Lecture 6. Tight-binding model

Similar documents
VALENCE Hilary Term 2018

Lecture 3. Energy bands in solids. 3.1 Bloch theorem

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

CHAPTER 10 Tight-Binding Model

Lecture 15 From molecules to solids

Kronig-Penney model. 2m dx. Solutions can be found in region I and region II Match boundary conditions

Electrons in a periodic potential

Theoretical Concepts of Spin-Orbit Splitting

Lecture 4: Basic elements of band theory

Lecture 9: Molecular Orbital theory for hydrogen molecule ion

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

Electron States of Diatomic Molecules

CHEMISTRY Topic #1: Bonding What Holds Atoms Together? Spring 2012 Dr. Susan Lait

PROJECT C: ELECTRONIC BAND STRUCTURE IN A MODEL SEMICONDUCTOR

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

2 Electronic structure theory

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

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 9, February 8, 2006

Chapter 1: Chemical Bonding

Introduction to Electronic Structure Theory

Chemistry 2000 Lecture 1: Introduction to the molecular orbital theory

Electrons in Crystals. Chris J. Pickard

1 Reduced Mass Coordinates

ECE 535 Theory of Semiconductors and Semiconductor Devices Fall 2015 Homework # 5 Due Date: 11/17/2015

6.730 Physics for Solid State Applications

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

Chemistry 2. Lecture 1 Quantum Mechanics in Chemistry

ECE440 Nanoelectronics. Lecture 07 Atomic Orbitals

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

Applications of Quantum Theory to Some Simple Systems

CHAPTER 11 MOLECULAR ORBITAL THEORY

Applied Nuclear Physics (Fall 2006) Lecture 3 (9/13/06) Bound States in One Dimensional Systems Particle in a Square Well

Ch 125a Problem Set 1

Symmetry III: Molecular Orbital Theory. Reading: Shriver and Atkins and , 6.10

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

σ u * 1s g - gerade u - ungerade * - antibonding σ g 1s

Outline Spherical symmetry Free particle Coulomb problem Keywords and References. Central potentials. Sourendu Gupta. TIFR, Mumbai, India

Three Most Important Topics (MIT) Today

Structure of diatomic molecules

Electronic Structure of Surfaces

Molecular Orbital Theory

ECE 487 Lecture 6 : Time-Dependent Quantum Mechanics I Class Outline:

Energy bands in two limits

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

1 One-dimensional lattice

Nearly Free Electron Gas model - II

ECE 487 Lecture 5 : Foundations of Quantum Mechanics IV Class Outline:

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

Quantum Condensed Matter Physics

Chemistry 3502 Physical Chemistry II (Quantum Mechanics) 3 Credits Fall Semester 2003 Christopher J. Cramer. Lecture 25, November 5, 2003

The Quantum Heisenberg Ferromagnet

7.1 Variational Principle

Introduction to Quantum Mechanics and Spectroscopy 3 Credits Fall Semester 2014 Laura Gagliardi. Lecture 27, December 5, 2014

Density Functional Theory: from theory to Applications

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 17, March 1, 2006

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

Electronic Structure Models

Electronic structure of correlated electron systems. Lecture 2

Lecture 10. Central potential

Lecture 3, January 9, 2015 Bonding in H2+

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

Optical Lattices. Chapter Polarization

The Simple Harmonic Oscillator

A. F. J. Levi 1 EE539: Engineering Quantum Mechanics. Fall 2017.

Tight Binding Method: Linear Combination of Atomic Orbitals (LCAO)

From molecules to solids

The 3 dimensional Schrödinger Equation

Lecture 4 Quantum mechanics in more than one-dimension

Lecture 10. Born-Oppenheimer approximation LCAO-MO application to H + The potential energy surface MOs for diatomic molecules. NC State University

Quantum mechanics can be used to calculate any property of a molecule. The energy E of a wavefunction Ψ evaluated for the Hamiltonian H is,

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

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

The broad topic of physical metallurgy provides a basis that links the structure of materials with their properties, focusing primarily on metals.

Basis 4 ] = Integration of s(t) has been performed numerically by an adaptive quadrature algorithm. Discretization in the ɛ space

Ψ g = N g (1s a + 1s b ) Themes. Take Home Message. Energy Decomposition. Multiple model wave fns possible. Begin describing brain-friendly MO

Molecular Structure Both atoms and molecules are quantum systems

Section 10: Many Particle Quantum Mechanics Solutions

Quantum Mechanics: The Hydrogen Atom

Topic 9-1: Bloch Theorem and the Central Equation Kittel Pages:

Lecture 12. The harmonic oscillator

2m 2 Ze2. , where δ. ) 2 l,n is the quantum defect (of order one but larger

Introduction to DFTB. Marcus Elstner. July 28, 2006

r 2 dr h2 α = 8m2 q 4 Substituting we find that variational estimate for the energy is m e q 4 E G = 4

Lecture 19: Building Atoms and Molecules

(1/2) M α 2 α, ˆTe = i. 1 r i r j, ˆV NN = α>β

X-Ray transitions to low lying empty states

Chemistry 120A 2nd Midterm. 1. (36 pts) For this question, recall the energy levels of the Hydrogenic Hamiltonian (1-electron):

We can model covalent bonding in molecules in essentially two ways:

1.1 A Scattering Experiment

We start with some important background material in classical and quantum mechanics.

we have to deal simultaneously with the motion of the two heavy particles, the nuclei

Nearly Free Electron Gas model - I

CHEM-UA 127: Advanced General Chemistry I

Lecture 19: Building Atoms and Molecules

P3317 HW from Lecture and Recitation 7

+E v(t) H(t) = v(t) E where v(t) is real and where v 0 for t ±.

Lecture 10. Transition probabilities and photoelectric cross sections

Schrödinger equation for central potentials

Lecture 4 Quantum mechanics in more than one-dimension

Approximation Methods in QM

Transcription:

Lecture 6 Tight-binding model In Lecture 3 we discussed the Krönig-Penny model and we have seen that, depending on the strength of the potential barrier inside the unit cell, the electrons can behave like in a FEG (in the limit µ = 0) or like electrons in atoms (in the limit µ = ). In Lecture 4 we introduced the nearly-free electron model, which describes reasonably situations where the electrons feel a weak tal potential. In this lecture we want to look at the other end of the spectrum, where µ is large but not infinite. This corresponds to a situation where electrons are tightly bound to their nuclei. While in the FEG the starting point are travelling electron waves, in the tight binding (TB) model the starting point are atomic wavefunctions. Figure 6.1 illustrates schematically this conceptual difference. In order to understand TB model we start with the H 2 molecule and then we move on to the simplest 3D model of a solid, the cubium. Figure 6.1: Schematic view of the theoretical models that we can use to describe solids depending on the strength of the tal potential. For very weak potentials the electrons are essentially travelling waves and we can use the NFEG approximation. For very strong potentials it could be more convenient, at least conceptually, to start from an atomic-like description as given by the tight-binding approximation. While in the NFEG approximation we write the wavefunctions as linear combinations of planewaves, in TB we write the wavefunctions as linear combinations of atomic wavefunctions.

48 6 Tight-binding model 6.1 Hydrogen molecule We start by considering the H atom. The Schrödinger equation for the ground state 1s wavefunction is: 2 φ 1s (r) + v(r)φ 1s (r) = E 1s φ 1s (r), (6.1) with E 1s = 13.6 ev, e2 v(r) = 4πɛ 0 r, (6.2) φ 1s (r) = 1 π a 3/2 0 e r /a 0, (6.3) and a 0 = 0.53 A. The hydrogen molecule is formed by putting together two H atoms. If we have atom A and atom B in the positions A and B, respectively, the total potential seen by the electrons is: V (r) = v(r A ) + v(r B ), (6.4) and the Schrödinger equation becomes: 2 ψ(r) + V (r)ψ(r) = Eψ(r), (6.5) 2 ψ(r) + [v(r A ) + v(r B )]ψ(r) = Eψ(r). (6.6) The tight-binding approximation consists of assuming that the electrons in this molecule remain tightly bound to their nuclei. If this is the case, then we can find a solution which is very similar to the two original 1s wavefunctions. The simplest guess is to look for a solution such as: ψ(r) = c A φ 1s (r A ) + c B φ 1s (r B ), (6.7) i.e. a linear combination of atomic orbitals (LCAO). In practice, while in the NFEG we were looking for a solution which is a linear combination of planewaves, here we use a linear combination of local orbitals: we expand the wavefunctions in a localized basis set. In order to find the wavefunctions of the H 2 molecule we now replace Eq. (6.7) inside Eq. (6.5). Before doing this, however, it is convenient to simplify the notation. We use the following definitions: φ A (r) = φ 1s (r A ), (6.8) φ B (r) = φ 1s (r B ), (6.9) v A (r) = v(r A ), (6.10) v B (r) = v(r B ). (6.11)

Hydrogen molecule 49 We can now perform the substitution and obtain: 2 (c A φ A + c B φ B ) + (v A + v B )(c A φ A + c B φ B ) = E(c A φ A + c B φ B ). (6.12) From Eq. (6.1) we already know that: 2 φ A + v A φ A = E 1s φ A, (6.13) 2 φ B + v B φ B = E 1s φ B, (6.14) therefore we can use these equalities to simplify Eq. (6.12): c A (E 1s E + v B )φ A + c B (E 1s E + v A )φ B = 0. (6.15) In order to determine the coefficients c A and c B we need two equations. These equations can be found as follows: we multiply both sides of Eq. (6.15) by φ A and integrate over the whole space. This gives us the first equation relating c A and c B. The second equation can be found by multiplying both sides of Eq. (6.12) by φ B and integrating over the whole space. The integration leads to the following equations: c A [(E 1s E)S AA + γ] + c B [(E 1s E)S AB + β] = 0, (6.16) having defined: S AA = S AB = γ = β = dr φ A (r) 2, (6.17) dr φ A (r)φ B (r), (6.18) dr φ A 2 v B (r), (6.19) dr φ A (r)v A (r)φ B (r). (6.20) Now S AA = 1 since the H wavefunction is normalized. S AB is called the overlap integral, and can be neglected because the H wavefunctions decay exponentially fast. The term γ is called the tal field shift, and we are going to neglect it because the 1s wavefunction of the atom A is negligibly small on the nuclear site of atom B. Finally the term β is called the bond integral, and cannot be neglected since the 1/r potential on site A diverges, hence the overlap between v A φ A and φ B is not small. We note also that β < 0 because both φ A and φ B are positive while the potential v A is negative. Taking into account the above considerations we can simplify Eq. (6.16) as follows: c A (E 1s E) + β c B = 0. (6.21)

50 6 Tight-binding model The second equation that we need for determining the coefficients c A and c B can be obtained similarly and reads: β c A + (E 1s E)c B = 0. (6.22) Be rewriting the previous two equations in matrix notation we obatin: [ ] [ ] E1s E β ca = 0. (6.23) β E 1s E The solutions of the secular equation (E 1s E) 2 = β 2 are therefore: and the final wavefunctions can be written as: c B E b = E 1s β, E a = E 1s + β, (6.24) ψ b (r) = 1 2 [φ 1s (r A ) + φ 1s (r B )], (6.25) ψ a (r) = 1 2 [φ 1s (r A ) φ 1s (r B )], (6.26) where the prefactor comes from the normalization condition. The lowest energy solution is the bonding combination of the two hydrogenic wavefunctions, while the highest energy solution is the antibonding combination. The difference between these two solutions is illustrated in Fig. 6.2. Figure 6.2: Electron eigenstates of the H 2 molecule within the tight-binding approximation. The bonding state piles up electronic charge in the middle of the H-H bond and lowers the total energy of the system. An electron in this bonding state stabilizes the molecule through the formation of a covalent bond. The antibonding state piles up electronic charge around the nuclear sites, thereby increasing the total energy of the system.

Cubium 51 6.2 Cubium The cubium is a toy-model of a material which consists of a simple cubic lattice with one s orbital per site. For instance we can imagine an ideal sc lattice of H atoms. The Schrödinger equation for this system is: 2 ψ(r) + V (r)ψ(r) = Eψ(r). (6.27) In analogy with the H 2 molecule the potential is given by the sum of the individual atomic potentials: V (r) = v(r ), (6.28) where the sum runs over all the direct lattice vectors = a(n x u x + n y u y + n z u z ). As in the case of the H 2 molecule we look for a solution given by a linear combination of atomic orbitals: ψ(r) = c() φ 1s (r ), (6.29) with φ 1s (r) given by Eq. (6.3). In the case of the H 2 molecule we found the eigenstates and eigenvalues by performing a diagonalization. In this case the matrix would be huge, because we have as many coefficients c() as the number of lattice sites. It is therefore convenient to take an alternative approach, and start by requiring that our solution for a given wavevector k satisfies the Bloch theorem: ψ k (r + ) = ψ k (r) e ik. (6.30) We now replace Eq. (6.29) in the last equation in order to determine a condition for the coefficients c k () (the following steps are very tedious): c k ( ) φ 1s (r + ) = c k ( ) φ 1s (r ) e ik. [c k ( + ) c k ( ) e ik ] φ 1s (r ) = 0. (6.31) c k ( + ) φ 1s (r ) = c k ( ) φ 1s (r )e ik. The last equality must be satisfied for every r, therefore the only possibility is that the term between square brackets vanishes identically: If we rewrite this for the particular case = 0 we obtain: c k ( + ) = c k ( ) e ik. (6.32) c k () = c k (0) e ik. (6.33)

52 6 Tight-binding model We can now rewrite our Bloch wavefunction Eq. (6.29) as follows: ψ k (r) = c k (0) e ik φ 1s (r ), = e ik φ 1s (r ), (6.34) having set c k (0) = 1 in order to have the wavefunction normalized in the unit cell. At this point we have found the eigenstates ψ k (r) and we only need to find the associated eigenvalues. In order to simplify the notation it is convenient to introduce the Hamiltonian operator as follows: Ĥ = 2 + v(r ). (6.35) Using this notation Eq. (6.27) can be rewritten in a more compact form as: Ĥ(r)ψ k (r) = E k ψ k (r). (6.36) In order to find the eigenvalues E k we multiply both sides of the last equation by ψk (r) and integrate over the tal volume, obtaining: dr ψ k(r)ĥ(r)ψ k(r) = E k dr ψk(r)ψ k (r) = N c E k, (6.37) N c being the number of unit cells in the tal. The last equality follows from the normalization of the wavefunctions. We can now replace Eq. (6.34) inside Eq. (6.37) (the following steps are also very tedious): E k = 1 dr ψ N k(r)ĥ(r)ψ k(r) c = 1 e ik ( ) dr φ 1s (r N )Ĥ(r) φ 1s(r ) c = 1 N c = 1 N c = e ik e ik ( ) e ik ( ) dr φ 1s (r )Ĥ(r ) φ 1s [r + ( )] dr φ 1s (r)ĥ(r) φ 1s[r ( )] dr φ 1s (r)ĥ(r) φ 1s(r ). (6.38) In the second and third steps we made use of the fact that the Hamiltonian operator is translationally invariant (because the kinetic energy and the potential are both translationally invariant). The integral in the last equation needs to be computed by writing

Cubium 53 explicitly the Hamiltonian operator: dr φ 1s (r)ĥ(r) φ 1s(r ) = [ = dr φ 1s (r) 2 + ] v(r ) φ 1s (r ) ] = dr φ 1s (r) [ 2 + v(r) φ 1s (r ) + dr φ 1s (r)v(r ) φ 1s (r ) 0 = E 1s dr φ 1s (r)φ 1s (r ) + dr φ 1s (r)v(r ) φ 1s (r ). 0 = E 1s δ,0 + dr φ 1s (r)v(r ) φ 1s (r ), (6.39) 0 with δ,0 = 1 if = 0, and 0 otherwise. At this stage we need to introduce the tight-binding approximations, in analogy with the case of the H 2 molecule. When = 0 the integral dr φ 1s(r)v(r ) φ 1s (r ) corresponds to the tal field shift (cf. Sec. 6.1). Since 0 in the sum, we can neglect this term because φ 1s (r) is localized around = 0. Now we are left with the case 0 and 0. If = we have a bond integral and we call it β(). If and both are nonzero, then φ 1s (r), v(r ), and φ 1s (r ) are located at three different lattice sites. This term is called a three-centre integral. Since the atomic orbitals are localized, the product of these three functions is always very small and we can neglect this term. By combining these results we have: dr φ 1s (r)ĥ(r) φ 1s(r ) = E 1s δ,0 + β()(1 δ,0 ). (6.40) We can finally rewrite Eq. (6.38) using this result: E k = e ik [E 1s δ,0 + β()(1 δ,0 )]. = E 1s + 0 e ik β(). (6.41) With reference to Fig. 6.3 we now make the approximation that only the bond integrals β() between nearest neighbor sites are non negligible. This means that the only sites we need to consider are ±au x, ±au y, and ±au z. The bond integrals between nearest neighbor sites are all the same by symmetry, and we will call them simply β. Within

54 6 Tight-binding model Figure 6.3: 2D representation of a simple cubic lattice. In the tight-binding approximation we keep only the bond integrals between nearest-neighbor lattice sites (NN in the figure). The rationale for this choice is that atomic wavefunctions are localized around their nuclei and decay exponentially with the distance. Sometimes it may be necessary to include interactions up to the next-nearest neighbors (NNN in the figure), and even possibly further neighbors, depending on the application. such approximation the energy vs momentum dispersion relations in Eq. (6.41) simplifies to: E k = E 1s + β[e i(k ux)a + e i(k ux)a + e i(k uy)a + e i(k uy)a + e i(k uz)a + e i(k uz)a ] = E 1s + β[e ikxa + e ikxa + e ikya + e ikya + e ikza + e ikza ] = E 1s + 2β(cos k x a + cos k y a + cos k z a). (6.42) It is useful to remember that β < 0 and rewrite this equation as E k = E 1s 2 β (cos k x a + cos k y a + cos k z a). (6.43) Equation (6.43) gives the tight-binding s-band for a simple cubic lattice. Figure 6.4 shows the band structure of cubium and the associated density of states.

Cubium 55 Figure 6.4: Band structure and DOS of cubium. The wavevector k is in units of π/a. The energy is given in dimensionless units (E E 1s )/ β. The highest energy state is found for k = (1, 1, 1)π/a and corresponds to the wavefunction ψ 111 (r) = eiπ(nx+ny+nz ) φ 1s (r ). The phase factor exp[iπ(n x +n y +n z )] alternates in sign when we move across nearest-neighbor lattice sites, therefore the state ψ 111 corresponds to a checkerboard configuration of φ 1s (r) and φ 1s (r) functions on each lattice site. The lowest energy state is found for k = (0, 0, 0) and corresponds to the wavefunction ψ 000 (r) = φ 1s(r ). Therefore in the state ψ 000 all the φ 1s (r) functions carry the same sign. The width of the band is 12 β and gives a measure of the bond integral β.