<<

Concepts in : Tutorial I

Tight Binding and The Hubbard Model

Everything should be made as simple as possible, but no simpler  A. Einstein

1 Introduction

The Hubbard Hamiltonian (HH) oers one of the most simple ways to get insight into how the interactions between give rise to insulating, magnetic, and even novel superconducting eects in a . It was written down [1, 2, 3, 4] in the early 1960's and initially applied to the behavior of the transition-metal monoxides (FeO, NiO, CoO), compounds which are antiferromagnetic insulators, yet had been predicted to be metallic by methods which treat strong interactions less carefully. Over the intervening years, the HH has been applied to many systems, from heavy fermions and the Cerium volume collapse transition in the 1980's, to high temperature superconductors in the 1990's. Indeed, it is an amazing feature of the HH that, despite its simplicity, it exhibits behavior relevant to many of the most subtle and beautiful properties of solid state systems. We focus here for the most part on the single-band HH. Multi-band variants like the Periodic Anderson Model (PAM) allow one to introduce other fundamental concepts in many-body physics, such as the competition between magnetic order and singlet formation. Randomness can be simply introduced into the HH, so it can be used as a starting point for investigations of the interplay of interactions and disorder in metal-insulator transitions and, recently, many-body localization. Textbook discussions of the HH can be found in Refs. [5, 6, 7, 8] and a recent celebration of its 50th anniversary [9] emphasizes the resurgence of interest due to optical lattice emulation experiments.

1 2 Tight binding method

Consider a lattice of static Ions creating a periodic potential,

X   X R~ VLattice (~r) = VIon ~r − R~ = VIon (~r) , (1) R~ R~

R~   where R~ points to the lattice sites and VIon (~r) = VIon ~r − R~ is the potential of the Ion positioned at R~. In the limit of innitely far away Ions we can solve the system by solving the Ions independently,

 ~p2  + V R~ (~r) φR~ (~r) = R~ φR~ (~r) , (2) 2m Ion n n n

  where R~ ~ is the eigenstate of the th orbital of the Ions positioned at ~, φn (~r) = φn ~r − R n R with corresponding energy R~ . The total energy and wavefunction are: n

N X ~ E = Ri , (3a) ni i=1   n ~ oN Ri (3b) Φ = Slater φn , i i=1 where N is the total number of electrons. Upon bringing the Ions closer together there are two main dierences to consider:

1. Each site R~ start feeling the potential due to the other sites R~ 0. This leads to a small correction to the single Ion potential: P R~ 0 . ∆VR~ (~r) = R~ 06=R~ VIon (~r)

2. The wavefunctions corresponding to dierent Ions, positioned at R~ and R~ 0, begin to overlap: R~0 R~ . h φm | φn i 6= 0 Therefore, we do perturbation theory in , which to rst order gives a matrix element ∆VR~ for tunnelling between sites and orbitals,

R,~ R~ 0 R~ R~ 0 (4) tn,m ≡ − h φn | ∆VR~ 0 | φm i .

It should be noted that, as addressed in point number two above, the set of wavefunctions R~ are not orthogonal and extra care should be taken to properly derive the tunneling | φn i

2 matrix element R,~ R~ 0 . However, we will not dwell on this more in this course. The tight- tn,m binding method will be taught in a rigorous manner in the next course, Solid State 2, and can be found also in Ashcroft and Mermin's book [10]. We have so far disregarded the electron spin completely. Adding it is straightforward though, since the potential does not depend on it at all. The eigenstates will simply carry also a spin index, on which the energies are degenerate, and the hopping matrix element t will be diagonal in spin. In 2nd quantization we dene creation and annihilation operators, Z c† = d3rφ (~r)ψ†(~r) , (5a) R,n,σ~ R,n,σ~

Z 3 ∗ (5b) cR,n,σ~ = d rφR,n,σ~ (~r)ψ(~r) ,

n † o satisfying c ~ , c = δ ~ ~ 0 δ δ 0 , and the Hamiltonian takes the form R,n,σ R~0,m,σ0 R,R n,m σ,σ

X X  ~ ~ 0  H =  c† c − tR,R c† c + h.c. . (6) R,n~ R,n,σ~ R,n,σ~ n,m R,n,σ~ R~0,m,σ R,n,σ~ R,~ R~0,n,m,σ

Before adding electron-electron interactions we will make some simplications. Some are very widely used, and some are just for our own convenience:

1. We will consider only a single orbital per site, and drop the index n. This is not justied in all systems, but will be enough for our purposes.

2. Since the matrix element for electron hopping, t , decays exponentially with R~ − R~ 0 , R,~ R~ 0 we will truncate it after nearest neighbours (n.n.). Namely, does not vanish only tR,~ R~ 0 for n.n. sites, denoted by hR,~ R~ 0i.

3. We will assume that all Ions are identical, and set for all ~ as well as R~ ≡ µ R thR,~ R~ 0i ≡ t for all n.n. R~ and R~ 0. Systems where this assumption is wrong are called inhomo- geneous. In reality all systems are inhomogeneous, but usually the inhomogeneity is weak, and can be neglected or treated in perturbation theory.

Armed with these simplications, and specializing to one dimension (1d), the system is described by the Hamiltonian,

  X † X † † (7) H = µ cj,σcj,σ − t cj,σcj+1,σ + cj+1,σcj,σ . j,σ j,σ

3 For the rest of the tutorial we will work at half lling, namely one electron per site. Notice that for a xed number of electrons the µ term is just a constant, and so we will drop it. To solve this Hamiltonian we use the (discrete) Fourier transform1

1 X −ikxj ck,σ = √ e cj,σ , (8a) N j

1 X ikxj cj,σ = √ e ck,σ , (8b) N k where is the number of sites and 2πn , where is the length of the system. N kn = L L = Na 2π Notice that c 2π = ck,σ, so k-space is periodic and it is enough to consider the region k+ a ,σ a  π π , namely the rst Brillouin Zone (BZ1). Plugging Eq.(8) into Eq.(7) (dropping k ∈ − a , a the µ term) we diagonalize the Hamiltonian

X h i 1 X H = −tc† c (eiq + e−ik) e−i(k−q)xj k,σ q,σ N k,q,σ j h i X † iqa −ika (9) = −tck,σcq,σ(e + e ) δk,q k,q,σ X † = [−2t cos (ka)] ck,σck,σ , k,σ where we used the fact that 1 P −i(k−q)xj . The solution is depicted in Fig.1. In N j e = δk,q order to count the number of states we notice that they are spaced 2π apart spanning ∆k = Na 2π . The number of states is then given by, kmax − kmin = a

kmax − kmin 2 = 2N, (10) ∆k where the factor of 2 is for spin. At half lling (one electron per site, i.e., N electrons) the ground state has the lowest N states occupied. The system is therefore gapless, meaning that it can be excited above the ground state by any innitesimal energy injection (or temperature). A gapless system of electron is called metallic, since it conducts. In order to introduce a gap, consider a dimerization of the lattice such that all even sites

1 † One can use various denitions for the Fourier transform, but the symmetric √1 convention ensures c N k,σ and ck,σ are indeed creation and annihilation operators, i.e., have appropriate anti-commutation relations n o † . ck,σ, cq,σ0 = δk,qδσ,σ0

4 E 2t 1

0.5

ka -π - π π π 2 2

-0.5

-1

Figure 1: Spectrum of a free fermion on a 1d lattice in the tight-binding approximation. The energy band is doubly degenerate in the spin degree of freedom. At half lling all state below zero (between −π/2 and π/2) are lled. are slightly displaced from their position2. Instead of changing the distance between Ions we will use a trick and change the tunneling matrix element,

 j tj,j+1 = t 1 + ∆(−1) , (11) where ∆ is governed by the displacement of the Ions. Indeed doing so captures essentially the same physics since, as can be seen in Eq.(4), the size of t relates to the distance between Ions. Notice that now the symmetry for translation by a is broken, and one must dene a unit cell of size 2a with two Ions in the unit cell. We will denote the creation operators on odd sites by † and those on even by † . The Hamiltonian can now be written as Aj,σ Bj,σ     X † † X † † (12) H = −t(1 − ∆) Aj,σBj,σ + Bj,σAj,σ − t(1 + ∆) Aj+1,σBj,σ + Bj,σAj+1,σ . j,σ j,σ

We now transform again to k-space using the denitions (8), but with and N unit cells 2a 2 2In reality, such a dimerization is always energetically favorable in 1d system at half lling, since as the gap opens the (occupied) states near the middle of the gap lower their energy. This lower energy overcomes the energy cost of displacing the Ions. Therefore, the lattice will spontaneously break the translation symmetry from a to 2a, leading to an insulator rather than a metal. This eect is called Peierls instability, or Peierls transition.

5 to get

! X 0 −t0 − te˜ i2ka X H = C~ † C~ ≡ C~ † H(k)C~ , (13) k,σ 0 ˜ −i2ka k,σ k,σ k,σ k,σ −t − te 0 k,σ

 T ~ 0 where Ck,σ = Ak,σ Bk,σ , t = t(1 − ∆) and t˜ = t(1 + ∆). To completely solve the system we just need to diagonalize withing the unit cell, namely the two-by-two matrix H(k). This can be done easily by rst noticing that

~ H(k) = dk · ~τ = |dk| ~n · ~τ, (14)   where ~n = cos 2θk sin 2θk 0 is a unit vector, ~τ is a vector of Pauli matrices (here not ~ related to the spin degree of freedom but rather to pseudo spin), |dk| is the mudulus of dk h i and 1 t˜sin(2ka) . The energies and eigenstates are then θk = 2 arctan −t0−t˜cos(2ka)

q 2 2 E±(k) = ± |dk| = ±2t 1 − (1 − ∆ ) sin (ka) , (15a)

! 1 e−i2θk 1 −i2θk  Γ±(k) = √ = √ e Ak,σ ± Bk,σ , (15b) 2 ±1 2 and are depicted in Fig.2 for ∆ = 0.2. Notice that there are two bands now, corresponding to the two sites in the unit cell, and a gap between them. At half lling the lower band is completely full, while the upper is completely empty, and the system is gapped. Meaning that exciting the system requires overcoming a nite energy gap, in this case 2∆. For example, thermal excitation of the system is suppressed by a Boltzman factor of exp [−kBT/2∆].A gapped system of electrons is called insulating since it does not conduct.

3 The Hubbard Model

After playing around with the simplest non-interacting tight binding model we are ready to add interactions. We will do so in three dimensions (3d). Consider electron-electron interactions of the form Z X 3 3 0 0 † † 0 0 (16) Velel = d rd r V (~r − ~r )ψσ(~r)ψσ0 (~r )ψσ0 (~r )ψσ(~r) , σ,σ0

6 E 2t 1

0.5

ka - π - π π π 2 4 4 2

-0.5

-1

Figure 2: Spectrum of a free fermion on a dimerized 1d lattice in the tight-binding approx- imation. Both energy bands are doubly degenerate in the spin degree of freedom. At half lling the lower band is lled, while the upper is completely empty.

where V (~r−~r0) could for example be Coulomb repulsion. In the tight binding approximation, roughly speaking using Eq.(5), the interactions are given by

X X X F † † V˜elel = U n n + U n n + J c c c c , (17) R,~ R~ R,~ ↑ R,~ ↓ R,~ R~ 0 R~ R~ 0 R,~ R~ 0 R,σ~ R~ 0,σ0 R,σ~ 0 R~ 0,σ R~ R~ 06=R~ R~ 06=R,σ,σ~ 0 where P is the electron density on the site positioned at ~, is the density nR~ = σ nR,σ~ R nR,σ~ of electrons of spin σ on the site positioned at R~, and the various matrix elements are:

1 Z 3 3 2 2 (18a) U ~ ~ 0 = d r1d r2V (~r1 − ~r2) |φ ~ (~r1)| φ ~0 (~r2) , R,R 2 R R

1 Z J F = d3r d3r V (~r − ~r )φ (~r )φ∗ (~r )φ (~r )φ∗ (~r ) . (18b) R,~ R~ 0 2 1 2 1 2 R~ 1 R~ 0 1 R~ 0 2 R~ 2 Notice that again we have three free parameters reminiscent of the three channels. Here however, J F is very small compared to U , since it includes two overlap integrals of R,~ R~ 0 R,~ R~ 0 wavefunctions centered around dierent sites. By far the most important term in Eq.(17) is . Consequently, the other terms are commonly truncated3 resulting in the so-called UR,~ R~ Hubbard model

X  †  X HHubbard = −t c c + h.c. + U n n , (19) R,σ~ R~ 0,σ R,~ ↑ R,~ ↓ hR,~ R~ 0i,σ R~

3We will return to the J F term later on in this tutorial. R,~ R~ 0

7 where which again is assumed to be the same for all ~. Although we have omitted U = UR,~ R~ R many terms that in reality are present, the resulting simplied model is still surprisingly rich. Moreover, even though it is arguably the simplest interacting electron model one can write down, we do not know how to solve it exactly. Therefore, we will consider various limits. We already treated the U = 0 case in 1d. Generalizing it to 3d is straightforward, and we will skip it. One can treat using perturbation theory in U . It does not lead to U  t t dramatic eects, and we will skip that as well. Interesting results occur for the t = 0 and t  U limits, and so we will focus on these. For t = 0 the Hamiltonian reduces to

X (20) H0 = U nR,~ ↑nR,~ ↓ , R~ for which the (degenerate) ground states can be easily inferred for any number of electrons. At half lling the ground state energy is E = 0 and the ground states have all sites occupied by a single electron. Lowest excited states, those that have one doubly occupied site, have a nite energy U, so the system is again gapped, but now due to interactions (rather than band-structure). This is called a . As opposed to a band insulator, which is completely inert, in this Mott insulator there are still low energy degrees of freedom, namely the spins of the electrons µ 1 X † µ S = c σ 0 c . (21) R~ 2 R,s~ s,s R,s~ 0 s,s0 The many degenerate ground states correspond to all possible congurations of the spins of the electrons. This degeneracy is lifted by any nite t leading to a tunneling Hamiltonian,

X   H = −t c† c + h.c. , (22) t R,σ~ R~ 0,σ hR,~ R~ 0i,σ which will prefer certain spin congurations. We will treat Ht in perturbation theory and derive an eective Hamiltonian for the degenerate ground space. To second order in pertur- bation theory the eective Hamiltonian is given by

1  t3  e (23) H = P0HtP0 + P0HtP1 P1HtP0 + O 2 , E0 − H0 U where P is the projection operator to the ground space, and P0 = α∈g.s. | α i h α | P1 = P is the complementary projection. The rst term in Eq.(23) clearly 1 − P0 = a6∈g.s. | a i h a | vanishes since Ht connects states in P0 with states in P1 which are the orthogonal to states

8 in P0. The second term can be simplied by noting:

1. The g.s. energy is zero, namely E0 = 0.

2. Since Ht acting on states in P0 creates only one double occupied site, all states in

P1HtP0 have exactly one doubly occupied site.

We can then write:

X X 1  t3  e H = | α i h α | Ht | a i h a | | b i h b | Ht | β i h β | + O 2 −H0 U α,β∈g.s. a,b∈1 d.o. 1 X X  t3  = − | α i h α | H | a i h a | b i h b | H | β i h β | + O U t t U 2 α,β∈g.s. a,b∈1 d.o. (24) 1 X X  t3  = − | α i h α | H | a i h a | H | β i h β | + O U t t U 2 α,β∈g.s. a∈1 d.o. 1 X  t3  = − h α | H2 | β i | α i h β | + O , U t U 2 α,β∈g.s.

So we need to compute that Matrix elements 2 . Since P only h α | Ht | β i Ht = hR,~ R~ 0i(Ht)R,~ R~ 0 connect n.n. we concentrate on the spin conguration of two specic n.n. sites R~ and R~ 0. We denote the states in short | β i = | · · · , σ , σ0 , · · · i ≡ | σ , σ0 i , and look at the various R~ R~ 0 R~ R~ 0 cases:

(I) If both spins are oriented at the same direction, namely or | β i = | ↑R~ , ↑R~ 0 i | β i = , they will be annihilated by since the site ~ (or ~ 0) cannot be | ↓R~ , ↓R~ 0 i (Ht)R,~ R~ 0 R R occupied by two electrons with the same spin (Pauli exclusion principle):

(25) (Ht)R,~ R~ 0 | σR~ , σR~ 0 i = 0 .

(II) If the spins are opposite, namely or , then | β i = | ↑R~ , ↓R~ 0 i | β i = | ↓R~ , ↑R~ 0 i (Ht)R,~ R~ 0 leads to a doubly occupied site R~ or R~ 0:

(26) (Ht)R,~ R~ 0 | σR~ , σ¯R~ 0 i = −t | σR~ σ¯R~ , 0 i − t | 0, σR~ 0 σ¯R~ 0 i .

The second then splits them back onto both sites, by either going back to the (Ht)R,~ R~ 0 original state or by exchanging the electrons:

9 (a) Back to original state:

2 2 (27) (Ht)R,~ R~ 0 | σR~ , σ¯R~ 0 i = 2t | σR~ , σ¯R~ 0 i .

(b) Exchanged electrons:

2 2 (28) (Ht)R,~ R~ 0 | σR~ , σ¯R~ 0 i = −2t | σ¯R~ , σR~ 0 i .

The extra minus sign in this case comes from the exchange of the electrons, and can be seen explicitly:

  (H )2 | σ , σ¯ i = (H ) −tc† c − tc† c | σ , σ¯ i t R,~ R~ 0 R~ R~ 0 t R,~ R~ 0 R~ 0,↑ R,~ ↑ R,~ ↓ R~ 0,↓ R~ R~ 0 h †   †   †   † i = −tc c ~ 0 −tc c ~ + −tc c ~ −tc c ~ 0 | σ ~ , σ¯ ~ 0 i R,~ ↓ R ,↓ R~ 0,↑ R,↑ R~0,↑ R,↑ R,~ ↓ R ,↓ R R   = t2 c† c c c† + c c† c† c | σ , σ¯ i R,~ ↓ R,~ ↑ R~ 0,↓ R~ 0,↑ R,~ ↑ R,~ ↓ R~ 0,↑ R~ 0,↓ R~ R~ 0 = −2t2c† c c† c | σ , σ¯ i R,~ ↓ R,~ ↑ R~ 0,↑ R~ 0,↓ R~ R~ 0 2 = −2t | σ¯R~ , σR~ 0 i . (29)

Noticing that Sz = 1 [ | ↑ i h ↑ | − | ↓ i h ↓ | ], S+ = | ↑ i h ↓ | and S− = | ↓ i h ↑ | R~ 2 R~ R~ R~ R~ R~ R~ R~ R~ R~ R~ we write cases (I) and (II)(a) as

2t2 4t2 1 − [ | ↑ , ↓ i h ↑ , ↓ | + | ↓ , ↑ i h ↓ , ↑ | ] = (Sz Sz − ) , (30) U R~ R~ 0 R~ R~ 0 R~ R~ 0 R~ R~ 0 U R~ R~ 0 4 and case (II)(b) as

2t2 2t2 [ | ↑ , ↓ i h ↓ , ↑ | + | ↓ , ↑ i h ↑ , ↓ | ] = (S−S+ + S+S− ) . (31) U R~ R~ 0 R~ R~ 0 R~ R~ 0 R~ R~ 0 U R~ R~ 0 R~ R~ 0 In total, summing over all n.n. we get (dropping the constant factor) the eective Hamilto- nian   X z z 1 − + + − X He = J S S + (S S + S S ) = J S~ · S~ , (32) R~ R~ 0 2 R~ R~ 0 R~ R~ 0 R~ R~ 0 hR,~ R~ 0i hR,~ R~ 0i with 4t2 , namely a Heisenberg anti-ferromagnet. Recall now that earlier in the J = U > 0

10 tutorial we truncated the interaction term

F X † † H F = J c c c c , (33) J R,σ~ R~ 0,σ0 R,σ~ 0 R~ 0,σ hR,~ R~ 0i and the P . The latter is trivial in the ground space, namely , V hR,~ R~ 0i nR~ nR~ 0 P0nR~ nR~ 0 P0 = 1 but the former actually also gives a Heisenberg Hamiltonian

F X ~ ~ (34) P0HJF P0 = −2J SR~ · SR~ 0 . hR,~ R~ 0i

It can be shown that for Coulomb interaction (and also for a contact interaction) J F > 0, namely this interaction is ferromagnetic. Previously we could disregard it with respect to , but now that the eective Hamiltonian (32) is of order t2 the two can be comparable and U U should be taken together. Indeed dierent systems can have either one of them dominating over the other and consequently display either ferromagnetism or antiferromagnetism. This is the main mechanism by which magnetism arises in systems. The Hubbard model can also describe BCS . This requires changing of the sign of U from positive to negative, i.e., from repulsive to attractive. We will not dwell at this stage on how can electrons have an attractive density-density interaction (obviously Coulomb is not enough), but take this axiomatically and see how it leads to the formation of cooper pairs. Consider the following transformation on the creation and annihilation operators  c ~ → d ~ ↑,R ↑,R , (35) c → (−1)R~ d†  ↓,R~ ↓,R~ where (−1)R~ is the parity of the site at position R~, e.g. (−1)x/a+y/a+z/a on a 3d cubic lattice with unit cell of size a3. It is easy to check that the d operators are also creation and annihilation operators, and plugging this into the Hubbard model we nd

X  †  X HHubbard = −t d d + h.c. − U n n , (36) R,σ~ R~ 0,σ R,~ ↑ R,~ ↓ hR,~ R~ 0i,σ R~ namely the Hubbard model but with U → −U. Performing the same transformation on the

11 spin operators we nd  z S ~ → nR~  R S+ → (−1)R~ d†d† . (37) R~ ↑ ↓  ~ S− → (−1)Rd d R~ ↓ ↑ We conclude that there is a correspondence between magnetism in the repulsive Hubbard model and cooper pairs in the attractive Hubbard model. We will learn more about super- conductivity later in the course.

References

[1] M.C. Gutzwiller, Phys. Rev. Lett. 10, 159 (1963)

[2] J. Kanamori, Prog. Theor. Phys. 30, 275 (1963)

[3] J. Hubbard, Proc. R. Soc. A 276, 237 (1963)

[4] J. Hubbard, Proc. R. Soc. A 281, 401 (1964)

[5] P. Fazekas: Lecture Notes on Electron Correlation and Magnetism (World Scientic, Singapore, 1999)

[6] M. Rasetti (ed.): The Hubbard Model  Recent Results (World Scientic, Singapore, 1991)

[7] A. Montorsi (ed.): The Hubbard Model  A Reprint Volume (World Scientic, Singa- pore, 1992)

[8] F. Gebhard: The Mott Metal-Insulator Transition, Models and Methods (Springer, Heidelberg, 1997)

[9] Editorial: The Hubbard model at half a century, Nature Physics 9, 523 (2013)

[10] Chapter 10 of Solid State Physics by Ashcroft and Mermin

12