic acid binds more strongly to (100), leading to faster growth along directions [30]. The particle shape can be varied depending on the proportion of oleic acid and weakly binding oleyl amine [30]. Without complete passivation, there will be a chemically and magnetically inhomogeneous surface. In recent years attention has been paid to the inhomogeneity of oxygen distribution in iron oxide NPs. Many groups have observed ∼5% oxygen vacancies in “magnetite-like” NPs [25, 31]. While much closer to Fe3 O4 than to γ-Fe2 O3 in composition, these particles did not undergo a Verwey transition near 120 K that is characteristic of bulk magnetite. The electrical conductivity of Fe3 O4 is three orders of magnitude greater than that of γ-Fe2 O3 and is associated with charge hopping between Fe2+ and Fe3+ or polaron formation [32]. At high temperature, γ-Fe2 O3 has the structure of Fe3 O4 with oxygen vacancies and without Fe2+ ions. The addition of oxygen vacancies to a Fe3 O4 NP would change the chemistry and reduce the conductivity. Magnetite NPs that undergo a Verwey transition have been routinely prepared by aqueous methods [33], but here particles are rougher and less monodisperse than those from inert atmosphere methods. The latter approach has been modified using CO to generate small, smooth NPs that have a metal-insulator transition [34], and it is hoped that syntheses with less toxic reagents will be developed in the future. While most magnetic nanoparticles are nominally spherical, some degree of shape anisotropy is common. Bulk Fe3 O4 has [35] easy axis directions. However, a 5% distortion of a 10 nm magnetite sphere is sufficient to make it uniaxial due to shape anisotropy. Faceting is commonly observed in NPs. The rate of growth of (111) and (100) faces depends on the type and concentration of the surfactant. Truncated octahedra are frequently observed. While cube-shaped NPs have been prepared for a variety of materials [36–38], the cubes generally have rounded corners. Needle-shaped γ-Fe2 O3 nanoparticles where the shape anisotropy makes them uniaxial, with coercivity ∼800 Oe, were used as magnetic recording media for tapes and for the first hard disk drive (RAMAC). However, there is

    20 Magnetic Nanoparticles

    1017

    less research focus on elongated (acicular) magnetic nanoparticles today because thin films are used for hard disk media, and other research favors NPs that are either superparamagnetic – for biomedical applications and ferrofluids, or more monodisperse – for fundamental studies and self-assembly. Magnetic nanoparticles can also interact with each other. Quantum mechanical exchange requires distances of a few Ångstroms. Unless there is an intermediate magnetic phase to couple the nanoparticles, as in the case of nanocrystalline materials where an amorphous matrix surrounds Fe-rich nanocrystals [39], exchange interactions are usually ignored. However, the magnetic moment of a NP generates a dipolar magnetic field that can be up to tens of kAm−1 nearby. For a grain of magnetic recording media, this field is small relative to the switching field, but in composites or assemblies made from NPs with lower anisotropy, magnetostatic interactions can impact the collective response. There are no simple models of dipolar interactions in nanoparticle assemblies, since the dipolar field is long-range and anisotropic. In disordered assemblies or polydisperse samples, this leads to spin glass-like behavior with a complex energy barrier landscape. Deviations from noninteracting response are most pronounced in the magnetization dynamics. The imaginary susceptibility χ  shows differences in ferrofluids with even 1 volume percent of iron oxide NPs [40]. For the same reason, magnetic hyperthermia applications where the NPs are exposed to an AC magnetic field will depend on whether the particles are dilute or densely clustered. In the extreme limit of an ordered magnetic assembly with strong dipolar interactions, dipolar ferromagnetism is observed [41, 42]. Here there are multiparticle magnetic domains, and the magnetization is stable over time. In between the noninteracting and dipolar ferromagnet cases are superferromagnets, where the moment directions of a cluster of neighboring particles are correlated, but the direction changes over time [43]. Two opposing effects modify the magnetization curve for interacting NPs [2, 44, 45]. Disorder due to position or size introduces magnetic frustration that leads to a slower approach to saturation, relative to an assembly of noninteracting particles. Random crystallographic orientation reduces the effective anisotropy of the assembly, leading to saturation at reduced fields. When trying to understand experimental results for magnetic nanoparticles, it is therefore important to consider how they may differ from the idealized macrospin model, both in terms of the spin configuration within an individual NP and also in terms of magnetostatic interactions with other NPs.

    Magnetic Nanoparticle Preparation General Nanoparticle Formation According to the La Mer model, nanoparticles form when a nucleation threshold is researched in a supersaturated solution, above which clusters tend to grow rather than dissolve [46]. This can occur in either the gas or liquid phases. Most magnetic nanoparticles are synthesized by solution chemistry methods because it is easier to

    1018

    S. A. Majetich

    make the degree of supersaturation spatially homogeneous. Nucleation then occurs everywhere at the same time, leading to highly monodisperse nanoparticles. Solid-based routes to magnetic nanoparticles involve either lithography to make patterned elements, or else ball milling to crush a bulk-like material. Ball milling introduces structural damage in order to shatter larger particles and can introduce impurities from the milling material and walls of the container. The particle size distribution from ball milling is log normal, and this makes it difficult to analyze the magnetic behavior because a small number of large particles can dominate the magnetic properties. Nanoparticles made from solids will therefore not be discussed further in this review. Gas phase methods have been used but tend to have greater tradeoffs between the production rate and the particle quality. These particles are as close as possible to the “bare” NPs of theoretical calculations, but they must be trapped in a passivating matrix, or they will not stay pristine for long. Cluster beam sources have been used to measure the evolution of magnetic moment as a function of cluster size [47] and to prepare NPs of novel materials such as Co2 Si [48] and Fe nitride [49]. While it is possible to capture the clusters in a matrix of surfactant and then redisperse them in a carrier solvent, the throughput is low. In contrast, gas phase aerosol methods can produce kilograms of magnetic NPs. Here the primary particle size is nanoscale, but the particles are branched or aggregated, and cannot be broken apart. For a catalyst the large surface area is beneficial, but for ferrofluid, biomedical, or magnetic storage-related applications, agglomeration is undesirable.

    Liquid Phase Syntheses Liquid phase production of nanoparticles is typically on the milligram to gram scale. Liquids enable greater control of precursor homogeneity, though high dilution is still required to prevent coalescence of the growing clusters. Two methods are used to minimize the likelihood that colliding clusters will stick together: surface charge for Coulombic repulsion and organic surfactants for steric repulsion. The preferred synthetic route depends on the end use of the nanoparticles, and details specific to particular applications will be discussed later. Figure 3 shows the transmission electron microscopy images of a variety of NPs made by liquid phase synthesis.

    Nonaqueous Syntheses For studies of fundamental physics where monodispersity and high crystallinity are important, hydrophobic syntheses have been most promising. Here a high boiling point organic solvent is used to control the thermal decomposition of organometallic precursors and therefore the nucleation threshold. This is done in a dilute dispersion in the presence of surfactant molecules to prevent coalescence. The most commonly used surfactant for magnetic nanoparticles is oleic acid, which has a carboxylic acid group that binds to magnetic metal atoms at the NP surface. With Au and CdSe NPs, surfactants with thiol groups are used because they form strong bonds to Au or Cd. With magnetic nanoparticles there are no surfactants that form such strong

    20 Magnetic Nanoparticles

    1019

    Fig. 3 Transmission electron microscopy images of different types of Fe3 O4 nanoparticles. (a) Spherical NPs with a size of 13 nm prepared by high-temperature decomposition in nonaqueous solvent, (b) cube-shaped NPs, (c) NPs made by aqueous coprecipitation, and (d) magnetic beads containing NPs. (Note: distortion of the beads is due to particle melting of the polystyrene matrix by the electron beam)

    bonds, and there will be an equilibrium between adsorbed and desorbed surfactant on the NP surface. For this reason care should be taken when diluting a dispersion of magnetic NPs with a pure solvent. The NP size can be controlled to some extent by temperature, reaction time, and the precursor to surfactant ratio [50–52]. However, most of these syntheses have a sweet spot, and it is difficult to find a single synthesis that can be tuned over a wider size range. The upper limit is often set by the density of the compound or alloy. Using the same surfactant molecules, it is more difficult to sterically stabilize dense materials (e.g., Au or FePt) than lighter ones (e.g., Fe3 O4 or γ-Fe2 O3 ). Another problem is the occurrence of antiphase boundaries. Some of the iron oxide syntheses proceed through a FeO (wustite) NP precursor that is then transformed into Fe3 O4 . In larger FeO NPs, there can be multiple nucleation sites, which can lead to antiphase boundaries when the grains meet [25]. The Fe-O-Fe superexchange across the antiphase boundary is antiferromagnetic and can make the NPs multi-domain even when they are very small. This can lead to greatly

    1020

    S. A. Majetich

    reduced magnetic moments, and apparently higher effective anisotropy, as seen in the blocking temperature.

    Aqueous Syntheses Most magnetic NPs for biomedicine are spinel oxide NPs such as Fe3 O4 or γ-Fe2 O3 made in aqueous solution [53–56]. In this coprecipitation method, metal salts are dissolved, and a base such as ammonium hydroxide is used to generate a metal oxyhydroxide phase that loses water to form a spinel ferrite. The reaction temperature limits the crystallinity of the NPs, though improvements are achieved by calcining, heating near the boiling point of water, or through hydrothermal synthesis where the particles are heated under pressure in an autoclave. Once the NPs are formed, a surfactant is usually added in order to aid in stabilization. Since iron oxide is amphoteric, collisions between uncoated particles can neutralize the surface charge. Without surfactant coating a typical charge-stabilized dispersion will show noticeable settling within a day. Compared with nonaqueous solution syntheses , the coprecipitation and hydrothermal methods yield particles with greater polydispersity and lower crystallinity. Depending on the method and particle size, the specific saturation magnetization of these particles can be significantly lower than that of the bulk iron oxide. Nanoparticle Coatings and Extra Requirements of Biomedical Applications The NP surface coating prevents agglomeration and helps stabilize the particle dispersion, since the coating is made of lower density material than the core. To disperse in nonpolar solvents, the surfactant coating molecules should have nonpolar tail groups that extend into the solvent. Using dynamic light scattering, it is possible to measure the hydrodynamic diameter of the NPs and their coating and therefore to determine whether the particles have single or multiple inorganic cores. For virtually all biomedical uses, the NPs must form stable aqueous dispersions, often with a high salt concentration such as that found in blood. There are numerous ways to modify the as-made NP coating so that they can form aqueous dispersions [57, 58]. Ex vivo biomedical applications are less restrictive than those where particles enter the body. NPs must still be chemically inert and biocompatible and form stable dispersions, but here there is more flexibility to use materials with enhanced magnetic properties relative to those of iron oxide. A popular method to increase the magnetic moment while retaining dispersion stability is to encase many magnetic NPs, usually iron oxide, into a polymer bead that is typically 1–10 microns in diameter [59]. There are areas for improvement in the magnetic beads. One would be to improve the magnetic uniformity of the beads, so that they all have the same net magnetic moment. Ideally this is done with a high packing density, which not only increases the magnetic moment but also leads to superferromagnetism when the individual monodomain NPs have strong magnetostatic interactions within the bead. Compared with an individual superparamagnet, a superferromagnet has a larger magnetization at lower fields; i.e., it approaches saturation more rapidly. Another

    20 Magnetic Nanoparticles

    1021

    strategy would be to replace magnetite or maghemite with monodomain NPs of a higher magnetization material such as Fe or FeCo. So far the limitation for this has been the difficulty of chemical stabilization. NPs that will be used in vivo have the strictest requirements. They must be nontoxic, biocompatible, biodegradable, noncarcinogenic, non-mutagenic, and not trigger an immunoresponse. They cannot sediment or form large agglomerates that could clog arteries (>20 microns), which could potentially cause a stroke or a heart attack. For good manufacturing practices, all of the chemicals used in the synthesis and purification should also be nontoxic. The combination of these needs rules out anything made by ball milling (some particles will be too big), NPs of cobalt or nickel (carcinogenic), and NPs made by high-temperature decomposition (due to residual traces of the high boiling point organic solvents). For these reasons magnetic NPs intended for in vivo use are typically made in aqueous solution. Typically they have ∼10 nm particle cores and an organic coating, so they form a stable aqueous dispersion. In order to retain stability in biological media, which has a high salt concentration, they are sterically stabilized rather than charge stabilized, as in an aqueous ferrofluid. The NP surface coating affects how the body responds to the NP. Protein molecules quickly adsorb, and depending on the charge, this can signal macrophages to remove the NPs from circulation. It is important for diagnostic and therapeutic applications that the NPs are not quickly cleared by the reticuloendothelial system (RES). Because Fe3 O4 will oxidize to γ-Fe2 O3 over time in aqueous dispersions, there is often a mixture of phases [60]. It is significant if either Fe2+ from a NP is solubilized, since this can generate reactive oxygen species (ROS) catalytically. The onset of numerous neurodegenerative diseases is associated with an upset in the balance of ROS. The US Food and Drug Administration (FDA) has approved only a few types of iron oxide (magnetite and maghemite), for in vivo use, under the trade names Resovist, Feridex, and Endorem. The coatings are typically polysaccharides. Other magnetic materials may have superior magnetic properties, but getting approval for in vivo use is a long and expensive process. The European Commission has been active in sponsoring studies related to health consequences of nanomaterials [61, 62]. There are no simple answers or sweeping generalizations, partly because so many factors concerning nanoparticles , including their size, coating, and degree of dispersion as well, and their chemical composition all contribute to the biological response.

    Applications of Magnetic Nanoparticles Important Magnetic Characteristics Most applications of magnetic nanoparticles rely on their small size, which enables them to be dispersed in liquids and moved with a magnetic field. These applications depend on several magnetic properties, but the most central is the magnetic moment → | − μ |= Ms V . The magnitude of μ depends on the material the particle is made

    1022

    S. A. Majetich

    of through Ms and the particle volume V. The maximum V can be limited by the maximum monodomain size (Eq. 1), or the dispersion stability, which can be problematic for large sizes or if dense coating materials such as Au are used. If particle clumping is undesirable, the superparamagnetic limit (Eq. 3) restricts the maximum size, though the effective moment can be increased by embedding many magnetic nanoparticles in a polymer to create a magnetic bead, which is typically microns in diameter. In magnetic resonance imaging (MRI) contrast agents, the spatially inhomogeneous magnetic field generated by μ is given by μ cos θ − → 2μ cos θ H = θˆ , rˆ + 3 r r3

    (7)

    where θ is the angle between the moment μ and the point of interest. This local field shifts the resonance frequency of nearby water molecules and enables the removal of the background signal from water molecules so that the response of protons (1H) in tissue can be seen more clearly. In magnetic separation, a spatially varying external magnetic field is used to apply a force on the magnetic moment: F = (μ · ∇) H

    (8)

    A strong permanent magnet causes a small deflection of a US one dollar bill, due to iron oxide NPs embedded in the cellulose fibers in certain regions. Magnetic paper [63], copying machine toner [64], and hydrogel composites [65, 66] have used magnetic field gradients to manipulate materials containing magnetic nanoparticles. Table 1 summarizes the magnetic properties that are most important for different applications. While there are exceptions, this table is intended as a guide to understand whether NPs should be superparamagnetic (not stable) and whether performance could potentially be improved by increasing the magnetic moment (to generate a larger field and to manipulate in a field or field gradient).

    Table 1 Important magnetic characteristics for applications of magnetic nanoparticles Application Magnetic separation, manipulation Magnetic hyperthermia Magnetic particle imaging Magnetic resonance imaging Ferrofluids Magnetic recording media

    Generates field Y (for detection)

    Moves in field Y

    Magnetically stable N

    N

    N

    Y (for detection) Y

    N N

    N (for DC) but Y (for AC field used) N N

    Y (to form chains) Y (for reading)

    Y N

    N Y

    20 Magnetic Nanoparticles

    1023

    Magnetic Separation and Manipulation Here micron-sized magnetic beads are used instead of individually dispersed nanoparticles because they provide larger magnetic moment together with a reduced viscous drag force. There are numerous references that focus on methods to manipulate NPs [67–69] as well as the effect of NP shape [70]. For biomedical sensing applications, the surfaces of the polymer beads can be functionalized for selective binding to specific molecules or biomarkers, and a single bead can have different types of binding sites. In most assays the beads will have two types of selective binding sites, one to bind to the analyte of interest and the other to bond to the detection area. The beads are mixed into a complex dispersion, such as blood, where they diffuse and bind with the analyte. They are then passed near a functionalized surface where they can bind selectively. This could be as simple as a glass slide with a permanent magnet on the opposite surface, so that blood cells infected with malarial parasites could be collected for optical microscopy analysis [71] or as complex as a magnetoresistive sensor coated with a thin Au coating that is functionalized to bind selectively with the beads [72, 73]. In the latter case, beads that are not covalently bound are washed off, in order to improve the signal-to-noise ratio. The magnetoresistance of the sensor changes depending on the magnetic field generated by the beads. The interaction between the magnetic NPs and the free layer of the magnetoresistive sensor is purely magnetostatic, so the beads must be very close. They must also cover a large fraction of the sensor area, and they must have high magnetic moment, so that the field from the particle penetrating into the free layer changes the sensor resistance. Other types of magnetic manipulation with NPs are moving toward in vivo application and focus on therapeutics rather than detection. There have been ex vivo studies of magnetic vascular repair, in which NPs are forced by a radial magnetic field gradient to collect on the inner surface of an artery [74, 75]. Here the magnetic NP delivers a gene for transduction into damaged endothelial cells that line the arteries. If successful this could lead to a noninvasive method to control cardiovascular disease. In magnetic drug delivery, a field gradient would be used to guide the magnetically tagged drug, thereby decreasing the dose required with systemic administration. At this time, magnetically guided targeting remains a challenge for locations deep within the body. The surface coating must prevent rapid elimination by the RES, and while monoclonal antibodies could be attached for selective binding, targeting through general circulation is very inefficient ( T and EC /e > V ), sequential tunnelling processes are blocked by the Coulomb charging effects. Nevertheless, a small current may still flow, even in the ideal case of no leakages, by considering higher order processes [176], the most commonly discussed of which is known as cotunnelling. It was predicted by Takahashi and Maekawa that in this cotunnelling regime the TMR can be substantially enhanced when the magnetisation of the central island is switched against the outer electrodes. This is because this coherent process leads to the overall resistance of the double arising from the product of the individual junction resistances, whereas it follows the usual series resistor sum

    21 Artificially Engineered Magnetic Materials

    1067

    rule in the sequential regime. It can be shown within the Jullière model that the TMR in the cotunnelling regime should be enhanced by a factor of 2/(1 − P2 ), where P is the electrode spin polarisation. These effects were explored in some detail using a conventional double junction stack with granular CoFe deposited between two outer pinned CoFe electrodes [177]. At 7 K the TMR ratio was 24 % at zero bias, roughly double the value measured beyond the gap at V = 100 mV. The effective spin polarisation P was estimated to be 32 % from the non-enhanced TMR, leading to a prediction from the 2/(1 − P2 ) expression of an enhanced TMR of 23 %, in agreement with the observed value. Another magnetic configuration is double magnetic tunnel junctions with conventional free and pinned outer electrodes, in which superparamagnetic NiFe nanoparticles form the central electrode sandwiched between alumina barriers [178]. Whilst the TMR in the CB regime is a little lower than that for a control sample lacking the NiFe islands, the cotunnelling enhancement in the TMR was clearly observed at biases small enough to lie within the Coulomb gap for the double MTJ. This shows that spin information can be propagated through the fluctuating NiFe island moments. This is reasonable since typical tunnelling times are of the order of femtoseconds, much shorter than typically superparamagnetic fluctuation lifetimes, which are ∼ 0.1-1 ns.

    Kondo effect The Kondo effect arises when the spin of a magnetic impurity interacts with the surrounding free electrons to form a singlet state. The earliest studies of the Kondo effect considered the situation of doping a conductor, such as copper, with a magnetic impurity [179]. As a result, the conductivity saturates when reducing the temperature. The maximum value of the conductivity is achieved for the so called Kondo temperature (TK ) [180]. This value directly depends on the number of defects introduced in the conductor. In single electtron devices, the a net spin in the artificial atom can interact with the electrons in the leads to form the Kondo singlet [181]. In tunnel junctions with magnetic nanoclusters in the insulating layer, the Kondo effect can produce an enhancement of the resistance or its reduction depending on the exact location of the clusters within the layer. For instance, Kondo resonances were suggested to be be the cause of a rise in the resistance at low temperature and low bias when introducing thin Cr(< 0.4 nm)/Co(< 0.6 nm) impurities in one of the electrode/barrier interfaces of a magnetic tunnel junctions [182]. On the other hand, the Kondo effect produces an increase of the conductance if the insulating tunnel barrier is doped with a magnetic material [183] or some magnetic nanoclusters are placed within this layer [184]. In the CB regime, where nanoclusters are intentionally placed within the insulator, some theoretical work [185, 186], and recent experimental results [187], show that the Kondo effect in TMR can be important even if the magnetic moments have large magnetic anisotropy [188]. A theoretical discussion of the Kondo effect in single-electron devices can be found in Ref. [189]. The crossover between Kondo TMR suppression and co-tunneling enhancement of TMR in double magnetic tunnel junctions with CoFe nanoclusters within the MgO barrier is discussed in [187]. For all CoFe thicknesses the tunnelling is dominated by sequential tunnelling at high temperature

    1068

    C. H. Marrows

    and bias voltage. However, when the temperature is reduced, some typical Kondo signatures can be observed: a zero bias anomaly (ZBA) in the conductance versus bias voltage curve; the temperature dependence of the ZBA; and suppression of the TMR at low bias below a particular temperature. From each of these experimental features, the Kondo temperature can be determined. The crossover from the two effects, Kondo and CB, was found to be correlated with the fluctuations of the magnetic moments of the nanoclusters. A related experiment using high-anisotropy CoPt nanoclusters found a gradual competition between cotunneling enhancement of the TMR and the TMR suppression due to the Kondo effect, with both effects having been found to coexist even in the same sample. It is possible to tune between these two states with temperature [190].

    Chemical potential effects Modifications of the chemical potential μ in nanodots can have a significant effect on transport through them. This is, after all, the basis of the operation of a gate electrode in a SET. The magneto-Coulomb effect, discovered by Ono et al. [191, 192] was studied by theoretically [193]. The central issue is that the application of a magnetic field H can modulate the μ in a ferromagnet through the flux density B it gives rise to. CB anisotropic magnetoresistance is another effect related to manipulation of μ in a ferromagnetic dot [194], in this case one with a large spin-orbit coupling. There, the chemical potential shifts are related to the uniaxial magnetocrystalline anisotropy, which arises from spin-orbit coupling, and the magnetisation angle plays the role of a gate voltage. Another related experiment studied transport through a single Au nanoparticle connected to Co leads [165]. The I(V ) characteristics of this device could be fitted well using the orthodox theory of CB, with the commonplace requirement of a local charge offset Q0 reflecting the local electrostatic environment of the nanoparticle. Again, this can be seen as a chemical potential shift μ = Q0 e/C, with C the capacitance of the dot and e the electronic charge. Remarkably, this charge was found to vary simply with the direction of magnetisation of the two Co electrodes in a saturating field: the moment direction is coupled through spin-orbit interactions to the chemical potential of the dot and hence charge transfer from the leads. A 90◦ rotation corresponded to a change in background charge ΔQ0 = 0.033e, giving rise to large conductance changes near the Coulomb steps in the I(V ) curve. As with the other magneto-Coulomb effects described above, this essentially allow spintronic SETs to be constructed where the transistor action is gated by a magnetic, rather than an electric, field.

    Single atom manipulation and measurements Whilst it is a commonplace that a scanning tunnelling microscope (STM) can image surfaces with atomic resolution, the ability to transfer an single atom from the tip to a surface [195] and to position individual atoms on a surface [196] mean that it is a tool that can also be used to manipulate matter with atomic resolution, albeit

    21 Artificially Engineered Magnetic Materials

    1069

    painstakingly. This means that materials can be engineered at the fundamental limit imposed by the discreteness of matter.

    Quantum corrals and wave function imaging In one of the iconic experiments of nanotechnology, Crommie et al. used these capabilities of the STM to build ‘quantum corrals’ [197]. In this first experiment, electrons in the Shockley surface state of (111) copper were confined within a circular barrier constructed from 48 iron adatoms. These formed a circular corral of radius 7.13 nm, within which scanning tunnelling spectroscopy revealed that the two-dimensional wavefunction takes the form expected for a particle in a circular box, illustrated in Fig. 11. In the original experiments the magnetic nature of the adatoms was not relevant, but the fact that a Co adatom possesses a magnetic moment was of critical importance to another classic experiment in this field: the observation of a ‘quantum mirage’ by Manoharan et al. [199]. Here the physics of quantum corrals is combined with the Kondo effect, which describes the many-body response of an electron gas to a magnetic impurity. In this experiment, a ellipse-shaped corral was formed, with a single Co atom placed at one of the two foci. The spatially localized spectroscopic response of the Kondo impurity was found at the other focus, where there was in fact no Kondo impurity. The mirage experiment achieved this by taking advantage of both the locally modified electron density in the corral and the scattering properties of a Kondo impurity. A scattering theory has been developed that is remarkably successful in reproducing every detail of the experiments including the electron standing-wave patterns, the energies and widths of corral states, and all features of the quantum mirage [200].

    Fig. 11 An STM image of a circular quantum corral constructed from 48 Fe atoms. The diameter of the corral is 143 Å. The circular pattern in the center of the corral is the density distribution due to 3 nearly degenerate quantum states of the corral. [198]

    1070

    C. H. Marrows

    Magnetism at the atomic scale The STM can also be used to investigate magnetism at the atomic scale, for instance by measuring the magnetisation curve of a single atom [201]. Doing so for individual Fe atoms on a Cu (111) surface, combined with inelastic scanning tunneling spectroscopy revealed that the typical moment of such an atom is 3.5 Bohr magnetons with a magnetic excitation lifetime of 200 fs [202]. The findings were quantitatively explained by the decay of the magnetization excitation into Stoner modes of the itinerant electron system. Longer timescales of between 50 and 250 ns were found using STM-based pump-probe methods [203]. Combining these measurement methods with the ability to engineer nanostructures specified atom-by-atom, including the exchange coupling between the moments on neighbouring atoms (by controlling the interatom distance to yield the selected value of the oscillatory indirect exchange) between them [201], has led to the ability to make all-spin logic gates from atomic chains [204], control the ground states of chains (depending on even or odd numbers of constituent atoms), engineer magnetic frustration within arrays [205], and excite coupled arrays with spin transfer torques arising from spin-polarised currents injected from a magnetic STM tip [206]. Quantum tunnelling between Néel states of antiferromagnetic chains has been observed [207]. Building a chain of Fe atoms on a superconducting Pb surface has allowed for the detection of Majorana fermion-like collective excitations in a condensed matter system [208], manifesting themselves as zero-energy states at the ends of the chain.

    Anisotropy of individual atoms The thermal stability of any magnetic moment arises from its magnetic anisotropy energy . At the atomic scale, this comes from the ligand crystal fields (Chapter ANI), the electrostatic energy arising from the shape of the orbitals of both the magnetic adatom and its neighbours in the substrate. Spin-orbit coupling locks the adatom spin moment to its orbital moment, which is subject to these ligand fields, generating the anisotropy. From this we see that the three factors needed for a strong anisotropy are strong spin-orbit coupling , a large orbital moment, and a strong ligand field. The last two are usually inversely correlated to each other, since the ligand field tends to quench the orbital moment. Single cobalt atoms deposited onto platinum (111) were found to have a giant magnetic anisotropy energy of 9 meV per atom, determined using x-ray magnetic circular dichroism [209]. Here the large unquenched orbital moment of 1.1 Bohr magnetons and strong spin-orbit coupling induced by the heavy metal substrate combine to yield this large value. The use of STM-based methods described above show that even such large anisotropy energies are not enough to thermally stabilise individual atomic moments on a metallic substrate [210, 203, 202], due to

    21 Artificially Engineered Magnetic Materials

    1071

    spin-flip scattering of electrons in the substrate. Partially isolating the the adatom electronically from the susbtrate using interlayers such as copper nitride [211] or magnesium oxide [212] was found to boost the anisotropy considerably: indeed in the latter case a record anisotropy energy of 60 meV for Co, close to the free atom value, was achieved. Tuning the strength of the interactions between adatom and substrate allows the strength of the anisotropy, in turn, to be controlled [213].

    References 1. Y. Yoshizawa, S. Oguma, K. Yamauchi, New Fe-based soft magnetic alloys composed of ultrafine grain structure. J. Appl. Phys. 64, 6044 (1988) 2. M. Manivel Raja, K. Chattopadhyay, B. Majumdar, A. Narayanasamy, Structure and soft magnetic properties of FINEMET alloys. J. Alloys Compounds 297, 199 (2000) 3. G. Herzer, Grain structure and magnetism of nanocrystalline ferromagnets. IEEE Trans. Magn. 25, 3327 (1989) 4. R. Alben, J.J. Becker, M.C. Chi, Random anisotropy in amorphous ferromagnets. Journal of Applied Physics 49, 1653 (1978) 5. W.H. Meiklejohn, C.P. Bean, New magnetic anisotropy. Phys. Rev. 105, 904 (1957) 6. J. Nogués, I.K. Schuller, Exchange bias. J. Magn. Magn. Mater. 192, 203 (1999) 7. R.L. Stamps, Mechanisms for exchange bias. J. Phys. D: Appl. Phys. 33, R247 (2000) 8. M. Kiwi, Exchange bias theory. J. Magn. Magn. Mater. 234, 584 (2001) 9. B. Dieny, V.S. Speriosu, S.S.P. Parkin, B.A. Gurney, D.R. Wilhoit, D. Mauri, Giant magnetoresistive in soft ferromagnetic multilayers. Phys. Rev. B 43, 1297 (1991) 10. P. Steadman, M. Ali, A.T. Hindmarch, C.H. Marrows, H.B. J., S. Langridge, R.M. Dalgliesh, S. Foster, Exchange bias in spin-engineered double superlattices. Phys. Rev. Lett. 89, 077201 (2002) 11. D. Mauri, H.C. Siegmann, P.S. Bagus, E. Kay, Simple model for thin ferromagnetic films exchange coupled to an antiferromagnetic substrate. J. Appl. Phys. 62, 3047 (1987) 12. A.P. Malozemoff, Random-field model of exchange anisotropy at rough ferromagneticantiferromagnetic interfaces. Phys. Rev. B 35, 3679 (1987) 13. E. Fulcomer, S.H. Charap, Thermal fluctuation aftereffect model for some systems with ferromagnetic-antiferromagnetic coupling. J. Appl. Phys. 43, 4190 (1972) 14. C. Park, K. Lee, D. Hwang, S. Lee, M. Kim, J. Rhee, Exchange coupling between antiferromagnetic NiO and ferromagnetic films. Journal of the Korean Physical Society 31, 508 (1997) 15. R.F.C. Farrow, R.F. Marks, S. Gider, A.C. Marley, S.S.P. Parkin, D. Mauri, Mnx Pt1−x : A new exchange bias material for Permalloy. J. Appl. Phys. 81, 4986 (1997) 16. R. Jungblut, R. Coehoorn, M.T. Johnson, J. aan de Stegge, A. Reinders, Orientational dependence of the exchange biasing in molecular-beam-epitaxy-grown Ni80 Fe20 /Fe50 Mn50 bilayers (invited). J. Appl. Phys. 75, 6659 (1994) 17. R. Nakatani, H. Hoshiya, K. Hoshino, Y. Sugita, Relationship between film structure and exchange coupling in Mn-Ir/Ni-Fe films. J. Magn. Magn. Mater. 173, 321 (1997) 18. M. Ali, P. Adie, C.H. Marrows, D. Greig, B.J. Hickey, R.L. Stamps, Exchange bias using a spin glass. Nat. Mater. 6, 70 (2007) 19. C. Maurain, Sur une action magnétisante de contact et son rayon d’activité. J. Phys 1, 90 (1902) 20. J.C. Bruyére, G. Clerc, O. Massenet, D. Paccard, R. Montmory, L. Néel, J. Valin, A. Yelon, Indirectly coupled films. IEEE Trans. Magn. 1, 174 (1965) 21. O. Massenet, F. Biraget, H. Juretschke, R. Montmory., A. Yelon, Orgin of coupling in multilayered films. IEEE Trans. Magn. 2, 553 (1966)

    1072

    C. H. Marrows

    22. S.S.P. Parkin, Systematic variation of the strength and oscillation period of indirect magnetic exchange coupling through the 3d, 4d, and 5d transition metals. Phys. Rev. Lett. 67, 3598 (1991) 23. P. Grünberg, R. Schreiber, Y. Pang, M.B. Brodsky, H. Sowers, Layered magnetic structures: Evidence for antiferromagnetic coupling of Fe layers across Cr interlayers. Phys. Rev. Lett. 57, 2442 (1986) 24. S.S.P. Parkin, N. More, K.P. Roche, Oscillations in exchange coupling and magnetoresistance in metallic superlattice structures: Co/Ru, Co/Cr, and Fe/Cr. Phys. Rev. Lett. 64, 2304 (1990) 25. C. Kittel, Indirect exchange interactions in metals. Solid State Phys. 22, 1 (1968) 26. R. Coehoorn, Period of oscillatory exchange interactions in Co/Cu and Fe/Cu multilayer systems. Phys. Rev. B 44, 9331 (1991) 27. Y. Yafet, Ruderman-Kittel-Kasuya-Yosida range function of a one-dimensional free-electron gas. Phys. Rev. B 36, 3948 (1987) 28. P. Bruno, C. Chappert, Ruderman-Kittel theory of oscillatory interlayer exchange coupling. Phys. Rev. B 46, 261 (1992) 29. D.M. Edwards, J. Mathon, R.B. Muniz, M.S. Phan, Oscillations of the exchange in magnetic multilayers as an analog of de Haas-van Alphen effect. Phys. Rev. Lett. 67, 493 (1991) 30. J.E. Ortega, F.J. Himpsel, Quantum well states as mediators of magnetic coupling in superlattices. Phys. Rev. Lett. 69, 844 (1992) 31. P. Bruno, Interlayer exchange coupling: a unified physical picture. J. Magn. Magn. Mater. 121, 248 (1993) 32. K.B. Hathaway, J.R. Cullen, A free-electron model for the exchange coupling of ferromagnets through paramagnetic metals. J. Magn. Magn. Mater. 104-107, 1840 (1992) 33. M. Pinarbasi, H.A.A. Santini, E.E. Marinero, P.C. Arnett, R.W. Olson, R. Hsiao, M.L. Williams, R.N. Payne, R.H.H. Wang, J.O. Moore, B.A. Gurney, T. Lin, R.E.E. Fontana, 12 Gb/in2 recording demonstration with SV read heads and conventional narrow pole-tip write heads. IEEE Trans. Magn. 35, 689 (1999) 34. J. Slaughter, R. Dave, M. Durlam, G. Kerszykowski, K. Smith, K. Nagel, B. Feil, J. Calder, M. DeHerrera, B. Garni, S. Tehrani, in Electron Devices Meeting, 2005. IEDM Technical Digest. IEEE International (2005), p. 873. doi: https://doi.org/10.1109/IEDM.2005.1609496 35. S.H. Yang, K.S. Ryu, S.S.P. Parkin, Domain-wall velocities of up to 750 ms−1 driven by exchange-coupling torque in synthetic antiferromagnets. Nature Nanotech. 10, 221 (2015) 36. L. Néel, L’anisotropie superficielle des substances ferromagnétiques. Compt. Rend. 237, 1468 (1953) 37. P.F. Carcia, A.D. Meinhaldt, A. Suna, Perpendicular magnetic anisotropy in Pd/Co thin film layered structures. Appl. Phys. Lett. 47, 178 (1985) 38. W.J.M. de Jonge, P.J.H. Bloemen, F.J.A. den Broeder, Experimental Investigations of Magnetic Anisotropy (Springer, Berlin, 1994), Ultrathin Magnetic Structures, vol. 1, chap. 2.3, p. 65 39. G.H.O. Daalderop, P.J. Kelly, F.J.A. den Broeder, Prediction and confirmation of perpendicular magnetic anisotropy in Co/Ni multilayers. Phys. Rev. Lett. 68, 682 (1992) 40. S. Ikeda, K. Miura, H. Yamamoto, K. Mizunuma, H.D. Gan, M. Endo, S. Kanai, J. Hayakawa, F.M. andH. Ohno, A perpendicular-anisotropy CoFeB-MgO magnetic tunnel junction. Nature Mater. 9, 721 (2010) 41. I. Dzyaloshinsky, A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. J. Phys. Chem. Solids 4, 241 (1958) 42. T. Moriya, New mechanism of anisotropic superexchange interaction. Phys. Rev. Lett. 4, 228 (1960) 43. T. Moriya, Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120, 91 (1960) 44. A. Fert, Magnetic and transport properties of metallic multilayers. Mater. Sci. Forum 59, 439 (1991) 45. A. Crépieux, C. Lacroix, Dzyaloshinsky-Moriya interactions induced by symmetry breaking at a surface. J. Magn. Magn. Mater. 182, 341 (1998)

    21 Artificially Engineered Magnetic Materials

    1073

    46. A. Fert, V. Cros, J. Sampaio, Skyrmions on the track. Nat. Nanotech. 8, 152 (2013) 47. M. Bode, M. Heide, K. von Bergmann, P. Ferriani, S. Heinze, G. Bihlmayer, A. Kubetzka, O. Pietzsch, S. Blügel, R. Wiesendanger, Chiral magnetic order at surfaces driven by inversion asymmetry. Nature 447, 190 (2007) 48. A. Thiaville, S. Rohart, Émilie Jué, V. Cros, A. Fert, Dynamics of Dzyaloshinskii domain walls in ultrathin magnetic films. EPL (Europhysics Letters) 100, 57002 (2012) 49. A. Kubetzka, O. Pietzsch, M. Bode, R. Wiesendanger, Spin-polarized scanning tunneling microscopy study of 360◦ walls in an external magnetic field. Phys. Rev. B 67, 020401 (2003) 50. G. Chen, T. Ma, A.T. N’Diaye, H. Kwon, C. Won, Y. Wu, A.K. Schmid, Tailoring the chirality of magnetic domain walls by interface engineering. Nat. Commun. 4, 2671 (2013) 51. M.J. Benitez, A. Hrabec, A.P. Mihai, T.A. Moore, G. Burnell, D. McGrouther, C.H. Marrows, S. McVitie, Magnetic microscopy and topological stability of homochiral Néel domain walls in a Pt/Co/AlOx trilayer. Nat. Commun. 6, 8957 (2015) 52. A. Bogdanov, A. Hubert, Thermodynamically stable magnetic vortex states in magnetic crystals. J. Magn. Magn. Mater. 138, 255 (1994) 53. R. Wiesendanger, Nanoscale magnetic skyrmions in metallic films and multilayers: a new twist for spintronics. Nat. Reviews Mater. 1, 16044 (2016) 54. S. Heinze, K. von Bergmann, M. Menzel, J. Brede, A. Kubetzka, R. Wiesendanger, G. Bihlmayer, S. Blügel, Spontaneous atomic-scale magnetic skyrmion lattice in two dimensions. Nat. Phys. 7, 713 (2011) 55. G. Chen, A. Mascaraque, A.T. N’Diaye, A.K. Schmid, Room temperature skyrmion ground state stabilized through interlayer exchange coupling. Appl. Phys. Lett. 106, 242404 (2015) 56. C. Moreau-Luchaire, C. Moutafis, N. Reyren, J. Sampaio, C.A.F. Vaz, N.V. Horne, K. Bouzehouane, K. Garcia, C. Deranlot, P. Warnicke, P. Wohlhüter, J.M. George, M. Weigand, J. Raabe, V. Cros, A. Fert, Additive interfacial chiral interaction in multilayers for stabilization of small individual skyrmions at room temperature. Nat. Nanotech. 11, 444 (2016) 57. O. Boulle, J. Vogel, H. Yang, S. Pizzini, D. de Souza Chaves, A. Locatelli, T.O. Mente¸s, A. Sala, L.D. Buda-Prejbeanu, O. Klein, M. Belmeguenai, Y. Roussigné, A. Stashkevich, S.M. Chérif, L. Aballe, M. Foerster, M. Chshiev, S. Auffret, I.M. Miron, G. Gaudin, Roomtemperature chiral magnetic skyrmions in ultrathin magnetic nanostructures. Nat. Nanotech. 11, 449 (2016) 58. W. Jiang, P. Upadhyaya, W. Zhang, G. Yu, M.B. Jungfleisch, F.Y. Fradin, J.E. Pearson, Y. Tserkovnyak, K.L. Wang, O. Heinonen, S.G.E. te Velthuis, A. Hoffmann, Blowing magnetic skyrmion bubbles. Science 349, 283 (2015) 59. S. Woo, K. Litzius, B. Krüger, M.Y. Im, L. Caretta, K. Richter, M. Mann, A. Krone, R.M. Reeve, M. Weigand, P. Agrawal, I. Lemesh, M.A. Mawass, P. Fischer, M. Kläui, G.S.D. Beach, Observation of room-temperature magnetic skyrmions and their current-driven dynamics in ultrathin metallic ferromagnets. Nat. Mater. 15, 501 (2016) 60. W. Shockley, G.L. Pearson, Modulation of conductance of thin films of semi-conductors by surface charges. Phys. Rev. 74, 232 (1948) 61. K. von Klitzing, G. Dorda, M. Pepper, New method for high-accuracy determination of the fine-structure constant based on quantized hall resistance. Phys. Rev. Lett. 45, 494 (1980) 62. D.C. Tsui, H.L. Stormer, A.C. Gossard, Two-dimensional magnetotransport in the extreme quantum limit. Phys. Rev. Lett. 48, 1559 (1982) 63. K.S. Novoselov, Z. Jiang, Y. Zhang, S.V. Morozov, H.L. Stormer, U. Zeitler, J.C. Maan, G.S. Boebinger, P. Kim, A.K. Geim, Room-temperature quantum hall effect in graphene. Science 315, 1379 (2007) 64. A. Ohtomo, H.Y. Hwang, A high-mobility electron gas at the LaAlO3 /SrTiO3 heterointerface. Nature 427, 423 (2004) 65. A.D. Caviglia, S. Gariglio, C. Cancellieri, B. Sacépé, A. Fête, N. Reyren, M. Gabay, A.F. Morpurgo, J.M. Triscone, Two-dimensional quantum oscillations of the conductance at laalo3 /srtio3 interfaces. Phys. Rev. Lett. 105, 236802 (2010)

    1074

    C. H. Marrows

    66. N. Reyren, S. Thiel, A.D. Caviglia, L.F. Kourkoutis, G. Hammerl, C. Richter, C.W. Schneider, T. Kopp, A.S. Rüetschi, D. Jaccard, M. Gabay, D.A. Muller, J.M. Triscone, J. Mannhart, Superconducting interfaces between insulating oxides. Science 317, 1196 (2007) 67. S. Thiel, G. Hammerl, A. Schmehl, C.W. Schneider, J. Mannhart, Tunable quasi-twodimensional electron gases in oxide heterostructures. Science 313, 1942 (2006) 68. A.D. Caviglia, S. Gariglio, N. Reyren, D. Jaccard, T. Schneider, M. Gabay, S. Thiel, G. Hammerl, J. Mannhart, J.M. Triscone, Electric field control of the LaAlO3 /SrTiO3 interface ground state. Nature 456, 624 (2008) 69. A. Brinkman, M. Huijben, M. van Zalk, J. Huijben, U. Zeitler, J.C. Maan, W.G. van der Wiel, G. Rijnders, D.H.A. Blank, H. Hilgenkamp, Magnetic effects at the interface between nonmagnetic oxides. Nat. Mater. 6, 493 (2007) 70. S. Seri, L. Klein, Antisymmetric magnetoresistance of the SrTiO3 /LaAlO3 interface. Phys. Rev. B 80, 180410 (2009) 71. Ariando, X. Wang, G. Baskaran, Z.Q. Liu, J. Huijben, J.B. Yi, A. Annadi, A.R. Barman, A. Rusydi, S. Dhar, Y.P. Feng, J. Ding, H. Hilgenkamp, T. Venkatesan, Electronic phase separation at the LaAlO3 /SrTiO3 interface. Nat. Commun. 2, 188 (2011) 72. J.A. Bert, B. Kalisky, C. Bell, M. Kim, Y. Hikita, H.Y. Hwang, K.A. Moler, Direct imaging of the coexistence of ferromagnetism and superconductivity at the LaAlO3 /SrTiO3 interface. Nat. Phys. 7, 767 (2011) 73. D.A. Dikin, M. Mehta, C.W. Bark, C.M. Folkman, C.B. Eom, V. Chandrasekhar, Coexistence of superconductivity and ferromagnetism in two dimensions. Phys. Rev. Lett. 107, 056802 (2011) 74. B. Kalisky, J.A. Bert, B.B. Klopfer, C. Bell, H.K. Sato, M. Hosoda, Y. Hikita, H.Y. Hwang, K.A. Moler, Critical thickness for ferromagnetism in LaAlO3 /SrTiO3 heterostructures. Nat. Commun. 3, 922 (2012) 75. J.E. Moore, The birth of topological insulators. Nature 464, 194 (2010) 76. C.L. Kane, E.J. Mele, Z2 topological order and the quantum spin Hall effect. Phys. Rev. Lett. 95, 146802 (2005) 77. L. Fu, C.L. Kane, Topological insulators with inversion symmetry. Phys. Rev. B 76, 045302 (2007) 78. D. Hsieh, D. Qian, L. Wray, Y. Xia, Y.S. Hor, R.J. Cava, M.Z. Hasan, A topological dirac insulator in a quantum spin hall phase. Nature 452, 970 (2008) 79. C.H. Li, O.M.J. van ’t Erve, J.T. Robinson, Y. Liu, L. Li, B.T. Jonker, Electrical detection of charge-current-induced spin polarization due to spin-momentum locking in Bi2 Se3 . Nature Nanotech. 9, 218 (2014) 80. Experiments provide first direct signatures of a topological insulator - a new phase of quantum matter (2009). URL http://www-ssrl.slac.stanford.edu/content/science/highlight/2009-03-30/ experiments-provide-first-direct-signatures-topological-insulator-new. SSRL Science Highlight 81. R. Yu, W. Zhang, H.J. Zhang, S.C. Zhang, X. Dai, Z. Fang, Quantized anomalous hall effect in magnetic topological insulators. Science 329, 61 (2010) 82. C.Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo, K. Li, Y. Ou, P. Wei, L.L. Wang, Z.Q. Ji, Y. Feng, S. Ji, X. Chen, J. Jia, X. Dai, Z. Fang, S.C. Zhang, K. He, Y. Wang, L. Lu, X.C. Ma, Q.K. Xue, Experimental observation of the quantum anomalous Hall effect in a magnetic topological insulator. Science 340, 167 (2013) 83. D.A. Allwood, G. Xiong, C.C. Faulkner, D. Atkinson, D. Petit, R.P. Cowburn, Magnetic domain-wall logic. Science 309, 1688 (2005) 84. S.S.P. Parkin, M. Hayashi, L. Thomas, Magnetic domain-wall racetrack memory. Science 320, 190 (2008) 85. A. Fert, L. Piraux, Magnetic nanowires. J. Magn. Magn. Mater. 200, 338 (1999) 86. E. Ferain, R. Legras, Track-etched membrane: dynamics of pore formation. Nucl. Instrum. Methods B 84, 331 (1994) 87. H. Masuda, K. Fukuda, Ordered metal nanohole arrays made by a two-step replication of honeycomb structures of anodic alumina. Science 268, 1466 (1995)

    21 Artificially Engineered Magnetic Materials

    1075

    88. K. Nielsch, F. Müller, A.P. Li, U. Gösele, Uniform nickel deposition into ordered alumina pores by pulsed electrodeposition. Adv. Mater. 12, 582 (2000) 89. A. Blondel, B. Doudin, J.P. Ansermet, Comparative study of the magnetoresistance of electrodeposited Co/Cu multilayered nanowires made by single and dual bath techniques. J. Magn. Magn. Mater. 165, 34 (1997) 90. M. Alper, K. Attenborough, R. Hart, S.J. Lane, D.S. Lashmore, C. Younes, W. Schwarzacher, Giant magnetoresistance in electrodeposited superlattices. Appl. Phys. Lett. 63, 2144 (1993) 91. D. AlMawlawi, N. Coombs, M. Moskovits, Magnetic properties of Fe deposited into anodic aluminum oxide pores as a function of particle size. J. Appl. Phys. 70, 4421 (1991) 92. J.C. Lodder, L. Cheng-Zhang, Reversal behaviour in perpendicular iron particle arrays (alumite media). IEEE Trans. Magn. 25, 4171 (2002) 93. T.M. Whitney, J.S. Jiang, P.C. Searson, C.L. Chien, Fabrication and magnetic properties of arrays of metallic nanowires. Science 261, 1316 (1993) 94. L. Piraux, S. Dubois, E. Ferain, R. Legras, K. Ounadjela, J.M. George, J.L. Maurice, A. Fert, Anisotropic transport and magnetic properties of arrays of sub-micron wires. J. Magn. Magn. Mater. 165, 352 (1997) 95. L. Piraux, J.M. George, J.F. Despres, C. Leroy, E. Ferain, R. Legras, K. Ounadjela, A. Fert, Giant magnetoresistance in magnetic multilayered nanowires. Appl. Phys. Lett. 65, 2484 (1994) 96. L. Piraux, S. Dubois, A. Fert, Perpendicular giant magnetoresistance in magnetic multilayered nanowires. J. Magn. Magn. Mater. 159, L287 (1996) 97. L. Piraux, S. Dubois, A. Fert, L. Belliard, The temperature dependence of the perpendicular giant magnetoresistance in Co/Cu multilayered nanowires. Eur. Phys. J. B 4, 413 (1998) 98. J.E. Wegrowe, D. Kelly, X. Hoffer, P. Guittienne, J.P. Ansermet, Tailoring anisotropic magnetoresistance and giant magnetoresistance hysteresis loops with spin-polarized current injection. J. Appl. Phys. 89, 7127 (2001) 99. X. Huang, L. Tan, H. Cho, B.J.H. Stadler, Magnetoresistance and spin transfer torque in electrodeposited Co/Cu multilayered nanowire arrays with small diameters. J. Appl. Phys. 105, 07D128 (2009) 100. C.L. Kane, E.J. Mele, Quantum spin Hall effect in graphene,. Phys. Rev. Lett. 95, 226801 (2005) 101. B.A. Bernevig, S.C. Zhang, Quantum spin Hall effect. Phys. Rev. Lett. 96, 106802 (2006) 102. M. König, S. Wiedmann, C. Brüne, A. Roth, H. Buhmann, L.W. Molenkamp, X.L. Qi, S.C. Zhang, Quantum spin Hall insulator state in HgTe quantum wells. Science 318, 766 (2007) 103. K. Suzuki, Y. Harada, K. Onomitsu, K. Muraki, Edge channel transport in the InAs/GaSb topological insulating phase. Phys. Rev. B 87, 235311 (2013) 104. L. Du, I. Knez, G. Sullivan, R.R. Du, Robust helical edge transport in gated InAs/GaSb bilayers. Phys. Rev. Lett. 114, 096802 (2015) 105. R.P. Cowburn, D.K. Koltsov, A.O. Adeyeye, M.E. Welland, D.M. Tricker, Single-domain circular nanomagnets. Phys. Rev. Lett. 83, 1042 (1999) 106. R.P. Cowburn, A.O. Adeyeye, M.E. Welland, Controlling magnetic ordering in coupled nanomagnet arrays. New J. Phys. 1, 16 (1999) 107. R.P. Cowburn, M.E. Welland, Room temperature magnetic quantum cellular automata. Science 287, 1466 (2000) 108. J. Shibata, Y. Otani, Magnetic vortex dynamics in a two-dimensional square lattice of ferromagnetic nanodisks. Phys. Rev. B 70, 012404 (2004) 109. C. Behncke, M. Hänze, C.F. Adolff, M. Weigand, G. Meier, Band structure engineering of two-dimensional magnonic vortex crystals. Phys. Rev. B 91, 224417 (2015) 110. S.T. Bramwell, M.J.P. Gingras, Spin ice state in frustrated magnetic pyrochlore materials. Science 294, 1495 (2001) 111. J.D. Bernal, R.H. Fowler, A theory of water and ionic solution, with particular reference to hydrogen and hydroxyl ions. J. Chem. Phys. 1, 515 (1933) 112. L. Pauling, The structure and entropy of ice and of other crystals with some randomness of atomic arrangement. J. Am. Chem. Soc. 57, 2680 (1935)

    1076

    C. H. Marrows

    113. C. Castelnovo, R. Moessner, S.L. Sondhi, Magnetic monopoles in spin ice. Nature 451, 42 (2008) 114. M. Tanaka, E. Saitoh, H. Miyajima, T. Yamaoka, Y. Iye, Domain structure and magnetic iceorder in NiFe nano-network with honeycomb structure. J. Appl. Phys. 97, 10J710 (2005) 115. R.F. Wang, C. Nisoli, R.S. Freitas, J. Li, W. McConville, B.J. Cooley, M.S. Lund, N. Samarth, C. Leighton, V.H. Crespi, P. Schiffer, Artificial ‘spin ice’ in a geometrically frustrated lattice of nanoscale ferromagnetic islands. Nature 439, 303 (2006) 116. Y. Han, Y. Shokef, A.M. Alsayed, P. Yunker, T.C. Lubensky, A.G. Yodh, Geometric frustration in buckled colloidal monolayers. Nature 456, 898 (2008) 117. M.L. Latimer, G.R. Berdiyorov, Z.L. Xiao, F.M. Peeters, W.K. Kwok, Realization of artificial ice systems for magnetic vortices in a superconducting MoGe thin film with patterned nanostructures. Phys. Rev. Lett. 111, 067001 (2013) 118. A. Ortiz-Ambriz, P. Tierno, Engineering of frustration in colloidal artificial ices realized on microfeatured grooved lattices. Nature Commun. 7, 10575 (2016) 119. C. Nisoli, R. Moessner, P. Schiffer, Colloquium: Artificial spin ice: Designing and imaging magnetic frustration. Rev. Mod. Phys. 85, 1473 (2013) 120. C. Nisoli, R. Wang, J. Li, W.F. McConville, P.E. Lammert, P. Schiffer, V.H. Crespi, Ground state lost but degeneracy found: the effective thermodynamics of ‘artificial spin ice’. Phys. Rev. Lett. 98, 217103 (2007) 121. X. Ke, J. Li, C. Nisoli, P.E. Lammert, W. McConville, R.F. Wang, V.H. Crespi, P. Schiffer, Energy minimization and ac demagnetization in a nanomagnet array. Phys. Rev. Lett. 101, 037205 (2008) 122. C. Nisoli, J. Li, X. Ke, D. Garand, P. Schiffer, V.H. Crespi, Effective temperature in an interacting vertex system: Theory and experiment on artificial spin ice. Phys. Rev. Lett. 105, 047205 (2010) 123. Y. Qi, T. Brintlinger, J. Cumings, Direct observation of the ice rule in an artificial kagome spin ice. Phys. Rev. B 77, 094418 (2008) 124. S. Ladak, D.E. Read, G.K. Perkins, L.F. Cohen, W.R. Branford, Direct observation of magnetic monopole defects in an artificial spin-ice system. Nature Phys. 6, 359 (2010) 125. E. Mengotti, L.J. Heyderman, A. Fraile Rodríguez, F. Nolting, R.V. Hügli, H.B. Braun, Realspace observation of emergent magnetic monopoles and associated Dirac strings in artificial kagome spin ice. Nature Phys. 7, 68 (2011) 126. J.P. Morgan, A. Stein, S. Langridge, C.H. Marrows, Thermal ground-state ordering and elementary excitations in artificial magnetic square ice. Nature Phys. 7, 75 (2011) 127. J.M. Porro, A.Bedoya-Pinto, A. Berger, P. Vavassori, Exploring thermally induced states in square artificial spin-ice arrays. New J. Phys. 15, 055012 (2013) 128. S. Zhang, I. Gilbert, C. Nisoli, G.W. Chern, M.J. Erickson, L. O’Brien, C. Leighton, P.E. Lammert, V.H. Crespi, P. Schiffer, Crystallites of magnetic charges in artificial spin ice. Nature 500, 553 (2013) 129. Y. Perrin, B. Canals, N. Rougemaille, Extensive degeneracy, Coulomb phase and magnetic monopoles in artificial square ice. Nature 540, 410 (2016) 130. G. Möller, R. Moessner, Magnetic multipole analysis of kagome and artificial ice dipolar arrays. Phys. Rev. B 80, 140409 (2009) 131. G.W. Chern, P. Mellado, O. Tchernyshyov, Two-stage ordering of spins in dipolar spin ice on the kagome lattice. Phys. Rev. Lett. 106, 207202 (2011) 132. L. Anghinolfi, H. Luetkens, J. Perron, M.G. Flokstra, O. Sendetskyi, A. Suter, T. Prokscha, P.M. Derlet, S.L. Lee, L.J. Heyderman, Thermodynamic phase transitions in a frustrated magnetic metamaterial. Nature Commun. 6, 8278 (2015) 133. I. Gilbert, C. Nisoli, G.W. Chern, S. Zhang, L. O’Brien, B. Fore, P. Schiffer, Emergent ice rule and magnetic charge screening from vertex frustration in artificial spin ice. Nature Phys. 10, 670 (2014) 134. R.F. Wang, C. Nisoli, R.S. Freitas, J. Li, W. McConville, B.J. Cooley, M.S. Lund, N. Samarth, C. Leighton, V.H. Crespi, P. Schiffer, Demagnetization protocols for frustrated interacting nanomagnet arrays. J. Appl. Phys. 101, 09J104 (2007)

    21 Artificially Engineered Magnetic Materials

    1077

    135. J.P. Morgan, A. Bellew, A. Stein, S. Langridge, C.H. Marrows, Linear field demagnetization of artificial magnetic square ice. Frontiers in Physics 1, 28 (2013) 136. C. Nisoli, J. Li, X. Ke, D. Garand, P. Schiffer, V.H. Crespi, Effective temperature in an interacting vertex system: Theory and experiment on artificial spin ice. Phys. Rev. Lett. 105, 047205 (2010) 137. J. Morgan, J. Akerman, A. Stein, C. Phatak, R.M.L. Evans, S. Langridge, C. Marrows, Real and effective thermal equilibrium in artificial square spin ices. Phys. Rev. B 87, 024405 (2013) 138. P.E. Lammert, X. Ke, J. Li, C.Nisoli, D.M. Garand, V.H. Crespi, P. Schiffer, Direct entropy determination and application to artificial spin ice. Nature Phys. 6, 786 (2010) 139. V. Kapaklis, U.B. Arnalds, A. Harman-Clarke, E.T. Papaioannou, M. Karimipour, P. Korelis, A. Taroni, P.C.W. Holdsworth, S.T. Bramwell, B. Hjörvarsson, Melting artificial spin ice. New J. Phys. 14 (2012) 140. I.A. Chioar, B. Canals, D. Lacour, M. Hehn, B. Santos Burgos, T.O. Mente¸s, A. Locatelli, F. Montaigne, N. Rougemaille, Kinetic pathways to the magnetic charge crystal in artificial dipolar spin ice. Phys. Rev. B 90, 220407 (2014) 141. J. Drisko, S. Daunheimer, J. Cumings, FePd3 as a material for studying thermally active artificial spin ice systems. Phys. Rev. B 91, 224406 (2015) 142. A. Farhan, P.M. Derlet, A. Kleibert, A. Balan, R.V. Chopdekar, M. Wyss, L. Anghinolfi, F. Nolting, L.J. Heyderman, Exploring hyper-cubic energy landscapes in thermally active finite spin-ice systems. Nature Phys. 9, 375 (2013) 143. A. Farhan, P.M. Derlet, A. Kleibert, A. Balan, R.V. Chopdekar, M. Wyss, J. Perron, A. Scholl, F. Nolting, L.J. Heyderman, Direct observation of thermal relaxation in artificial spin ice. Phys. Rev. Lett. 111, 057204 (2013) 144. V. Kapaklis, U.B. Arnalds, A. Farhan, R.V.C.A. Balan, A. Scholl, L.J. Heyderman, B. Hjörvarsson, Thermal fluctuations in artificial spin ice. Nature Nano. 9, 514 (2014) 145. S.A. Morley, A. Stein, M.C. Rosamond, D. Alba Venero, A. Hrabec, P.M. Shepley, M.Y. Im, P. Fischer, M.T. Bryan, D.A. Allwood, P. Steadman, S. Langridge, C.H. Marrows, Temperature and magnetic-field driven dynamics in artificial magnetic square ice. Proc. SPIE 9551, 95511Q (2015) 146. M.J. Morrison, T.R. Nelson, C. Nisoli, Unhappy vertices in artificial spin ice: new degeneracies from vertex frustration. New J. Phys. 15, 045009 (2013) 147. I. Gilbert, Y. Lao, I. Carrasquillo, L. O’Brien, J.D. Watts, M. Manno, C. Leighton, A. Scholl, C. Nisoli, P. Schiffer, Emergent reduced dimensionality by vertex frustration in artificial spin ice. Nature Phys. 12, 162 (2016) 148. B. Canals, I.A. Chioar, V.D. Nguyen, M. Hehn, D. Lacour, F. Montaigne, A. Locatelli, T.O. Mente¸s, B.S. Burgos, N. Rougemaille, Fragmentation of magnetism in artificial kagome dipolar spin ice. Nature Commun. 7, 11446 (2016) 149. B. Farmer, V.S. Bhat, A. Balk, E. Teipel, N. Smith, J. Unguris, D.J. Keavney, J.T. Hastings, L.E.D. Long, Direct imaging of coexisting ordered and frustrated sublattices in artificial ferromagnetic quasicrystals. Phys. Rev. B 93, 134428 (2016) 150. Y.L. Wang, Z.L. Xiao, A. Snezhko, J. Xu, L.E. Ocola, R. Divan, J.E. Pearson, G.W. Crabtree, W.K. Kwok, Rewritable artificial magnetic charge ice. Science 352, 962 (2016) 151. M.A. Kastner, Artificial atoms. Physics Today 46, 24 (1993) 152. R.C. Ashoori, Electrons in artificial atoms. Nature 379, 413 (1996) 153. K.K. Likharev, Single-electron devices and their applications. Proc. IEEE 87, 606 (1999) 154. P. Seneor, A. Bernand-Mantel, F. Petroff, Nanospintronics: when spintronics meets single electron physics. J. Phys.: Cond. Matt. 19, 165222 (2007) 155. K.J. Dempsey, D. Ciudad, C.H. Marrows, Single electron spintronics. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 369, 3150 (2011) 156. D.V. Averin, Y.V. Nazarov, in Single Charge Tunneling: Coulomb Blockade Phenomena in Nanostructures, NATO ASI Series B: Physics, vol. 294, ed. by H. Grabert, M.H. Devoret (Plenum Press, New York, 1992), chap. 6, pp. 217–247

    1078

    C. H. Marrows

    157. K.K. Likharev, Correlated discrete charge transfer of single electrons in ultrasmall tunnel junctions. IBM. J. Res. Dev. 32, 144 (1988) 158. D.V. Averin, Y.V. Nazarov, Virtual electron diffusion during quantum tunneling of the electric charge. Phys. Rev. Lett. 65, 2446 (1990) 159. J. Barna´s, I. Weymann, Spin effects in single-electron tunnelling. J. Phys.: Cond. Matt. 20, 423202 (2008) 160. J. Barna´s, A. Fert, Magnetoresistance oscillations due to charging effects in double ferromagnetic tunnel junctions. Phys. Rev. Lett. 80, 1058 (1998) 161. J. Barna´s, A. Fert, Effects of spin accumulation on single-electron tunneling in a double ferromagnetic microjunction. Europhys. Lett. 44, 85 (1998) 162. M. Jullière, Tunneling between ferromagnetic films. Phys. Lett. A 54, 225 (1975) 163. K. Yakushiji, F. Ernult, H. Imamura, K. Yamane, S. Mitani, K. Takanashi, S. Takahashi, S. Maekawa, H. Fujimori, Enhanced spin accumulation and novel magnetotransport in nanoparticles. Nature Materials 4, pp57 (2005) 164. O.D. Miller, B. Muralidharan, N. Kapur, A.W. Ghosh, Rectification by charging: Contactinduced current asymmetry in molecular conductors. Phys. Rev. B 77, 125427 (2008) 165. A. Bernand-Mantel, P. Seneor, K. Bouzehouane, S. Fusil, C. Deranlot, F. Petroff, A. Fert, Anisotropic magneto-Coulomb effects and magnetic single-electron-transistor action in a single nanoparticle. Nature Physics 5, 920 (2009) 166. A. Brataas, Y.V. Nazarov, J. Inoue, G.E.W. Bauer, Spin accumulation in small ferromagnetic double-barrier junctions. Phys. Rev. B 59, 93 (1999) 167. J. Barna´s, A. Fert, Interplay of spin accumulation and Coulomb blockade in double ferromagnetic junctions. J. Magn. Magn. Mater. 192, 391 (1999) 168. H. Imamura, S. Takahashi, S. Maekawa, Spin-dependent Coulomb blockade in ferromagnet/normal-metal/ferromagnet double tunnel junctions. Phys. Rev. B 59, 6017 (1999) 169. R.C. Temple, C.H. Marrows, Single-electron spin interplay for characterization of magnetic double tunnel junctions. Physical Review B 88 (2013) 170. R.C. Temple, M. McLaren, R.M.D. Brydson, B.J. Hickey, C.H. Marrows, Long spin lifetime and large barrier polarisation in single electron transport through a CoFe nanoparticle. Scientific Reports 6, 28296 (2016) 171. F. Ernult, K. Yamane, S. Mitani, K. Yakushiji, K. Takanashi, Y.K. Takahashi, K. Hono, Spindependent single-electron-tunneling effects in epitaxial Fe nanoparticles. Appl. Phys. Lett. 84, 3106 (2004) 172. A. Bernand-Mantel, P. Seneor, N. Lidgi, M.M. noz, V. Cros, S. Fusil, K. Bouzehouane, C. Deranlot, A. Vaures, F. Petroff, A. Fert, Evidence for spin injection in a single metallic nanoparticle: A step towards nanospintronics. Appl. Phys. Lett. 89, 062502 (2006) 173. Y.G. Wei, C.E. Malec, D. Davidovi´c, Saturation of spin-polarized current in nanometer scale aluminum grains. Phys. Rev. B 76, 195327 (2007) 174. Y. Yafet, in Solid State Physics: Advances in Research and Applications, vol. 14, ed. by F. Seitz, D. Turnbull (Academic Press, New York, 1963), chap. 1, pp. 1–98 175. S. Mitani, Y. Nogi, H. Wang, K. Yakushiji, F. Ernult, K. Takanashi, Current-induced tunnel magnetoresistance due to spin accumulation in Au nanoparticles. Appl. Phys. Lett. 92, 152509 (2008) 176. L.J. Geerligs, D.V. Averin, J.E. Mooij, Observation of macroscopic quantum tunneling through the Coulomb energy barrier. Phys. Rev. Lett. 65, 3037 (1990) 177. H. Sukegawa, S. Nakamura, A. Hirohata, N. Tezuka, K. Inomata, Significant magnetoresistance enhancement due to a cotunneling process in double tunnel junction with single discontinuous ferromagnetic layer insertion. Phys. Rev. Lett. 94, 068304 (2005) 178. K.J. Dempsey, A.T. Hindmarch, H.X. Wei, Q.H. Qin, Z.C. Wen, W.X. Wang, G. VallejoFernandez, D.A. Arena, X.F. Han, C.H. Marrows, Cotunneling enhancement of magnetoresistance in double magnetic tunnel junctions with embedded superparamagnetic nife nanoparticles. Phys. Rev. B 82, 214415 (2010) 179. C.M. Hurd, Eletrons in Metals (Wiley, New York, 1975) 180. J. Kondo, Resistance minimum in dilute magnetic alloys. Prog. Theor. Phys. 32, 37 (1964)

    21 Artificially Engineered Magnetic Materials

    1079

    181. D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U. Meirav, M.A. Kastner, Kondo effect in a single-electron transistor. Nature 391, 156 (1998) 182. J. Kohlhepp, P. LeClair, H. Swagten, W. de Jonge, Interfacial sensitivity and zero-bias anomalies in magnetic tunnel junctions. Phys. Stat. Sol. (a) 189, 585 (2002) 183. S. Bermon, C.K. So, Conductance peaks produced by Kondo scattering from O2 in M-I-M tunnel junctions. Solid State Comms. 27, 723 (1978) 184. S.Y. Bae, S.X. Wang, Transport in magnetically doped magnetic tunnel juctions. IEEE Trans. Magn. 38, 2721 (2002) 185. C. Romeike, M. Wegewijs, W. Hofstetter, H. Schoeller, Kondo-transport spectroscopy of single molecule magnets. Phys. Rev. Lett. 97, 206601 (2006) 186. C. Romeike, M. Wegewijs, W. Hofstetter, H. Schoeller, Quantum-tunneling-induced Kondo effect in single molecular magnets. Phys. Rev. Lett. 96, 196601 (2006) 187. H. Yang, S.H. Yang, S.S.P. Parkin, Crossover from Kondo-assisted suppression to cotunneling enhancement of tunneling magnetoresistance via ferromagnetic nanodots in MgO tunnel barrier. Nano Lett. 8, 340 (2008) 188. M. Reyes Calvo, J. Fernández-Rossier, J.J. Palacios, D. Jacob, D. Natelson, C. Untiedt, The Kondo effect in ferromagnetic atomic contacts. Nature 458, 1150 (2009) 189. J. Weis, in CFN Lectures on Functional Nanostructures vol. 1, Lecture Notes in Physics, vol. 658, ed. by K. Busch, A.K. Powell, C. Röthig, G. Schön, J. Weissmüller (Springer, Berlin, 2005), chap. 5, pp. 87–122 190. D. Ciudad, Z.C. Wen, A.T. Hindmarch, E. Negusse, D.A. Arena, X.F. Han, C.H. Marrows, Competition between cotunneling, kondo effect, and direct tunneling in discontinuous highanisotropy magnetic tunnel junctions. Phys. Rev. B 85 (2012) 191. K. Ono, H. Shimada, Y. Ootuka, Enhanced magnetic valve effect and magneto-Coulomb oscillations in ferromagnetic single electron transistor. J. Phys. Soc. Japan 66, 1261 (1997) 192. K. Ono, H. Shimada, Y. Ootuka, Spin polarization and magneto-Coulomb oscillations in ferromagnetic single electron devices. J. Phys. Soc. Japan 67, 2852 (1998) 193. S.J. van der Molen, N. Tombros, B.J. van Wees, Magneto-Coulomb effect in spin-valve devices. Phys. Rev. B 73, 220406 (2006) 194. J. Wunderlich, T. Jungwirth, B. Kaestner, A.C. Irvine, A.B. Shick, N. Stone, K.Y. Wang, U. Rana, A.D. Giddings, C.T. Foxon, R.P. Campion, D.A. Williams, B.L. Gallagher, Coulomb blockade anisotropic magnetoresistance effect in a (Ga, Mn) As single-electron transistor. Phys. Rev. Lett. 97, 077201 (2006) 195. R.S. Becker, J.A. Golovchenko, B.S. Swartzentruber, Atomic-scale surface modifications using a tunnelling microscope. Nature 325, 419 (1987) 196. D.M. Eigler, E.K. Schweizer, Positioning single atoms with a scanning tunnelling microscope. Nature 344, 524 (1990) 197. M.F. Crommie, C.P. Lutz, D.M. Eigler, Confinement of electrons to quantum corrals on a metal surface. Science 262, 218 (1993) 198. D. Eigler, in Nanostructures and Quantum Effects, Springer Series in Materials Science, vol. 31 (Springer, Berlin Heidelberg, 1994), pp. 311–314 199. H.C. Manoharan, C.P. Lutz, D.M. Eigler, Quantum mirages formed by coherent projection of electronic structure. Nature 403, 512 (2000) 200. G.A. Fiete, E.J. Heller, Colloquium: Theory of quantum corrals and quantum mirages. Rev. Mod. Phys. 75, 933 (2003) 201. F. Meier, L. Zhou, J. Wiebe, R. Wiesendanger, Revealing magnetic interactions from singleatom magnetization curves. Science 320, 82 (2008) 202. A.A. Khajetoorians, S. Lounis, B. Chilian, A.T. Costa, L. Zhou, D.L. Mills, J. Wiebe, R. Wiesendanger, Itinerant nature of atom-magnetization excitation by tunneling electrons. Physical Review Letters 106 (2011) 203. S. Loth, M. Etzkorn, C.P. Lutz, D.M. Eigler, A.J. Heinrich, Measurement of fast electron spin relaxation times with atomic resolution. Science 329, 1628 (2010) 204. A.A. Khajetoorians, J. Wiebe, B. Chilian, R. Wiesendanger, Realizing all-spin-based logic operations atom by atom. Science 332, 1062 (2011)

    1080

    C. H. Marrows

    205. A.A. Khajetoorians, J. Wiebe, B. Chilian, S. Lounis, S. Blügel, R. Wiesendanger, Atom-byatom engineering and magnetometry of tailored nanomagnets. Nature Physics 8, 497 (2012) 206. A.A. Khajetoorians, B. Baxevanis, C. Hubner, T. Schlenk, S. Krause, T.O. Wehling, S. Lounis, A. Lichtenstein, D. Pfannkuche, J. Wiebe, R. Wiesendanger, Current-driven spin dynamics of artificially constructed quantum magnets. Science 339, 55 (2013) 207. S. Loth, S. Baumann, C.P. Lutz, D.M. Eigler, A.J. Heinrich, Bistability in atomic-scale antiferromagnets. Science 335, 196 (2012) 208. S. Nadj-Perge, I.K. Drozdov, J. Li, H. Chen, S. Jeon, J. Seo, A.H. MacDonald, B.A. Bernevig, A. Yazdani, Observation of Majorana fermions in ferromagnetic atomic chains on a superconductor. Science 346, 602 (2014) 209. P. Gambardella, S. Rusponi, M. Veronese, S.S. Dhesi, C. Grazioli, A. Dallmeyer, I. Cabria, R. Zeller, P.H. Dederichs, K. Kern, C. Carbone, H. Brune, Giant magnetic anisotropy of single cobalt atoms and nanoparticles. Science 300, 1130 (2003) 210. A.J. Heinrich, Single-atom spin-flip spectroscopy. Science 306, 466 (2004) 211. C.F. Hirjibehedin, C.Y. Lin, A.F. Otte, M. Ternes, C.P. Lutz, B.A. Jones, A.J. Heinrich, Large magnetic anisotropy of a single atomic spin embedded in a surface molecular network. Science 317, 1199 (2007) 212. I.G. Rau, S. Baumann, S. Rusponi, F. Donati, S. Stepanow, L. Gragnaniello, J. Dreiser, C. Piamonteze, F. Nolting, S. Gangopadhyay, O.R. Albertini, R.M. Macfarlane, C.P. Lutz, B.A. Jones, P. Gambardella, A.J. Heinrich, H. Brune, Reaching the magnetic anisotropy limit of a 3d metal atom. Science 344, 988 (2014) 213. J.C. Oberg, M.R. Calvo, F. Delgado, M. Moro-Lagares, D. Serrate, D. Jacob, J. FernándezRossier, C.F. Hirjibehedin, Control of single-spin magnetic anisotropy by exchange coupling. Nature Nanotechnology 9, 64 (2013)

    Christopher H. Marrows is Professor of Condensed Matter Physics at the University of Leeds. He received his PhD there in 1997 and was a Research Fellow of the Royal Commission for the Exhibition of 1851 from 1998 to 2000. He was the 2011Wohlfarth Lecturer. His research interests span the fields of nanomagnetism and spintronics.

    Part III Methods

    Magnetic Fields and Measurements

    22

    Oliver Portugall, Steffen Krämer, and Yurii Skourski

    Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Field Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Permanent Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electromagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . High-Field Magnet Facilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Field Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . General Technical Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Field Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bulk Magnetic Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Stray Field Mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Calibration and Metrology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    1084 1084 1085 1087 1091 1104 1105 1107 1109 1130 1139 1141 1143

    O. Portugall () LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA Toulouse, Univ. Toulouse 3, Toulouse, France e-mail: [email protected] S. Krämer LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA Toulouse, Univ. Toulouse 3, Grenoble, France e-mail: [email protected] Y. Skourski Hochfeld-Magnetlabor Dresden (EMFL-HLD), Helmholtz-Zentrum Dresden-Rossendorf, 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_24

    1083

    1084

    O. Portugall et al.

    Abstract

    Magnetic fields play a crucial role in manipulating and characterizing the electronic and magnetic properties of matter. In the present chapter, we discuss the generation of magnetic fields in a laboratory environment, their measurement, and the measurement of magnetic properties. As part of the magnet section, we explain fundamental limitations for generating high magnetic fields, the principal technical strategies to cope with them, and the implementation of different concepts in state-of-the-art high-field facilities. The measurement section starts with a brief review of physical phenomena that can be used to measure a magnetic field, followed by general technical considerations regarding noise, signal treatment, and basic sensor requirements. A detailed presentation of individual sensor types, their operating principle, performance, and application is given afterward. The subsequent discussion of techniques to measure magnetic properties essentially refers to the same sensors integrated in dedicated setups. Stray field methods and questions concerning metrology and calibration are briefly mentioned at the end of the chapter.

    Introduction Magnetic fields and, more generally, magnetism are associated with a vast number of physical phenomena as well as a number of misbeliefs and superstitions that continue to defy scientific evidence right into the twenty-first century. It is therefore appropriate to recall that magnetism basically just describes interactions involving two specific properties of matter, charge and spin. Moving charges and spins are both at the origin of, and affected by, magnetic fields. Their physics thus determines the construction of magnets as well as magnetic measurements that are the principal subjects of the present chapter.

    Magnetic Field Generation The co-existence of two distinct properties of matter associated with magnetic fields, charge and spin, finds its counterpart in the existence of two distinct classes of magnets, electric and permanent. Each class consists of numerous types and has its specific practical strengths and weaknesses as far as energy consumption, field strength, and the modulation of the magnetic field in time and space are concerned. The specific restrictions of different magnets nevertheless have a common origin, namely, the limited capability to procure, sustain, and contain the magnetic energy stored in the field. In this respect, the foremost difference between permanent and electromagnets is that the first provide the magnetic energy by themselves, whereas the second depend on an external source.

    22 Magnetic Fields and Measurements

    1085

    Permanent Magnets Permanent magnets make use of hard magnetic materials that generate a stable magnetic flux distribution and, consequently, magnetic energy in the surrounding free space without the need of an external energy source. In terms of applications, they are widely used as sources of uniform and non-uniform magnetic fields that can be static or time dependent. With the currently available materials, permanent magnets are fully competitive with electromagnets for magnetic fields up to 2 T with record values near 5 T [27]. They are most commonly used as part of a magnetic circuit composed of permanent magnets with optional soft ferromagnetic materials that channel the flux to a region of free space where the usable magnetic flux is generated. The latter is often called the air gap. Therefore, the natural description of permanent magnets uses the magnetic field strength or flux density B, since flux is conserved and interactions of electric charges and magnetic moments all depend on B [27, 39].

    Physical and Material Properties of Permanent Magnets When a permanent magnet is operated in a magnetic circuit in a quasi-equilibrium state at a given temperature and in the presence of an optional additional magnetic field, its performance is limited by the fact that the sum of all relevant solid state interactions plus the energy stored in the magnetic field must represent an energetically favorable state. Important key parameters of permanent magnets are their spontaneous polarization Js , their coercive field Hc , and their maximum energy product (BH )max over the volume of the magnet, which is proportional to the energy of the magnetic field generated by the magnet in the surrounding free space. This can be derived from basic magnetostatic relations starting from a situation where the magnet is part of a magnetic circuit with a magnetic flux density B and a demagnetizing field H . It can be shown that the energy product never exceeds μ0 Ms2 /4, where Ms is the saturation magnetization of the material used for the permanent magnet [39]. Permanent magnets are always based on anisotropic ferromagnetic and ferrimagnetic materials, since anisotropy is essential to stabilize the magnetization direction. The basic characterization of magnets involves recording of the M(H ) and M(B) hysteresis loops that provide various macroscopic indicators such as remanent magnetization, remanent induction, coercive fields for magnetization and induction, squareness coefficient of the loop, and the maximum magnetic energy product [39]. These properties originate from the geometry of the magnet (defining the demagnetization field), the processing (defining the microstructure), as well as the fundamental microscopic properties of the material like their band structures, the exchange interactions, and the spin-orbit coupling (the most important contribution to anisotropy) [156]. Other important parameters are the Curie temperature of the material and the temperature dependence of the hysteresis loops. Depending on the type of application, the magnet geometry and the working point (for static applications) or working range (for dynamic applications) of the magnet in

    1086

    O. Portugall et al.

    the B-H hysteresis loop are chosen. Permanent magnets are either available as oriented sintered magnets, which show the best performance but are expensive and complicated to manufacture, or bonded magnets that involve the usage of a nonmagnetic matrix. The latter are cheaper to produce, but their performance is limited. Materials for permanent magnets can be divided into four groups according to their maximum energy product and their cost [39]: Ferrites: Discovered in the 1950s, ferrites are used for traditional and mass market applications (90% in weight), since they are easily available at low cost (about 1 US$ per Joule of magnetic energy). They provide good longterm stability and corrosion resistance. Their disadvantage is their low magnetic energy product due to their low remanent magnetization. Neodymium-iron-boron (NdFeB): Discovered in the 1980s, this family is used for high-performance and miniaturized applications, as they provide the highest magnetic energy product (above 400 kJ/m3 ; for comparison, lodestone, or natural magnetite, has a magnetic energy product of 0.7 kJ/m3 ) at still reasonable price (less than 10 US$ per Joule). Limitations are low Curie temperature and easy oxidation in air that cause grains to become noncoercive. As these magnets have the greatest potential for wider applications, intense research is conducted to improve their properties [156]. Aluminum-nickel-cobalt alloys (AlNiCo): This type of magnets became available in the late 1930s. Their production cost is almost ten times higher in comparison to ferrites, yet they provide excellent thermal stability and high Curie temperatures making them ideal for high-temperature and high-precision applications. Samarium-cobalt magnets (SmCo): SmCo-based magnets were discovered in the 1960s and are high-performance magnets due to their high-temperature behavior and their compatibility with corrosive environments. SmCo magnets are widely used in spatial and military applications. Their disadvantage is their high price.

    Applications of Permanent Magnets As Flux Sources Permanent magnets nowadays serve a wide range of purposes including static field sources for magnetic sensors and measurement techniques requiring a magnetic field (magnetization, Hall effect, magnetic resonance techniques); traditional industrial products such as motors, generators, microphones, and loudspeakers; and miniaturized devices in hard disks and sensors for automotive applications. Depending on their usage and technical principle, three types of permanent magnet applications can be distinguished: Static magnetic field sources: Homogeneous and inhomogeneous field profiles up to 2 T can be obtained by suitable spatial arrangements of permanent magnets. The magnetic field can be calculated by superposing the contributions of individual magnets whose mutual perturbation is small due to the rigidity of their magnetization. In comparison to solenoids, permanent magnets are better adapted for small systems (mm sized), since the generation of equivalent magnetic inductions would require impossible electric current strengths or number of windings [27]. Important applications include Halbach magnet arrays

    22 Magnetic Fields and Measurements

    1087

    that provide highly homogeneous magnetic fields for nuclear magnetic resonance (NMR) spectroscopy and imaging [133, 31]; single-sided magnets for ex situ high-resolution NMR spectroscopy [120] and other sensor applications; and motion and position sensing using magnetic induction or Hall probes [39]. As permanent magnets generate strong magnetic field gradients, they can also be used for stray field magnetic resonance imaging in solids (STRAFI). Here, the magnetic field gradient used for spatial encoding is also generated by the main magnet, in contrast to conventional magnetic resonance imaging techniques. In complement to STRAFI methods in solenoids [146,82], their use with permanent magnets has recently opened the possibility of compact and mobile NMR [31], as permanent magnets require no external energy source and provide strong gradients (up to 100 T/m) [27]. Magneto-mechanical applications: Magneto-mechanical assemblies are characterized by forces and torques produced by direct interaction between different magnets or between magnets on the one hand and magnetic materials on the other hand. As the magnetic force is given by the spatial gradient of the magnetic energy, its value depends on the square of the spontaneous polarization Js . Important magneto-mechanical applications are magnetic gears, bearings, shock absorbers, magnetic levitation, as well as magnetic separation systems for paramagnetic and ferromagnetic particles [27]. Electro-mechanical applications: Electro-mechanical applications are characterized by forces and torques produced by interaction between magnets and devices carrying external or induced electrical currents. The interaction strength of the forces and torques is proportional to I J , where I is the current and J is the polarization of the magnet. Electro-mechanical applications represent the biggest application area, since they comprise motors, loudspeakers, and actuators involving external currents as well as devices involving eddy currents as separators for non-magnetic metallic particles [27] and counters for speed or current consumption [39]. Another important electro-mechanical application is the use of permanent magnets in beamlines and accelerators in order to guide the charged particles or to generate curved paths in wigglers and undulators in order to generate narrowband intense synchrotron radiation [20]. For further discussions of permanent magnets and their applications, see also  Chap. 28, “Permanent Magnet Materials and Applications,” of this book.

    Electromagnets Electromagnets are current driven and thus governed by the Biot-Savart law relating the current density j (r) and the magnetic field strength B(r). B(r) =

    μ0 4π

     V

    j (r  ) ×

    r − r 3  d r. |r − r  |3

    (1)

    1088

    O. Portugall et al.

    As the field varies linearly with the current density, electromagnets are particularly useful when a magnetic field has to be switched, swept, or tuned. They can also generate higher fields than permanent magnets and tend to be more practical to cover larger volumes. The performance of electromagnets is fundamentally limited by dissipation and magnetic pressure. These problems as well as advanced technical strategies to overcome them will be discussed in the next section on high-field magnet facilities. Here, we limit our attention to common laboratory magnets whose operating regime gives rise to thermal and mechanical constraints that can be controlled with relatively basic means. This is the case for resistive magnets providing fields up to about 5 T and superconducting magnets reaching 23–25 T. Laboratory magnets are used in physics, chemistry, biology, and material science. They permit measurements of field-dependent physical properties of matter like magnetization, Hall effect, magnetoresistance, optical properties, specific heat, thermal transport, and magnetic resonance. Other possible uses include the alignment or separation of magnetic objects or domains and the manipulation of chemical reactions and biological processes. Typical industrial applications comprise motors, loudspeakers, generators, and actuators. The design of laboratory magnets is subject to application-related requirements such as the desired maximum field value, homogeneity, a particular spatial field profile, or geometric constraints like the need of openings for optical access. The basic approach is to start from the Biot-Savart law (Eq. 1) and to place current sections such that the desired field profile is obtained. The usage of different circuits operating at different current levels, of ferromagnetic materials, and of permanent magnets provides additional degrees of freedom for field shaping. Whereas simple magnets or magnet assemblies can be calculated by analytical methods [163, 56], numerical approaches including finite-element techniques have to be used for more complex arrangements. In addition, constraints given by the magnetic pressure and, in the case of resistive electromagnets, the energy dissipation may be modelled numerically. The most commonly used laboratory electromagnets are resistive and superconducting solenoids for variable fields, Helmholtz coils providing high homogeneity, and solenoids equipped with a magnetic core to amplify and bundle magnetic flux. In the following, these three types will be briefly discussed:

    Resistive and Superconducting Solenoids In solenoids, the current that generates the magnetic field flows along a helical path on a cylinder surface. In a wider sense, the expression is also used for wirewound coils that are composed of successive cylindrical layers, generally of equal length. Solenoids are called “long” when this length is substantially larger than their diameter and the magnetic field in the center of the inner region, called the bore, becomes very homogeneous. They are ideally suited for high-voltage and highfrequency applications. Resistive solenoids are mostly used in combination with magnetic materials. Technical applications include relays as well as linear and rotary actuators, while

    22 Magnetic Fields and Measurements

    1089

    the most typical laboratory devices consist of a pair of electromagnets with iron pole caps and an air gap. These will be discussed separately below (see Sect. “Electromagnets with Magnetic Core”). With the availability of superconducting NbTi or Nb3 Sn alloys, superconducting magnets have become a standard laboratory tool providing field strengths up to 23.5 T for commercial high-resolution NMR applications (1 GHz) [123] and still higher fields up to 24 T in dedicated research laboratories [66]. Various research and development projects are underway to extend this limit by using high-Tc superconductors [181, 191]. As of 2019, a commercial 28.2 T (1200 MHz) highresolution NMR magnet is close to regular operation [152]. Moreover, the high magnetic field facilities in Tallahassee and Grenoble have successfully operated prototypes of full-scale high-Tc magnet inserts in superconducting or resistive background fields providing total magnetic fields of 32 T and 32.5 T, respectively [111, 160, 161]. In Tallahassee, the potential of high-Tc superconductors has been furthermore demonstrated by ramping up a scaled-down test insert until it quenched at 45.5 T [62]. In contrast to electromagnets made of resistive materials, superconducting magnets exhibit no strong heat dissipation when operated at high magnetic fields and hence require no high electric power. However, magnetic pressure and the resulting forces still play a role. In addition, for superconducting magnets, there are other physical phenomena that limit their maximum field and their operating conditions. The principal operation constraint is given by the upper critical field Bc2 (all technical superconductors are type-II), the critical temperature Tc , and the critical current density jc as shown in Fig. 1. For typical operating temperatures at 4.2 K or near 2.0 K, jc currently limits the maximum available field strengths of NbTi magnets to 11–12 T and of Nb3 Sn magnets to 23.5 T (cf., Fig. 1). Moreover, since all technical superconductors are type-II where resistive losses can occur, care has to be taken that a stable superconducting operating state is maintained under field changes as well as current and temperature variations [187]. Equally, the presence of stress induced by magnetic pressure has to be taken into account, as it can alter the critical current of the conductor. Finally, it has to be ensured that in case of a breakdown of the superconducting state (quench), the stored magnetic energy between 0.5 MJ for a NbTi-based magnet operating at 12 T and 4.2 K and 26 MJ for an NMR magnet operating at 23.5 T and 2 K [153] is safely evacuated. This is achieved by embedding the superconducting wire in a highly conductive Cu matrix and by appropriate protection circuits including diodes and heating elements. It should be noted here that superconducting magnets allow a commercial usage of high magnetic fields for NMR spectroscopy and magnetic resonance imaging (MRI). The design of magnets for NMR and MRI is subject to additional constraints with respect to homogeneity (ΔB/B ≤ 10−9 for a sphere of about 30 mm diameter for NMR and ΔB/B ≤ 10−5 for a cylinder of about 50 cm diameter and 30 cm length for MRI) and a time stability of better than 10−8 /h. This is obtained by a set of superconducting and resistive correction coils (shim coils) and magnet operation in persistent mode at current densities far below jc .

    1090

    O. Portugall et al.

    Fig. 1 Left: graph showing the relation between upper critical field Bc2 , the critical temperature Tc , and the critical current density jc of a technical type-II superconductor. The superconducting phase is limited to points below the surface. Right: comparison of critical current densities (at 4.2 K unless otherwise stated) available in strands and tapes of more than 100 m. (Adapted from [93])

    Helmholtz Coils Helmholtz coils are a special form of electromagnets that generate a highly homogeneous field. They consist of a pair of identical solenoids that are thin and short and placed on a common axis in a distance that equals their radius. These conditions ensure that the series expansion of the magnetic field in the geometric center only starts with a fourth-order term. Main laboratory applications of Helmholtz coils using DC currents are low-field NMR and electron paramagnetic resonance (EPR) experiments as well as sensor calibration. Alternatively, Helmholtz coils may also be driven with AC currents in order to study the effect of time-varying magnetic fields, for example, on biological systems. Arrangements of three orthogonal Helmholtz coils are furthermore used to calibrate three-axis sensors and to cancel unwanted magnetic fields and in particular the earth magnetic field for applications such as zero-field NMR. Electromagnets with Magnetic Core Magnetic fields generated by electric currents can be amplified and bundled with magnetic materials exhibiting a high permeability. This is achieved with high-susceptibility metals and alloys involving 3d elements like iron. For many applications, it is desirable to use low-hysteresis materials in order to avoid energy losses when the field is removed. Annealed soft iron has interesting properties in

    22 Magnetic Fields and Measurements

    1091

    this respect, as it exhibits a high saturation field of 2.16 T at room temperature and does not remain magnetized at zero field. Magnetic cores are used in various devices such as laboratory magnets, transformers, electro-motors, magnetic switches, and recording heads. For most of these applications, the maximum field is about 2 T. However, rare-earth materials have also been used to boost high-field resistive magnets [144]. A widely used scheme for magnets with magnetic core consists of two solenoid coils filled with iron cylinders that are aligned on a common axis. Close to the symmetry plane, the iron cylinders are shaped into pole pieces in order to guide magnetic flux into the air gap where the experiment is placed. The shape of the pole pieces can be optimized to provide a suitable compromise between the highest possible and the most homogeneous magnetic field in the air gap [48, 42, 13, 14]. Magnets of this type are commercially available up to 5 T and require electric power of the order of 10 kW and water cooling. They are extensively used for a variety of laboratory applications including most notably electron paramagnetic resonance (EPR) measurements in the L-, S-, and X-Bands (1–2 GHz/0.06 T, 2–4 GHz/0.12 T, and 8–10 GHz/0.3 T). For this application, high-class and high-stability power supplies are required, typically with a stability of less than 10 ppm. A very useful method for the design of electromagnets with magnetic core is provided by  the concept of the magnetic circuit analogy to the electric circuit [78]. The term C H · dl = S J · dS (Ampere’s law) defines a magnetomotive force F (MMF) that takes the role of a “voltage.” In the case of a coil with N turns and a current I , the MMF becomes NI. The “current” of the magnetic circuit is represented by the magnetic flux Φ = S B · dS. The equivalent of the resistance is the reluctance R = F /Φ. For a magnetic circuit in a magnetic material of permeability μ that has a length l and a cross section A, the reluctance R equals l/(Aμ). Using this concept, one can find analogs of Kirchhoff’s law in order to describe circuits containing electromagnets and magnetic materials.

    High-Field Magnet Facilities To obtain the highest possible fields in a practically useful volume requires substantial amounts of energy, respectively power, as well as the possibility to control them. In this section, we discuss these limitations together with the principal technical strategies to overcome them and their implementation in large high-field magnet facilities.

    Performance Limitations and Magnet Classification The performance of electromagnets is fundamentally limited by dissipation and magnetic pressure. The first is part of Poynting’s theorem and hence related to energy conservation in a system composed of electromagnetic fields and particles. The second refers in the same way to momentum conservation (cf., Maxwell stress tensor). Both turn out to be proportional to the square of the magnetic field which explains the substantial technological difficulties in generating high magnetic fields.

    1092

    O. Portugall et al.

    The expression B 2/2μ0 quantifies both the local magnetic energy density associated with a field B and the pressure exerted on a boundary separating a parallel field on one side from a field-free region on the other. It is implicitly understood that such boundary has to be magnetic or contain a net current in order to satisfy the macroscopic Maxwell’s equations. Magnetic pressure is primarily a practical quantity that provides a simple upper limit for the net force a magnet has to sustain. More detailed analyses of local mechanical constraints make use of the force density that can be evaluated either via the divergence of the Maxwell stress tensor f = ∇σ where σij = (2Bi Bj − B 2 δij )/2μ0 or, more commonly, via the Lorentz-force density f = j × B that refers explicitly to a conductor segment carrying a current density j . In a solenoid, f gives rise to axial compression and tangential hoop stress j rB where r is the radius of the current loop. The hoop stress attains its maximum in midplane and represents the primary limitation for the mechanical integrity of solenoidal magnets. Under realistic conditions, its magnitude is comparable with the magnetic pressure of the center field which explains why the latter is often used as a simplified criterion in design studies. Figure 2 gives an impression of the order of magnitude of the magnetic pressure and energy density encountered in advanced high-field magnets. Dissipation makes it necessary to apply an electric field E that maintains the current density j in a magnet. The respective power density j · E = j 2 of a conductor with resistivity ρ gives rise to Joule heating which is quadratic in j and hence B. Figure 3 gives an idea of the actual power conversion inside a highfield electromagnet, i.e., the electrical power that has to be provided to maintain the field and the thermal power that has to be either evacuated (DC magnets) or temporarily stored (pulsed magnets). The total magnetic energy stored in the coil,

    Fig. 2 Conversion scale between magnetic field (lower), pressure, and energy density (both upper). Arrows indicate the maximum field for different types of magnets (lower) and, for comparison with the magnetic pressure, the ultimate tensile strength of high-strength materials (upper). Horizontal bars indicate the maximum energy density obtained with electrostatic (capacitors), electrochemical (batteries), and chemical (fuel) storage. For electrochemical storage, the absolute maximum obtained in non-rechargeable Li batteries is distinguished from typical values for largescale storage in Pb-acid batteries. Although the magnetic pressure and the maximum hoop stress inside a magnet are not a priori equal, both tend to be similar for most practical magnet geometries

    22 Magnetic Fields and Measurements

    1093

    Fig. 3 Dissipation and the energetic requirements of electromagnets featuring constant current density, a quadratic cross section parallel to the axis, and a bore size of either 2 cm (solid lines) or 1/10 (dashed lines) of the outer diameter: (a) required power as a function of magnetic field and coil size (assuming the resistivity of Cu at 300 K); (b) respective power density and associated heating if the dumped energy is not evacuated (assuming the heat capacity of Cu at 300 K); (c) total field/inductive energy stored in the magnet. (a) and (c), respectively, permit the dimensioning of power supplies for DC magnets and energy storage for pulsed magnets. (b) permits estimates of the cooling requirements for DC magnets and the maximum field duration for pulsed magnets depending on the admissible temperature. As a guide to the eye, blue and red circles mark the approximate parameter space for state-of-the-art DC and pulsed high-field magnets. For real magnets, power requirements have to be corrected to account for finite filling factors (28) 30 24 20

    P u l s e d Buse /Btest (T) 35/35.8 43.5 64 60/65 60 60/62 65 65 65/67 70/71 70/74 75/85.8 80/82 83/85 85/90.6 90/91 93/95.6 95/98.8 95/100.8

    Bore (mm) 22 30 22 33 18 28 21 20 15.5 24 13 15 13 16 12 8 12 8.5 10.5

    Bore (mm) 10 6-10

    Fac.

    Bitter Bitter Polyh. Bitter Bitter Bitter Polyh. Bitter Bitter Bitter SCH Hybrid Hybrid Hybrid Hybrid Hybrid

    Ta Nij Gre Hef Tsu Ta Gre Nij Hef Ta Ta Tsu Hef Ta Gre Nij

    Florida-Bitter principle

    m a g n e t s

    FWHM (ms) 1500 1000 10* 300 35 90 80 10 20.1 59 55 4 40 5 23 13 4 11.8 10.5

    M e g a g a u s s Buse (T) 150/300 650/1200

    Type

    Cool (min) 120 240 30 120 20 75 35 60 37 240 50 20 240 240 240 90 240 120 180

    Type Fac. mono mono flat-t mono mono mono mono mono mono mono mono mono mono dual dual dual dual triplex multi

    Wu To Wu Wu To Tls Wu Dre LA Dre Tls To Tls Dre Wu Tls Dre Tls LA

    Polyhelix principle

    Distributed fibre reinforcement

    g e n e r a t o r s FWHM (µ s) 5-7 2

    Type

    Fac.

    STC EMFC

    To,Tls LA

    To

    Single-turn coil

    22 Magnetic Fields and Measurements

    1097

    Bitter-type magnets consist of radially slotted copper disks that are vertically stacked and separated by insulation foils enabling an electrical contact only via a small segment next to the slit. As the position of the slit and the contact changes by an azimuth angle along the stacking direction, a series connection of all disks is established that gives rise to a helical current and hence a magnetic field. Cooling is ensured by holes that allow a vertical flow of water [15,151]. The shape of the holes is optimized so as to provide the best possible compromise between heat exchange and flow resistance. Their radial distribution furthermore anticipates local heating due to the non-homogeneous current density, while their azimuthal arrangement close to the bore avoids the radial transmission of forces [9, 10]. Polyhelix magnets consist of a set of (up to 14) concentric cylinders with individual wall thicknesses of about 1 cm. Each cylinder has a helical slit that is produced by electroerosion and subsequent gluing with an insulating material. The entire set of cylinders is connected in series and generates a magnetic field along the axis of the cylinder. Cooling is ensured by narrow hydraulic channels between two neighbouring cylinders [149, 33]. Polyhelix magnets are more difficult to fabricate than Bitter magnets but provide more design freedom as far as the spatial control of the current density is concerned. Resistive high-field magnets have reached values between 36 and 41.5 T (cf., Table 1) giving rise to stresses up to 400 MPa, a value which exceeds the tensile limit of Cu (cf., Fig. 2) and comes close to that of other currently available low-resistance materials featuring conductivities of 45 to 58 MS/m. Although the midplane tensile stress undoubtedly represents the most crucial limitation for such magnets, their design also has to pay heed to axial compression that may in turn create shear forces. Further complications are due to the mechanical robustness of insulating materials that have to sustain at least part of the applied forces. Taking into account the finite filling factor of real magnets, i.e., the fraction of the volume actually occupied by conductor material, one obtains slightly larger power values than those in Fig. 3. Large facilities therefore operate power supplies providing 15 to 25 MW giving rise to power densities of typically 2 W/mm3 inside a magnet. This causes heat flux densities up to 5 W/mm2 at their surface, a value about one order of magnitude higher than in pressurized vessels of steam generators. The  Table 1 (continued) of each table-cell relates the displayed value to the maximum in the respective class of magnets. Blue and green colours are used to highlight groups of magnets with similar features. The last column indicates the facility as explained below. DC-magnet table: SCH stands for series-connected hybrid. Magnets under construction are in italics. Pulsed magnet table: Buse /Btest distinguishes fields that are available for users from records obtained with the same magnet. For nested coils (dual, triplex, multi) the pulse duration (FWHM) refers to the shortest pulse of the innermost coil. The starred value for flat-top magnets (flat-t, see explanation in the text) indicates the duration of the plateau. Megagauss table: user fields depend on minimum bore size. Dre: HLD Dresden [73], Gre: LNCMI Grenoble [92], Hef: HMFL Hefei [71], LA: NHMFL Los Alamos [110], Nij: HFML Nijmegen [70], Ta: NHMFL Tallahassee [110], Tls: LNCMI Toulouse [92], To: ISSP Tokyo/Kashiwa [80], Tsu: NIMS Tsukuba [173], Wu: WHMFC Wuhan [188]

    1098

    O. Portugall et al.

    dissipated energy is evacuated by a constant flow of liquid water at 300 K generating a typical temperature rise by 30 K. Ideally, the respective cooling power is exactly equivalent to the applied electrical power, thus permitting continuous operation at the highest field values. In practice, all but one [150] of the high-field facilities use pre-cooled water reservoirs to enable operation at electric power levels above the average cooling power for a limited amount of time. A resistive DC-magnet with its hydraulic tubing is depicted in Fig. 5. Hybrid magnets. Although it would be technically possible to generate fields of the order of 50 T in resistive magnets [5], controlling the thermal and mechanical constraints would require a substantial increase of their size and a reduction of their filling factor to provide additional space for cooling and mechanical support [8]. Therefore, higher DC magnetic fields are generated with hybrid magnets that combine a resistive and a superconducting part. Table 1 provides an overview of operational and projected hybrids. Further details can be found in the literature [11, 21, 49, 114, 113, 129, 185]. Hybrid magnets are able to generate magnetic field strengths above 30 T with only 10 MW of electric power and above 40 T with 20 MW. The resistive and superconducting parts are usually operated by different power supplies. This is due to the preferred operation mode for hybrids: the superconducting part is first ramped to maximum field and kept there; then, field ramps are only performed with the resistive part. Another reason for separate power supplies is the different electric power consumptions and characteristics of the two types of magnets: the resistive part typically consumes 20 MW and exhibits low inductance values (below 50 mH), whereas the superconducting part has a large inductance (up to 3 H) and requires only a very small fraction of the total power. However, there also exists a special variation of the hybrid design, the so-called series-connected hybrid, where the superconducting and resistive parts are connected in series to the same power supply [36]. This design provides enhanced stability and resolution, since fluctuations of the resistive magnet and the power supply are damped by the superconducting part. Series connected hybrids therefore enable high-resolution NMR in the 30–40 T field range [55]. Owing to the critical field that limits superconductivity, hybrid magnets always consist of a superconducting outsert and a resistive insert. Since the field of the outsert exerts additional forces on the insert, the field strength of the latter is reduced with respect to its maximum value in stand-alone operation. Mechanical stability is also a problem for the superconducting part, as strong magnetic forces between the insert and the outsert may occur in case of asymmetries due to imperfect mounting or partial defects in the resistive part. The design and operation of hybrid magnets furthermore requires special considerations regarding the stability and safety of the superconducting part in the presence of a rapidly ramped resistive insert. Current hybrid projects are aiming to obtain magnetic fields up to 45 T using a combination of low-temperature superconductors and resistive magnets. Future plans aim to reach a magnetic field of the order of 60 T. However, using presentday technology, such field values would require large electric power (40–50 MW) and a superconducting outsert operating near 20 T with a stored energy of the order

    22 Magnetic Fields and Measurements

    1099

    Fig. 5 (a) Historical image of the Bellevue electromagnet built in 1928 to generate 5–7 T (©CNRS photothèque, with kind permission); (b) modern triplex-pulsed magnet with 700 mm outer diameter producing close to 100 T (©LNCMI, with kind permission); (c) hydraulics tubes around the chassis of a resistive DC magnet (©Steffen Krämer); (d) 50 MJ capacitor bank for generating pulsed fields (©HZDR, with kind permission)

    of 1 GJ [129]. Consequently, next-generation hybrid magnets will have to involve high-temperature superconducting materials with their capacity to operate beyond 30 T [111, 160, 161, 62]. Design studies are underway to reach this ambitious goal [11, 128].

    Pulsed Magnets and Facilities Pulsed magnets [67] provide higher fields than DC magnets, are cost-efficient in terms of investment and operation, and can be constructed with relatively basic technical know-how. In Western Europe alone, small pulsed magnet facilities have been, or are, operated in at least 15 different laboratories. The creation of large infrastructures is a recent phenomenon. As of 2017, fully operational facilities making routine use of energies in the 10 MJ range exist in Dresden (EMFLHLD, Germany), Los Alamos (NHMFL, United States), Toulouse (EMFL-LNCMI, France), and Wuhan (WHMFC, China). As outlined before, in pulsed magnets, the dissipation problem is solved by reducing the field duration. For all practical purposes, this provides satisfactory solutions despite the fact that shorter pulse durations also give rise to additional heating by eddy currents or, in an alternative description, the finite currentpenetration depth that effectively reduces the available conductor cross section. This

    1100

    O. Portugall et al.

    penetration depth is related to, but not identical with, the ordinary skin effect that only applies for stationary oscillations. More recently, pulsed magnets have also been equipped with cooling channels, not to evacuate heat during shots but in between [53]. This gives rise to considerably enhanced duty cycles which represent an important factor for applications in user facilities. Realistic estimates for the heating of pulsed magnets can be obtained by integrating both sides of the expression j 2 (t) dt = D(T ) cp (T )/ρ(T ) dT that associates the application of a current density j for a short interval dt with a temperature rise dT in a conductor that is characterized by its temperature-dependent mass density D, specific heat cp , and resistivity ρ. For most conductors, this so-called action integral becomes more favorable at low temperatures which is one of the reasons why advanced pulsed magnets are pre-cooled with liquid nitrogen. A base temperature of 77 K also leaves a larger margin with respect to the relatively limited heat tolerance of many insulating materials. The heating of pulsed magnets during a shot can easily exceed 100 K. In complete analogy to resistive DC magnets whose heat limitation motivates the integration of cooling channels, pulsed magnets require reinforcing elements to contain the applied magnetic pressure. First-generation pulsed magnets were therefore reinforced externally. Their design was based on analytic calculations and the destructive testing of downsized prototype coils followed by the application of scaling rules to build larger specimens to be operated below the destructive limit. The main shortcoming of this approach is that the magnet is regarded as a mechanical continuum rather than a complex structure composed of helical windings, isolations, reinforcements, and impregnations which cannot be easily scaled simultaneously. With one exception [89], modern pulsed magnets or pulsed magnet inserts attaining the highest available fields are all based on the same basic concept of distributed fiber reinforcement [17]. The principle is that of a wire-wound magnet whose conducting layers are individually stabilized with layers of high-strength fibers, a technique that has the additional benefit of electrically insulating adjacent conductor sections. The material of choice is a synthetic polymer fiber commonly known as Zylon™ whose tensile strength comes close to that of carbon fiber albeit without the disadvantage of being conductive. The thickness of each reinforcement section is adjusted such that neither a substantial transmission of forces between successive layers occurs nor a separation when the coil is elastically deformed. In recent years, the perfection of finite-element methods has shifted the design focus from the comparably simple question of how to contain maximum stresses in midplane to problems related to aging and local forces caused by the imperfect cylinder symmetry of real magnets. In the latter respect, current terminals and the transition points at either end of the magnet where wire is guided from one layer to the next are potentially weak spots. Aging, on the other hand, can be due to either repeated deformation beyond the elastic limit if a magnet is operated close to its maximum, or thermal cycles and, in some cases, also the electric or mechanical deterioration of insulating materials.

    22 Magnetic Fields and Measurements

    1101

    Table 1 gives an overview of key parameters of advanced pulsed magnets. The relatively large variation of parameters such as bore size, pulse duration, and cooling time reflects technical limitations of available energy sources as well as deliberate choices to account for specific experimental requirements. Nested coils marked “dual,” “triplex,” or “multi” resemble DC-hybrid magnets insofar as their architecture permits different technical choices closer to the bore where fields are highest and further away where most of the energy is stored and needed. In nested systems, the innermost coils thus dispose of the strongest reinforcement which reduces their filling factor, i.e., the relative amount of conductor material, and hence their heat capacity. As a consequence, they require shorter pulse durations and are powered independently. An example of a pulsed triplex-coil with 3 independent coaxial current terminals is shown in Fig. 5. With few exceptions, pulsed field installations use capacitor banks as energy source, cf., Fig. 5. Figure 6 shows the operation principle depending on whether a simple serial RLC arrangement with a primitive closing switch is used or a modern setup with a thyristor switch and diode crowbar. The use of solid-state devices makes the switching practically noise-free and allows for some control of the discharge. A full crowbar circuit thus avoids a voltage reversal that ultimately deteriorates capacitors and thyristors, reduces the heating of the magnet, and produces a smooth field decay that is preferable for field-dependent measurement applications. Losses notwithstanding, the energy stored in the capacitor bank 1/2 CU 2 provides the magnetic energy, cf., Fig. 3c, and hence limits either the coil size for a given peak field or vice versa. The capacitance C imposes further restrictions since together with the inductance of the electrical circuit – primarily the coil – it determines the pulse duration and hence the heating as demonstrated in Fig. 3b. The application of high voltages is therefore unavoidable. In practice, capacitor banks for conventional

    Fig. 6 Voltage, current, and coil heating for different configurations of capacitor-driven pulsed field generators. Violet, generator with a simple closing switch S and no crowbar CB; Red, thyristor switch but no crowbar; Orange, modern crowbar circuit. The thyristor-and-crowbar option has the advantage that the capacitor bank is discharged at the end of the pulse. The smooth decay is also favorable for measurement applications, and the heating can be tuned via the crowbar resistance

    1102

    O. Portugall et al.

    pulsed magnets are limited to 25 kV, a range where corona and other partial discharge phenomena can still be controlled with relatively simple technical means and where self-healing metallized film capacitors for pulsed power applications are widely available on the market. Although state-of-the-art capacitor banks in large facilities feature energies exceeding 10 MJ, a variety of smaller sizes exists ranging from transportable systems in shipping containers over mid-sized installations for special applications [2, 52, 81] down to tabletop and even microscopic generators [96]. Special architectures have also been developed to produce repetitive [79] and flat-top magnetic fields [85]. As far as alternative energy sources are concerned, inductive storage has been successfully tested in the past [6], but the project has been discontinued. Flywheel generators originally built for other high-power applications have been recycled by some facilities in order to obtain energies approaching the gigajoule range and the possibility to implement controlled waveforms [147, 90]. Although generating the highest possible fields for scientific research remains at the center of development efforts, pulsed magnet technology has started to expand into other directions. Recent innovations for scientific purposes include split coils providing lateral access in order to combine high fields with intense radiation sources [40, 81] and pulsed dipole magnets for investigating the quantum vacuum [7]. Pulsed magnets have also found their way into industrial applications where they are used for magnetizing permanent magnets, electromagnetic forming, and magnetic pulse welding.

    Megagauss Magnetic Fields Megagauss generators provide magnetic fields well beyond the current technological limit for conventional pulsed magnets, i.e., 100 T or 1 megagauss in cgs units. Both their production and their use in scientific applications require highly specialized equipment that is available in only a few laboratories worldwide. In 2019, Megagauss fields for scientific research were generated on a regular basis in Los Alamos (NHMFL, United States), Tokyo (ISSP, Japan), and Toulouse (EMFLLNCMI, France). Often coined destructive techniques because coils are inevitably destroyed, Megagauss generators make explicit use of inertia that resists the displacement of conductor elements by a magnetic force. Interestingly, Newton’s historical distinction between objects that are either initially at rest or in a state of uniform motion finds its counterpart in two distinct techniques for generating Megagauss fields. Single-turn coils have been reported to produce fields up to 355 T [154], yet their principal interest lies in their capacity to generate 150 to 200 T in a useful diameter of 10 mm without causing excessive peripheral damage [68,109,125,103]. The latter applies for thin-walled single-turn coils that basically consist of a rectangular copper strip bent to form a single coil winding (Table 1). When the coil disintegrates, the magnetic pressure in the bore propels fragments radially outward, thus protecting equipment in the center. As of 2015, single-turn coils with almost identical geometry

    22 Magnetic Fields and Measurements

    1103

    and typical dimensions of 2h = 2a1 = 10 ∼ 15 mm and a2 − a1 = 2 ∼ 3 mm were operated by three facilities worldwide to perform regular scientific experiments. Although their physics is characterized by a complex interplay of electrodynamic, thermodynamic, and mechanical processes [67, 91], the basic working principles of single-turn coils can be derived relatively simply. Most importantly, a relevant timescale can be determined by assuming the magnetic pressure associated with a sinusoidal field, neglecting all elastic forces and integrating the equation of motion dF = p dA = dm r¨ for the radial displacement r of a conductor element. A sensible limit of