Lecture 15 From molecules to solids

Similar documents
From molecules to solids

Structure of diatomic molecules

14. Structure of Nuclei

Chemistry 483 Lecture Topics Fall 2009

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

Lecture 4 Quantum mechanics in more than one-dimension

Lecture 4 Quantum mechanics in more than one-dimension

2 Electronic structure theory

Atomic Structure and Atomic Spectra

Electron States of Diatomic Molecules

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,

Chemistry 881 Lecture Topics Fall 2001

ECE440 Nanoelectronics. Lecture 07 Atomic Orbitals

VALENCE Hilary Term 2018

Atoms, Molecules and Solids (selected topics)

CHAPTER 11 MOLECULAR ORBITAL THEORY

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

Atoms, Molecules and Solids (selected topics)

Lecture 9 Electronic Spectroscopy

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

Lecture 19: Building Atoms and Molecules

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

Lecture 4: Basic elements of band theory

r R A 1 r R B + 1 ψ(r) = αψ A (r)+βψ B (r) (5) where we assume that ψ A and ψ B are ground states: ψ A (r) = π 1/2 e r R A ψ B (r) = π 1/2 e r R B.

Lecture 6. Tight-binding model

Molecular orbitals, potential energy surfaces and symmetry

Quantum Condensed Matter Physics

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

Preliminary Quantum Questions

Introduction to Electronic Structure Theory

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

Molecular Physics. Attraction between the ions causes the chemical bond.

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

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

On the Uniqueness of Molecular Orbitals and limitations of the MO-model.

Lecture 19: Building Atoms and Molecules

Chem 442 Review of Spectroscopy

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

Introduction to Computational Chemistry

Electronic Structure of Surfaces

CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions

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

Intermission: Let s review the essentials of the Helium Atom

eigenvalues eigenfunctions

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

Chapter 10: Multi- Electron Atoms Optical Excitations

P. W. Atkins and R. S. Friedman. Molecular Quantum Mechanics THIRD EDITION

64-311/5: Atomic and Molecular Spectra

221B Lecture Notes Many-Body Problems II Molecular Physics

Quantum Physics II (8.05) Fall 2002 Outline

Chemistry 2000 Lecture 1: Introduction to the molecular orbital theory

UGC ACADEMY LEADING INSTITUE FOR CSIR-JRF/NET, GATE & JAM PHYSICAL SCIENCE TEST SERIES # 4. Atomic, Solid State & Nuclear + Particle

CHEM-UA 127: Advanced General Chemistry I

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

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

MOLECULES. ENERGY LEVELS electronic vibrational rotational

QUANTUM MECHANICS AND MOLECULAR STRUCTURE

e L 2m e the Bohr magneton

Quantum Mechanics II Lecture 11 ( David Ritchie

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

Tentative content material to be covered for Exam 2 (Wednesday, November 2, 2005)

Molecular-Orbital Theory

Modern Physics for Frommies IV The Universe - Small to Large Lecture 4

The general solution of Schrödinger equation in three dimensions (if V does not depend on time) are solutions of time-independent Schrödinger equation

Modern Physics for Scientists and Engineers International Edition, 4th Edition

Chapter 1: Chemical Bonding

( ( ; R H = 109,677 cm -1

Electrons in a periodic potential

Exercises 16.3a, 16.5a, 16.13a, 16.14a, 16.21a, 16.25a.

NERS 311 Current Old notes notes Chapter Chapter 1: Introduction to the course 1 - Chapter 1.1: About the course 2 - Chapter 1.

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

X-Ray transitions to low lying empty states

221B Lecture Notes Many-Body Problems I

Luigi Paolasini

Quantum Condensed Matter Physics Lecture 4

Quantum Theory of Many-Particle Systems, Phys. 540

Potential energy, from Coulomb's law. Potential is spherically symmetric. Therefore, solutions must have form

PHYSICAL CHEMISTRY I. Chemical Bonds

Quantum Mechanics Solutions

CHEM J-5 June 2014

Physics 221A Fall 1996 Notes 19 The Stark Effect in Hydrogen and Alkali Atoms

3.024 Electrical, Optical, and Magnetic Properties of Materials Spring 2012 Recitation 8 Notes

The general solution of Schrödinger equation in three dimensions (if V does not depend on time) are solutions of time-independent Schrödinger equation

Approximation Methods in QM

5.62 Physical Chemistry II Spring 2008

Spectrum of atomic hydrogen

Quantum Chemistry. NC State University. Lecture 5. The electronic structure of molecules Absorption spectroscopy Fluorescence spectroscopy

Basic Physical Chemistry Lecture 2. Keisuke Goda Summer Semester 2015

Vibrations and Rotations of Diatomic Molecules

( ) electron gives S = 1/2 and L = l 1

From Atoms to Solids. Outline. - Atomic and Molecular Wavefunctions - Molecular Hydrogen - Benzene

Lecture 4: Nuclear Energy Generation

Principles of Molecular Spectroscopy

CHAPTER 10 Tight-Binding Model

Chemistry 6 (9 am section) Spring Covalent Bonding

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Lecture #21: Hydrogen Atom II

The 3 dimensional Schrödinger Equation

Basic Quantum Mechanics

Transcription:

Lecture 15 From molecules to solids

Background In the last two lectures, we explored quantum mechanics of multi-electron atoms the subject of atomic physics. In this lecture, we will explore how these concepts translate into many-atom systems, from simple molecular structures to solid state. Here we will explore H + 2 and H 2, a simple elemental chain, and a simple two-dimensional crystal (e.g. graphene).

Molecular physics: background A molecule consists of electrons moving in the complex potential set up by the charges of the atomic nuclei and other electrons. Even in classical mechanics, it would be extremely difficult to solve equations of motion. Fortunately, for most purposes, we can treat motion of electrons and nuclei separately, due to their very different masses: Since forces that act on nuclei are comparable to those acting on electrons, their momenta are comparable and their velocities are different. Therefore, (a) in studying the motion of electrons, we can treat the nuclei as being nailed down and (b) in studying nuclear motion (vibrations/rotations) we can assume that electrons adjust instantly to changes in molecular conformation: basis of Born-Oppenheimer approximation.

Born-Oppenheimer approximation Wavefunction of a molecule, Ψ({r n }, {R N }, t) determined by many-body Schrödinger equation [ ˆp 2 n + ] ˆp 2 N + V ({r n }, {R N }) Ψ({r n }, {R N }, t) 2m n e 2m N N = E Ψ({r n }, {R N }, t) Since k.e. of nuclei small as compared to electrons, we may drop N ˆp 2 N 2m N and focus on electron component of wavefunction: Here ψ k, k =0, 1, 2, denote electron wavefunction with nuclei nailed down at positions R a, R b,... As conformation varied, ground state E 0 ({R N }) traces a manifold molecular potential energy; minimum equilibrium structure.

Born-Oppenheimer approximation Wavefunction of a molecule, Ψ({r n }, {R N }, t) determined by many-body Schrödinger equation [ ˆp 2 n + ] ˆp 2 N + V ({r n }, {R N }) ψ k ({r n }, {R N }, t) 2m n e 2m N N E k ({R N }) ψ k ({r n }, {R N }, t) Since k.e. of nuclei small as compared to electrons, we may drop N ˆp 2 N 2m N and focus on electron component of wavefunction: Here ψ k, k =0, 1, 2, denote electron wavefunction with nuclei nailed down at positions R a, R b,... As conformation varied, ground state E 0 ({R N }) traces a manifold molecular potential energy; minimum equilibrium structure.

The H + 2 ion To apply these ideas, consider simplest molecule, the hydrogen ion H + 2 two protons (R a, R b ) and one electron (r). H + 2 is found in abundance in interstellar gas clouds. In Born-Oppenheimer approximation, electron Schrödinger equation: [ ( )] ˆp 2 e2 1 2m e 4πɛ 0 r R a + 1 ψ(r; R a, R b )=Eψ(r; R a, R b ) r R b Although equation can be solved analytically(!), more instructive to look for more general approximation scheme.

The H + 2 ion To apply these ideas, consider simplest molecule, the hydrogen ion H + 2 two protons (R a, R b ) and one electron (r). H + 2 is found in abundance in interstellar gas clouds. In Born-Oppenheimer approximation, electron Schrödinger equation: [ ( )] ˆp 2 e2 1 2m e 4πɛ 0 r R a + 1 ψ(r; R a, R b )=Eψ(r; R a, R b ) r R b Although equation can be solved analytically(!), more instructive to look for more general approximation scheme.

The H + 2 ion [ ˆp 2 2m e e2 4πɛ 0 ( )] 1 r R a + 1 ψ(r; R a, R b )=Eψ(r; R a, R b ) r R b Since Hamiltonian does not have a small parameter, we adopt variational approach. If electron is close to one proton, expect other to exert only a small influence here ψ mirrors hydrogen atomic orbital. Therefore, in seeking ground state, take trial wavefunction that is linear combination of 1s wavefunctions centred on two protons, ψ(r; R a, R b )=αψ a (r; R a )+βψ b (r; R b ), ψ a,b (r) = e r R a,b /a 0 where coefficients α and β are taken as real. (πa 3 0 )1/2

The H + 2 ion ψ(r; R a, R b )=αψ a (r; R a )+βψ b (r; R b ) Variational ground state energy: E = ψ Ĥ ψ ψ ψ = α2 H aa + β 2 H bb +2αβH ab α 2 + β 2 +2αβS where H aa = ψ a Ĥ ψ a = H bb, and H ab = ψ a Ĥ ψ b = H ba. Since ψ a and ψ b not orthogonal, we have to introduce overlap integral, S = ψ a ψ b. Since potential is symmetric, wavefunction must be either symmetric or antisymmetric, i.e. α = ±β, and E 0 E = H aa ± H ab 1 ± S, ψ(r; R a, R b )=α(ψ a ± ψ b )

The H + 2 ion E 0 E = H aa ± H ab 1 ± S, ψ(r; R a, R b )=α(ψ a ± ψ b ) Two possible wavefunctions for H + 2 ion, ψ g = ψ a + ψ b 2(1 + S), ψ u = ψ a ψ b 2(1 S) with energies E g = H aa+h ab 1+S, E u = H aa H ab 1 S. Subscript g (gerade even) used in molecular physics to denote state symmetric under inversion (without exchanging nuclei) The odd (ungerade) state denoted by u. In chemistry, orthogonal states, ψ g and ψ u, are molecular orbitals and general methodology is known as the linear combination of atomic orbitals (LCAO) approach.

The H + 2 ion ψ g = ψ a + ψ b 2(1 + S), ψ u = ψ a ψ b 2(1 S) The state ψ g has lower energy, while ψ u represents an excited state of the molecular ion. For ψ g (ψ u ) state, the two atomic wavefunctions interfere constructively (destructively) in region between protons. For ψ g, enhanced electron density in region where electron is attracted by both protons screens protons from each other.

The H + 2 ion: molecular potential E( R a R b ) As expected, variational approximation provides upper limit on ground state energy. Since E u + Ry does not have a minimum, suggests that odd wavefunction ψ u does not correspond to bound molecular state. E/eV 4 3 2 1 0 1 2 3 E u + Ry 100 200 300 R/pm LCAO E g + Ry Exact E g + Ry To improve approximation, could introduce further orbitals, e.g. at small R wavefunction should approach He +.

The H + 2 ion: remarks Although not very accurate, LCAO approximation for ψ g does exhibit correct features of true bonding wavefunction, σ g : (i) it is even with respect to inversion; (ii) there is constructive interference which leads to enhanced probability of finding electron in the region between nuclei. ψ u is characteristic of anti-bonding state, σ u.

The H 2 molecule At first sight, might expect H 2 is simple extension of H + 2 ion; but several new features arise. In Born-Oppenheimer approximation, for two electrons r 1,2 and two protons, R a,b, Ĥ = Ĥ0 + Ĥ1 where Ĥ 0 = n=1,2 [ ] ˆp 2 n + V (r n ), Ĥ 1 = e2 2m e 4πɛ 0 ( 1 1 ) r 12 r ab ( ) where V (r n )= e2 1 4πɛ 0 r ab 1 r 1a 1 r 1b and r 1b = r 1 R b, etc., i.e. sum of two H + 2 ions (Ĥ0) and an additional term Ĥ1. Since 1 r 12 1 r ab can treat Ĥ1 as a perturbation.

The H 2 molecule Ĥ 0 = n=1,2 [ ] ˆp 2 n + V (r n ) 2m e Neglecting Ĥ1, there are four ways of filling two orbitals σ g and σ u, ψ g (r 1 )ψ g (r 2 ), ψ g (r 1 )ψ u (r 2 ), ψ u (r 1 )ψ g (r 2 ), ψ u (r 1 )ψ u (r 2 ) Of these, expect ψ g (r 1 )ψ g (r 2 ) to be ground state. However, at this stage, we have given no consideration to constraints imposed by particle statistics. Since electrons are identical fermions, total wavefunction must be antisymmetric under exchange. Taking account of spin degrees of freedom, for both electrons to occupy the bonding σ g orbital, they must occupy spin singlet state, X 0,0 = 1 2 (χ + (1)χ (2) χ (1)χ + (2))

The H 2 molecule Ĥ 0 = n=1,2 [ ] ˆp 2 n + V (r n ) 2m e Neglecting Ĥ1, there are four ways of filling two orbitals σ g and σ u, ψ g (r 1 )ψ g (r 2 ), ψ g (r 1 )ψ u (r 2 ), ψ u (r 1 )ψ g (r 2 ), ψ u (r 1 )ψ u (r 2 ) Of these, expect ψ g (r 1 )ψ g (r 2 ) to be ground state. However, at this stage, we have given no consideration to constraints imposed by particle statistics. Since electrons are identical fermions, total wavefunction must be antisymmetric under exchange. Taking account of spin degrees of freedom, for both electrons to occupy the bonding σ g orbital, they must occupy spin singlet state, X 0,0 = 1 2 (χ + (1)χ (2) χ (1)χ + (2))

The H 2 molecule If we compute energy of state ψ g (r 1 )ψ g (r 2 )X 0,0 as a function of R, minimum occurs at R 0 = 85pm with binding energy of 2.7eV but true molecule is smaller and more tightly bound. Allowing for more variation in atomic orbitals, variable effective charge, etc., gives equilibrium R 0 much closer to experiment, but a binding energy that is still not high enough. The reason is that σ 2 g configuration alone is not a very good representation of the ground state. Why...?

The H 2 molecule In the LCAO approximation, σ 2 g wavefunction has a strange form, ψ g (r 1 )ψ g (r 2 ) [ψ a (r 1 )ψ b (r 2 )+ψ b (r 1 )ψ a (r 2 )] +[ψ a (r 1 )ψ a (r 2 )+ψ b (r 1 )ψ b (r 2 )] First term involves two electrons shared between two hydrogen atoms a covalent bond Second term involves both electrons assigned to same atom an ionic bond. Since equal coefficients, ionic and covalent contributions are equal when protons pulled apart, ground state just as likely to consist of H + and H ion as two neutral atoms! implausible. If we drop ionic part of wavefunction valence bonding approximation binding energy and nuclear separation improved. Including a variational parameter for amplitude of ionic component, find optimal value λ 1/6 only ca. 3% ionic.

The H 2 molecule In the LCAO approximation, σ 2 g wavefunction has a strange form, ψ g (r 1 )ψ g (r 2 ) [ψ a (r 1 )ψ b (r 2 )+ψ b (r 1 )ψ a (r 2 )] +[ψ a (r 1 )ψ a (r 2 )+ψ b (r 1 )ψ b (r 2 )] First term involves two electrons shared between two hydrogen atoms a covalent bond Second term involves both electrons assigned to same atom an ionic bond. Since equal coefficients, ionic and covalent contributions are equal when protons pulled apart, ground state just as likely to consist of H + and H ion as two neutral atoms! implausible. If we drop ionic part of wavefunction valence bonding approximation binding energy and nuclear separation improved. Including a variational parameter for amplitude of ionic component, find optimal value λ 1/6 only ca. 3% ionic.

The H 2 molecule In the LCAO approximation, σ 2 g wavefunction has a strange form, ψ g (r 1 )ψ g (r 2 ) [ψ a (r 1 )ψ b (r 2 )+ψ b (r 1 )ψ a (r 2 )] +[ψ a (r 1 )ψ a (r 2 )+ψ b (r 1 )ψ b (r 2 )] First term involves two electrons shared between two hydrogen atoms a covalent bond Second term involves both electrons assigned to same atom an ionic bond. Since equal coefficients, ionic and covalent contributions are equal when protons pulled apart, ground state just as likely to consist of H + and H ion as two neutral atoms! implausible. If we drop ionic part of wavefunction valence bonding approximation binding energy and nuclear separation improved. Including a variational parameter for amplitude of ionic component, find optimal value λ 1/6 only ca. 3% ionic.

The H 2 molecule In the LCAO approximation, σ 2 g wavefunction has a strange form, ψ g (r 1 )ψ g (r 2 ) [ψ a (r 1 )ψ b (r 2 )+ψ b (r 1 )ψ a (r 2 )] +[ψ a (r 1 )ψ a (r 2 )+ψ b (r 1 )ψ b (r 2 )] First term involves two electrons shared between two hydrogen atoms a covalent bond Second term involves both electrons assigned to same atom an ionic bond. Since equal coefficients, ionic and covalent contributions are equal when protons pulled apart, ground state just as likely to consist of H + and H ion as two neutral atoms! implausible. If we drop ionic part of wavefunction valence bonding approximation binding energy and nuclear separation improved. Including a variational parameter for amplitude of ionic component, find optimal value λ 1/6 only ca. 3% ionic.

The H 2 molecule In the LCAO approximation, σ 2 g wavefunction has a strange form, ψ g (r 1 )ψ g (r 2 ) [ψ a (r 1 )ψ b (r 2 )+ψ b (r 1 )ψ a (r 2 )] +[ψ a (r 1 )ψ a (r 2 )+ψ b (r 1 )ψ b (r 2 )] First term involves two electrons shared between two hydrogen atoms a covalent bond Second term involves both electrons assigned to same atom an ionic bond. Since equal coefficients, ionic and covalent contributions are equal when protons pulled apart, ground state just as likely to consist of H + and H ion as two neutral atoms! implausible. If we drop ionic part of wavefunction valence bonding approximation binding energy and nuclear separation improved. Including a variational parameter for amplitude of ionic component, find optimal value λ 1/6 only ca. 3% ionic.

The H 2 molecule In the LCAO approximation, σ 2 g wavefunction has a strange form, ψ g (r 1 )ψ g (r 2 ) [ψ a (r 1 )ψ b (r 2 )+ψ b (r 1 )ψ a (r 2 )] + λ[ψ a (r 1 )ψ a (r 2 )+ψ b (r 1 )ψ b (r 2 )] First term involves two electrons shared between two hydrogen atoms a covalent bond Second term involves both electrons assigned to same atom an ionic bond. Since equal coefficients, ionic and covalent contributions are equal when protons pulled apart, ground state just as likely to consist of H + and H ion as two neutral atoms! implausible. If we drop ionic part of wavefunction valence bonding approximation binding energy and nuclear separation improved. Including a variational parameter for amplitude of ionic component, find optimal value λ 1/6 only ca. 3% ionic.

From molecules to solids Having established basic principles of molcular structure, we now consider how methodology can be applied to crystalline solids. In the Born-Oppenheimer approximation, for elemental solid, Ĥ = n ˆp n 2m e e2 4πɛ 0 n,n Z r nn m<n In the physical system, we would have to take into account the influence of relativistic corrections (viz. spin-orbit interaction). To address the properties of such a complex interacting system, we will have to draw upon many of the insights developed previously. 1 r mn

From molecules to solids To proceed, helpful to partition electrons into those which are bound to core and those which are unbound. Tightly-bound electrons screen charge leading to a modified nuclear potential, V eff (r). Focussing on those electrons which are free (itinerant), Ĥ n Ĥ n + e2 4πɛ 0 m<n 1 r mn where Ĥ n = ˆp2 n 2m e + V eff (r n ) represents single-particle Hamiltonian experienced by each electron. Ĥ n describes motion of an electron in a periodic lattice potential, V eff (r) =V eff (r + R) with R belonging to set of lattice vectors. If electrons remain itinerant, they screen each other and diminish the effect of electron-electron interaction.

From molecules to solids Droping Coulomb interaction between electrons, we can apply molecular orbital theory using variational LCAO scheme: i.e. build trial wavefunction by combining orbital states of single ion, V ion (r), where V eff (r) = R V ion(r). As with hydrogen molecule, Hamiltonian for individual nuclei, Ĥ 0 = ˆp 2 2m e + V ion (r), associated with a set of atomic orbitals, ψ q, with quantum numbers, q. In atomic limit when atoms are far-separated, these states mirror simple hydrogenic wavefunctions. To find variational ground state of the system, we can then build trial state from a linear combination of these atomic orbitals. Taking only lowest orbital q = 0 into account, ψ(r) = R α R ψ(r R) where, as before, α R represent set of variational coefficients.

From molecules to solids Droping Coulomb interaction between electrons, we can apply molecular orbital theory using variational LCAO scheme: i.e. build trial wavefunction by combining orbital states of single ion, V ion (r), where V eff (r) = R V ion(r). As with hydrogen molecule, Hamiltonian for individual nuclei, Ĥ 0 = ˆp 2 2m e + V ion (r), associated with a set of atomic orbitals, ψ q, with quantum numbers, q. In atomic limit when atoms are far-separated, these states mirror simple hydrogenic wavefunctions. To find variational ground state of the system, we can then build trial state from a linear combination of these atomic orbitals. Taking only lowest orbital q = 0 into account, ψ(r) = R α R ψ(r R) where, as before, α R represent set of variational coefficients.

From molecules to solids Variational state energy, E = ψ Ĥ0 ψ ψ ψ = R,R α R H RR α R R,R α R S RR α R where H RR = d d r ψ (r R)Ĥ0ψ(r R ) denote matrix elements of orbital wavefunction and S RR = d d r ψ (r R)ψ(r R ) represent overlap integrals. Minimizing energy with respect to αr, obtain secular equation, If basis functions were orthogonal, this would just be an eigenvalue equation.

From molecules to solids Variational state energy, E = ψ Ĥ 0 ψ ψ ψ = R,R α R H RR α R R,R α R S RR α R where H RR = d d r ψ (r R)Ĥ0ψ(r R ) denote matrix elements of orbital wavefunction and S RR = d d r ψ (r R)ψ(r R ) represent overlap integrals. Minimizing energy with respect to αr, obtain secular equation, δe R H = RR α R R,R α R,R α R S R H RR α R R S RR α R RR α R R,R α R S RR α R R,R α R S RR α R δα R If basis functions were orthogonal, this would just be an eigenvalue equation.

From molecules to solids Variational state energy, E = ψ Ĥ0 ψ ψ ψ = R,R α R H RR α R R,R α R S RR α R where H RR = d d r ψ (r R)Ĥ 0 ψ(r R ) denote matrix elements of orbital wavefunction and S RR = d d r ψ (r R)ψ(r R ) represent overlap integrals. Minimizing energy with respect to αr, obtain secular equation, δe R H = RR α R R S R,R α R S E RR α R RR α R R,R α R S RR α R δα R If basis functions were orthogonal, this would just be an eigenvalue equation.

From molecules to solids Variational state energy, E = ψ Ĥ0 ψ ψ ψ = R,R α R H RR α R R,R α R S RR α R where H RR = d d r ψ (r R)Ĥ 0 ψ(r R ) denote matrix elements of orbital wavefunction and S RR = d d r ψ (r R)ψ(r R ) represent overlap integrals. Minimizing energy with respect to αr, obtain secular equation, δe R [H = RR ES RR ] α R! R,R α R S =0 RR α R δα R If basis functions were orthogonal, this would just be an eigenvalue equation.

From molecules to solids Variational state energy, E = ψ Ĥ0 ψ ψ ψ = R,R α R H RR α R R,R α R S RR α R where H RR = d d r ψ (r R)Ĥ0ψ(r R ) denote matrix elements of orbital wavefunction and S RR = d d r ψ (r R)ψ(r R ) represent overlap integrals. Minimizing energy with respect to αr, obtain secular equation, R [H RR ES RR ] α R =0 If basis functions were orthogonal, this would just be an eigenvalue equation.

Example: periodic one-dimensional system R [H RR ES RR ] α R =0 If atoms are well-separated, overlap integrals and matrix elements decay exponentially with separation. Dominant contribution then derives from coupling neighbouring states Hückel approximation. In 1d, secular equation: (ε E)α n (t + ES)(α n+1 + α n 1 )=0, for each n where H nn = ε is atomic orbital energy, H n,n+1 = H n+1,n = t < 0 denotes matrix element between neighbouring states, S n,n = 1 and S n+1,n = S n+1,n = S.

Example: periodic one-dimensional system (ε E)α n (t + ES)(α n+1 + α n 1 )=0 cf. equation of motion of discrete classical N atom chain : m φ n = k s (φ n+1 φ n )+k s (φ n φ n 1 ), φ n+n = φ n with spring constant k s and masses m. With φ n (t) =e iωt φ n, (mω 2 +2k s )φ n k s (φ n+1 + φ n 1 )=0 Normal vibrational modes: φ n = 1 N e ikna where k = 2πm Na denote N discrete reciprocal lattice vectors with N/2 < m N/2.

Example: periodic one-dimensional system (ε E)α n (t + ES)(α n+1 + α n 1 )=0 For N lattice sites and periodic boundary condition, α n+n = α n, solution given by α n = 1 N e ikna, E = E k = ε 2t cos(ka) 1+S cos(ka) i.e. reciprocal lattice vector k parameterizes a band of electronic states.

Example: periodic one-dimensional system According to LCAO approximation, for a single electron, lowest energy state predicted to be uniform α n = 1 N with E 0 = ε 2t 1+S. For more than one electron, we must consider influence of Pauli exclusion and particle indistinguishability. Since electrons are identical fermions, each state k can host a maximum of two electrons in a spin singlet configuration. Lowest energy state obtained by adding electrons sequentially into states of increasing k. If maximum k value Fermi level lies within band, excitations cost vanishingly small energy metal. If each atom contributes an even integer number of electrons, Fermi wavevector may lie at a band gap band insulator.

Example: graphene Recently, much attention has been paid to graphene, a single layer of graphite. Flakes of graphene can be prepared by running graphite a pencil! over adhesive layer. Resulting electron states of single layer compound have been of enormous interest to physicists. To understand why, let us implement LCAO technology to explore electronic structure of graphene.

Example: graphene Graphene forms periodic two-dimensional honeycomb lattice structure with two atoms per unit cell. With electron configuration (1s 2 )(2s 2 )(2p 2 ), two 1s electrons are bound tightly to nucleus. 2s and 2p orbitals hybridize into three sp 2 orbitals which form covalent σ bonding orbitals and constitute honeycomb lattice. Remaining electrons (1/atom), which occupy out-of-plane p z orbital, is then capable of forming an itinerant band of electron states. It is this band which we now address.

Example: graphene Suppose that wavefunction of band involves the basis of p z orbitals, ψ(r) = R [α R ψ 1 (r R)+β R ψ 2 (r R)] Taking into account matrix elements involving only nearest neighbours, trial wavefunction translates to secular equation, (ε E)α R (t + ES)(β R + β R a1 + β R a2 ) = 0 (ε E)β R (t + ES)(α R + α R+a1 + α R+a2 ) = 0 where the primitive lattice vectors a 1 =( 3/2, 1/2)a and a 2 =( 3/2, 1/2)a, with a the lattice spacing.

Example: graphene (ε E)α R (t + ES)(β R + β R a1 + β R a2 ) = 0 (ε E)β R (t + ES)(α R + α R+a1 + α R+a2 ) = 0 With the plane wave ansatz, α R = α k N e ik R and β R = β k N e ik R (ε E)α k (t + ES)f k β k =0 (ε E)β k (t + ES)f k α k =0 where f k = 1 + 2e i 3k x a/2 cos(k y a/2). If we neglect overlap integral S (for simplicity), E = E k = ε ± f k t

Example: graphene electronic structure At half-filling, where each atom contributes one electron to band, Fermi level lies precisely at centre where dispersion, E k is point-like.

Example: graphene Doping electrons into (or removing electrons from) the system results in (two copies) of a linear dispersion, E k c k, where c is a constant (velocity). Such a linear dispersion relation is the hallmark of a relativistic particle (cf. a photon). Although electrons are not moving at relativistic velocities, their properties mirror behaviour of relativistic particles.

From molecules to solids But what about (neglected) electron-electron interactions? In principle, we could develop Hartree-Fock scheme to address effects on band structure in perturbation theory. However, it is a surprising, yet robust, feature of Fermi systems that properties of non-interacting ground state remain qualitatively correct over an unreasonably wide range. Rigidity can be attributed to constraints implied by nodal structure of wavefunction encapsulated by Landau s Fermi liquid theory. However, electron interactions can induce striking modifications in ground state reflected in novel experimental behaviour Electron localization Mott transition; local moment and itinerant magnetism; quantum Hall fluids; and superconductivity. Such phases, which by their nature, lie outside any perturbative scheme built around the non-interacting ground state, underpin the field of modern quantum condensed matter and solid state physics.

Molecular spectra Having established a basic formalism to describe molecular structure, we turn now to consider excitations and (radiative) transitions. These can include transitions between electron states (typically O(eV ) optical) as well as rotational/vibrational excitations. Usually electronic transition induces motion of nuclei as well. Energies of rotational states, O( 2 /2I ), are much smaller than those of electron excitations. Typical rotational energies are O(10 4 ev) (far IR) and vibrational excitations are O(10 1 ev) (cf. Greenhouse effect ). All three types of transitions can occur radiatively, i.e. through emission or absorption of a photon of frequency ν = E/h.

Molecular spectra: transitions As with atoms, most probable radiative transitions are electric dipole, i.e. usual selection rule applies: J =0, ±1, but not 0 0 accompanied by change in the parity of the molecular state. In a gas or liquid, transitions can also be produced by collisions. Such non-radiative transitions do not have to obey selection rules, i.e. molecule in a metastable state can be de-excited by collision. Collisions lead to thermal distribution of molecular energy levels, n i g i exp [ E i k B T where g i is degeneracy and E i the energy. At room temperature, k B T 2 10 2 ev, many rotational states of molecules are excited, but not electronic or vibrational. ]

Molecular spectra: transitions As with atoms, most probable radiative transitions are electric dipole, i.e. usual selection rule applies: J =0, ±1, but not 0 0 accompanied by change in the parity of the molecular state. In a gas or liquid, transitions can also be produced by collisions. Such non-radiative transitions do not have to obey selection rules, i.e. molecule in a metastable state can be de-excited by collision. Collisions lead to thermal distribution of molecular energy levels, n i g i exp [ E i k B T where g i is degeneracy and E i the energy. At room temperature, k B T 2 10 2 ev, many rotational states of molecules are excited, but not electronic or vibrational. ]

Born-Oppenheimer approximation revisited How can rotational/vibrational excitation spectrum be computed? Armed with Born-Oppenheimer approximation for electron states, ψ k, we can exploit completeness to express full stationary state as Ψ({r n }, {R N })= k φ k ({R N })ψ k ({r n }, {R N }) where φ k represents the nuclear part of the wavefunction. Substituting into full time-independent Schrödinger equation,

Born-Oppenheimer approximation revisited How can rotational/vibrational excitation spectrum be computed? Armed with Born-Oppenheimer approximation for electron states, ψ k, we can exploit completeness to express full stationary state as Ψ({r n }, {R N })= k φ k ({R N })ψ k ({r n }, {R N }) where φ k represents the nuclear part of the wavefunction. Substituting into full time-independent Schrödinger equation, ˆp 2 N + 2m N ˆp 2 n + V ({r n }, {R N }) Ψ({r n }, {R N }) 2m e N=a,b,... n=1,2,... = EΨ({r n }, {R N })

Born-Oppenheimer approximation revisited How can rotational/vibrational excitation spectrum be computed? Armed with Born-Oppenheimer approximation for electron states, ψ k, we can exploit completeness to express full stationary state as Ψ({r n }, {R N })= k φ k ({R N })ψ k ({r n }, {R N }) where φ k represents the nuclear part of the wavefunction. Substituting into full time-independent Schrödinger equation, ˆp 2 N + ˆp 2 n + V ({r n }, {R N }) φ k ψ k 2m N 2m e k N=a,b,... n=1,2,... = E k φ k ψ k

Born-Oppenheimer approximation revisited How can rotational/vibrational excitation spectrum be computed? Armed with Born-Oppenheimer approximation for electron states, ψ k, we can exploit completeness to express full stationary state as Ψ({r n }, {R N })= k φ k ({R N })ψ k ({r n }, {R N }) where φ k represents the nuclear part of the wavefunction. Substituting into full time-independent Schrödinger equation, ˆp 2 N + E k ({R N }) φ k ψ k 2m N k N=a,b,... = E k φ k ψ k where E k are the electron energy levels.

Born-Oppenheimer approximation revisited k N=a,b,... ˆp 2 N + E k ({R N }) φ k ψ k = E 2m N k φ k ψ k Now, since dependence of ψ k ({r n }, {R N }) on {R N } is weak compared with that of the nuclear part φ k ({R N }), we can write 2 Nφ k ψ k ψ k 2 Nφ k Using the orthogonality of ψ k, we can pick out the k = 0 term, ψ k 2 2 N + E 0 ({R N }) φ k = E ψ k φ k 2m N k N=a,b,... k i.e. φ 0 satisfies Schrödinger equation in which V ({r n }, {R N }) is replaced by molecular potential energy E 0 ({R N }), i.e. electrons assumed to react instantly to changes in molecular conformation.

Born-Oppenheimer approximation revisited k N=a,b,... ˆp 2 N + E k ({R N }) φ k ψ k = E 2m N k φ k ψ k Now, since dependence of ψ k ({r n }, {R N }) on {R N } is weak compared with that of the nuclear part φ k ({R N }), we can write 2 Nφ k ψ k ψ k 2 Nφ k Using the orthogonality of ψ k, we can pick out the k = 0 term, ψ0 ψ k 2 2 N + E 0 ({R N }) φ k = E ψ0 ψ k φ k 2m N k N=a,b,... i.e. φ 0 satisfies Schrödinger equation in which V ({r n }, {R N }) is replaced by molecular potential energy E 0 ({R N }), i.e. electrons assumed to react instantly to changes in molecular conformation. k

Born-Oppenheimer approximation revisited k N=a,b,... ˆp 2 N + E k ({R N }) φ k ψ k = E 2m N k φ k ψ k Now, since dependence of ψ k ({r n }, {R N }) on {R N } is weak compared with that of the nuclear part φ k ({R N }), we can write 2 Nφ k ψ k ψ k 2 Nφ k Using the orthogonality of ψ k, we can pick out the k = 0 term, N=a,b,... 2 2 N + E 0 ({R N }) φ 0 = Eφ 0 2m N i.e. φ 0 satisfies Schrödinger equation in which V ({r n }, {R N }) is replaced by molecular potential energy E 0 ({R N }), i.e. electrons assumed to react instantly to changes in molecular conformation.

Molecular rotation Schrödinger equation for nuclear motion has many solutions, which give various molecular energy levels for given electron configuration. For diatomic molecules, E 0 (R 1, R 2 )=E 0 (R) where R = R 1 R 2. Separating variables, Schrödinger equation of relative motion, ] [ 2 2µ 2 R + E 0 (R) φ 0 = Eφ 0 where µ = m 1m 2 m 1 +m 2 is reduced mass. E 0 (R) acts as a central potential, and the usual separation into angular and radial equations can be carried out.

Molecular rotation ] [ 2 2µ 2 R + E 0 (R) φ 0 = Eφ 0 In lowest radial state, rotational energy levels described by usual spherical harmonic functions Y J,mJ, with E J = 2 2I J(J + 1) where I = µr 2 0 is moment of inertia of molecule, and R 0 is equilibrium bond length. Since molecular dimensions set by Bohr radius a 0, I m N a 2 0 E J 2 /m N a 2 0. For the electron states, p e /a 0 and electron energies are ca. 2 /m e a 2 0, a factor of m N/m e 10 4 greater.

Molecular rotation: transitions To bring about a radiative rotational transition, an emitted or absorbed photon must interact with the electric dipole moment of the molecule. Since the initial and final electronic states are the same, this state needs to have a permanent electric dipole moment. Therefore, can have rotational radiative transitions in heteronuclear diatomic molecules (e.g. HCl and CO), which have permanent dipole moments, but not in homonuclear (e.g. H 2 and O 2 ). Usual electric dipole selection rules apply: J = ±1, 0; but requirement for parity change excludes J = 0. E J J+1 = 2 2I [(J + 1)(J + 2) J(J + 1)] = 2 I (J + 1)

Molecular vibrations ] [ 2 2µ 2 R + E 0 (R) φ 0 = Eφ 0 For a diatomic molecule, can Taylor expand molecular potential E 0 (R) around the equilibrium nuclear separation R 0, E 0 (R) =E 0 (R 0 )+ 1 2 (R R 0) 2 2 RE 0 R0 +... Leads to approximate harmonic oscillator with ω =( 1 µ 2 R E 0 R0 ) 1/2, E = E 0 (R 0 )+(n +1/2) ω, n =0, 1, 2,... Vibrational excitations typically larger than rotational by m N /m e and smaller than electronic excitations by m e /m N (see notes). Vibrational transitions: dipole selection rule (exercise), n = ±1, i.e. only a single energy in the spectrum, E =(E n+1 E n )= ω

Molecular vibrations ] [ 2 2µ 2 R + E 0 (R) φ 0 = Eφ 0 For a diatomic molecule, can Taylor expand molecular potential E 0 (R) around the equilibrium nuclear separation R 0, E 0 (R) =E 0 (R 0 )+ 1 2 (R R 0) 2 2 RE 0 R0 +... Leads to approximate harmonic oscillator with ω =( 1 µ 2 R E 0 R0 ) 1/2, E = E 0 (R 0 )+(n +1/2) ω, n =0, 1, 2,... Vibrational excitations typically larger than rotational by m N /m e and smaller than electronic excitations by m e /m N (see notes). Vibrational transitions: dipole selection rule (exercise), n = ±1, i.e. only a single energy in the spectrum, E =(E n+1 E n )= ω

Example: Rotation-vibration spectrum of HCl E = ω + 2 I (J + 1)