Theoretical Concepts of Spin-Orbit Splitting

Similar documents
Optical Lattices. Chapter Polarization

Tight-Binding Model of Electronic Structures

PROJECT C: ELECTRONIC BAND STRUCTURE IN A MODEL SEMICONDUCTOR

Electrons in a periodic potential

Lecture 6. Tight-binding model

The 3 dimensional Schrödinger Equation

Problem Set 2 Due Thursday, October 1, & & & & # % (b) Construct a representation using five d orbitals that sit on the origin as a basis:

Refering to Fig. 1 the lattice vectors can be written as: ~a 2 = a 0. We start with the following Ansatz for the wavefunction:

PHYSICS 4750 Physics of Modern Materials Chapter 5: The Band Theory of Solids

3: Many electrons. Orbital symmetries. l =2 1. m l

Physics 221A Fall 2018 Notes 22 Bound-State Perturbation Theory

Energy bands in two limits

1 One-dimensional lattice

Problem 1: Spin 1 2. particles (10 points)

Physics 211B : Problem Set #0

Electronic Structure of Surfaces

Structure of diatomic molecules

PH 451/551 Quantum Mechanics Capstone Winter 201x

Modeling Transport in Heusler-based Spin Devices

Lecture. Ref. Ihn Ch. 3, Yu&Cardona Ch. 2

Answers Quantum Chemistry NWI-MOL406 G. C. Groenenboom and G. A. de Wijs, HG00.307, 8:30-11:30, 21 jan 2014

Angular momentum. Quantum mechanics. Orbital angular momentum

IV. Electronic Spectroscopy, Angular Momentum, and Magnetic Resonance

DFT EXERCISES. FELIPE CERVANTES SODI January 2006

Problem Set 2 Due Tuesday, September 27, ; p : 0. (b) Construct a representation using five d orbitals that sit on the origin as a basis: 1

The symmetry properties & relative energies of atomic orbitals determine how they react to form molecular orbitals. These molecular orbitals are then

CHAPTER 10 Tight-Binding Model

Lecture 13: The Actual Shell Model Recalling schematic SM Single-particle wavefunctions Basis states Truncation Sketch view of a calculation

POEM: Physics of Emergent Materials

Three Most Important Topics (MIT) Today

ECEN 5005 Crystals, Nanocrystals and Device Applications Class 20 Group Theory For Crystals

Rotational Motion. Chapter 4. P. J. Grandinetti. Sep. 1, Chem P. J. Grandinetti (Chem. 4300) Rotational Motion Sep.

Physics 221A Fall 1996 Notes 16 Bloch s Theorem and Band Structure in One Dimension

Salmon: Lectures on partial differential equations

Quantum Theory of Angular Momentum and Atomic Structure

The Simple Harmonic Oscillator

Curvilinear coordinates

Crystal field effect on atomic states

Chem 442 Review for Exam 2. Exact separation of the Hamiltonian of a hydrogenic atom into center-of-mass (3D) and relative (3D) components.

Spin Interactions. Giuseppe Pileio 24/10/2006

Chapter 3. The (L)APW+lo Method. 3.1 Choosing A Basis Set

Strongly correlated systems in atomic and condensed matter physics. Lecture notes for Physics 284 by Eugene Demler Harvard University

26 Group Theory Basics

One-electron Atom. (in spherical coordinates), where Y lm. are spherical harmonics, we arrive at the following Schrödinger equation:

Nuclear Shell Model. Experimental evidences for the existence of magic numbers;

Physics 541: Condensed Matter Physics

Physics of Semiconductors (Problems for report)

I. CSFs Are Used to Express the Full N-Electron Wavefunction

SUPPLEMENTARY INFORMATION

PHYS 3313 Section 001 Lecture # 22

Section 3 Electronic Configurations, Term Symbols, and States

Semiconductor Physics and Devices Chapter 3.

Electronic structure of correlated electron systems. G.A.Sawatzky UBC Lecture

structure of graphene and carbon nanotubes which forms the basis for many of their proposed applications in electronics.

QUANTUM MECHANICS AND ATOMIC STRUCTURE

Supplementary Materials for

The Quantum Heisenberg Ferromagnet

Chapter 2 Approximation Methods Can be Used When Exact Solutions to the Schrödinger Equation Can Not be Found.

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

2.3 Band structure and lattice symmetries: example of diamond

Effective theory of quadratic degeneracies

Spins and spin-orbit coupling in semiconductors, metals, and nanostructures

Electronic structure theory: Fundamentals to frontiers. 1. Hartree-Fock theory

Physics 228 Today: April 22, 2012 Ch. 43 Nuclear Physics. Website: Sakai 01:750:228 or

1.6. Quantum mechanical description of the hydrogen atom

Lecture contents. A few concepts from Quantum Mechanics. Tight-binding model Solid state physics review

Supplementary Discussion 1: Site-dependent spin and orbital polarizations of the highestenergy valence bands of diamond, Si, Ge, and GaAs

Lecture 4: Basic elements of band theory

Phys 622 Problems Chapter 5

Classical Mechanics. Luis Anchordoqui

Contents. 1 Solutions to Chapter 1 Exercises 3. 2 Solutions to Chapter 2 Exercises Solutions to Chapter 3 Exercises 57

5.5. Representations. Phys520.nb Definition N is called the dimensions of the representations The trivial presentation

Lecture 45: The Eigenvalue Problem of L z and L 2 in Three Dimensions, ct d: Operator Method Date Revised: 2009/02/17 Date Given: 2009/02/11

CHEM-UA 127: Advanced General Chemistry I

Lecture 5. Hartree-Fock Theory. WS2010/11: Introduction to Nuclear and Particle Physics

Bonding in solids The interaction of electrons in neighboring atoms of a solid serves the very important function of holding the crystal together.

Schrödinger equation for the nuclear potential

Physical Chemistry Quantum Mechanics, Spectroscopy, and Molecular Interactions. Solutions Manual. by Andrew Cooksy

Quantum Mechanics Solutions

A Quantum Mechanical Model for the Vibration and Rotation of Molecules. Rigid Rotor

Chapter 4 Symmetry and Chemical Bonding

Chemistry 431. Lecture 14. Wave functions as a basis Diatomic molecules Polyatomic molecules Huckel theory. NC State University

An introduction to Solid State NMR and its Interactions

Schrödinger equation for central potentials

PHY492: Nuclear & Particle Physics. Lecture 6 Models of the Nucleus Liquid Drop, Fermi Gas, Shell

Nearly Free Electron Gas model - II

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

arxiv: v3 [cond-mat.mes-hall] 31 Oct 2015

Degenerate Perturbation Theory. 1 General framework and strategy

Tight-Binding Approximation. Faculty of Physics UW

Physics 221A Fall 1996 Notes 13 Spins in Magnetic Fields

Time part of the equation can be separated by substituting independent equation

C2: Band structure. Carl-Olof Almbladh, Rikard Nelander, and Jonas Nyvold Pedersen Department of Physics, Lund University.

SECOND PUBLIC EXAMINATION. Honour School of Physics Part C: 4 Year Course. Honour School of Physics and Philosophy Part C C3: CONDENSED MATTER PHYSICS

Errata 1. p. 5 The third line from the end should read one of the four rows... not one of the three rows.

Chapter 4 Symmetry and Chemical Bonding

5.61 Physical Chemistry Lecture #36 Page

Electronic structure of correlated electron systems. Lecture 2

The 1+1-dimensional Ising model

Transcription:

Chapter 9 Theoretical Concepts of Spin-Orbit Splitting 9.1 Free-electron model In order to understand the basic origin of spin-orbit coupling at the surface of a crystal, it is a natural starting point to consider a free-electron model. The model describes a 2D electron gas (2D-EG) with spin-orbit interaction, applicable to surface states that are localized in the near surface plane [83]. The model Hamiltonian will be written as with the Rashba Hamiltonian H = p2 2m + H R, (9.1) H R = α ( e z k) σ. (9.2) Solutions of the Schrödinger equation with the Hamiltonian of Eq. 9.1 can be found analytically. We choose a Cartesian coordinate system with the z- axis perpendicular to the 2D plane of electron motion. It can be shown (see Appendix C) that the free-electron parabolic dispersion will now be split in two parabola, shifted in k direction with energies E 1,2 = 2 k 2 2m The associated wave functions have the form ± α k (9.3) ψ 1 (r) = e i k r ( + ie iθ k ) for E 1 = 2 k 2 2m + α k, (9.4) ψ 2 (r) = e i k r (ie iθ k + ) for E 2 = 2 k 2 α k. (9.5) 2m The spin functions, denote spin-up and spin-down electron states, respectively, with respect to the z direction; θ k is the angle between k and x- axis, according to geometry illustrated in Fig. 9.1. The linear combination of 85

86 CHAPTER 9. THEORETICAL CONCEPTS Figure 9.1: Geometry considered in the free-electron model. spin-up and spin-down states in the eigenfunctions 9.4 and 9.5 describe a spin polarization within the xy-plane (see Appendix C), which is oriented perpendicular to the direction of electron motion. This means that, the electrons with the same k vector and opposite spins will have different energies. The results derived from this model are illustrated in Fig. 9.2. In the lower part, the Rashba spin-orbit split energy dispersion of a (free-electron like) surface state is presented. The two concentric circles in the upper part represent the corresponding Fermi surfaces, with arrows indicating the in-plane spin orientations that are always perpendicular to the electron momentum. The most important results of the model are: (1) it correctly describes the nature of the splitting and, (2) it shows that Rashba type spin-orbit interaction will orient the spins of propagating electrons. However, the size of the splitting, estimated by using the surface potential gradient, is much too small compared with experimental values. The surface potential gradient can be roughly approximated by ( V ) z Φ/λ F, with the work function Φ and the Fermi wavelength λ F [84]. Applied to the case of Au(111) (λ F = 5Å and Φ = 4.3 ev), this leads to an estimated value of E = 10 6 ev, which is several orders of magnitude smaller than the experimentally observed splitting. This deficit is inherent in the free electron model, which does not account for regions of steep nuclear potential gradients near atomic cores. Shockley-type surface states penetrate as deep as several atomic layers into the bulk [85]. It has been shown by G. Bihlmayer [70] for the example of the gold surface state, that regions near atomics cores have also to be taken into account when describing the Rashba splitting of a surface state. The potential gradients close to the nucleus of an atom, leading to atomic spin-orbit splitting, are orders of magnitude larger than the gradient of surface barrier, and its consideration leads to an increase of the Rashba splitting. In order to obtain a better understanding of the atomic contributions to spin-orbit interaction at the surface, a tight-binding model shall be considered next.

9.2. TIGHT-BINDING MODEL 87 Figure 9.2: Energy dispersion of free-electron state with Rashba splitting bottom part together with the corresponding Fermi surfaces (upper part). Arrows indicate spin orientations of electrons at the Fermi surface. 9.2 Tight-binding model The free-electron model is built upon the assumption that electrons propagate almost freely and that their wave functions can be approximated by plane waves. As an opposite starting point, one can use localized atomic orbitals as a basis set to perform band-structure calculations; this is the tight binding (TB) or linear combination of atomic orbitals (LCAO) approach. This approach was developed in molecular physics and has been extended to the description of electronic states in solids. As a starting point, we consider two identical but separate atoms with their (atomic) wave function for a single valence electron. As we bring them together, the interaction between the two valence charges will lead to the formation of new orbitals, the so-called bonding and antibonding orbitals. The energy of the bonding orbital is lowered by an amount that is determined by the interaction Hamiltonian, and the energy of the antibonding orbital is raised by the same amount. The matrix elements of the interaction Hamiltonian between atomic orbitals are usually referred to as overlap parameters. They have a simple physical interpretation as interaction strength between electrons on adjacent atoms. The concept of bonding and antibonding orbitals is easily extendable to crystals, if we assume that atomic orbitals of each atom in the crystal overlap with whose of its nearest neighbors only. This is a reasonable approximation for many solids. As a result of the overlap (during crystal formation) bonding and antibonding orbitals broaden forming valence and conductions bands, respectively. Of course, the crystal

88 CHAPTER 9. THEORETICAL CONCEPTS structure strongly affects the overlap between the atomic orbitals. Surface states at a crystal surface form a two-dimensional manifold of energy eigenstates. For the present approach, we take Wannier functions to form a basis set. Surface states are localized states that decay exponentially into the bulk. To represent the essential physics of spin-orbit interaction, we use p-orbitals as an example, keeping the model simple and transparent, and to reveal the important ingredients of the Rashba effect. For sp-derived surface states, like those at the Au(111) surface, this is sufficient, since s-states do not contribute to spin-orbit interaction (zero angular momentum). For Gd and Tb, however, it is not fully correct to consider p-states, since the surface state in this case has d z 2 character (derived from the bulk d-band). Construction of the model with d-orbitals included can be done in full analogy to the case with p-orbitals as a basic set. However, including d-states will increase the number of basis functions and will probably conceal the main idea of this consideration, i.e. to obtain a transparent and clear physical picture of the problem. Instead, we will rather compare with results of density-functional-theory (DFT) calculations, see below. The Gd(0001) and Tb(0001) surfaces considered in this work have a sixfold symmetry at the surface. Accordingly, surface states are described in the model with sheet of hexagonally arranged atoms using the three atomic orbital functions p x, p y, p z, where the in-plane orbitals p x and p y are coupled in the usual way by the help of the directional cosine. Atomic orbitals can be expressed as the product of a radial wave function and spherical harmonics Yl m (θ,φ), with the atomic nucleus chosen as the origin. In case of two atoms, it is convenient to choose coordinate axes such that the z-axis is parallel to the vector d connecting the atoms, and the azimuthal angles are the same. The interaction Hamiltonian will have cylindrical symmetry with respect to d and therefore cannot depend on φ. Symmetry consideration permits to gradually reduce the number of overlap parameters, which have to be found by sorting out zero and equals integrals. For a crystal lattice, it is convenient to choose crystallographic axes as coordinate axes, and the spherical harmonics Yl m (θ,φ) of the atomic orbitals are defined with respect to a fixed coordinate system. In calculating the overlap parameter for any pair of neighboring atoms, one expands the spherical harmonics defined with respect to d in terms of Yl m (θ,φ). Figure 9.3 illustrates this expansion for p-orbitals. If one considers the surface layer as an isolated system, p x and p y would not couple to the p z orbital. Inversion symmetry z z would lead to an equally large overlap of a p x, p y lobe with the positive and negative p z lobe, resulting in a zero total overlap. At the surface, this symmetry is obviously broken due to the surface potential, V (z), which breaks the symmetry of the electron density distribution of surface states with respect to the surface plane. Only the p x, p y orbitals will retain the symmetry of their atomic counterparts. This effect can be included by actually coupling p x, p y, and p z states. As was found by Petersen and Hedegard [83], the simplest tight-binding Hamiltonian that allows us to discuss the influence of SOC (on the 2D electronic structure) has the form

9.2. TIGHT-BINDING MODEL 89 Figure 9.3: Projection of the overlap parameter between p y orbitals onto directions along the vector d joining the two atoms and perpendicular to d. From Ref. [86]. H 0 = t αβ (R i R j ) p α (R i ),σ p β (R j ),σ, (9.6) where t αβ (R i R j ) is the overlap matrix element given by w cos 2 (θ ij ) + δ sin 2 (θ ij ) (α,β) = (x,x) (w + δ)cos(θ ij )sin(θ ij ) (α,β) = (x,y) = (y,x) w sin 2 (θ t αβ (R i R j ) = ij ) δ cos 2 (θ ij ) (α,β) = (y,y) γ cos(θ ij ) (α,β) = (x,z) = (z,x) γ sin(θ ij ) (α,β) = (y,z) = (z,y) δ (α,β) = (z,z) for R i and R j pointing towards nearest-neighbor lattice sites only; all other matrix elements are zero. θ ij is the angle between the vector (R i R j ) and the x-axis. Parameter w denotes the V ppσ and δ is a shorthand for V ppπ. V ppσ and V ppπ are visualized in Fig. 9.3; their physical interpretation is given by the matrix elements of the interaction Hamiltonian (9.6) that have been calculated [87, 88] for different lattices and can be found in the literature. The parameter γ= p z (R) V p n (R + x) (n = x,y) is essentially a measure of the surface potential gradient; it has a similar meaning as V z in the free electron model discussed above. The case γ = 0 corresponds to a 2D model with inversion symmetry. By forming Bloch waves with overlap matrix elements written above and by diagonalizing the resulting matrix, we can solve the problem given by the Hamiltonian (9.6). Without taking into account the electron spin, the resulting matrix will be only 3 3. From here, with γ/w = 0.1 and δ/w = 0.3, we obtain the band structure shown in Fig. 9.4. The band with a minimum at the Γ point is primarily of p z character. The true surface states correspond to the free-electron-like part of this band around the Γ point. Also, there are strongly dispersive bands with minima at the M-point and along the Γ-K direction; they have primarily p x, p y character. The gaps appear due to the broken symmetry with respect to reflection z z, which is controlled by the γ parameter; the

90 CHAPTER 9. THEORETICAL CONCEPTS Figure 9.4: Tight-binding-model band structure for surface states in a twodimensional hexagonal lattice, derived from p-orbitals without spin-orbit interaction. The true surface states belong to the free-electron-like band around the Γ point with primarily p z character; from Ref. [83]. gaps disappear when γ = 0, and symmetry is restored. Only the free-electronlike band around the Γ point resides inside a gap of the projected bulk band structure in a real crystal and therefore can be identified as a true surface state. To introduce spin-orbit interaction into our model, we take the Hamiltonian H soc = αl S = α 2 (L+ σ + L σ + + L z σ z ), (9.7) where L ± are angular momentum step-up/step-down operators and σ ± are the same for the spin, α is the atomic spin-orbit constant. Now, when taking into account the spin parts of the atomic wave functions used as basis set { p x,, p x,, p y,, p y,, p z,, p z, }, the resulting matrix becomes H soc = α 2 0 i 0 0 0 1 i 0 0 0 0 1 0 0 0 1 i 0 0 0 1 0 i 0 0 0 i i 0 0 1 i 0 0 0 0. (9.8) Inspection of the H soc matrix directly reveals that, for spin-orbit interaction to become effective, it is necessary to have p x and p y states in the model. Otherwise, as can be seen from the 2 2 block of zeros in the lover right corner of the matrix that corresponds to p z states, no spin-orbit interaction will be present.

9.2. TIGHT-BINDING MODEL 91 Figure 9.5: Band structure with spin-orbit splitting for surface states within the tight-binding model constructed from p-orbitals in a two-dimensional hexagonal lattice. The true surface states belong to the free-electron-like band around the Γ point. The splitting is determined by the atomic spin-orbit coupling strength, α. By including H soc in the Hamiltonian (9.6), one can derive the spin-orbitsplit band structure. The result of the calculation with α = 0.2w is presented in Fig. 9.5. The idea of this section is not only to demonstrate that the free-electron model and the tight-binding model can provide some insight into the role of spinorbit interaction in the surface electronic structure, but also to demonstrate the important contribution of atomic spin-orbit coupling to the Rashba splitting of surface states. For this purpose, we will limit ourselves to the subspace of the Hilbert space that is spanned by p z -like states. A simple truncation of the Hamiltonian to this subspace will give no spin-orbit interaction, since this requires breaking parity which is only achieved by an admixture of p x,y states, i.e. by a coupling between p z and p x,y orbitals. To account for these bands, these virtual transitions to them should be included to second order in the coupling. This is possible as long as the p z and the p x, p y bands are well separated in energy. From a mathematical point of view, we follow the work by Petersen and Hedegard [83]. One writes the resolvent (ǫ H) 1, which is the Laplace transformation of the time evolution operator, and consequently includes all necessary information. Since we consider only the subspace spanned by p z and we are interested in the time evolution of the low-energy bands, our projected resolvent is P(ǫ H) 1 P, where P is the projection operator onto the p z subspace. From linear algebra one obtains [83]:

92 CHAPTER 9. THEORETICAL CONCEPTS P 1 ǫ H P = 1 ǫ PHP PHQ 1 ǫ QHQ QHP, where Q projects onto the complement of P, i.e. Q = 1 P. Therefore the effective Hamiltonian for the p z subspace takes the form: 1 H eff = PHP + PHQ ǫ QHQ QHP. For ǫ and QHQ in the denominator, we make the following approximation. Since we are interested in k points close to the Γ point, ǫ ǫ 0 z (k = 0) and QHQ Qǫ 0 x,y (k = 0) will be a good approximation, where ǫ0 i (k)... denotes the unperturbed band structure. This procedure, performed to second order in α,γ, and k, yields the following expression for the subspace spanned by the p z orbitals: ( 6δ + ( 3 H eff = 2 δ + 9γ2 /w)k 2 ) 6i(k x ik y )αγ/w 6i(k x + ik y )αγ/w 6δ + ( 3 2 δ + 9γ2 /w)k 2. We can now directly identify the contributions to the Hamiltonian that describe the system with spin-orbit interaction. The diagonal terms are similar to the free-electron-model case, with the effective mass determined by δ. The offdiagonal term corresponds to the Rashba spin-orbit term, and the parameter α R is equal to 6αγ/w. This explicitly demonstrates that the Rashba spin-orbit interaction at the surface depends on the atomic spin-orbit parameter (α), as well as on the potential gradient at the surface, described by the parameter γ. When setting γ to zero, which corresponds to a switching-on of the inversion symmetry, the Rashba splitting will be removed. Also, the disappearing of the atomic spin-orbit interaction will lead to a vanishing of the Rashba splitting. With this consideration, we have achieved a connection between the freeelectron model and the tight-binding model. Finally, we like to mention that atomic spin-orbit interaction is responsible for a pronounced Z-dependence of the Rashba surface-state splitting. From the present tight-binding model calculations (Ref. [83]) it can be understood why the Rashba splitting should be of the same order of magnitude as the atomic splitting, and not much smaller as was estimated earlier using the free-electron approach.