Quick viewing(Text Mode)

Arxiv:0906.1640V7 [Physics.Chem-Ph] 21 Feb 2010

Arxiv:0906.1640V7 [Physics.Chem-Ph] 21 Feb 2010

An introduction to effective low-energy Hamiltonians in condensed matter physics and chemistry

B. J. Powell Centre for Organic Photonics and Electronics, School of Mathematics and Physics, The University of Queensland, Queensland 4072, Australia∗

I. AIMS AND SCOPE II. A BRIEF INTRODUCTION TO SECOND QUANTISATION NOTATION These lecture notes introduce some simple effective Hamiltonians (also known as semi-empirical models) that The models discussed in these notes are easiest to un- have widespread applications to solid state and molecular derstand if one employs the so-called second quantisation systems. They are aimed as an introduction to a begin- formalism. In this section we briefly and informally in- ning graduate student. I also hope that it may help to troduce this formalism. More details can be found in break down the divide between the physics and chemistry many textbooks, (e.g. Refs. 3 and 4). Readers already literatures. familiar with this notation may wish to skip this section, however, the last two paragraphs do define some nomen- After a brief introduction to second quantisation no- clature that is used throughout these notes. tation (section II), which is used extensively, I focus of the “four H’s”: the H¨uckel (or tight binding; section III), Hubbard (section IV), Heisenberg (section V) and Hol- A. The simple harmonic oscillator stein (section VII) models. These models play central roles in our understanding of condensed matter physics, Let us begin by considering a particle of mass m mov- particularly for materials where electronic correlations ing in a one-dimensional harmonic potential: are important, but are less well known to the chem- istry community. Some other related models, such as the 1 V (x) = kx2. (1) Pariser-Parr-Pople model, the extended , 2 multi-orbital models and the ionic Hubbard model, are This may be familiar as the potential of an ideal spring also discussed in section VI. As well as their practical displaced from its equilibrium position by a distance x, in applications these models allow us to systematically in- which context k is known as the spring constant (5). Eq. vestigate electronic correlations by ‘turning on’ various 1 is also the potential felt by an as it is displaced (by interactions in the Hamiltonian one at a time. Finally, a small amount) from its equilibrium position in molecule in section VIII, I discuss the epistemological basis of ef- (6). Classically this problem is straightforward to solve fective Hamiltonians and compare and contrast this ap- (5) and, as well as the trivial solution, one finds that proach with ab initio methods before discussing the prob- the particle may oscillate with a resonant frequency ω = lem of the parameterisation of effective Hamiltonians. pk/m. The time-independent Schr¨odingerequation for As these notes are intended to be introductory, I will a simple harmonic oscillator is therefore, not attempt to make frequent comparisons to the latest  2  research problems, rather I compare the predictions of pˆ 1 Hˆ ψ ≡ + mω2xˆ2 ψ = E ψ , (2) model Hamiltonians with simple systems chosen for ped- sho n 2m 2 n n n agogical reasons. Likewise, references have been chosen ~ ∂ for their pedagogical and historical value rather than on wherep ˆ = i ∂x is the particle’s momentum and ψn is the the basis of scientific priority. nth wavefunction or , which has energy, or Given the similarity in the problems addressed by eigenvalue, En. theoretical chemistry and theoretical condensed matter This problem is solved in many introductory texts on arXiv:0906.1640v7 [physics.chem-ph] 21 Feb 2010 physics, there are surprisingly few advanced texts dis- (7) using the standard methods of cussing the interface of two subjects. This, unfortu- ‘first quantised’ quantum mechanics. However, a more nately, leads to many cultural differences between the elegant way to solve this problem is to introduce the ‘lad- fields. Nevertheless, some textbooks do try to bridge the der operator’, gap, and the reader in search of more than the introduc- rmω pˆ tory material presented here is referred to Refs. 1 and aˆ ≡ xˆ + i√ (3a) 2. 2~ 2m~ω and its hermitian conjugate r † mω pˆ ∗ aˆ ≡ xˆ − i√ . (3b) Electronic address: [email protected] 2~ 2m~ω One of the most important features of quantum me- B. Second quantisation for light and matter chanics is that momentum and position do not commute (7), i.e., [ˆp, xˆ] ≡ pˆxˆ − xˆpˆ = −i~. From this commutation We can extend the second quantisation formalism to relation it is straightforward to show that light and matter. Let us first consider bosons, which are not subject to the Pauli exclusion principle, e.g.,  1 phonons, photons, deuterium nuclei, 4He , etc. We Hˆ = ω aˆ†aˆ + (4) sho ~ 2 define the bosonic ‘field operator’ ˆb†(r) as creating a bo- son at position r, similarly, ˆb(r) annihilates a boson at and position r. The bosonic field operators obey the commu- tation relations [ˆb(r), ˆb(r0)] = 0, [ˆb†(r), ˆb†(r0)] = 0, and [ˆa, aˆ†] ≡ aˆaˆ† − aˆ†aˆ = 1. (5) [ˆb(r), ˆb†(r0)] = δ(r − r0). (10) ˆ † 1 One can also show that [Hsho, aˆ] = [~ω(ˆa aˆ + 2 ), aˆ] = ω[ˆa†, aˆ]ˆa = − ωaˆ in a similar manner. Therefore This is just the generalisation of Eq. 5 for the field op- ~ ~ erators. We can create any state by acting products, or [Hsho, aˆ]ψn = −~ωaψˆ n and hence sums of products, of the ˆb†(r) on the vacuum state, i.e., Hˆ aψˆ = (E − ω)ˆaψ . (6) the state that does not contain any bosons, which is usu- sho n n ~ n ally denoted as |0i. Many body wavefunctions for fermions, e.g. , Eq. 6 tells us thataψ ˆ is an eigenstate of Hˆ with n sho protons, neutrons, 3He atoms, etc., are complicated by energy E − ω, providedaψ ˆ 6= 0. That is, the operator n ~ n the need for the antisymmetrisation of the wavefunction, aˆ moves the system from one eigenstate to another whose i.e., the wavefunction must change sign under the ex- energy is lower by ω, thusa ˆ is known as the lowering or ~ change of any two fermions. Therefore, if we introduce destruction operator. ˆ† ˆ Note that for any wavefunction, φ, hφ|pˆ2|φi ≥ 0 and the fermionic field operators ψ (r) and ψ(r), which, re- 2 spectively, create and annihilate fermions at position r, hφ|xˆ |φi ≥ 0. Therefore, it follows from Eq. 2 that En ≥ 0 for all n. Hence, there is a lowest energy state, or we must make sure that any wavefunction that we can make by acting some set of these operators on the vac- ground state, which we will denote as ψ0. Therefore there is a limit to how often we can keep lowering the energy of uum state is properly antisymmetrised. This is ensured (9) if one insists that the field operators anti-commute, the state, i.e.,aψ ˆ 0 = 0. We can now calculate the ground state energy of the harmonic oscillator, i.e., if ˆ ˆ† 0 ˆ ˆ† 0 ˆ† 0 ˆ 0   {ψ(r), ψ (r )} ≡ ψ(r)ψ (r ) + ψ (r )ψ(r) = δ(r − r ), ˆ † 1 1 Hshoψ0 = ω aˆ aˆ + ψ0 = ω. (7) (11a) ~ 2 2~ {ψˆ(r), ψˆ(r0)} = 0, (11b) In the same way as we derived Eq. 6, one can easily ˆ† ˆ† 0 † † † {ψ (r), ψ (r )} = 0. (11c) show that Hshoaˆ ψn = (En + ~ω)ˆa ψn. Thereforea ˆ moves us up the ladder of states thata ˆ moved us down. This guarantee of an antisymmetrised wavefunction is Hencea ˆ† is known as a raising or creation operator. Thus one of the most obvious advantages of the second quanti- we have sation formalism as it is much easier than having to deal √ with the Slater determinants that are typically used to † aˆ ψn = n + 1 ψn+1, (8a) ensure the antisymmetrisation of the many-body wave- √ aψˆ n = n ψn−1 (8b) function in the first quantised formalism (3). For any practical calculation one needs to work with a where the terms inside the radicals are required for the particular basis set, {φi(r)}. The field operators can be correct normalisation of the wavefunctions (8). Therefore expanded in an arbitrary basis set as ψ = √1 (ˆa†)nψ and n n! 0 ˆ X ψ(r) = cˆiφi(r), (12a)   i 1 X En = ~ω n + . (9) ˆ† † ∗ 2 ψ (r) = cˆi φi (r). (12b) i

Notice that above we solved the simple harmonic oscil- (†) Thusc ˆ annihilates (creates) a fermion in the state lator, i.e., calculated the energies of all of the eigenstates, i φ (r). These operators also obey fermionic anticommu- without needing to find explicit expressions for any of the i tation relations, first quantised , ψn. This general feature of the second quantised approach is extremely advanta- † {cˆi, cˆj} = δij, (13a) geous when we are dealing with the complex many-body {cˆ , cˆ } = 0, (13b) wavefunctions typical in condensed matter physics and i j † † chemistry. {cˆi , cˆj} = 0. (13c)

2 As fermions obey the Pauli exclusion principle there where can be at most one fermion in a given state. We will th Z  2 2  denote a state in which the i basis function contains 3 ∗ ~ ∇ tij = − d r φi (r) − + U(r) φj(r) (19) zero (one) particles by |0ii (|1ii). Therefore 2m Z Z 3 3 ∗ cˆi|0ii = 0c ˆi|1ii = |0ii Vijkl = d r1 d r2 φi (r1)φj(r1) (20) † † (14) cˆi |0ii = |1ii cˆi |1ii = 0. ∗ ×V (r1 − r2)φk(r2)φl(r2),

It is important to realise that the number 0 is very dif- and the labels i, j, k and l are taken to define the spin ferent from the state |0ii. as well as the basis function. This is exact provided we Any operator acting on a system of fermions can be have an infinite complete basis. But practical calcula- expressed in terms of thec ˆ operators. A particularly tions require the use of finite basis sets and often use † important example is the ‘number operator’,n ˆi ≡ cˆi cˆi, incomplete basis sets. The simplest approach is to just which simply counts the number of particles in the state ignore this problem and calculate tij and Vijkl directly i - as can be confirmed by explicit calculation from from the finite basis set. However, this is often not the Eqs. 14. The total number of particles in the system best approach. We will delay a detailed discussion of why is therefore simply the expectation value of the operator this is and of the deep philosophical issues raised by this ˆ P P † N = i nˆi = i cˆi cˆi. Importantly, because we can write until section VIII. We also delay discussion of how to cal- any operator in terms of thec ˆ operators, we can calcu- culate these parameters until section VIII. Until then we late any observable from the expectation value of some will simply assume that tij, Vijkl and other similar pa- set ofc ˆ operators. Thus we have access to a complete rameters required are known and instead focus on how to description of the system from the second quantisation perform practical calculations using models of the form formalism. Further, we can always write the wave func- of Eq. 18 and closely related Hamiltonians. tion in terms thec ˆ operators if an explicit description In what follows we will assume that the states created of the wavefunction is required. For example the sum of † by thec ˆi operators form an orthonormal basis. This Slater determinants, greatly simplifies the mathematics, but differs from the approach usually taken in introductory chemistry text-

φ1(r1) φ2(r1) φ3(r1) φ4(r1) books as most quantum chemical calculations are per- Ψ(r1, r2) = α + β , φ1(r2) φ2(r2) φ3(r2) φ4(r2) formed in non-orthogonal bases for reasons of computa- (15) tional expedience. describes the same state as III. THE HUCKEL¨ OR TIGHT-BINDING MODEL |Ψi = (αcˆ1cˆ2 + βcˆ3cˆ4) |0i, (16) The simplest model with the form of Eq. 18 is usual where |0i = |01, 02, 03, 04,... i is the vacuum state, as called the H¨uckel model in the context of molecular sys- Ψ(r1, r2) = hr1, r2|Ψi (cf. Ref. 8). tems (10) and the tight-binding model in the context of Often, in order to describe solid state and chemical crystals (11). In these models one makes the approxi- systems, one needs to describe a set of N electrons whose mation that Vijkl = 0 for all i, j, k, and l. Therefore behaviour is governed by a Hamiltonian of the form these models explicitly neglect the interactions between electrons. Both models are identical, but slightly differ-   N 2 2 ent notation is standard in the different contexts. We X ∇n 1 X H = −~ + U(r ) + V (r − r ) ,(17) assume that our basis set consists of orbitals centred on  2m n 2 n m  n=1 m6=n particular sites, as we will in all of the models considered in these notes. These sites might be, for example, atoms in a molecule or solid, chemical groups in a molecule, p-d where V (rn − rm) is the potential describing the interac- hybrid states in a transition metal oxide, entire molecules tions between electrons and U(ri) is an external potential (including interactions with ions or nuclei, which may of- in a molecular crystal, or even larger strucutures. Clearly ten be considered to be stationary on the time scales the simplest problem has only one orbital per spin state relevant to electronic processes - although we will discuss on each site, in which case, effects due to the displacement of the nuclei in section ˆ X † VII). In terms of our second quantisation operators this Htb = − tijcˆiσcˆjσ, (21) Hamiltonian may be written as ijσ

X 1 X (†) Hˆ = − t cˆ†cˆ + V cˆ†cˆ† cˆ cˆ , (18) wherec ˆ annihilates (creates) an with spin σ ij i j 2 ijkl i k l j iσ ij ijkl in an orbital centred on site i.

3 A. Molecules (the H¨uckel model)

The standard notation in this context is tii = −αi, tij = −βij if site i and site j are connected by a ; one assumes that tij = 0 otherwise. Note that the subscripts on α and β are also often dropped, but they are usually implicit; if the molecule contains more than one species of atom the αs will clearly be different on the different species and the βs will depend on the species of each of the atoms that the electron is hopping between. Therefore,

ˆ X † X † HH¨uckel = αicˆiσcˆiσ + βijcˆiσcˆjσ, (22) iσ hijiσ where hiji serves to remind us that the sum is only over FIG. 1 The energy levels of the atomic and molecular or- those pairs of atoms joined by a chemical bond. Note bitals in the H¨uckel description of H2. The bonding orbital that βij is typically negative. is |β| lower in energy than the , whereas the antibonding orbital is |β| higher in energy than the atomic orbital. Therefore, neutral H2 is stabilised by 2|β| relative to 1. Molecular hydrogen 2H.

Clearly, in H2 there is only a single atomic species. In this case one can set αi = α for all i without loss of generality. Further, as there is also only a single bond, we may also choose βij = β giving respecting the Pauli exclusion principle. If the two pro- tons are infinitely separated β = 0 and the system has X X   Hˆ = α (ˆn +n ˆ ) + β cˆ† cˆ +c ˆ† cˆ , total energy Nα, where N is the total number of elec- H¨uckel 1σ 2σ 1σ 2σ 2σ 1σ trons. H+ has only one electron, which, in the ground σ σ 2 state, will occupy the bonding orbital, and so H+ has a (23) 2 binding energy of β.H2 has two electrons; in the ground where we have labelled the two atomic sites 1 and 2. state these electrons have opposite spin and therefore can both occupy the bonding orbital. Thus H2 has a binding This Hamiltonian has two eigenstates: one is known as − the bonding state, energy of 2β.H2 has three electrons, so while two can occupy the bonding state one must be in the antibond- 1 † † ing state, therefore the binding energy is only β. Finally, |ψbσi = √ (ˆc1σ +c ˆ2σ)|0i, (24) 2− 2 H2 has four electrons so one finds two in the each molec- ular orbital. Therefore the bonding energy is zero: the and the other is known as antibonding state, molecule is predicted to be unstable.

1 † † |ψaσi = √ (ˆc − cˆ )|0i. (25) Thus the H¨uckel model makes several predictions: neu- 2 1σ 2σ tral H2 is predicted to be significantly more stable than The bonding state has energy α + β, whereas the an- any of the ionic states; the two singly ionic species are tibonding state has energy α − β, recall that β < 0. predicted to be equally stable; the doubly cationic species Therefore every electron in the bonding state stabilises is predicted to be unstable. Further, the lowest optical the molecule by an amount |β|, whereas electrons in the absorption is expected to correspond to the transition be- antibonding state destabilise the molecule by an amount tween the bonding orbital and the antibonding orbital. |β|, hence the nomenclature.1 This is sketched in Fig. 1. The energy gap for this transition is 2|β|. Therefore, the lowest optical absorption is predicted to be the same in Because Vijkl = 0 the electrons are non-interacting and so the molecular orbitals are not dependent on the occu- the neutral species and the singly cationic species. Fur- pation of other orbitals. Therefore to calculate the total ther, this absorption is predicted to occur at a frequency energy of the ground state of the molecule one simply fills with the same energy as the heat of formation for the up the states starting with the lowest energy states and neutral species. While these predictions do capture qual- itatively what is observed experimentally, they are cer- tainly not within chemical accuracy (i.e. within kBT ∼ 1 kcal mol−1 ∼ 0.03 eV for T = 300 K). For example the 1 experimentally determined binding energies (10) are 2.27 Note that in a non-orthogonal basis the antibonding orbital may + − 2− be destabilised by a greater amount than the bonding orbital is eV for H2 , 4.74 eV for H2, 1.7 eV for H2 , while H2 is stabilised. indeed unstable.

4 and

1  † † † † † †  B2g 2|β| |ψB i = √ cˆ − cˆ +c ˆ − cˆ +c ˆ − cˆ |0i, 2g 6 1σ 2σ 3σ 4σ 5σ 6σ

E2u |β| where  = eiπ/3. These wavefunctions are sketched in

Fig. 2. The energies of these states are EA2u = α − 2|β|, E = E0 = α − |β|, E = E0 = α + |β| and 0 E1g E1g E2u E2u

EB2g = α + 2|β|. The subscripts are symmetry labels (12; 13) for the group D6h and one should recall that, E1g -|β| because we are dealing with π-orbitals, all of the orbitals sketched here are antisymmetric under reflection through A2u -2|β| the plane of the page. The degenerate (E1g and E2u) orbitals are typically written/drawn rather differently (cf. Ref. 10). However, any linear combination of degenerate eigenstates is also an eigenstate; this representation was FIG. 2 The molecular orbitals for benzene from π-H¨uckel the- chosen as it highlights the symmetry of the problem. For ory. Different colours indicate a change in sign of the wave- a more detailed discussion of this problem see Ref. 14. function. In the neutral molecule the A2u and both E1g states are occupied, while the B2g and E2u states are virtual. Note that we have taken real superpositions (10) of the two-fold degenerate states in order to facilitate these plots. 3. Electronic interactions and the parameterisation of the H¨uckel model

2. π-H¨uckel theory of benzene As noted above the H¨uckel model does not explicitly include interactions between electrons. This leads to se- rious qualitative and quantitative failures of the model, For many organic molecules a model known as π- some of which we have seen above and which will discuss H¨uckel theory is very useful. In π-H¨uckel theory one further below. However, given the (mathematical and considers only the π electrons. A simple example is a conceptual) simplicity and the computational economy benzene molecule. The hydrogen atoms have no π elec- of the method one would like to improve the method trons and are therefore not represented in the model. as far as possible. So far we have treated the theory This leaves only the carbon atoms, so again we can set as parameter free. However, if we treat the model as a α = α and β = β. Because of the ring geometry of i ij semi-empirical method instead one can include some of benzene (and assuming that the molecule is planar) the the effects due to electron-electron interactions without Hamiltonian becomes greatly increasing the computational cost of the method. X X   For example, one can make α dependent on the charge Hˆ = α nˆ + β cˆ† cˆ +c ˆ† cˆ , (26) H¨uckel iσ iσ i+1σ i+1σ iσ on the atom. This is reasonable, as the more electrons iσ iσ we put on an atom the harder it is to add another due where the addition in the site index is defined modulo six, to the additional Coulomb repulsion from the extra elec- i.e., site number seven is site number one. For benzene trons. The simplest way to account for this is the ‘ω we have six solutions per spin state, which are technique’ (10) where one replaces 0 αi → α = αi + ω(q0 − qi)β, (28) 1  † † † † † †  i |ψA i = √ cˆ +c ˆ +c ˆ +c ˆ +c ˆ +c ˆ |0i, 2u 6 1σ 2σ 3σ 4σ 5σ 6σ where qi is the charge on atom i, q0 is a (fixed) reference |ψE1g i charge and ω is a parameter. The ω technique suppresses 1   the unphysical fluctuations of the electron density, which = √ cˆ† + cˆ† + 2cˆ† − cˆ† − cˆ† − 2cˆ† |0i, 6 1σ 2σ 3σ 4σ 5σ 6σ are often predicted by the H¨uckel model (cf. the dis- cussion of H above). Similar techniques can also be |ψ0 i 2 E1g applied to β. These parameterisations only slightly com- 1   = √ cˆ† − 2cˆ† − cˆ† − 2cˆ† + cˆ† +c ˆ† |0i, plicate the model and do not lead to a major inflation 6 1σ 2σ 3σ 4σ 5σ 6σ of the computational cost, but can significantly improve the accuracy of the predictions of the H¨uckel model (15). |ψE2u i 1   = √ cˆ† + 2cˆ† − cˆ† +c ˆ† + 2cˆ† − cˆ† |0i, 6 1σ 2σ 3σ 4σ 5σ 6σ B. Crystals (the tight binding model) |ψ0 i E2u 1   For infinite systems it is necessary to work with a fixed = √ cˆ† − cˆ† + 2cˆ† +c ˆ† − cˆ† + 2cˆ† |0i 6 1σ 2σ 3σ 4σ 5σ 6σ chemical potential rather than a fixed particle number.

5 Therefore before we discuss the tight binding model we electrons, i.e., one is working in the grand canonical en- will briefly review the chemical potential (also see Ref. 6 semble (17). Thus, the Fermi distribution for the system for a discussion of the chemical potential in a chemical is given by context). 1 f(E,T ) = (E−µ) (31) 1 + e kB T 1. The chemical potential Therefore at T = 0 all of the states with energies lower When one is dealing with a large system keeping track than the chemical potential are occupied, and all of of the number of particles can become difficult. This the states with energies greater than the chemical po- is particularly true in the thermodynamic limit where tential are unoccupied. Therefore, the Fermi energy, ˆ the number of electrons Ne ≡ hNi → ∞ and the vol- EF = µ(T = 0). Note that as F is temperature depen- ume of the system V → ∞ in such a way so as to en- dent Eq. 30 shows that, in general, µ will also be tem- sure that the electronic density, ne = Ne/V , remains perature dependent.2 Nevertheless Eq. 31 gives a clear constant. Lagrange multipliers (16) are a powerful and interpretation of the chemical potential at any tempera- general method for imposing constraints on differential ture: µ(T ) is the energy of a state with a 50% probability equations (such as the Schr¨odingerequation) without of occupation at temperature T . requiring the solution of integro-differential equations. Briefly, consider a function, f(x, y, z, . . . ) that we wish to extremise (minimise or maximise) subject to a con- straint that means that x, y, z . . . are no longer inde- 2. The tight binding model pendent. In general we may write the constraint in the form φ(x, y, z, . . . ) = 0. This allows us to define the For periodic systems (crystals) one usually refers to the function g(x, y, z . . . , λ) ≡ f(x, y, z . . . ) + λφ(x, y, z . . . ), H¨uckel model as the tight binding model. Often one con- where λ is known as a Lagrange multiplier. One may siders models with only ‘nearest neighbour’ terms, that show (16) that the extremum of g(x, y, z . . . , λ) with re- is one takes tii = −i, tij = t if i and j are at nearest spect to x, y, z . . . and λ is the extremum of f(x, y, z . . . ) neighbour sites, and tij = 0 otherwise. Thus, for nearest with respect to x, y, z . . . subject to the constraint that neighbour hopping only, φ(x, y, z, . . . ) = 0. ˆ ˆ X † X † Typically the problem we wish to solve in chemistry Htb − µN = −t cˆiσcˆjσ + (i − µ)ˆciσcˆiσ, (32) and condensed matter physics is to minimise the free en- hijiσ iσ ergy, F , (which reduces to the energy, E, at T = 0) subject to the constraint of having a fixed number of where µ is the chemical potential and hiji indicates that electrons (determined by the chemistry of the material the sum is over nearest neighbours only. Further, if we in question). This suggests that one should simply in- consider materials with only a single atomic species we troduce a Lagrange multiplier to resolve the difficulty of can set i = 0 yielding constraining the number of electrons in the thermody- X X namic limit. A suitable constraint could be introduced Hˆ − µNˆ = −t cˆ† cˆ − µ cˆ† cˆ . (33) ˆ tb iσ jσ iσ iσ by adding the term λ(N0 −N) to the Hamiltonian, where hijiσ iσ N0 is the chemically required number of electrons, and requiring the the free energy is an extremum with respect to λ. However, one can also impose the same constraint and achieve additional physical insight by subtracting the 3. The one dimensional chain term µNˆ from the Hamiltonian and requiring that The simplest infinite system is a chain with nearest ∂F neighbour hopping only. As we are on a chain the sites N = − . (29) 0 ∂µ have a natural ordering and the Hamiltonian may be written as The chemical potential (for electrons), µ, is then given ˆ ˆ X  † †  X † by Htb − µN = −t cˆiσcˆi+1σ +c ˆi+1σcˆiσ − µ cˆiσcˆiσ, iσ iσ ∂F (34) µ = − . (30) ∂Ne We can solve this model exactly by performing a lattice . We begin by introducing the recipro- Therefore, specifying a system’s chemical potential is equivalent to specifying the number of electrons, but pro- vides a far more powerful approach for bulk systems. 2 Physically this approach is equivalent to thinking of In contrast, as EF is only defined at T = 0 it is not temperature the system as being attached to to an infinite bath of dependent.

6 2t b) 2t

t t k k

ε 0 ε 0

-t -t

-2t -2t -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 k a k a c) 2t d) 2t

t t k k

ε 0 ε 0

-t -t

-2t -2t -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 k a k a

FIG. 3 (a) The dispersion relation, εk = −2t cos(ka), of the one dimensional tight binding chain with nearest neighbour hopping only. (b) Shaded area shows the filled states for µ = −t. (c) Shaded area shows the filled states for µ = 0. (d) Shaded area shows the filled states for µ = t.

cal space creation and annihilation operators: the energy is just the sum of εk for the states kσ that are occupied, and we have solved the problem. We plot 1 X ikRi cˆiσ = √ cˆkσe , (35a) the dispersion relation in Fig. 3a. For a tight binding N k model calculating the dispersion relation is equivalent to 1 X solving the problem. cˆ† = √ cˆ† e−ikRi , (35b) iσ N kσ The chemical potential, µ, must be chosen to ensure k that there are the physically required number of elec- where k is the lattice wavenumber or crystal momentum trons. Changing the chemical potential has the effect of th moving the Fermi energy up or down the band and hence and Ri is the position of the i lattice site. Therefore, changing the number of electrons in the system. For ex- 1 0 X † i(k −k)Ri ample (see Fig. 3b-d), in the above problem the half Hˆ − µNˆ = cˆ cˆ 0 e tb N kσ k σ filled band corresponds to µ = 0; the quarter filled band ikk0σ corresponds to µ = −t; and the three quarters filled band h 0 i × − t(eik a + e−ika) − µ ,(36) corresponds to µ = t. where a is the lattice constant, i.e., the distance between 0 1 P i(k −k)Ri 0 neighbouring sites Ri and Ri+1. N i e = δ(k − k) (18); therefore 4. The square, cubic and hypercubic lattices h i ˆ ˆ X † † In more than one dimension the notation becomes Htb − µN = − 2t cos(ka)c ˆkσcˆkσ − µcˆkσcˆkσ kσ slightly more complicated, but the mathematics does not, X necessarily, become any more difficult. The simplest gen- = (ε − µ)ˆc† cˆ , (37) k kσ kσ eralisation of the chain we have solved above is the two kσ dimensional square lattice where where εk = −2t cos(ka) is known as the dispersion rela- tion. Notice that Eq. 37 is diagonal, i.e., it only depends ˆ ˆ X † X † Htb − µN = −t cˆiσcˆjσ − µ cˆiσcˆiσ. (38) † on the number operator terms, nkσ =c ˆkσcˆkσ. Therefore hijiσ iσ

7 Recall that hiji indicates that the sum is over nearest models that include interactions can be solved exactly in neighbours only. To solve this problem we simply gener- infinite dimensions as we will discuss in section IV.D.2. alise our operators to 1 X ik·Ri cˆiσ = √ cˆkσe , (39a) 5. The hexagonal and honeycomb lattices N k 1 † X † −ik·Ri Even if the bonds are not all mutually perpendicu- cˆiσ = √ cˆkσe , (39b) N k lar the solution to the tight-binding model can still be found by Fourier transforming the Hamiltonian. Three where k = (kx, ky) is the lattice wavevector or crystal important examples of such lattices are the hexagonal th momentum and Ri = (xi, yi) is the position of the i lattice (which is often referred to as the triangular lat- lattice site. We then simply repeat the process we used tice, although this is formally incorrect), the anisotropic to solve the one dimensional chain. As the lattice only triangular lattice, and the honeycomb lattice, which are contains bonds in perpendicular directions the calcula- sketched in Fig. 4. For each lattice the solution is of the tions for the x and y directions go through independently form of Eq. 40. For the hexagonal lattice and one finds that √ !   X † 3 kxax Hˆtb − µNˆ = (εk − µ)ˆc cˆkσ, (40) ε = −2t cos(k a ) − 4t cos k a cos . kσ k x x 2 y y 2 kσ (41) where the dispersion relation is now εk = For the anisotropic triangular lattice −2t[cos(kxax) + cos(kyay)] and aν is the lattice constants in the ν direction. h i 0 A three dimensional cubic lattice is not any more dif- εk = −2t cos(kxax) + cos(kyay) − 2t cos(kxax + kyay). ficult. In this case k = (kx, ky, kz) and the solution is (42) of the form of Eq. 40 but with εk = −2t[cos(kxax) + The honeycomb lattice has an important additional cos(kyay) + cos(kzaz)]. Indeed so long as we keep all the subtlety, that there are two inequivalent types of lattice bonds mutually perpendicular one can keep generalising site (cf. Fig. 4c), which it is worthwhile to work through. this solution to higher dimensions. This may sound some- We begin by introducing new operators,c ˆiνσ, which an- what academic as no materials live in more than three nihilate an electron with spin σ on the νth sublattice in dimensions, but the infinite dimensional hypercubic lat- the ith unit cell, where ν = A or B. Therefore we can tice has become important in recent years because many rewrite Eq. 38 as

ˆ X † † Htb = −t cˆiAσcˆjBσ +c ˆjBσcˆiAσ hijiσ † X  cˆ   0 1   cˆ  = −t iAσ iAσ (43) cˆiBσ 1 0 cˆiBσ hijiσ  †     X cˆkAσ 0 hk cˆkAσ = −t ∗ , cˆkBσ hk 0 cˆkBσ kσ √ √ ikxa −i(kx+ 3ky )a/2 −i(kx− 3ky )a/2 where hk = e + e + e . Therefore v u √ !   u √ 3kya 3kxa ε = ±t |h | = ±tt3 + 2 cos( 3k a) + 4 cos cos (44) k k y 2 2

We plot this dispersion relation in Fig. 5. In order to see why these points are interesting consider The most interesting features of this band structure a point K+q in the neighbourhood of K. Recalling that 1 2 are called the ‘Dirac points’. The Dirac points are lo- cos(K + q) = cos K − q sin K + 2 q cos K + ... one finds 0 cated at k = nK√+ mK where n and m are integers,√ K = (2π/3a, 2π/3 3a), and K0 = (2π/3a, −2π/3 3a).

8 FIG. 4 The (a) hexagonal (triangular), (b) anisotropic triangular, (c) honeycomb and (d) kagome lattices. The hexagonal lattice contains two inequivalent types of lattice site, some of which are labelled A and B. The sets of equivalent sites are referred to as sublattices.

Ek = ~c|k|. Thus the low-energy electronic excitations on a honeycomb lattice behave as if they are massless relativistic particles, with the Fermi velocity playing the role of the speed of light in the theory. Therefore much excitement (19) has been caused by the recent synthe- sis of atomically thick sheets of (20), in which carbon atoms form a honeycomb lattice. In graphene 6 −1 vF ' 1 × 10 ms , two orders smaller than the speed of light in the vacuum. This has opened the possibility of exploring and controlling ‘relativistic’ effects in a solid state system (19).

IV. THE HUBBARD MODEL

So far we have neglected electron-electron interactions. FIG. 5 The Dirac dispersion of the honeycomb lattice In real materials the electrons repel each other due to the Coulomb interaction between them. The most obvious extension to the tight binding model that describes some that, for small |q|, of the electron-electron interactions is to allow only on- ε = v |q| + ... (45) site interactions, i.e., if Vijkl 6= 0 if and only if i, j, k and K+q ~ F l all refer to the same orbital. For one orbital per site we where vF = 3ta/2~ is known as the Fermi velocity. then have the Hubbard model, This result should be compared with the relativistic result ˆ X † X † † HHubbard = −t cˆiσcˆjσ + U cˆi↑cˆi↑cˆi↓cˆi↓, (47) 2 2 4 2 2 2 hijiσ i Ek = m c + ~ c |k| , (46) where m is a particles rest mass and c is the speed of where we have assumed nearest neighbour hopping only. light. This reduces to the famous E = mc2 for k = 0, It follows from Eq. 21 that U > 0, i.e., electrons repel but for massless particles, such as photons, one finds that one another.

9 A. The two site Hubbard model: molecular hydrogen n + δn, where n (m) is the mean value of n (m) and δn (δm) are the fluctuations about the mean, which are The two site Hubbard model is a nice context in which assumed to be small, one notes that to consider some of the basic properties of the chemical bond. The two body term in the Hubbard model greatly complicates the problem relative to tight binding model. mn = (m + δm)(n + δn) Therefore the Hubbard model also presents a nice context = m n + m δn + δm n + δm δn in which to introduce one of the most important tools in theoretical physics and chemistry: mean-field theory. ≈ m n + m δn + δm n. (48)

1. Mean-field theory, the Hartree-Fock approximation & Thus mean-field approximations neglect terms that are theory quadratic in the fluctuations. To construct a mean-field theory of any two, as yet Hartree theory is a mean-field in the electron density, unspecified, physical quantities, m = m + δm and n = i.e.,

† † h † † † ih † † † i cˆαcˆβcˆγ cˆδ = hcˆαcˆβi + cˆαcˆβ − hcˆαcˆβi hcˆγ cˆδi + cˆγ cˆδ − hcˆγ cˆδi † † † † † † ≈ hcˆαcˆβicˆγ cˆδ +c ˆαcˆβhcˆγ cˆδi − hcˆαcˆβihcˆγ cˆδi. (49) However, it was quickly realised that this does not allow for electron exchange, i.e., one should also include averages † such as hcˆαcˆδi, therefore a better mean-field theory is Hartree-Fock theory, which includes these terms. However, because of the limited interactions included in the Hubbard model the Hartree theory is identical to the Hartree-Fock † theory if one assumes that spin-flip terms are negligible, i.e., that hcˆi↑cˆi↓i = 0, which we will. The Hartree-Fock approximation to the Hubbard Hamiltonian is therefore

ˆ X † X h † † † † † † i HHF = −t cˆiσcˆjσ + U hcˆi↑cˆi↑icˆi↓cˆi↓ +c ˆi↑cˆi↑hcˆi↓cˆi↓i − hcˆi↑cˆi↑ihcˆi↓cˆi↓i hijiσ i X † X h † † i = −t cˆiσcˆjσ + U ni↑cˆi↓cˆi↓ + ni↓cˆi↑cˆi↑ − ni↑ni↓ , (50) hijiσ i

† where niσ = hcˆiσcˆiσi. Thus we have a Hamiltonian for a single electron moving in the mean-field of the other electrons. Note that this Hamiltonian is equivalent to the ω method parameterisation of the H¨uckel model (cf. section III.A.3 and particularly Eq. 28) if we set ω = U/β. Thus the ω method is just a parameterisation of the Hubbard model solved in the Hartree-Fock approximation. The Hubbard model with two sites and two electrons can be taken as a model for molecular hydrogen. In the 0 Hartree-Fock ground state, |ΨHFi, the two electrons have opposite spin and each occupy the bonding state, which we found to be the ground state of the H¨uckel model in section III.A.1: 1 |Ψ0 i = |ψ i ⊗ |ψ i = (ˆc† +c ˆ† )(ˆc† +c ˆ† )|0i (51a) HF b↓ b↑ 2 1↑ 2↑ 1↓ 2↓ 1 = (ˆc† cˆ† +c ˆ† cˆ† − cˆ† cˆ† +c ˆ† cˆ† )|0i. (51b) 2 1↑ 1↓ 1↑ 2↓ 1↓ 2↑ 2↑ 2↓

0 + − Notice that |ΨHFi is just a product of two single particle (H +H ). This is not what is observed experimentally. wavefunctions (one for the spin up electron and another In reality the former is far more likely. for the spin down electron; cf. Eq. 51a). Thus we say that the wavefunction is uncorrelated and that the two electrons are unentangled. 2. The Heitler-London wavefunction & An important prediction of the Hartree-Fock theory is that if we pull the protons apart we are equally likely Just a year after Schr¨odinger wrote down his wave to get two hydrogen atoms (H+H) or two hydrogen ions equation (21), Heitler and London (22) proposed a theory

10 of the chemical bond based on the new quantum mechan- symmetry,3 which swaps the site labels 1 ↔ 2, whereas ics. Explaining the nature of the chemical bond remains |Ψct-i is odd under inversion symmetry. As the Hamil- one of the greatest achievement of quantum mechanics. tonian is symmetric under inversion the eigenstates will Heitler and London’s theory led to the valence bond the- have a definite parity so |Ψct-i is an eigenstate, with en- ory of the chemical bond (23). The two site Hubbard ergy Ect- = U. The other two singlet states are not model of H2 is the simplest context in which to study distinguished by any symmetry of the Hamiltonian and this theory. so they do couple, yielding the Hamiltonian matrix The Heitler-London wavefunction is  hΨ |Hˆ |Ψ i hΨ |Hˆ |Ψ i  H = HL Hubbard HL HL Hubbard ct+ 0 1 † † † † hΨct+|HˆHubbard|ΨHLi hΨct+|HˆHubbard|Ψct+i |ΨHLi = √ (ˆc1↑cˆ2↓ − cˆ1↓cˆ2↑)|0i. (52) 2  0 −2t  = . (53) −2t U Notice that the wavefunction is correlated as it cannot be written as a product of a wavefunction for each of the √ This has eigenvalues, E = 1 (U − U 2 + 16t2) and particles. Equivalently one can say that the two electrons √ CF 2 1 2 2 E 2 = (U + U + 16t ). The corresponding eigen- are entangled. The Heitler London wavefunction overcor- S 2 rects the physical errors in the Hartree-Fock molecular states are orbital wavefunction as it predicts zero probability of H2 dissociating to an ionic state, but is, nevertheless, a sig- |ΨCFi = cos θ|ΨHLi + sin θ|Ψct+i cos θ   nificant improvement on . = √ cˆ† cˆ† − cˆ† cˆ† 2 1↑ 2↓ 1↓ 2↑  sin θ  † † † †  + √ cˆ1↓cˆ1↑ +c ˆ2↓cˆ2↑ |0i(54a) 3. Exact solution of the two site Hubbard model 2 |ΨS2 i = sin θ|ΨHLi + cos θ|Ψct+i The Hilbert space of the two site, two electron Hubbard sin θ   = √ cˆ† cˆ† − cˆ† cˆ† model is sufficiently small that we can solve it analyti- 2 1↑ 2↓ 1↓ 2↑ cally; nevertheless this problem can be greatly simplified  cos θ  † † † †  by using the symmetry properties of the Hamiltonian. + √ cˆ1↓cˆ1↑ +c ˆ2↓cˆ2↑ |0i(54b), Firstly, note that the total spin operator commutes with 2 the Hamiltonian 47, as none of the terms in the Hamilto- √ 2 2 nian cause spin flips. Therefore the energy eigenstates where tan θ = (U − U + 16t )/4t. For U > 0, as is must also be spin eigenstates. For two electrons this required physically, the state |ΨCFi is the ground state means that all of the eigenstates will either be singlets for all values of U/t. |ΨCFi is often called the Coulson- (S = 0) or triplets (S = 1). Fischer wavefunction. m Inspection of Eq. 54a reveals that for U/t → ∞ the Let us begin with the triplet states, |Ψ1 i. Con- 1 Coulson-Fischer state tends to the Heitler-London wave- sider a state with two spin up electrons, |Ψ1i. Be- cause there is only one orbital per site the Pauli ex- function, while for U/t → 0 we regain the molecular or- clusion principal ensures that there will be exactly one bital picture (Hartree-Fock wavefunction). 1 † † electron per site, i.e., |Ψ1i =c ˆ1↑cˆ2↑|0i. The electrons cannot hop between sites as the presence of the other electron and the Pauli principle forbid it. Therefore, B. Mott insulators & the Mott-Hubbard metal-insulator 1 † 1 1 † 1 transition hΨ1|(−t cˆ1σcˆ2σ)|Ψ1i = hΨ1|(−t cˆ2σcˆ1σ)|Ψ1i = 0 for σ =↑ or ↓. There is exactly one electron on each site so In 1949 Mott (24) asked an apparently simple ques- 1 P † † 1 hΨ1|U i cˆi↑cˆi↑cˆi↓cˆi↓|Ψ1i = 0. Thus the total energy of tion with a profound and surprising answer. As we 1 this state is E1 = 0. have seen above, for the two site Hubbard model both −1 The same chain of reasoning shows that |Ψ1 i = the molecular orbital (Hartree-Fock) and valence bond † † −1 (Heitler-London) wavefunctions are just approximations cˆ1↓cˆ2↓|0i and E1 = 0. It then follows from spin ro- † † † † to the exact (Coulson-Fischer) wavefunction. Mott asked tation symmetry that |Ψ0i = √1 (ˆc cˆ +c ˆ cˆ )|0i and 1 2 1↑ 2↓ 1↓ 2↑ whether the equivalent statement is true in an infinite 0 E1 = 0. As the Hilbert space contains six states, this leaves three singlet states. A convenient basis of these is formed 3 by the Heitler-London state and the two charge trans- It may not be immediately obvious that |ΨHLi is even under 1 † † † † fer states: |ΨHLi = √ (ˆc cˆ − cˆ cˆ )|0i, |Ψct+i = inversion symmetry, but this is easily confirmed as I|ΨHLi = 2 1↑ 2↓ 1↓ 2↑ 1 † † † † 1 † † † † † † † † † † † † √ (ˆc cˆ − cˆ cˆ )|0i = √ (−cˆ cˆ +c ˆ cˆ )|0i = |ΨHLi, √1 (ˆc cˆ +ˆc cˆ )|0i and |Ψ i = √1 (ˆc cˆ −cˆ cˆ )|0i. 2 2↑ 1↓ 2↓ 1↑ 2 1↓ 2↑ 1↑ 2↓ 2 1↑ 1↓ 2↑ 2↓ ct- 2 1↑ 1↓ 2↑ 2↓ where I is the inversion operator, which swaps the labels 1 and Note that |ΨHLi and |Ψct+i are even under ‘inversion’ 2.

11 solid, and, surprisingly, found that the answer is no. Fur- large U/t ther, Mott showed that the Hartree-Fock and Heitler- small U/t London wavefunctions predict very different properties for crystals. One of the most important properties of a crystal is its conductivity. In a metal the conductivity is high

E and increases as the temperature is lowered. Whereas in a semiconductor or an insulator the conductivity is low and decreases as the temperature is lowered. These behaviours arise because of fundamental differences be- tween the electronic structures of metals and semiconduc- tors/insulators (11). In metals there are excited states 0 0.2 0.4 0.6 0.8 1 at arbitrarily low energies above the Fermi energy. This n means that, even at the lowest temperatures, electrons p can move in response to an applied electric field. In semi- FIG. 6 Sketch of Mott’s proposal for the energy of the Hub- conductors and insulators there is an energy gap between bard model as a function of the number of holon-doublon the highest occupied electronic state and the lowest unoc- pairs, np, at low (zero) temperature(s) for large and small cupied electronic state at zero temperature. This means U/t. that a thermal activation energy must be provided if elec- trons are to move in response to an applied field. The difference between semiconductors and insulators is sim- doublon pairs “it no longer follows that work must nec- ply the size of the gap; therefore we will not distinguish essarily be done to form some more”. This is because the between the two below and will refer to any material with holon and doublon now interact via a screened potential, a gap as an insulator. V (r) = −(e2/κr) exp(−qr), where q is the Thomas-Fermi Consider a Hubbard model at ‘half-filling’, i.e., with wavevector (cf. Ref. 11). For sufficiently large q there the same number of electrons as lattice sites. In order will be no bound states and the molecular orbital theory for a macroscopic current to flow, an electron must move predicts that the half-filled Hubbard model is metallic. from one lattice site (leaving an empty site with a net Thus, Mott argued that their are two (local) minima of positive charge) to a distant site (creating a doubly oc- the free energy in a crystal (cf. Fig. 6). One of the min- cupied site with a net negative charge). The net charges ima corresponds to a state with no holon-doublon pairs may move through collective motions of the electrons. that is well approximated by a valence bond wavefunction One could keep track of this by describing the movement and is now known as the Mott insulating state. The sec- of all of the electrons, but it is easier to introduce an ond minimum corresponds to a state with many doublon- equivalent description where we treat the net charges as holon pairs that is well approximated by a molecular or- particles moving in a neutral background. Therefore, we bital wavefunction and is metallic. As we saw above, will refer to the positive charge as a holon and the neg- valence bond theory works well for U  t and molecular ative charge as a doublon. In the ground state of the orbital theory works well for U  t. Therefore, in the valence bond theory all of the sites are neutral and there half-filled Hubbard model we expect a for are no holons or doublons (cf. Eq. 52). However, it is large U/t and a metal for small U/t. Further the ‘double reasonable to postulate that there are low lying excited well’ structure of the energy predicted by Mott’s argu- states and hence thermal states that contain a few dou- ment (Fig. 6) suggests that there is a first order metal- blons and holons. These doublons and holons would in- insulator phase transition, known as the Mott transition. teract via the Coulomb potential, V (r) = −e2/κr, where Mott predicted that this metal-insulator transition can κ is the dielectric constant of the crystal. We know from be driven by applying pressure to a Mott insulator. This the theory of the hydrogen atom (or, better, positron- has now been observed in a number of systems; perhaps ium, cf. Ref. 8) that this potential gives rise to bound the purest examples are the organic charge transfer salts states. Therefore one expects that, in the valence bond (BEDT-TTF)2X (25). theory, holons and doublons are bound and separating It is interesting to note that this infusion of chemical holon-doublon pairs costs a significant amount of energy. ideas into condensed matter physics has remained im- Thus one expects the number of distant holon-doublon portant in studies of the Mott transition. Of particular pairs to decrease as the temperature is lowered. There- note is Anderson’s resonating valence bond theory of su- fore, the valence bond theory predicts that the half-filled perconductivity in the high temperature superconductors Hubbard model is an insulator. (26; 27), which describes superconductivity in a doped In contrast the molecular orbital theory has large num- Mott insulator in terms of a generalisation of the valence bers of holons and doublons (cf. Eq. 51b, which sug- bond theory discussed above. This theory can also be gests that for an N-site model there will be N/2 neu- modified to describe superconductivity on the metallic tral sites, N/4 empty sites, and N/4 doubly occupied side of the Mott transition for a half-filled lattice. This sites). Mott reasoned that if there are many holon- theory then provides a good description of the supercon-

12 not predict a Mott insulating state. Thus weakly cor- 2t εk↑ related theories make the qualitatively incorrect predic- εk↓ tion that materials such as NiO, V2O3, La2CuO4 and ε0 t k κ-(BEDT-TTF)2Cu[N(CN)2]Cl are metals, whereas ex- perimentally all are insulators. 0 We will discuss a quantitative theory of the Mott tran- sition is section IV.C.2. -t

-2t C. Mean-field theories for crystals

-3 -2 -1 0 1 2 3 1. Hartree-Fock theory of the Hubbard model: Stoner k a ferromagnetism FIG. 7 The dispersion relations for spin-up and spin-down electrons in the Hartree-Fock theory of the Hubbard chain In a similar manner to that in which we constructed the (Stoner model of ferromagnetism) with m = 0.8t/U. Hartree-Fock mean-field theory for the two site Hubbard model in section IV.A.1 we can also construct a Hartree- Fock theory of the infinite lattice Hubbard model. Again, ductivity observed in the (BEDT-TTF) X salts (28). we simply replace the number operators in the two body 2 † Note that theories, such as Hartree-Fock theory or den- term by their mean values, niσ ≡ hcˆiσcˆiσi, plus the fluctu- †  sity functional theory (29), that do not include the strong ations about the mean, cˆiσcˆiσ − niσ , and neglect terms electronic correlations present in the Hubbard model do that are quadratic in the fluctuations, viz.,

X † † X h † ih † i U cˆi↑cˆi↑cˆi↓cˆi↓ = U ni↑ + cˆi↑cˆi↑ − ni↑ ni↓ + cˆi↓cˆi↓ − ni↓ i i X h † † i ' U ni↓cˆi↑cˆi↑ + ni↑cˆi↓cˆi↓ − ni↑ni↓ . (55) i

If we make the additional approximation that niσ = nσ for all i, i.e., that the system is homogeneous and does not spontaneously break translational symmetry, we find that the Hartree-Fock Hamiltonian for the Hubbard model is

ˆ ˆ X † X † HHF − µN = −t cˆiσcˆjσ + (Unσ − µ)ˆciσcˆiσ − UNn↑n↓, (56) hijiσ iσ where N is the number of lattice sites and σ is the opposite spin to σ. It is convenient to write this Hamiltonian in terms of the total electron density, n = n↑ + n↓ and the magnetisation density, m = n↑ − n↓, which gives,

ˆ ˆ X † X † HHF − µN = −t cˆiσcˆjσ − µ cˆiσcˆiσ hijiσ iσ X h1 1 1 i +U (n − m)ˆc† cˆ + (n + m)ˆc† cˆ − (n + m)(n − m) , 2 i↑ i↑ 2 i↓ i↓ 4 i X  Um  Un X NU = ε0 + σ nˆ − µ − nˆ − (n2 − m2) (57) k 2 kσ 2 kσ 4 kσ kσ

0 where εk is the dispersion relation for U = 0 and σ = netisation density is non-zero the dispersion relation for ±1 =↑↓. The last term is just a constant and will not spin-up electrons is different from that for spin-down elec- concern us greatly. The penultimate term is the ‘renor- trons (cf. Fig. 7). It is important to note that the malised’ chemical potential, i.e., the chemical potential, Hartree-Fock approximation has reduced the problem to µ, of the system with U = 0 is decreased by Un/2 due to a single particle (single determinant) theory. Thus we the interactions. The first term is just the renormalised dispersion relation, in particular we find that if the mag-

13 2 D (0)U>1 Guess value of m D0 (0)U<1 f(m)=m 1.5

Calculate Dσ(ε) from εkσ for current m Set m=(1-α)m+αf(m) 1 Calculate f(m) f(m)

Does f(m)=m to within the 0.5 specified tolerance? No

Yes 0 0 0.5 1 1.5 2 m=f(m) is a self consistent solution m

FIG. 9 Graphical solution of the self-consistency equation FIG. 8 How to find the self consistent solution of Eq. 59. If (Eq. 59) for the Stoner model of ferromagnetism. the convergence works well one can take α = 1, but for some problems convergence can be reached more reliably with a small value of α (often a value as small as ∼ 0.05 is used). Fig. 9. Furthermore, the m 6= 0 solutions typically have lower energy than the m = 0 solution and therefore for UD0(0) > 1 the ground state is ferromagnetic. UD0(0) ≥ can write 1 is known as the Stoner condition for ferromagnetism. X NU In order for the Stoner condition to be satisfied a system Hˆ − µNˆ = (ε∗ − µ∗)n ˆ − (n2 − m2), (58) HF kσ kσ 4 must have narrow bands [small t, and hence large D(0)] kσ and strong interactions (large U). There are three elemental ferromagnets, Fe, Co and Ni, where ε∗ = ε0 − 1 σUm and µ∗ = µ − 1 Un. kσ k 2 2 each of which is also metallic. As the Hartree-Fock theory We can now calculate the magnetisation density: of the Hubbard model predicts metallic magnetism if the Stoner criterion is satisfied and these materials have nar- m = n↑ − n↓ row bands of strongly interacting electrons it is natural to Z 0 ∗ ∗ ask whether this is a good description of these materials. = d [D↑ ( − µ ) − D↓ ( − µ )] −∞ However, if one extends the above treatment to finite Z 0   1 1  temperatures (30) one finds that the the Hartree-Fock = d D0  − Um + Un − µ theory of the Hubbard model does not provide a good −∞ 2 2   theory of the three elemental magnets. The Curie tem- 1 1 peratures, T , (i.e., the temperature at which the mate- −D0  + Um + Un − µ C 2 2 rial becomes ferromagnetic) of Fe, Co and Ni are ∼1000 2 ≡ f(m) = D0(0)Um + O(m ), (59) K (see, e.g., table 33.1 of Ref. 11). The Hartree-Fock theory predicts that Tc ∼ Um0, where m0 is the mag- where D0() = ∂N0()/∂| is the (DOS; netisation at T = 0. If the parameters in the Hubbard cf. Ref. 11) per spin for U = 0, N0() is the num- model are chosen so that Hartree-Fock theory reproduces 0 ber of electrons (per spin species) for which εk ≤  for the observed m0 then the predicted critical temperature U = 0, Dσ() = ∂Nσ()/∂| is the full interacting DOS is ∼10,000 K. This, order-of-magnitude, disagreement for spin σ electrons, and Nσ() is the number of elec- with experiment results from the failure of the mean-field trons with spin σ for which εkσ ≤ . The standard way Hartree-Fock approximation to properly account for the to solve mean-field theories, known as the method of self fluctuations in the local magnetisation. This is closely consistent solution, is illustrated in Fig. 8. The major related to the (incorrect) prediction of the Hartree-Fock difficulty with self consistent solutions is that it is not approximation that there are no local moments above Tc. possible to establish whether or not one has found all of (Experimentally local moments are observed above Tc.) the self consistent solutions and therefore it is not possi- However, for weak ferromagnets, such as ZrZn2 (TC ∼ 30 ble to establish whether or not one has found the global K) the Hartree-Fock theory of the Hubbard model pro- minimum. Therefore it is prudent to try a wide range vides an excellent description of the observed behaviour of initial guesses for m (or whatever variable the initial (31). guess is made in). The effects missed by Hartree-Fock theory are referred Clearly m = 0 is always a solution of Eq. 59, and to as electronic correlations. The dramatic failure of for UD0(0) < 1 this turns out to be the only solution. Hartree-Fock theory in Fe, Co and Ni shows that elec- But, for UD0(0) > 1 there are additional solutions with tron correlations are very important in these materials, m 6= 0. This is easily understood from the sketch in as do other comparisons of theory and experiment (32).

14 However, it is important to note that mean-field theory is which ensures that there is exactly one boson per site and not limited to Hartree-Fock theory (although the terms therefore that each site is either empty, partially occupied are often, but incorrectly, used synonymously). Rather or doubly occupied, and Hartree-Fock theory is the mean-field theory of the elec- † † ˆ† ˆ tronic density. By constructing mean-field theories of cˆiσcˆiσ − pˆiσpˆiσ − di di = 0, (62b) other properties it is possible to construct mean-field the- ories that capture (some) electronic correlations. We will which ensures that if a site contains a spin σ electron now consider an example of a rather different mean-field then it is either singly occupied (with spin σ) or doubly theory. occupied. Writing the Hubbard Hamiltonian in terms of the slave bosons yields 2. The Gutzwiller approximation, slave bosons & the ˆ X † † X ˆ† ˆ Brinkman-Rice metal-insulator transition HHubbard = −t zˆiσcˆiσcˆjσzˆjσ + U di di, (63) hijiσ i In 1963 Gutzwiller (33) proposed a variational wave- † † ˆ function for the Hubbard model: wherez ˆjσ =e ˆjpˆjσ +p ˆjσdj. Y We now make a mean-field approximation and replace |ΨGi = (1 − αnˆi↑nˆi↓)|Ψ0i the bosonic operators by the expectation values: heii = e, i hpi↑i = hpi↓i = p, hdii = d. Note that we have addition- ! X ally assumed that the system is homogeneous (the ex- = exp −g nˆi↑nˆi↓ |Ψ0i, (60) pectation values do not depend on i) and paramagnetic i (hpi↑i = hpi↓i). Therefore the constraints reduce to where g = − ln(1 − α) is a variational parameter and |e|2 + 2|p|2 + |d|2 = 1 (64a) |Ψ0i is the ground state for uncorrelated electrons. One should note that the Gutzwiller wavefunction is closely and related to the ansatz (1), which is widely n used in both physics and chemistry. Gutzwiller used |p|2 + |d|2 = hcˆ† cˆ i = , (64b) iσ iσ 2 this ansatz to study the problem of itinerant ferromag- netism. This leads to an improvement over the Hartree- where n is the average number of electrons per site. This Fock theory discussed above. However, in 1970 Brinkman amounts to only enforcing the constraints on average. and Rice (34) showed that this wavefunction also de- This theory does not reproduce the correct result for scribes a metal-insulator transition, now referred to as U = 0. However, this deficiency can be fixed ifz ˆjσ is a Brinkman-Rice transition. Rather than studying this replaced by the ‘renormalised’ quantity,z ˜jσ, defined such wavefunction in detail we will instead use an equivalent that technique known as ‘slave bosons’. This has the advan- n − |d|2   tage of making it clear that the Brinkman-Rice transition † 2 p 2 hz˜jσz˜jσi = n n d + 1 − n + |d| . (65) is just a mean-field description of the Mott transition. (1 − 2 ) 2 The ith site in a Hubbard model has four possible states: the site can be empty, |e i; contain a single spin Let us specialise to a ‘half filled’ band, n = 1. The i constraints now allow us to eliminate |p|2 = 1 − |d|2 and σ (=↑ or ↓) electron |σ i; or two electrons, |d i. The 2 i i 2 2 Kotliar-Ruckenstein slave boson technique introduces an |e| = |d| . Thus we find that over-complete description of these states: X 1 2 4 † 2 HˆHubbard ' −t (|d| − 2|d| )ˆc cˆjσ + UN0|d| † 8 iσ |eii =e ˆi |0ii, (61a) hijiσ |σ i =p ˆ† cˆ† |0 i, (61b) 1 X i iσ iσ i = (|d|2 − 2|d|4) ε0 nˆ + UN |d|2, (66) 8 k kσ 0 and kσ ˆ† † † where ε0 is the dispersion for U = 0 and N is the num- |dii = d cˆ cˆ |0ii, (61c) k i i↑ i↓ 2 † 2 ber of lattice sites. Recall that |d| = hdi dii, i.e., |d| is † † ˆ† the probability of site being doubly occupied. We con- wheree ˆi ,p ˆiσ, and di are bosonic creation operators which correspond to empty, partially filled, and doubly struct a variational theory by ensuring that the energy is occupied sites. |0ii is a state with no fermions and no minimised with respect to |d|, which yields bosons on site i; note that this is not a physically real- ∂E 1 X isable state. This transformation is not only kosher, but = (|d| − 4|d|3) ε0 hnˆ i + 2UN |d| = 0. (67) ∂|d| 4 k kσ 0 also exact, so long as we also introduce the constraints kσ † X † ˆ† ˆ Eq. 67 allows one to solve the problem self consistently, eˆi eˆi + pˆiσpˆiσ + di di = 1, (62a) σ cf. Fig. 8. For small U this equation has more than one

15 minimum and the lowest energy state has |d|2 > 0, which non-local correlations, have led to further insights into corresponds to a correlated metallic state (the details of strongly correlated materials (40). Considerable success 0 this minimum depend on εk). But, above some critical has also been achieved by combining DMFT with density U the ground state has |d|2 = 0, which corresponds to no functional theory (41). doubly occupied states, i.e. the Mott insulator. Thus the dependence of the energy on the number holon-doublon 2 pairs (np = |d| ) calculated from the mean-field slave bo- 3. The Nagaoka point son theory is exactly as Mott predicted on rather general grounds (shown in Fig. 6). The Nagaoka point in the phase diagram of the Hub- bard model is the U → ∞ limit when we add one hole to a half filled system. Nagaoka rigourously proved (42; 43) D. Exact solutions of the Hubbard model that at this point the state which maximises the total spin of the system (i.e., the state with hS i = (N − 1)/2, 1. One dimension z for an N site lattice) is an extremum in energy, i.e., either the ground state or the highest lying excited state. On Lieb and Wu (35) famously solved the Hubbard chain most bipartite lattices (cf. Fig. 11) one finds that this at T = 0 using the Bethe ansatz (36; 37). Lieb and ‘Nagaoka state’ is indeed the ground state (43). However, Wu found that the half-filled Hubbard chain is a Mott on frustrated lattices (cf. Fig. 11) the Nagaoka state is insulator for any non-zero U. Nevertheless the Bethe typically only the ground state for one sign of t (44). ansatz solution is not straightforward to understand and It is quite straightforward to understand why the Na- weighty textbooks have been written on the subject (36; gaoka state is often the ground state. As we are consid- 37). ering the U → ∞ limit there will be strictly no double occupation of any sites. One therefore need only consider 2. Infinite dimensions: dynamical mean-field theory the subspace of states with no double occupation. As none of these states contain any potential energy (i.e., As one increases the dimension of a lattice the coor- terms proportional to U), the ground state will be the dination number (the number of nearest neighbours for state that minimises the (the term pro- each lattice site) also increases. In infinite dimensions portional to t). Clearly the ground state is the state that each lattice site has infinitely many nearest neighbours. maximises the magnitude of the kinetic energy with a For a classical model mean-field theory becomes exact in negative sign. In the Nagaoka state all of the electrons infinite dimensions, as the environment (the infinite num- align, this means that the holon can hop unimpeded by ber of nearest neighbours) seen by each site is exactly the the Pauli exclusion principle, thus maximising the mag- same as the mean-field. However, quantum mechanically nitude of the kinetic energy. It is then a simple matter things are complicated by the internal dynamics of the to check whether this is the ground state or the highest site. In the Hubbard model each site can contain zero, lying excited state as we just compare the energy of the one or two electrons, and a dynamic equilibrium between Nagaoka state with that of any other state satisfying the the different charge and spin states is maintained. How- constraint of no double occupation. ever, the environment is still described by a mean-field, Nagaoka’s rigourous treatment has not been extended even though the dynamics are not. Therefore, although to doping by more than one hole and it remains an out- the Hartree-Fock theory of the Hubbard model does not standing problem to further understand this interesting become exact in infinite dimensions, it is possible to con- phenomenon, which shares important features with the struct a theory that treats the on-site dynamics exactly magnetism observed in the elemental magnets (39) and and the spatial correlations at the mean-field level; this many strongly correlated materials (44). theory is known as dynamical mean-field theory (DMFT) (38). The importance of DMFT is not in the, somewhat aca- V. THE HEISENBERG MODEL demic, limit of infinite dimensions. Rather, DMFT has become an important approximate theory in the finite Like the Stoner ferromagnetism we discussed above in numbers of dimensions relevant to real materials (38). It the context of the Hartree-Fock solution for the Hub- has been found that DMFT captures a great deal of the bard model (section IV.C.1) and Hund’s rules (which we physics of strongly correlated electrons. Typically the will discuss in section VI.B), the Heisenberg model is an most important correlations are on-site and therefore are important paradigm for understanding magnetism. The correctly described by DMFT. These include the corre- Heisenberg model does not provide a realistic descrip- lations that are important in metallic magnetism (39) tion of the three elemental ferromagnets (Fe, Co and and many other strongly correlated materials (25; 38). Ni) as they are metals, whereas the Heisenberg model Cluster extensions to DMFT, such as cellular dynami- only describes insulators. However, as we will see in sec- cal mean-field theory (CDMFT) and the dynamical clus- tion V.C, the Heisenberg model is a good description of ter approximation (DCA), which capture some of the Mott insulators such as La2CuO4 (the parent compound

16 of the high temperature superconductors) and κ-(BEDT- We now note that the Hilbert space of the two site TTF)2Cu[N(CN)2]Cl (the parent compound for the or- Heisenberg model is spanned by four states (the spin on ganic superconductors). The Heisenberg model also plays each site may be up or down; in general for an N site an important role in the valence bond theory of the chem- Heisenberg model the Hilbert space is 2N -dimensional). ical bond (45). Further notice that the total spin of the model (Sˆ = In the Heisenberg model one assumes that there is a Sˆ1 + Sˆ2) commutes with the Hamiltonian, therefore the single (unpaired) electron localised at each site, and that eigenstates will also be eigenstates of the total spin. Thus the charge cannot move. Therefore, the only degrees of the four eigenstates must be a singlet, freedom in the Heisenberg model are the spins of each 1 site (the model can also be generalised to spin > 2 ). The 1 1 |Ψsi = √ (| ↑1i| ↓2i − | ↓1i| ↑2i) ≡ √ (| ↑↓i − | ↓↑i) , Hamiltonian for the Heisenberg model is 2 2 (72) ˆ X ˆ ˆ HHeisenberg = JijSi · Sj, (68) and a triplet, ij |Ψ+i = | ↑ i| ↑ i ≡ | ↑↑i (73a) ˆ ˆx ˆy ˆz 1 P † t 1 2 where Si = (Si , Si , Si ) = 2 αβ cˆiα~σαβcˆiβ is the spin 0 1 operator on site i, ~σ = (σ , σ , σ ) is the vector of Pauli |Ψ i = √ (| ↑1i| ↓2i + | ↓1i| ↑2i) x y z t 2 matrices, and Jij is the ‘exchange energy’ between sites 1 i and j. ≡ √ (| ↑↓i + | ↓↑i) (73b) 2 − |Ψt i = | ↓1i| ↓2i ≡ | ↓↓i. (73c) A. Two site model: classical solution It is now straightforward to calculate the total energy In the classical Heisenberg model one replaces the spin of the model for these states, operator, Sˆi, with a classical spin, i.e., a real vector, Si. Thus on two sites, with J12 = J, the energy of the model Es = JhΨs|Sˆ1 · Sˆ2|Ψsi is J   1    = h↑↓ | − h↓↑ | Sˆ+Sˆ− + Sˆ−Sˆ+ + SˆzSˆz (2) 2 2 i j i j i j EHeisenberg = JS1 · S2 = J|S1||S2| cos φ, (69)   where φ is the angle between the two spins (vectors). | ↑↓i − | ↓↑i The classical energy is minimised by φ = π for J > 0 3J = − (74) and φ = 0 for J < 0. Thus for J > 0 the lowest energy 4 solution is for the two spins to point antiparallel (i.e., in opposite directions to one another); we will refer to this and as the antiferromagnetic solution. For J < 0 the lowest + ˆ ˆ + energy solution is for the two spins to point parallel to Et = JhΨt |S1 · S2|Ψt i one another; we will refer to this as the ferromagnetic 0 ˆ ˆ 0 − ˆ ˆ − = JhΨt |S1 · S2|Ψt i = JhΨt |S1 · S2|Ψt i solution. Note that the difference in energy between the 1    antiferromagnetic solution and the ferromagnetic solu- = Jh↓↓ | Sˆ+Sˆ− + Sˆ−Sˆ+ + SˆzSˆz | ↓↓i 2 i j i j i j tion is 2J|S1||S2|; so for S = |S1| = |S2| = 1/2 the J energy difference is J/2. = + (75) 4

B. Two site model: exact quantum mechanical solution Thus we find that the singlet-triplet splitting for the quantum mechanical two site Heisenberg model is J. In order to solve the quantum mechanical version of the two site Heisenberg model it is useful to define the spin raising and lowering operators C. The Heisenberg model as an effective low-energy theory of the Hubbard model ˆ+ ˆx ˆy † Si ≡ Si + iSi =c ˆi↑cˆi↓ (70a) Consider a two site Hubbard model with two electrons ˆ− ˆx ˆy † Si ≡ Si − iSi =c ˆi↓cˆi↑. (70b) and U  t, which is known as the atomic limit. U  t implies θ → 0 in Eq. 54a and that the ground state is Let us denote the state with spin up on site i as | ↑ i and i the Heitler-London state, which is a singlet. The two the state with spin down on site i as | ↓ i. Therefore, i other singlet eigenstates are the charge transfer states, ˆ+ ˆ+ ˆ− ˆ− Si | ↑ii = 0, Si | ↓ii = | ↑ii, Si | ↑ii = | ↓ii and Si | ↓i which have energy ∼ U and so will not participate in i = 0. Further, it is straightforward to confirm that any low-energy processes, i.e., will not be involved in the 1   interesting physics or chemistry. Therefore we can ‘inte- Sˆ · Sˆ = Sˆ+Sˆ− + Sˆ−Sˆ+ + SˆzSˆz. (71) 1 2 2 i j i j i j grate out’ the charge transfer states and derive a simpler

17 model with a smaller Hilbert space. A model derived in this manner is known as an ‘effective low-energy Hamil- tonian’ (see section VIII). In this case we will use second order perturbation theory to derive our effective effective low-energy Hamiltonian. We start by writing the two site Hubbard model as

 t  Hˆ = U Hˆ + Hˆ , (76) Hubbard 0 U 1

ˆ P ˆ P † where H0 = i nˆi↑nˆi↓ and H1 = hijiσ cˆiσcˆjσ. Thus it is clear that the small parameter for our perturbation the- ory is t/U. For t = 0 the ground state is four-fold degen- erate; the four states involved being the Heitler-London state and the triplet states. Formally one should there- fore use degenerate perturbation theory. But, the per- turbation, Hˆ1, does not connect any of the four ground states, therefore, to second order (which is all we will consider), non-degenerate perturbation theory will yield FIG. 10 Sketch of the superexchange processes that lead to the same results. As it simplifies the discussion, we will the effective antiferromagnetic Heisenberg coupling between frame our discussion in terms of non-degenerate pertur- nearest neighbours in the large U/t limit of the half-filled Hub- bard model. These processes lower the energy of the singlet bation theory. It is left as an exercise to the reader to state by 4t2/U as their are four paths, the matrix element show that on adding the appropriate projection opera- between the ground state and the intermediate states is −t, tors (cf. Ref. 46) to perform degenerate perturbation and the intermediate states are higher in energy by U. The theory the result is unchanged. energy of the triplet states is unchanged to by perturbations Consider the related Hamiltonian with Hˆ0 → Hˆ0 + to second order in t/U as the Pauli exclusion principle pre- − η|ΨHLihΨHL| in the limit η → 0 (i.e., as η tends to 0 vents two electrons with the same spin from occupying the from below) this has the same properties as Eq. 76, ex- same site. cept that |ΨHLi is the true ground state and we may use ˆ non-degenerate perturbation theory. hΨHL|H1|ΨHLi = 0 them is non-zero. For an arbitrary lattice to second order so there is no correction to the ground state energy to J = 4|t |2/U. first order in t/U. The second order change in the ground ij ij (2) As the Heisenberg model is the large U/t limit of the state energy, ∆E is given by half-filled Hubbard model electronic correlations are vi- 2 tally important in the physics of materials described by ˆ hΨHL|H1|ψsi the Heisenberg model. Therefore weakly correlated theo- (2) X ∆E = − , (77) ries, such as Hartree-Fock and density functional theory, Es − EHL s6=HL give qualitatively incorrect results. where the sum over s runs over all states except the ground state. Note that, as is true in general, the second D. Frustration: the solution of the three and four site order contribution to the ground state energy is negative. classical Heisenberg models Evaluating the matrix elements (cf. Fig. 10) one finds that Before considering the classical three site model, let us spend a moment discussing the classical four site model. 4|t|2 ∆E(2) = − . (78) We assume that the four sites are situated on the ver- U tices of a square and there is an exchange interaction J In contrast if we add an infinitesimal term to make one of between nearest neighbours (i.e., along the sides of the − 1 1 square), but no interactions between next nearest neigh- the triplet states the true ground state, e.g. 0 |Ψ1ihΨ1|, we find that the Pauli exclusion principle ensures that bours (i.e., along diagonals of the square). The energy of 1 ˆ the model is hΨ1|H1|ψsi = 0 for all s. Thus there is no change in the energy of the triplet state to second order in t/U. (4) X X E = J S · S = J |S ||S | cos θ Therefore it is clear that for U/t → ∞ the half-filled Heisenberg i j i j ij Hubbard model reduces to the Heisenberg model, i.e., hiji hiji the eigenstates and energies are the same, if we set J X = cos θij, (79) J = 4|t|2/U. This result is not a special property of 4 hiji the two site model and is true to second order for an ar- bitrary lattice (47) as second order perturbation theory where θij is the angle between the spins on sites i and j, th only couples sites i and j if the hopping integral between Si is the spin on the i lattice site and in the last equality

18 (3) energy EHeisenberg > −3J/4 and thus one expects the difference in energy between this state and the ferromag- netic state to be < JNz/4. The concept of frustration can also be generalised to itinerant systems where a sim- ilar reduction in the bandwidth of the itinerant electrons is found (44). Having outlined our expectations, let us now consider the three site Heisenberg model more carefully. The en- ergy is given by FIG. 11 Examples of classical spins on (a) a bipartite cluster (3) X and (b) a frustrated cluster. On the square cluster (a) one can E = J S · S . (80) arrange all of the spins antiferromagnetically, i.e., so that each Heisenberg i j spin is antiparallel to all of its nearest neighbours. The same hiji is true for the square lattice. This cannot be accomplished on Without loss of generality we can choose S1 = either the triangular cluster (b) or the triangular lattice. In S1(1, 0, 0); S2 = S2(cos φ2, sin φ2, 0); and S3 = each panel there is an exchange interaction, J, between any S3(cos θ3 cos φ3, cos θ3 sin φ3, sin θ3). Thus, for S1 = S2 = two spins joined by a line. Modified from Ref. 82. S3 = 1/2, (3) EHeisenberg = (81) we have specialised to the case |S | = 1/2 for all i. Notice J h i i cos φ + cos θ cos(φ − φ ) + cos θ cos φ . that the Hamiltonian is just a sum over the ‘bonds’ (sides 4 2 3 2 3 3 3 of the square). As for the two site model (section V.A), Physically we seek the minimum energy, which yields the the solution depends on the sign of J. For J < 0 the conditions lowest energy state is ferromagnetic (all of the spins align (4) (3) parallel to one another) and has energy, E = J. ∂EHeisenberg J h i Heisenberg = sin θ3 cos(φ2 − φ3) + cos φ3 = 0, For J > 0 the lowest energy state is antiferromagnetic ∂θ3 4 (each spin aligns antiparallel to its nearest neighbour; (82a) see Fig. 11a). Thus the four site cluster is split into two (3) ‘sublattices’ with all of the spins parallel to one another ∂EHeisenberg J h i = cos θ3 sin(φ2 − φ3) − sin φ3 = 0 within the same sublattice and antiparallel to spins on ∂φ3 4 the other sublattice. The antiferromagnetic arrangement (82b) of spins therefore has energy E(4) = −J. Thus we Heisenberg and find that the energy difference between the ferromagnetic (3) and antiferromagnetic arrangements is 2J. ∂EHeisenberg J h i = − cos θ3 sin(φ2 − φ3) + sin φ2 = 0. A lattice that can be split, as described above, into ∂φ2 4 two sublattices such that all nearest neighbours are on (82c) different sublattices is referred to as bipartite. Both the four site (square) and two site lattices are bipartite. For For J < 0 the global minimum is, unsurprisingly, θ3 = bipartite lattices the energy difference between the ferro- φ2 = φ3 = 0, i.e., ferromagnetism. The energy of the magnetic and antiferromagnetic arrangements is JNz/4, ferromagnetic state is 3J/4. For J > 0 there are sev- where z is the coordination number of the lattice and N eral degenerate minima, which all show the same physics. is the number of lattice sites. This is because the energy For simplicity we will just consider the minimum θ3 = 0, of each bond can be optimised regardless of what hap- φ2 = 2π/3, and φ3 = 4π/3. In this solution each of o pens to other bonds for either sign of J. It can be seen the spins points 120 away from each of the other spins, o from Fig. 11 that the triangular lattice is not bipartite; hence this is known as the 120 state. It is left as an ex- this leads to significant differences in its physics. ercise to the reader to identify the other solutions of Eqs. 82, to show that there are none with lower energy than Before analysing this model mathematically let us con- those discussed above, and to show that all of the degen- sider some of those differences. Clearly for J < 0 erate solutions are physically equivalent. The energy of it is straightforward to arrange the spins ferromagneti- the 120o state is −3J/8 and hence the energy difference cally. Further, as the energy of each of the three bonds between the ferromagnetic state and 120o state is just will be optimised in this arrangement we expect the to- (3) 9J/8, i.e., less than we would expect (JNz/4 = 3J/2 for tal energy of this state to be EHeisenberg = 3J/4 for N = 3, z = 2) for a bipartite lattice. S = 1/2. But, as is shown in Fig. 11b, one cannot arrange three spins antiferromagnetically on a triangular lattice. Thus, for J > 0, we cannot optimise the energy E. Three site model: exact quantum mechanical solution of each bond individually. When this is the case one says that the lattice is ‘frustrated’. For a frustrated lattice Group theory, the mathematics of symmetry, allows with S = 1/2 we expect the solution for J > 0 to have one to solve the quantum spin 1/2 three site Heisen-

19 berg model straightforwardly. Unfortunately space does Each of these states have energy E = −5J/4 and they not permit an introduction to the relevant group theory. are the (degenerate) ground states for J > 0. Thus, Therefore the reader who is not familiar with the math- the energy difference between the highest spin state and ematics is advised either to refer to one of the many ex- the lowest spin state is 2J. From the solution to the cellent textbooks on the subject (e.g., Refs. 12; 13) or, two site model (section V.B) we expected each of the failing that, to simply check that the wavefunctions de- three bonds to yield an energy difference of J between the rived by the group theoretic arguments below are indeed lowest and highest spin states. Thus the frustration has eigenstates. a similar effect in both the quantum and classical models, The Hamiltonian is i.e. frustration lowers the energy difference between the ˆ(3) X ˆ ˆ highest spin and lowest spin states. HHeisenberg = J Si · Sj hiji X 1    F. The Heisenberg model on infinite lattices = J Sˆ+Sˆ− + Sˆ−Sˆ+ + SˆzSˆz (83) 2 i j i j i j hiji The Heisenberg model can be solved exactly in one di- We begin by noting that 2 ⊗ 2 ⊗ 2 = 2 ⊕ 2 ⊕ 4,4 i.e., a mension, and we will discuss this further below, but not system formed from three spin 1/2 particles will have two in any other finite dimension. However, in more than doublets (with two-fold degenerate spin 1/2 eigenstates) one dimension physics of the Heisenberg model is typ- and one quadruplet (with four-fold degenerate spin 3/2 ically very different from that in one dimension, so we eigenstates). will begin by discussing, qualitatively, the semi-classical There are only four possible quadruplet states consis- spin wave approximation for the Heisenberg model, which 5 captures many important aspects of magnetism. A quan- tent with the C3 point group symmetry of the model. Each of these belong to the A irreducible representation titative formulation of this theory can be found in many textbooks, e.g. Refs. 11; 30. of C3. They are In inelastic neutron scattering experiments a neutron E ψ3/2 = |↑↑↑i may have its spin flipped by its interaction with the mag- 3/2 net; this causes a spin 1 excitation in the material. The E 1 ψ1/2 = √ (|↓↑↑i + |↑↓↑i + |↑↑↓i) conceptually simplest spin 1 excitation would be to flip 3/2 3 1 one (spin- 2 ) spin; in a one dimensional ferromagnetic −1/2E 1 Heisenberg model this state has energy 2|J| greater than ψ = √ (|↑↓↓i + |↓↑↓i + |↓↓↑i) 3/2 3 the ground state. However, a much lower energy excita- E tion is a ‘spin wave’, where each spin is rotated a small ψ−3/2 = |↓↓↓i , 3/2 amount from its nearest neighbours (cf. Fig. 12). In a one dimensional ferromagnetic Heisenberg model spin where |αβγi = |Sz,Sz,Szi and α, β and γ =↑ or ↓. Each 1 2 3 waves have excitation energies of ω = 2|J|[1−cos(ka)], of these states have energy E = 3J/4 and they are the ~ k where a is the lattice constant (30). Note, in particular, (degenerate) ground states for J < 0. that the excitation energy vanishes for long wavelength We are left with the four doublet states. These belong (small k) spin waves. This spin wave spectrum can in- to the two dimensional E irreducible representation of C 3 deed be observed directly in neutron scattering exper- and, as the Hamiltonian is time reversal symmetric, all iments from suitable materials (48), and the spectrum four doublet states are degenerate. Explicitly the states is found to be in good agreement with the predictions of are the semi-classical theory in many materials. One can also E 1   ψ1/2 = √ |↓↑↑i + ei2π/3 |↑↓↑i + e−i2π/3 |↑↑↓i quantise the semi-classical theory by making a ‘Holstein- 1/2 3 Primakoff’ transformation (30). This yields a descrip- E 1   tion of the low-energy physics of the Heisenberg model ψ−1/2 = √ |↑↓↓i + ei2π/3 |↓↑↓i + e−i2π/3 |↓↓↑i 1/2 3 in terms of non-interacting bosons, which are known as ‘magnons’ and have the same dispersion relation as the 1/2E 1  −i2π/3 i2π/3  ψ˜ = √ |↓↑↑i + e |↑↓↑i + e |↑↑↓i classical spin waves. Similar spin wave and magnon de- 1/2 3 scriptions can be straightforwardly constructed for the E 1   ψ˜−1/2 = √ |↑↓↓i + e−i2π/3 |↓↑↓i + ei2π/3 |↓↓↑i antiferromagnetic Heisenberg model (30). 1/2 3 The effective low-energy physics of the one dimensional Heisenberg model is, as noted above, rather different from the semi-classical approximation. To understand this it is helpful to think of the Heisenberg model as a 4 In this notation the integers are the degeneracy of the state. special case of the ‘XXZ model’: 5 One might, reasonably, take the view that the model has either D3h or C3v. In fact the arguments in this section go through X x x y y  X z z almost identically for either of these symmetries (with appropri- HXXZ = Jxy Si Si+1 + Si Si+1 + Jz Si Si+1, ate changes in notation) due to the homomorphisms from these i i groups to C3. We will use C3 notation for simplicity. (86)

20 FIG. 12 Sketches of (a) the classical ground state of a ferromagnetic Heisenberg chain and (b) a spin wave excitation in the same model.

which reduces to the Heisenberg model for Jxy = Jz = J. For Jz < Jxy < 0 the model displays an exotic quantum phase known as a Luttinger liquid. (At Jxy = Jz the model undergoes a quantum phase transition from the Luttinger liquid to an ordered phase (49)). On the energy scales relevant to chemistry one does not need to worry about the fact that protons and neutrons are made up of smaller particles (quarks). This is be- cause the quarks are confined within the proton/neutron (50). Similarly, in a normal magnet it does not matter 1 that the material is made up of spin- 2 particles (elec- trons). As described above, on the energy scales relevant to magnets the spins are confined into spin one particles called magnons. However, magnons can be described in 1 terms of two spin- 2 spinons, which are confined inside the magnon. In the Luttinger liquid the spinons are de- confined, i.e., the spinons can move independently of one FIG. 13 Sketch of spinons in a 1D spin chain. (a) Local antiferromagnetic correlations. (b) A neutron scattering off another (cf. Fig. 13). As the magnon is a composite par- the chain causes one spin (circled) to flip. (c, d) Spontaneous ticle made from two spinons this is often referred to as flips of adjacent pairs of spins due to quantum fluctuations fractionalisation. A key prediction of this theory is that allow the spinons (circled) to propagate independently. A the spinons display a continuum of excitations in neutron key open question is: can this free propagation occur in 2D, scattering experiments (as opposed to the sharp disper- or do interactions confine the spinons? Modified from Ref. sion predicted for magnons). The two spinon continuum 82. has indeed been observed in a number of quasi-one di- mensional materials. (51) An open research question is: does fractionalisation occur in higher dimensions? Because of the success of spin wave theory (which implies confined spinons) in describing magnetically ordered materials one does not expect fractionalisation in materials with magnetic or- der. Therefore, one would like to investigate quasi-two or three dimensional materials whose low-energy physics is described by spin Hamiltonians (such as the Heisen- berg model), but that do not order magnetically even at the lowest temperatures. Such materials are collec- until very recently, (52) but there is now evidence for tively referred to as spin liquids. There is a long his- spin liquids in the triangular lattice compound κ-(BEDT- tory of theoretical contemplation of spin liquids, which TTF)2Cu(CN)3 (25; 53), the kagome lattice (cf. Fig. 4) suggests that frustrated magnets and insulating systems compound ZnCu3(OH)6Cl2 (54) and the hyperkagome near to the Mott transition are strong candidates to dis- lattice compound Na4Ir3O8 (55). It remains to be seen play spin liquid physics. However, evidence for real ma- whether any of these materials support fractionalised ex- terials with spin liquid ground states has been scarce citations.

21 VI. OTHER EFFECTIVE LOW-ENERGY B. Larger basis sets and Hund’s rules HAMILTONIANS FOR CORRELATED ELECTRONS Thus far we have focused mainly on models with one A. Complete neglect of differential overlap, the orbital per site. Often this is not appropriate, for exam- Pariser-Parr-Pople model & extended Hubbard models ple, if one were interested in chemical bonding or materi- als containing transition metals. Many of the models dis- We now consider another model for which the quan- cussed in these notes can be straightforwardly extended tum chemistry and condensed matter physics communi- to include more than one orbital per site. However, while ties have different names. These models belong to class writing down models with more than one orbital per site of models known as complete neglect of differential over- is not difficult, these models do contain significant addi- lap (CNDO). For a pair of orthogonal states, φ(x) and tional physics. Some of the most important effects are ψ(x), the integral over all space of the overlap of the known as Hund’s rules (1). These rules have important R ∞ two wavefunctions vansishes, i.e., −∞ φ(x)ψ(x)dx = 0. experimental consequences from atomic physics to biol- If the differential overlap vanishes then the overlap of ogy. In order to examine Hund’s rules let us consider the the two wavefunctions vanishes at every point in space, atomic limit (t = 0) of an extended Hubbard model with R x0+δ i.e., limδ→0 φ(x)ψ(x)dx = 0 for all x0. The CNDO two electrons in two orbitals per site: x0 approximation is simply to assume that the differential ˆ X ˆ ˆ overlap between all basis states is negligible. Thus CNDO HeH1s2o = U nˆµ↑nˆµ↓ + V0nˆ1nˆ2 + JH S1 · S2, (91) µ implies that Vijkl = Viikkδijδkl (cf. section II.B). Thus the general CNDO Hamiltonian is † where µ = 1 or 2 labels the orbitals,n ˆµσ =c ˆµσcˆµσ, P ˆ P † ˆ X † X nˆµ = σ nˆµσ, Sµ = αβ cˆµα~σαβcˆµβ, U is the Coulomb HCNDO = − tijcˆiσcˆjσ + Vijnˆiσnˆjσ0 , (87) repulsion between two electrons in the same orbital, V0 is 0 ijσ ijσσ the Coulomb repulsion between two electrons in different orbitals, and J is the ‘Hund’s rule coupling’ between † H where Vij ≡ Viijj and the number operatorn ˆiσ ≡ cˆiσcˆiσ. electrons in different orbitals. Notice that the Hund’s The Pariser-Parr-Pople (PPP) model is the CNDO ap- rule coupling is an exchange interaction between orbitals. proximation in a basis that only includes the π electrons. Further, if we compare the Hamiltonian with the defini- Often a H¨uckel-like notation is used with Vij = γij, thus tion given in Eq. 21 we find that Z Z ˆ X † X † X 3 3 ∗ HPPP = αicˆiσcˆiσ + βijcˆiσcˆjσ + γijnˆiσnˆjσ0 . − JH = d r1 d r2 φ1(r1)φ2(r1) iσ ijσ ijσσ0 ∗ (88) ×V (r1 − r2)φ2(r2) φ1(r2) Z Z The extended Hubbard model, as with the plain Hub- ∼ d3r d3r |φ (r )|2 V (r − r ) |φ (r )|2 bard model, is typically studied in a basis with one orbital 1 2 1 1 1 2 2 2 per site. Further, one often makes the approximation ≥ 0. (92) that Vii = U, Vij = V if i and j are nearest neighbours and Vij = 0 otherwise. This yields as V (r1−r2) is positive semidefinite. Therefore, typically, JH < 0, i.e., the Hund’s rule coupling favours the parallel X † X X alignment of the spins in a half-filled system. Hˆ = − t cˆ cˆ + U nˆ nˆ + V nˆ nˆ 0 . eH ij iσ jσ i↑ i↓ iσ jσ U is the largest energy scale in the problem, so, for sim- i 0 hijiσ hijiσσ plicity, let us consider the case U → ∞. For J = 0 there (89) H are four degenerate ground states: a singlet, √1 (| ↑↓ One can, of course, go beyond CNDO. The most gen- 2 eral possible model for two identical sites with a single i − | ↓↑i) (where the first arrow refers to the spin of orbital per site is the electron in orbital 1 and the second arrow refers to the spin in orbital 2), and a triplet, | ↑↑i, | ↓↓i and √1 (| ↑↓i − | ↓↑i). But, for J > 0 the energy of the triplet ˆ X  † †  2 HeH2 = − [t − X (ˆn1σ +n ˆ2σ)] cˆ1σcˆ2σ +c ˆ2σcˆ1σ states is JH lower than that of the singlet state. Indeed σ spin symmetry implies that even if we relax the condition X +U nˆi↑nˆi↓ + V nˆ1nˆ2 + JSˆ1 · Sˆ2 U → ∞ the triplet state remains lower in energy than the i singlet state as physically we require U > JH . One can   +P cˆ† cˆ† cˆ cˆ +c ˆ† cˆ† cˆ cˆ , (90) repeat this argument for any number of electrons in any 1↑ 1↓ 2↑ 2↓ 2↑ 2↓ 1↑ 1↓ number of orbitals and one always finds that the highest spin state has the lowest energy. However, if one studies P ˆ P † wheren ˆi = σ nˆiσ, Si = αβ cˆiα~σαβcˆiβ, ~σαβ is the vec- models with more than one site and moves away from tor of Pauli matrices, J is the direct exchange interaction, the atomic limit (t = 0) one finds that there is a sub- X is the correlated hopping amplitude, and P is the pair tle competition between the kinetic (hopping) term and hopping amplitude. the Hund’s rule coupling which means that the high spin

22 FIG. 14 Sketch of the toy model for a transition metal oxide, Hamiltonian 94, with two transition metal sites (1 and 2) and a single oxygen site (O). state is not always the lowest energy state. Many such interesting effects can be understood on the basis of a two site generalisation of this two orbital model (56).

C. The ionic Hubbard model

Thus far we have assumed that all sites are identical. Of course, this is not always true in real materials. In a compound more than one species of atom may con- tribute to the low-energy physics (57) or different atoms FIG. 15 Sketch of the processes described by Hamiltonian 94 of the same species may be found at crystallographical that give rise to the effective hopping integral between the two transition metal atom sites. distinct sites (44; 58). A simple model that describes this situation is the ionic Hubbard model:

X † X X as sketched in Fig. 14, which is just the ionic Hubbard HˆiH = −t cˆ cˆjσ + U nˆi↑nˆi↓ + inˆiσ (93) iσ model with U = 0 and ∆ = 1 − O = 2 − O > 0. hijiσ i iσ With three electrons in the system and t = 0 the ground state is four-fold degenerate, the ground states have two where  = t is the site energy, which will be taken i ii electrons on the O atom and the other electron on one to be different on different sites. Note that in the ionic of the metal atoms. If we now consider finite, but Hubbard model all sites are assumed to have the same small, t  ∆ we can construct a perturbation theory U. in t/∆. One finds that there is a splitting between An important application of the ionic Hubbard model † † † † the bonding, √1 (ˆc +c ˆ )ˆc cˆ |0i and antibonding, is in describing transition metal oxides (57). Typically 2 1σ 2σ O↑ O↓ 1 † † † † i is larger on the transition metal site than on the oxy- √ (ˆc − cˆ )ˆc cˆ |0i, states. The processes that lead 2 1σ 2σ O↑ O↓ gen site, therefore the oxygen orbitals are nearly filled. to this splitting are sketched in Fig. 15. Therefore our This means that there is a low hole density in the oxygen effective low-energy Hamiltonian is a tight binding model orbitals and, hence, that electronic correlations are less involving just the metal atoms: important for the electrons in the oxygen orbitals than for electrons in transition metal orbitals. If the difference ˆ ∗ X  † †  Heff = −t cˆ1σcˆ2σ +c ˆ2σcˆ1σ , (95) between i on the oxygen sites and i on the transition σ metal sites is large enough then the oxygen orbitals are completely filled in all low-energy states and therefore where, to second order in t/∆, the effective metal-to- need not feature in the low-energy description of the ma- metal hopping integral is given by terial. However, just because the oxygen orbitals do not t2 appear explicitly in the effective low-energy description t∗ = − . (96) of the material, does not mean that the oxygen does not ∆ have a profound effect on the low-energy physics. Note that, even though t is positive, t∗ < 0 (or, equiva- To see this consider a toy model with two metal sites lently, β∗ > 0), in contrast to our na¨ıve expectation that (labelled 1 and 2) and one oxygen site (labelled O), whose hopping integrals are positive (β < 0; cf. section III). Hamiltonian is

ˆ X  † † † †  HiH3 = −t cˆ1σcˆOσ +c ˆOσcˆ1σ +c ˆ2σcˆOσ +c ˆOσcˆ2σ VII. THE HOLSTEIN MODEL σ X ∆ + (ˆn +n ˆ − nˆ ) (94) So far we have assumed that the nuclei or ions form 2 1σ 2σ Oσ iσ a passive background through which the electrons move.

23 P † However, in many situations this is not the case. Atoms iµ ~ωiµaˆiµaˆiµ. Thus move and these lattice/molecular vibrations interact with ˆ X † X † the electrons via the electron-phonon/vibronic interac- HHolstein = −t cˆiσcˆjσ + ~ωiµaˆiµaˆiµ tion. One of the simplest models of such effects is hijiσ iµ the Holstein model, which we discuss below. Electron- X † † vibration interactions play important roles across sci- + giµ(ˆaiµ +a ˆiµ)ˆciσcˆiσ (100) ence. In physics electron-phonon interactions can give iσµ rise to superconductivity (59), spin and charge density waves (60), polaron formation (61) and piezoelectricity A. Two site Holstein model (59). In chemistry vibronic interactions impact electron- transfer processes (62), Jahn-Teller effects (63), spec- If we assume that there is only one electron and one troscopy (63), stereochemistry (63), activation of chem- mode per site then the Holstein model simplifies to ical reactions (63) and catalysis (63). In biology the vibronic interactions play important roles in photopro- ˆ X  † †  X † HHolstein = −t cˆ1σcˆ2σ +c ˆ2σcˆ1σ + ~ω aˆi aˆi tection (64), photosynthesis (65) and vision (66). It is σ i therefore clear that one of the central tasks for condensed X † matter theory and theoretical chemistry is to describe +g (ˆai +a ˆi)ˆni (101) electron-vibration interactions. i In general one may write the Hamiltonian of a system † on two symmetric sites, wheren ˆ = P nˆ = P cˆ cˆ . of electrons and nuclei as i σ iσ σ iσ iσ It is useful to change the basis in which we consider Hˆ = Hˆ + Hˆ + Hˆ , (97) the phonons√ to that of in phase (symmetric),s ˆ = e n en ˆ (ˆa1 +a ˆ2)/√ 2, and out of phase (antisymmetric), b = where Hˆe contains those terms that only effect the elec- (ˆa1 − aˆ2)/ 2, vibrations. In this basis one finds that trons, Hˆn contains those terms that only effect the nuclei HˆHolstein = Hˆs + Hˆbe (102a) and Hˆen describes the interactions between the electrons ˆ and the nuclei. He might be any of the Hamiltonians we where have discussed above. However, for the Holstein model ˆ ˆ † g † one assumes a tight-binding form for He. In the normal Hs = ~ωsˆ sˆ + √ (ˆs +s ˆ)(ˆn1 +n ˆ2) (102b) mode approximation (63), which we will make, one treats 2 molecular/lattice vibrations as harmonic oscillators (cf. and section II.A). As the ions carry a charge, any displace- ˆ X  † †  ˆ†ˆ ment of the ions from their equilibrium positions will Hbe = −t cˆ1σcˆ2σ +c ˆ2σcˆ1σ + ~ωb b change the potential felt by the electrons. The Holstein σ model assumes that each vibrational mode is localised g ˆ† ˆ +√ (b + b)(ˆn1 − nˆ2). (102c) on a single site. For this to be the case the site must 2 have some internal structure, i.e., the site cannot corre- Note thatn ˆ +n ˆ = N, the total number of electrons spond to a single atom. Therefore the Holstein model is 1 2 in the problem. As N is a constant of the motion the more appropriate for a molecular solids than for simple dynamics of the electrons cannot effect the symmetric crystals. For small displacements, x , of the µth mode iµ vibrations and vice versa. Hence all of the interesting of the ith lattice site we can perform a Taylor expansion effects are contained in Hˆ and we need only study this in the dimensionless normal coordinate of the vibration, be p Hamiltonian below. Qiµ = xiµ miµωiµ/~ where miµ and ωiµ are, respec- tively, the mass and the frequency of the µth mode on th the i site, and we find that 1. Diabatic limit, ~ω  t

X ∂tij   Hˆ = Q cˆ† cˆ +c ˆ† cˆ + .... (98) In the diabatic limit the vibrational modes are assumed en ∂Q iµ iσ jσ jσ iσ ijσµ iµ to instantaneously adapt themselves to the particle’s po- sition. Thus In the Holstein model one assumes that the derivative ˆ†ˆ g ˆ† ˆ ˆ†ˆ g ˆ† ˆ vanishes for i 6= j. We may quantise the vibrations in ~ωb b+ √ (b +b)(ˆn1 −nˆ2) = ~ωb b± √ (b +b). (103) 2 2 the usual way (cf. section II.A) which yields The plus sign is relevant when the electron is located on ˆ X † † site 1 and the minus sign is relevant when the electron Hen = giµ(ˆaiµ +a ˆiµ)ˆciσcˆiσ, (99) iσµ is on site 2. We now introduced the ‘displaced oscillator transformation’, wherea ˆ(†) destroys (creates) a quantised vibration in the iµ ˆ† ˆ† 1 g th th −1/2 ˆ b± = b ± √ . (104) µ mode on the i site, giµ = 2 ∂tii/∂Qiµ and Hn = 2 ~ω

24 Therefore we find that 2. Adiabatic limit, ~ω  t ˆ X  † †  Hbe = −t cˆ1σcˆ2σ +c ˆ2σcˆ1σ We begin by noting that, as there is only one electron σ the spin of the electron only leads to a trivial two-fold 2 ˆ† ˆ ˆ† ˆ  g degeneracy, and therefore can be neglected without loss +~ω b+b+ + b−b− − . (105) ~2ω2 of generality. A useful notational change is to introduce a pseudospin notation where we defineσ ˆ =c ˆ† cˆ −cˆ† cˆ ˆ ˆ z 1σ 1σ 2σ 2σ It is important to note that the operators b+ and b− † † andσ ˆx =c ˆ1σcˆ2σ +ˆc2σcˆ1σ. Therefore the one electron, two satisfy the same commutation relations as the ˆb operator, site Holstein model Hamiltonian becomes therefore they describe bosonic excitations. We define g ˆ ˆ ˆ†ˆ ˆ† ˆ the ground states of the displaced oscillators by b−|0−i = Hsb = −tσˆx + ~ωb b + √ (b + b)ˆσz. (111) ˆ 2 0 and b+|0+i = 0. Therefore which is often referred to as the spin-boson model. ˆ 1 g b|0+i = −√ |0+i (106) Let us now replace the bosonic operators by position 2 ~ω and momentum operators for the harmonic oscillator de- and hence fined as √ r ~  †  ˆ 2g xˆ = ˆb + ˆb (112a) b−|0+i = − |0+i; (107a) 2mω ~ω similarly and √ r m ω   ˆ 2g pˆ = i ~ ˆb† − ˆb . (112b) b+|0−i = |0−i, (107b) 2 ~ω √ ˆ Therefore i.e., |0±i is an eigenstate of b∓ with eigenvalue ∓ 2g/~ω. The eigenstates of bosonic annihilation operators are 2 r pˆ 1 2 mω known as coherent states (67). Eqs. 107 therefore show Hˆsb = −tσˆx + + mωxˆ + g xˆσˆz. (113) ˆ 2m 2 ~ that the ground state of one of the b± operators may be written as a coherent state of the other operator (68), The adiabatic limit is characterised by a sluggish i.e., bosonic bath that responds only very slowly to the mo- tion of the electron, i.e.,p ˆ2/2m → 0, which it is often " √  # 2g 1 ˆ† helpful to think of as the m → ∞ limit. Further, in |0±i = exp − ± b∓ |0∓i. (108) ~ω 2 the adiabatic limit the Born-Oppenheimer approxima- tion (3; 68) holds, which implies that the total wavefunc- Therefore tion of the system, |Ψi, is a product of a electronic (pseu- dospin) wavefunction, |φ i, and a vibrational (bosonic)  2  e g wavefunction, |ψ i, i.e., |Ψi = |φ i⊗|ψ i. Therefore, the h0+|0−i = exp − , (109) v e v ~2ω2 harmonic oscillator will be in a position eigenstate and we may replace the position operator,x ˆ, by a classical which is known as the Franck-Condon factor. position x, yielding The Franck-Condon factor describes the fact that, in the diabatic limit, the bosons cause a ‘drag’ on the elec- r ˆ mω 1 2 tronic hopping. That is, we can describe the solution of Hsb = −tσˆx + g xσˆz + mωx (114a) 2 the diabatic limit in terms of an effective two site tight ~  gp mω x −t  1 binding model if we replace t by = ~ + mωx2,(114b) −t −gp mω x 2  2  ~ ∗ g t = th0+|0−i = t exp − 2 2 . (110) where in the second line we have simply switched to the ~ ω matrix representation of the Pauli matrices. This is easily Thus the hopping integral is ‘renormalised’ by the inter- solved and one finds that the eigenvalues are actions of the electron with the vibrational modes (cf. r section VIII). This renormalisation is also found in the 1 2 2 mω 2 2 E± = mωx ± t + g x (115a) solution of an electron moving on a lattice in the diabatic 2 ~ limit. In this context the exponential factor is known as 1 mωg2x2 the polaronic band narrowing (61). The exponential fac- ≈ mωx2 ± ± t, (115b) 2 2 t tor results from the small overlap of the two displaced ~ operators, and may be thought of as an increase in the where Eq. 115b holds in the weak coupling limit, gx  t. effective mass of the electron. We plot the variation of these eigenvalues with x in this

25 E relativistic corrections are typically unimportant, thus all - of the interactions boil down to non-relativistic electro- 12 E+ magnetic effects. 10 Dirac’s worldview is realised in the ab initio approach 8 to . Wherein one starts from the Hartree-Fock solution to the full Schr¨odingerequation 6 in some small basis set. One then adds in correlations 4 via increasingly complex approximation schemes and in- creases the size of the basis set, in the hope that with a 2 sufficiently large computer one will find an answer that 0 is “sufficiently close” to the exact solution (full CI in an infinite complete basis set). -2 In the last few decades rapid progress has been made x in ab initio methods due to an exponential improvement in computing technology, methodological progress, and FIG. 16 The energies of the ground and excited states for a single electron in the two site Holstein model in the adiabatic the widespread availability of implementations of these methods (71). However, this progress is unsustainable: weak coupling limit (t  g  ~ω). Calculated from Eq. 115a. x is the position of the harmonic oscillator describing out of the complexity recognised by Dirac eventually limits the phase vibrations. accuracy possible from ab initio calculations. Indeed, solving the Hamiltonian given in Eq. 17 is known to be computationally hard. Feynman proposed building a limit in Fig. 16. Notice that for the electronic ground computer that uses the full power of quantum mechanics state, E−, the lowest energy states have x 6= 0. This to carry out quantum simulations (72). Indeed, the sim- is an example of spontaneous symmetry breaking (69), plest of all quantum chemical problems, the H2 molecule as ground state of a system has a lower symmetry than in a minimal basis set, has been solved on a prototype the Hamiltonian of the system. Thus the system must quantum computer (73). But, while even a rather small “choose” either the left well or the right well (but not scale quantum computer [containing just a few hundred both) in order to minimise its energy. qubits (73)] would provide a speed-up over classical com- putation, it is believed that the solution of Hamiltonian 17 remains hard even on a quantum computer (i.e., it is VIII. EFFECTIVE HAMILTONIAN OR SEMI-EMPIRICAL believed that even a quantum computer could not solve MODEL? Hamiltonian 17 in a time that grows only polynomially with the size of the system (74)). Further, simple exten- The models discussed in these notes are generally sions of these arguments provide strong reasons to believe known as semi-empirical models in a chemical context that there is no efficiently computable approximation to and as effective Hamiltonians in the physics community. the exact functional in density functional theory (74). Here the difference is not just nomenclature, but is also Therefore it appears that the equations will always re- indicative of an important difference in the epistemolog- main “too complex to be solved” directly. This suggests ical status awarded to these models by the two commu- that semi-empirical models will always be required for nities. In this section I will describe two different atti- large systems. tudes towards semi-empirical models/effective Hamilto- nians and discuss the epistemological views embodied in the work of two of the greatest physicists of the twentieth B. The Wilsonian project century. Typically one is only interested in a few low-energy states of a system, perhaps the ground state and the first A. The Diracian worldview few excited states. Therefore, so long as our model gives the correct energies for these low-energy states we should Paul Dirac famously wrote (70) that “the fundamental regard it as successful. This, apparently simple, realisa- laws necessary for the mathematical treatment of a large tion, particularly as embodied by Wilson’s renormalisa- part of physics and the whole of chemistry are thus com- tion group (75), has had profound implications through- pletely known, and the difficulty lies only in the fact that out modern physics from high energy particle physics to application of these laws leads to equations that are too condensed matter physics. complex to be solved.” There is clearly a great deal of The basic idea of renormalisation in remarkably sim- truth in the statement. In solid state physics and chem- ple. Imagine starting with some system that has a large istry we know that the Schr¨odingerequation provides an number of degrees of freedom. As we have noted, for extraordinarily accurate description of the observed phe- practical purposes we only care about about the lowest nomena. Gravity, the weak and strong nuclear forces and energy states. Therefore one might be tempted to sim-

26 plify the description of the system by discarding the high- the ground state energy. It is therefore clear why the est energy states. However, simply discarding such states minimal basis set gives such a poor result, it ignores all will cause a shift in the low-energy spectrum. Therefore, the higher order corrections to the total energy. one must remove the high energy states that complicate The failure of the simple minimal basis set calcula- the description and render the problem computationally tion does not, however, mean that the effective Hamil- intractable in such as way as to preserve the low-energy tonian approach also fails, despite the fact that the ef- spectrum. This is often referred to as ‘integrating out’ fective Hamiltonian is also in an extremely small basis the high energy degrees of freedom (because of the way set. Rather, one must realise that, as well as the first this process is carried out in the path integral formulation order contributions, U also contains contributions from of quantum mechanics (76)). Typically integrating out higher orders in perturbation theory. It is therefore pos- the high energy degrees of freedom causes the parame- sible, although extremely computationally demanding, to ters of the Hamiltonian to ‘flow’ or ‘run’, i.e., change their calculate the parameters for effective Hamiltonians from values. When this happens one says that the parameters this kind of perturbation theory (77). are renormalised. A more promising approach, which has been applied to A simple example is the Coulomb interaction between a number of molecular crystals (78; 79), is to use atom- the two electrons in a neutral Helium atom. For sim- istic calculations to parameterise an effective Hamilto- plicity lets imagine trying to calculate just the ground nian. For example, density functional theory gives quite state energy. We begin by analysing the problem in the reasonable values for the total energy of the ground state absence of a Coulomb interaction between the two elec- of many molecules. Therefore one approach to calcu- trons. In the ground state both electrons occupy the 1s lating the Hubbard U is to calculate the ionisation en- orbital. We would like to work in as small a basis set as ergy, I = E0(N − 1) − E0(N), and the electron affin- possible. The simplest approach is just to work in the ity, A = E0(N) − E0(N + 1), of the molecule, where minimal basis set, which, in this case, is just the two 1s E0(n) is the ground state energy of the molecule when spin-orbitals, φ1sσ(r). The total energy of a He atom ne- it contains n electrons and N is the filling correspond- glecting the inter-electron Coulomb interaction is -108.8 ing to half-filled band. One finds that U = I − A = eV (relative to the completely ionised state). Now we re- E0(N + 1) + E0(N − 1) − 2E0(N). A simple way to see store the Coulomb repulsion between electrons. A simple this is that if we assume the molecule is neutral when question is: how much does this change the total energy it contains N electrons then U corresponds to the en- of the He atom? In the minimal basis set the solution ergy difference in the charge disproportionation reaction seems straightforward: + − 2M M + M for two well separated molecules, M. A more extensive discussion of this approach is given in Z ∞ Z ∞ e2|φ |2|φ |2 h1s2|V |1s2i = d3r d3r 1s↑ 1s↓ Ref. 78. 1 2 4π |r − r | −∞ −∞ 0 1 2 It is worth noting that we have actually carried out this ' 34.0 eV. (116) program of parametrising effective Hamiltonians three times in the discussion above. In section V.C we showed Therefore it is tempting to conclude that we can model that the Heisenberg model is an effective low-energy the He atom by a one site Hubbard model with U = 2 2 model for the half-filled Hubbard model in the limit h1s |V |1s i. However, this yields a total energy for the t/U → 0. In section VI.C we derived an effective tight He atom of -74.8 eV, which is not particularly close to binding model that involved only the metal sites from an the experimental value of -78.975 eV (8). ionic Hubbard model of a transition metal oxide. Finally, Let us then continue to consider the problem in the in section VII.A.1 we showed that vibronic interactions basis set of the hydrogenic atom, which is complete due lead to an effective tight binding model describing the to the spherical symmetry of the Hamiltonian. One can low-energy physics of the Holstein model in the diabatic now straightforwardly carry out a perturbation theory limit, and that in this model the (electron- around the non-interacting electron solution where we like excitations) are polarons, a bound state of electrons take and vibrational excitations with a mass enhanced over 2  2 2 2  that of the bare electron. X ~ ∇i e H0 = − − , (117a) However, to date, the most important method for 2m π |r | i=1 0 i parametrising effective Hamiltonians has been to fit the parameters to a range of experimental data, whence the and name ‘semi-empirical’. Of course experimental data con- e2 tains all corrections to all orders therefore this is indeed H1 = . (117b) an extremely sensible thing to do. But, it is important 4π0|r1 − r2| to understand that empiricism is not a dirty word. In- A detailed description of this perturbation theory is given deed empiricism is what distinguishes science from other in chapter 18 of Ref. 8. However, for our discussion, belief systems. Further, this empirical approach is ex- the key point is that, in this perturbation theory, the actly the approach that the mathematics tells one to take. term h1s2|V |1s2i is simply the first order correction to It is also important to know that no quantum chemical

27 or solid state calculation is truly ab initio - the nuclear [2] Reimers, J. R. Computational Methods for Large systems: and electronic masses and the charge on the electron are Electronic Structure Approaches for Biotechnology and all measured rather than calculated. Indeed the modern Nanotechnology; Wiley: Hoboken, in press. view of the ‘standard model’ of particle physics is that [3] Schatz, G. C.; Ratner, M. A. Quantum mechanics in it too is an effective low-energy model (50). For exam- chemistry; Prentice Hall: Englewoods Cliffs, 1993. ple, in quantum electrodynamics (QED), the quantum [4] Mahan, G. D.; Many-particle physics; Kluwer Academic: New York, 2000. field theory of light and matter, the bare charge on the [5] Goldstein, H.; Poole, C.; Safko, J. Classical mechanics; electron is, for all practical purposes, infinite. But, the Addison Wesley: San Francisco, 2002. charge is renormalised to the value seen experimentally [6] Atkins, P.; de Paula, J. Atkin’s physical chemistry; Ox- in a manner analogous to the renormalisation of the Hub- ford University Press: Oxford, 2006. bard U of He discussed above. Therefore, as we do not, [7] See, for example, Rae, A. I. M. Quantum mechanics; In- at the time of writing, know the correct mathematical stitute of Physics Publishing: Bristol, 1996. description of processes at higher energies, all of theoret- [8] See, for example, Gasiorowicz, S. Quantum physics; Wi- ical science should, perhaps, be viewed as the study of ley: Hoboken, 2003. semi-empirical effective low-energy Hamiltonians (80). [9] Jordan P.; Wigner, E. Z. Phys. 1928, 47, 631-651. Finally, the most important point about effective [10] Lowe J. P.; Peterson, K. A. Quantum chemistry; Elsevier: Amsterdam, 2006. Hamiltonians is that they promote understanding. Ulti- [11] Ashcroft N.W.; Mermin, N.D. Solid state physics; Holt, mately the point of science is to understand the phenom- Rinehart and Winston: New York, 1976. ena we observe in the world around us. While the ability [12] Tinkham, M. Group theory and quantum mechanics; to perform accurate numerically calculations is impor- McGraw-Hill: New York, 1964. tant, we should not allow this to become our main goal. [13] Lax, M. Symmetry principles in solid state and molecular The models discussed above provide important insights physics; Wiley: New York, 1974. into the chemical bond, magnetism, polarons, the Mott [14] McWeeny, R. Coulson’s valence; Oxford University Press: transition, electronic correlations, the failure of mean Oxford, 1979. field theories, etc. All of these effects are much more diffi- [15] Brogli, F.; Heilbronner, E. Theor. Chim. Acta 1972, 26, cult to understand simply on the basis of atomistic calcu- 289-299. lations. Further, many important effects seen in crystals, [16] See, e.g., Arfken, G. Mathematical methods for physicists, 3rd ed.; Academic Press: Orlando, 1985. such as the Mott insulator phase, are not found methods [17] Mandl, F. Statistical physics; Wiley: Chichester, 1998. such as density functional theory or Hartree-Fock theory, [18] See pp 799-800 of Ref. (16). while post Hartree-Fock methods are not practical in in- [19] (a) Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; finite systems. Thus effective Hamiltonians have a vital Novoselov, K. S.; Geim, A. K. Rev. Mod. Phys. 2009, role to play in developing the new concepts that are re- 81, 109-162. (b) Castro Neto, A. H.; Guinea, F.; Peres, quired to understand the emergent phenomena found in N. M. R. Phys. World 2006, 19, 33-37. molecules and solids (81). [20] (a) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Gregorieva, I. V.; Firsov, A. A. Science 2004, 306, 666-669. (b) Choucair, M.; Acknowledgments Thordarson P.; Stride, J. A. Nature Nanotech. 2009, 4, 30-33. I would like to thank Balazs Gy¨orffy, who taught me [21] Schr¨odinger,E. Ann. Physik 1926, 79, 361-428. [22] Heitler, W.; London, F. Z. Phys. 1927, 44, 455-472. that “you can’t not know” many of things discussed [23] Pauling, L. The nature of the chemical bond and the above. I also thank James Annett, Greg Freebairn, Noel structure of molecules and crystals; Cornell Univ. Press: Hush, Anthony Jacko, Bernie Mostert, Seth Olsen, Jeff Ithaca, 1960. Reimers, Edan Scriven, Mike Smith, Eddy Yusuf, and [24] Mott, N.F. Proc. Roy. Soc. A 1949, 62, 416-422. particularly Ross McKenzie, for many enlightening con- [25] Powell, B. J.; McKenzie, R. H. J. Phys.: Condens. Mat- versations about the topics discussed above and for show- ter 2006, 18, R827-R865. ing me that chemistry is a beautiful and rich subject [26] (a) Anderson, P. W. Science 1987, 235, 1196-1198. (b) with many simplifying principles. I would also like to Zhang, F. C.; Gross, C.; Rice T. M.; Shiba, H. Supercond. thank Bernd Braunecker, Karl Chan, Anthony Jacko, Sci. Technol. 1988, 1, 36-46. Sergio Di Matteo, Ross McKenzie, Seth Olsen, Eddie [27] Anderson, P. W. Phys. Today 2008, 61 (4), 8-9. [28] Powell B. J.; McKenzie, R. H. Phys. Rev. Lett. 2005, 94, Ross and Kristian Weegink for their insightful comments 047004; Gan, J. Y.; Chen, Y.; Su, Z. B.; Zhang, F. C. on an early draft of these notes. I am supported by a ibid. 2005, 94, 067005; Liu, J.; Schmalian, J.; Trivedi, Queen Elizabeth II fellowship from the Australian Re- N. ibid. 2005, 94, 127003. search Council (project DP0878523). [29] Cohen, A. J.; Mori-Sanchez, P.; Yang, W. T. Science 2008, 321, 792-794. [30] R¨ossler,U. Solid state theory; Springer: Berlin, 2004. References [31] Mohn P.; Wohlfarth, E. P. J. Mag. Mag. Mat. 1987, 68, L283-L285. [1] Fulde, P. Electron correlations in molecules and solids; [32] Jacko, A. C.; Fjærestad, J. O.; Powell, B. J. Nature Phys. Springer: Berlin, 1995. 2009, 5, 422-425.

28 [33] Gutzwiller, M. C. Phys. Rev. Lett. 1963, 10, 159-162. [63] See, for example, Bersuker, I. B. The Jahn-Teller effect [34] Brinkmann W. F.; Rice, T. M. Phys. Rev. B 1970, 2, and vibronic interactions in modern chemistry; Plenum: 4302-4304. New York, 1984. [35] Lieb, E. H.; Wu, F. Y. Phys. Rev. Lett. 1968, 20, 1445- [64] (a) Olsen, S.; Riesz, J.; Mahadevan, I.; Coutts, A.; 1448. Bothma, J. P.; Powell, B. J.; McKenzie, R. H.; Smith, [36] Essler, F. H. L.; Frahm, H.; G¨ohmann,F.; Kl¨umper, S. C.; Meredith, P. J. Am. Chem. Soc. 2007, 129, A.; Korepin, V. E. The one-dimensional Hubbard model; 6672-6673. (b) Meredith, P.; Powell, B. J.; Riesz, J.; Cambridge University Press: Cambridge, 2005. Nighswander-Rempel, S.; Pederson, M. R.; Moore, E.; [37] Tsvelik, A. M. Quantum field theory in condensed matter Soft Matter 2006, 2, 37-44. physics; Cambridge University Press: Cambridge, 1996. [65] Reimers, J. R.; Hush, N. S. J. Am. Chem. Soc. 2004, [38] Kotliar G.; Vollhardt, D. Phys. Today 2004, 57 (3), 53- 126, 41324144. 59. [66] Hahn, S.; Stock, G. J. Phys. Chem. B 2000, 104, 1146- [39] Kollar, M.; Strack, R.; Vollhardt, D. Phys. Rev. B 1996, 1149. 53, 9225-9231. [67] Walls D. F.; Milburn, G. J. Quantum optics; Springer: [40] Maier, T.; Jarrell, M.; Pruschke, T.; Hettler, M. H. Rev. Berlin, 2006. Mod. Phys. 2005, 77, 1027-1080. [68] Weiss, U. Quantum dissipative systems; World Scientific: [41] Kotliar, G.; Savrasov, S. Y.; Haule, K.; Oudovenko, V. S.; Singapore, 2008. Parcollet, O.; Marianetti, C. A. Rev. Mod. Phys. 2006, [69] For an introductory discussion of broken symmetry see, 78, 865-951. for example, Blundell, S. J. Magnetism in condensed mat- [42] Nagaoka, Y. Phys. Rev. 1966, 145, 392-405. ter Oxford University Press, Oxford, 2001. For a more [43] Tian, G. J. Phys. A 1990, 23, 2231-2236. advanced discussion see, for example, Anderson, P. W. [44] Merino, J.; Powell, B. J.; McKenzie, R. H. Phys. Rev. B Basic notions of condensed matter physics Benjamin- 2006, 73, 235107. Cummings: Menlo Park, 1984. [45] Shaik, S.; Hiberty, P. C. Valence bond theory, its history, [70] Dirac, P. Proc. Roy. Soc. A 1929, 123, 714-733. fundamentals, and applications: a primer. In Reviews in [71] (a) Pople, J. A. Rev. Mod. Phys. 1999, 71, 1267-1274. computational chemistry; Lipkowitz, K. B., Larter, R., (b) Truhlar, D. G. J. Am. Chem. Soc., 2008, 130, 16824- Cundari, T. R., Eds.; Wiley-VCH: Hoboken, NJ, 2004; 16827. pp. 1-100. [72] Feynman, R. P. Int. J. Theor. Phys. 1982, 21, 467-488. [46] Sakurai, J. J. Modern quantum mechanics; Addison- [73] Lanyon, B. P.; Whitfield, J. D.; Gillet, G. G.; Goggin, M. Wesley: New York, 1994. E.; Almeida, M. P.; Kassal, I.; Biamonte, J. D.; Mohseni, [47] Chao, K. A.; Spa lek, J.; Ole´s,A. M. J. Phys. C 1977, M.; Powell, B. J.; Barbieri, M.; Aspuru-Guzik, A.; White, 10, L271-L276. A. G. Nature Chem. 2010, 2, 106-111. [48] Brockhouse, B. N. Slow neutron spectroscopy and [74] Schuch N.; Verstraete, F. Nature Phys. published online: the grand atlas of the physical world. In Nobel lec- 23 August 2009, doi:10.1038/nphys1370. tures in physics, 1991−1995; Ekspong, G., Ed.; [75] Goldenfeld, N. D. Lectures on Phase Transitions and the World Scientific: Singapore, 1997. Also available Renormalisation Group; Addison-Wesley, 1992. from http://nobelprize.org/nobel_prizes/physics/ [76] See, for example, Wen, X.-G. Quantum field theory of laureates/1994/brockhouse-lecture.html. many-body systems; Oxford Univ. Press: Oxford, 2004. [49] Zaliznyak, I. A. Nature Mat. 2005, 4, 273-275. [77] (a) Freed, K. F. Acc. Chem. Res. 1983, 16, 137-144. [50] Griffiths, D. Introduction to elementary particles; Wiley- (b) Gunnarsson, O. Phys. Rev. B 1990, 41, 514-518. (c) VCH: Weinheim, 2008. Iwata S.; Freed, K. F. J. Chem. Phys. 1976, 65, 1071- [51] (a) Coldea, R.; Tennant, D. A.; Tylczynski, Z. Phys. Rev. 1088. (d) Graham R. L.; Freed, K. F. J. Chem. Phys. B 2003, 68, 134424. (b) Lake, B.; Tennant, D. A.; Frost, 1992, 96, 1304-1316. (e) Martin C. M.; Freed, K. F. J. C. D.; Nagler, S. E. Nature Mat. 2005, 4, 329-334. Chem. Phys. 1994, 100, 7454-7470. (f) Stevens, J. E.; [52] Lee, P. A. Science 2008, 321, 1306-1307. Freed, K. F.; Arendt, F.; Graham, R. L. J. Chem. Phys. [53] Shimizu Y.; et al. Phys. Rev. Lett. 2003, 91, 107001. 1994, 101, 4832-4841. (g) Finley J. P.; Freed, K. F. J. [54] Helton J.; et al. Phys. Rev. Lett. 2007, 98, 107204. Chem. Phys. 1995, 102, 1306-1333. (h) Stevens, J. E.; [55] Okamoton Y.; et al. Phys. Rev. Lett. 2007, 99, 137207. Chaudhuri, R. K.; Freed, K. F. J. Chem. Phys. 1996, [56] Raczkowski, M.; Fr´esard,R.; Oles, A. M. J. Phys.: Con- 105, 8754-8768. (i) Chaudhuri, R. K.; Freed, K. F. J. dens. Matter 2006, 18, 7449-7469. Chem. Phys. 2003, 119, 5995-6002. (j) Chaudhuri, R. [57] Sarma, D. D. J. Sol. State Chem. 1990, 88, 45-52. K.; Freed, K. F. J. Chem. Phys. 2005, 122, 204111. [58] (a) Merino, J.; Powell, B. J.; McKenzie, R. H. Phys. Rev. [78] (a) Scriven, E.; Powell, B. J. J. Chem. Phys. 2009, 130, B 2009, 79, 161103(R); (b) Merino, J.; McKenzie, R. H.; 104508; (b) Phys. Rev. B 2009, 80, 205107. Powell, B. J. Phys. Rev. B 2009, 80, 045116. (c) Powell, [79] (a) Martin, R. L.; Ritchie, J. P. Phys. Rev. B 1993, 48, B. J.; Merino, J.; McKenzie, R. H. Phys. Rev. B 2009, 4845-4849. (b) Antropov, V. P.; Gunnarsson, O.; Jepsen, 80, 085113. O. Phys. Rev. B ,1992 46, 13647-13650. (c) Pederson, M. [59] See for example, Ziman, J. M. Electrons and phonons; R.; Quong, A. A. Phys. Rev. B 1992, 46, 13584-13591. Oxford University Press: Oxford, 1960. (d) Brocks, G.; van den Brink, J.; Morpurgo, A. F. Phys. [60] For a review see Gr¨uner,G. Density waves in solids; Rev. Lett. 2004, 93, 146405. (e) Cano-Cort´es,L.; Dolfen, Perseus Publishing: Cambridge, 1994. A.; Merino, J.; Behler, J.; Delley, B.; Reuter, K.; Koch, [61] See, for example, Alexandrov, A. S.; Mott, N. F. Polarons E. Eur. Phys. J. B 2007, 56, 173-176. and biploarons; World Scientific: Singapore, 1995. [80] For an accessible, and highly outspoken, discussion of [62] For a review see Marcus, R. A. Rev. Mod. Phys. 1993, these ideas see Laughlin, R. B.; Pines, D. Proc. Natl. 65, 599-610. Acad. Sci. 2000, 97, 28-31 and Laughlin, R. B. A differ-

29 ent universe Basic Books: New York, 2005. [82] Powell, B. J. Chem. Aust. 2009, 76, 18-21. [81] Anderson, P. W. Science 1972, 177, 393-396.

30