Classical and Quantum Spins
Classical and Quantum Spins
November 4, 2015
2
Contents
2 Mean-field Theory 15
2.1 Mean-field theory of Ising model . . . . . . . . . . . . 15
2.2 Ginzburg-Landau free energy from mean-field theory 18
2.3 Ising model coupled to magnetic field . . . . . . . . . 19
2.4 Mean-field theory of XY spins . . . . . . . . . . . . . 20
4 XY Model 33
4.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . 33
4.2 Spin-wave Theory . . . . . . . . . . . . . . . . . . . . 34
4.3 High-temperature expansion . . . . . . . . . . . . . . 36
4.4 Vortex Excitation and Kosterlitz-Thouless transition 37
4.5 Helicity Modulus . . . . . . . . . . . . . . . . . . . . 39
4.6 Binder cumulent . . . . . . . . . . . . . . . . . . . . 40
4.7 Villain transformation . . . . . . . . . . . . . . . . . 40
4.8 Frustrated XY model . . . . . . . . . . . . . . . . . . 44
5 Heisenberg Model 47
5.1 Saddle-point Method . . . . . . . . . . . . . . . . . . 47
5.2 Renormalization Group Analysis . . . . . . . . . . . . 48
5.3 Skyrmion Excitations . . . . . . . . . . . . . . . . . . 49
3
4 CONTENTS
7 Miscellaneous 59
7.1 Two-state Problem . . . . . . . . . . . . . . . . . . . 59
7.2 Three-state problem . . . . . . . . . . . . . . . . . . 59
7.3 d-orbitalogy . . . . . . . . . . . . . . . . . . . . . . . 59
7.4 p-orbitalogy . . . . . . . . . . . . . . . . . . . . . . . 60
10 Hubbard Model 69
10.1 The Model . . . . . . . . . . . . . . . . . . . . . . . . 69
10.2 Projection to truncated Hilbert space . . . . . . . . . 71
10.3 Mean-field Theory and Antiferromagnetism . . . . . . 74
10.3.1 Square lattice . . . . . . . . . . . . . . . . . . 74
10.3.2 Triangular lattice . . . . . . . . . . . . . . . . 77
10.4 Spin dependent hopping and DM interaction . . . . . 78
10.5 Orbitally degenerate Hubbard model . . . . . . . . . 80
10.6 Four-boson Theory . . . . . . . . . . . . . . . . . . . 81
10.6.1 Non-uniform Case . . . . . . . . . . . . . . . . 81
10.6.2 Uniform Case . . . . . . . . . . . . . . . . . . 82
11 Spin Interactions 87
11.1 Higher-order exchange processes . . . . . . . . . . . . 87
12.2.2 Antiferromagnet . . . . . . . . . . . . . . . . 97
12.2.3 Spiral ferromagnet . . . . . . . . . . . . . . . 98
12.2.4 Multiple spiral ferromagnet . . . . . . . . . . 98
12.3 Schwinger bosons . . . . . . . . . . . . . . . . . . . . 99
12.3.1 Ferromagnet . . . . . . . . . . . . . . . . . . . 100
12.3.2 Antiferromagnet . . . . . . . . . . . . . . . . 100
12.3.3 Spiral ferromagnet . . . . . . . . . . . . . . . 102
12.3.4 Multiple spiral ferromagnet . . . . . . . . . . 102
12.4 Slave boson . . . . . . . . . . . . . . . . . . . . . . . 102
12.5 Slave-fermion Schwinger boson . . . . . . . . . . . . . 103
Ginzburg-Landau Theory of
Critical Phenomena
an interesting new material to look into, none of these issues will be relevant, of course.
7
8CHAPTER 1. GINZBURG-LANDAU THEORY OF CRITICAL PHENOMENA
Xe−βEC
P
hXi = PC −βEC . (1.1)
Ce
The sum is over all allowed configurations C with a probability weight according
to the Boltzmann factor and β = 1/kB T .
We are dealing with spin systems in this lecture, so let’s try to identify
the right order parameter for them. At high temperature the system is de-
magnetized, meaning that magnetic moments can point in Pany arbitrary direc-
N
tion so that its average over the whole sample is zero: i=1 hSi i = 0. With
some N = 1023 atoms present in a typical sample we are looking at, the av-
erage
PN cannot of course be exactly zero. A more precise statement will be that
i=1 hSi i = O(1). That is, only a tiny, unobservable fraction of all the moments
survives the averaging.
It so happens that, by simply cooling the sample, the spins begin to talk to
each other and order. Mathematically, one expresses this fact by
N
X
Si = O(N ), (1.2)
i=1
where O(N ) means the sum is proportional to the number of spins, because all
the spins point more or less in
Pthe same direction. Appropriate order parameter
for the spin system is Stot = i Si . Its average is a macroscopic number if spins
are in an ordered phase, zero if disordered2 .
2 Not all transitions have a simple order parameter description. For example, a gas-to-liquid
transition (change of state of matter from disordered to less disordered form) is lacking an
obvious order parameter. Still, there is a well-defined transition temperature at which this
change of phase occurs.
1.3. ORDER PARAMETER EXPANSION OF FREE ENERGY 9
F = E − T S. (1.3)
As you know very well, E is the total energy (sum of kinetic, potential, and
whatever else is relevant) of the system, and S is the total entropy. In other
words, what you find in a box is a maximal harmony between the efforts to
minimize the energy (which lowers F ), and simultaneously try to maximize the
entropy (which also lowers F ). Loss of energy is often accompanied by loss of
order, so the two quantities, E and T S, are typically in competition with each
other. For a very high temperature T , entropically favorable state will be found
because T S E, and for a very low temperature we can largely ignore the
entropy effect and expect to find a minimum energy state being realized.
Here comes the great revelation of Landau and his famous disciple, Ginzburg3 .
They assumed that it is possible to write down the free energy F as a function
of its own order parameter which, for spin systems, is Stot . For convenience
we will introduce an average magnetization S ≡ Stot /N . Free energy is an ex-
tensive quantity so it will be proportional to N , but otherwise the free energy
density f ≡ F/N is assumed to depend solely on S. Further assume that f (S)
is an analytic function of S. All that means is that one can Taylor-expand f (S)
in powers of S for small S:
1
f (S) = f (0) + (∂α f )Sα + (∂α ∂β f )Sα Sβ + · · · . (1.4)
2
As it stands, however, it doesn’t look very useful. But, wait, let’s think a little
harder.
Say you are holding a piece of magnet with its magnetization pointing in
the +ẑ direction. You flip it, and the magnetization is pointing in −ẑ direction.
But it is the same piece of magnet after all! No matter how you rotate the
magnet on the palm of your hand it will be the same magnet, having the same
free energy f . In the mean time, the order parameter S has clearly changed its
orientation from +ẑ to −ẑ to whichever direction you rotated the magnet into.
This means that free energy cannot depend on the orientation of S, but only on
its magnitude. Then, f must be a function of S · S. Given this, we can re-write
the Taylor expansion as
t λ
f (S2 ) = f (0) + S2 + (S2 )2 + · · · . (1.5)
2 4
3 Ginzburg shared the 2003 Nobel prize for his development of Ginzburg-Landau theory.
10CHAPTER 1. GINZBURG-LANDAU THEORY OF CRITICAL PHENOMENA
the symbol (−1)ix +iy alternating between ±1 as the site i = (ix , iy ) changes
by one lattice constant. Two dimensional coordinates are used for convenience.
The magnetic moments are random as usual at high temperature, and begin to
take on the form, Eq. (1.6), below TN , which is called the Néel temperature
after Louis Néel who first discovered the existence of AFM.
What is the order parameter for P a Néel-ordered state? A naive application
of the prior definition S = (1/N ) i Si will give zero whether above or below
the transition. Rather, we should define a new order parameter
1 X
SN = (−1)i Si , (1.7)
N i
where the prefactor (−1)i = (−1)ix +iy has the effect to cancel out the alternat-
ing directions of spin alignment in Eq. (1.6). Expansion of the free energy in
terms of the Néel order parameter SN gives Eq. (1.5) with S → SN . From the
GL perspective, AFM and FM are the same entities5 .
Now imagine a peculiar kind of magnet, in which the magnetic moment
varies in the manner
2π 2π
Si = S cos ix x̂ + sin ix ŷ (1.8)
3 3
at low temperature. That is, spin direction rotates with a period of 3 lattice
constants when traversing the lattice along the x̂-direction. Once again we are
in 2 spatial dimensions. Can we write down an order parameter for such a
magnet and a free energy expansion in terms of it?
In defining an order parameter we should try to multiply each spin Si with a
suitable pre-factor so that, when ordered, each rotated spin S0i is pointing in the
same direction regardless of the site. Well, it’s pretty easy to guess the correct
answer in this case: you simply rotate each spin by the angle −2πix /3!
0 2πix 2πix
Six = cos Six + sin Siy
3 3
0 2πix 2πix
Siy = cos Siy − sin Six . (1.9)
3 3
The new spin S0i is always pointing in the x̂ direction in the ordered phase.
4 Some people call it Landau-Ginzburg (LG) free energy.
5 Laterwhen we study mean-field theory this equivalence between AFM and FM becomes
more exact.
1.4. ITS USES AND LIMITS 11
1 λ µ
f= t|S|2 + |S|4 + (S × S∗ )2 + · · · . (1.12)
2 4 2
time we will proceed to work out the consequences of the GL free energy we
have just derived.
In accordance with the principles of statistical mechanics, a system that’s
realized at a given temperature T is the one that minimizes f . For Eq. (1.5)
this condition gives
S[t + λS2 ] = 0. (1.13)
The solution of this is either S = 0 or, if t < 0, S2 = −t/λ. S = 0 means
there is no macroscopic average of spins, and implies a paramagnetic p phase.
If S2 = −t/λ, there is a net magnetic moment of magnitude −t/λ. This
is a sensible description of the ordered phase provided that −t were in fact
proportional to Tc −T . Then t changes sign precisely at the critical temperature
Tc and the magnetic moment develops6 . Whether t behaves as T − Tc is once
again up to the microscopic theory to verify. In the case of Ising model this can
be derived rather straightforwardly.
In summary, the assumption of the free energy expanded in terms of an
order parameter, plus the dependence of the coefficient t as T − Tc is sufficient
to guarantee that there will be √ a phase transition. Moreover, the expected
temperature dependence is |S| ∼ −t, which is subject to a ready experimental
test. Such square-root dependence of the order parameter on the temperature
is found in a wide variety of systems.
For the general case of Eq. (1.12) the analysis gets more complicated, but
not its spirit. One differentiates Eq. (1.12) with respect to S and S∗ , which are
independent quantities, and solve for the two coupled equations. If there are
several solutions, the one which gives the lowest free energy is the true solution.
The interesting exercise of working out the phase diagram for Eq. (1.12) is left
as an exercise.
So, as we have seen, the beauty of GL theory is that it is capable of producing
a qualitative phase diagram of the system being probed without knowing much
of the microscopic details about it. All that’s required is that you identify the
right order parameter.
A drawback of GL theory, on the other hand, is that it entirely ignores the
effects of fluctuations. When you measure, say, the magnetization, it does not
always give the exact same value from measurement to measurement. You might
think that’s always the case, but no, that’s not always the case. If you measure
the charge of an electron using some probe, you do so expecting to find always
the same number regardless of how, and when you measure it. If two persons
got different numbers of the charge, that’s because the measurements were not
perfect. With the magnetization, on the other hand, I’m saying that you should
expect to get a different number each time you measure even if your apparatus
were operating in perfect order!
Let me explain this using the total magnetization Sz as an example. Its
squared average, hSz i2 , is in general different from the average of the square,
hSz2 i, because there is nothing in the mathematics to guarantee that they ought
2 2
to be the same. This in turns implies that p h(Sz − hSz i) i = h(∆Sz ) i 6= 0.
2
The square root of this quantity, ∆Sz ≡ h(∆Sz ) represents the degree of
deviation in the measured value of Sz from its average value hSz i. If ∆Sz is
much greater than hSz i, the very concept of average loses its meaning.
6 For t < 0 one must compare the free energies of the two solution, S = 0 and S2 = −t/λ.
We saw that hSz i vanishes continuously to zero near Tc . Unless ∆Sz vanishes
faster than hSz i near the critical temperature, we have no reason to trust the
predictions of the GL theory. In reality, the fluctuation ∆Sz grows larger as Tc
is approached. So, at some temperature TG (after Ginzburg) below Tc we have
∆Sz /hSz i ∼ 1 and the GL theory breaks down.
Ginzburg himself noted this fact and was able to estimate the temperature
at which his theory begins to break down. His method of estimation is known
as the Ginzburg criterion. GL theory is only trustworthy for temperatures
below TG , not below Tc ! In the following chapter we will discuss how to derive
Ginzburg criterion for the Ising model. In the mean time let me assure you that
in certain systems Tc − TG is a very small fraction of Tc , so that one can in
practice never distinguish the two temperatures. In that case, it’s OK to ignore
the fluctuations.
two S’s and one ρ. This can be S · Sρ∗ , S∗ · S∗ ρ or |S|2 ρ. The last term is
forbidden by translational symmetry.
Under a lattice translation by R, we find S → e−iks ·R S, and ρ → e−ikc ·R ρ.
The combination |S|2 is invariant under this transformation, but |S|2 ρ becomes
e−ikc ·R |S|2 ρ. Unless kc equals one of the reciprocal lattice vectors, we cannot
use this term in the GL expansion. We are not considering Umklapp terms here,
and therefore kc must be zero. But that means there is no charge modulation!!
On the other hand, S · Sρ∗ transforms as ei(kc −2ks )·R S · Sρ∗ . Such a term is
OK provided the two modulation vectors are correlated: 2ks ≡ kc . Spins are
modulated along the same direction as the charge with a period which is twice
that of the charge.
In LSCO (La2−x Srx CuO4 ) family of cuprates, spin and charge modulations
are observed with the modulation periods in accordance with the predictions
of the GL theory. We were able to show from simple GL analysis that, should
there be a simultaneous ordering of spin and charge, and should there be any
interaction between the two, then it must happen that the modulation vectors
are related by the condition 2ks = kc .
The argument has a bit of a flavor of Dirac’s argument for the quantization
of electric charge. GL theory does not say there must exist a coupling between
spin and charge degrees of freedom (Dirac never said magnetic monopoles should
exist). But should there be an interaction between the two (If there is a magnetic
monopole somewhere), then they must obey the rule 2ks = kc (electric charge
must be quantized). The rule follows from translational symmetry (Dirac’s
quantization follows from gauge symmetry).
These days, more and more people are devoting their attention on com-
pounds that have complex ground states with ordering of spins, orbitals, and
even electric dipole moments. Microscopic description of this is undoubtedly
very hard. Even writing down the right Hamiltonian for such systems appears
to be very difficult, not to mention solving one. I believe GL theory for multiple
order parameters can do nicely here, and show some interesting results before
a heavier machinery can be brought to bear. And remember, now you know as
much about how to write down GL free energy as anybody else on earth!!
Chapter 2
Mean-field Theory
Classical spins are vectors, of fixed magnitude |Si | = S, but having an arbitrary
orientation on the 3-sphere. Often the magnitude S is absorbed in the definition
of Jij and the rotational degrees of freedom is expressed as a uni-modular vector,
Si · Si = 1. Furthermore, we treat the case of nearest-neighbor interaction only,
so that Jij is non-zero if and only if i and j are the two adjacent sites of
the lattice. The sign of J can be either negative, in which case the model is
ferromagnetic, or positive (antiferromagnetic).
The discussion of phase transition in the previous chapter was based on
a bunch of assumptions: First you needed to define an order parameter, then
assume that free energy can be expanded in terms of it. Secondly, the coefficient
in front of the square term in the expansion had to change sign at some finite
temperature in order to ensure that phase transition does occur. How do we
know that all these things happen in a realistic calculation of the free energy?
To dispel doubts and place GL theory on a firmer ground we must turn to a
microscopic derivation of the free energy, and be able to prove the existence of a
phase transition from the model such as the one given in Eq. (2.1). But before
we do that, let us simplify the model even further.
What do the other, non-magnetic atoms do in a compound? Well, the wave
functions of the surrounding non-magnetic ions can affect the spin dynamics of
the magnetic atoms in various ways. As a result, sometimes the environment
around each magnetic ion favors the spin direction to lie along, say, ẑ direction
much more than along the x̂ or ŷ. In that case, Eq. (2.1) is not quite correct,
because it says that two spins lying in the x̂ direction is just the same as the
spins lying along the ẑ direction as far as their interaction energy is concerned
1 Details of the mechanism for the spin-spin interaction is outside the scope of this lecture.
15
16 CHAPTER 2. MEAN-FIELD THEORY
with |Jz | |Jx |, |Jy |. Well, since Jx and Jy are tiny, we shall drop it altogether
in the first approximation and treat Jz only:
X
E = Jz Siz Sjz . (2.3)
ij
This is the celebrated Ising model, probably the most famous model in all of
many-body physics. From here and forward we shall drop the z altogether,
and treat the ferromagnetic case Jz = −J < 0 first. Due to the uni-modular
condition, Siz can only point either up (Siz = +1), or down (Siz = −1). It is
claimed that a simple model like the Ising model above is sufficient to capture
the qualitative features of the phase transition behavior.
Ising model is a many-body model in the sense that spin at site i interacts
with its neighboring spins at j each of which interacts with its own neighbors
at k, ad infinitum. In essence, two spins which are very very far apart manage
to affect, and be affected by, each other. In this regard, solving for the Ising
model is a lot like solving coupled, multi-variable equations. For the Ising
model, the number of variables equals the number of spins, which is a huge
number of Avogadroian scale. To get the answers you must know the values
of all the variables at once. In most cases, of course, exact, analytic answers
won’t be available, and one must rely on approximations. However, the nature
of approximation must be such that you can later make systematic improvement
should you wish to get a more accurate answer. And often times, a reasonable
approximation is the one based on sound physical picture. A general guiding
principle for reducing a hard, interacting problem to a simple, non-interacting
problem has been found a long time ago, and goes under the title of “mean-field
theory”2 .
Let’s look at the energy Eq. (2.3) from the perspective of a spin at one
particular site, i. Then the piece of the total energy that it cares about is given
by X
Ei = −Si (J Sj ). (2.4)
j∈i
Here we introduced j ∈ iPto indicate a sum over all nearest neighbors of i. For
a moment write hi ≡ J j∈i Sj , then Ei becomes −hi Si , which looks exactly
like the energy of a single, isolated spin subject to an external magnetic field
hi . Of course hi is not really an external field; it is a variable, that depends
on the spin values of sites j. If Sj changes, then so does hi . But rewritten in
this suggestive form, it shows that what the neighboring spins do is basically to
provide an effective magnetic field on the i-th spin.
Here comes the key logical loop:
2 Sometimes you hear words like “molecular field theory” or “Weiss theory” which mean the
same thing as the mean-field theory. I prefer to collect all the techniques, both classical and
quantum, which rely on reducing the many-body problem to a tractable one-body problem
under one roof, and call it mean-field theory.
2.1. MEAN-FIELD THEORY OF ISING MODEL 17
The nature of the argument, as you can see, is self-consistent, and the self-
consistency is one of the defining characteristics of any mean-field theory.
So, let’s do make the replacement
X X
hi = J Sj → J mj , (2.5)
j∈i j∈i
In this form, the spins are no longer interacting with each other. Partition
function can be readily derived,
X
Zi = e−βEi = 2 cosh[βhi ]
Si =±1
Y Y
Zmf = Zi = (2 cosh[βhi ]). (2.7)
i i
m = tanh(βJzm) (2.10)
where z is the coordination number indicating how many neighbors there are for
a given site. For a square lattice z equals 4, and for cubic, z = 6. For generality,
we keep z in the expression. The solution of this equation can be found by
graphical means, or by two short lines of programming on Mathematica. You
18 CHAPTER 2. MEAN-FIELD THEORY
while Eq. (2.6) without the adjustment will give twice this value as the energy.
So, Eq. (2.11) is right, and Eq. (2.6) is wrong. After correcting for this mistake,
free energy is derived from
X P Y
Zmf ≡ e−βFmf = e−βEmf = e−βJ ij mi mj (2 cosh[βhi ]), (2.13)
C i
and,
X X
Fmf = J mi mj − T ln[2 cosh(βhi )]
ij i
X X X
= J mi mj − T ln[2 cosh(βJ mj )]. (2.14)
ij i j∈i
This is the free energy derived in the mean-field theory. Taking advantage of
our knowledge about the ferromagnetic state we can set mi = m everywhere
and get (f ≡ F/N )
X 1X X
Emf = − hi Si + hi mi − Hi Si ,
i
2 i i
1X X
Fmf = hi mi − T ln (2 cosh[β(hi + Hi )]) ,
2 i i
X
mi = hSi i, hi = J mj . (2.18)
j∈i
m+h
m = tanh[βTc (m + H/Tc )] = tanh . (2.20)
t
where Tc = Jz stands for the mean-field critical temperature when H = 0, and
t = T /Tc the reduced temperature. For h = 0, the magnetization is nonzero
and grows with lowering temperature for T < Tc . For h > 0, the magnetization
remains positive and grows with lowering temperature. For h < 0, the magne-
tization remains negative and grows in magnitude with lowering temperature.
If one fixes the temperature at T > Tc and sweeps h from positive to negative,
the magnetization will change continuously from a positive, through zero, to
a negative value. If the temperature fixed is less than Tc , the magnetization
upon the field sweep remains positive for h > 0, then suddenly jumps over to a
negative value as h turns negative. The amount of magnetization jump is given
by twice the value determined by m = tanh[m/t] for a fixed t.
The magnetic susceptibility dm/dh is obtained from differentiating Eq. (2.20),
dm 1 1 − m2
= 2 = 2 . (2.21)
dh T cosh [(m + h)/t] − 1 m +t−1
For t > 1 the susceptibility never√diverges. For t < 1 the susceptibility diverges
at the point m2 = 1 − t, m = ± 1 − t.
This model is also known as the XY model for an obvious reason. A spin is free to
rotate within the XY plane. What are the possible phases and thermodynamic
properties of this model?
As a first step, we again use mean-field theory to reduce the energy functional
Eq. (2.22) to a non-interacting form. Writing Si = (Six , Siy ), we get
X X 1X
E=J Si · Sj → Emf = hi · Si − hi · mi (2.23)
ij i
2 i
P
where mi = hSi i, and hi = J j∈i mj . OnceP again we have to make adjustment
for the zero-point energy ∆Emf = −(1/2) i hi · mi . I hope these steps are
familiar to you now. Once we have obtained the non-interacting model this
way, it is trivial to get the partition function (I0 =modified Bessel function)
Z 2π
Zi = dθi e−βhi ·Si = 2πI0 (β|hi |). (2.24)
0
2.4. MEAN-FIELD THEORY OF XY SPINS 21
∂ ln Zi I1 (β|hi |)
mi = − =− ĥi ≡ −R(β|hi |)ĥi . (2.25)
∂(βhi ) I0 (β|hi |)
We get the magnetization pointing opposite to the local field ĥi ≡ hi /|hi | be-
cause of the antiferromagnetic nature of the coupling.
How are these results different from the Ising spins? Well, not very much. If
you plot Zi obtained above, it looks pretty similar to cosh[x] which was the parti-
tion function for the mean-field Ising solution. The function R(x) ≡ I1 (x)/I0 (x)
looks formidable, but if you plot it, you won’t see a dramatic difference from the
tanh(x) curve we found for the Ising model. Finally, self-consistency condition
expressing hi in terms of magnetization becomes
X
hi + J R(β|hj |)ĥj = 0. (2.26)
j∈i
We have all the ingredients to construct the mean-field free energy for the
XY spins:
X 1X X
F = ∆E − T hi · mi − T
ln Zi = − ln[I0 (β|hi |)]
i
2 i i
J XX X X
= − mi · mj − T ln[I0 (βJ| mi |)]
2 i j∈i i j∈i
Jz 2
→ N× m − T ln[I0 (βJzm)] . (2.27)
2
Fluctuations and
Excitations
23
24 CHAPTER 3. FLUCTUATIONS AND EXCITATIONS
below1 .
A single domain wall of the perimeter s contains approximately (s/4)2 num-
ber of flipped spins. The total number of down spins is then
X s 2
N↓ ∼ Ns P s , (3.1)
s
4
where Ns is the number of different ways in which domain walls of a given length
s can be formed, and Ps is the Boltzmann probability for the occurrence of such
a domain wall.
Figure 3.1: A domain wall configuration of spins in Ising model. Regions shaded
in blue have their spins reversed from the background spins.
A single domain wall costs an energy +2Js to excite, because there are a
total of s “bad bonds”, each of which costs an energy 2J. The probability of
thermally exciting such a configuration is given by Ps = e−β∆E = e−2βJs . This
is an exponentially small number at a sufficiently low temperature, so to a first
approximation we may assume that there is at most one such domain wall in
the entire system, leading to the estimate in Eq. (3.1).
The number of configurations Ns of a given perimeter s can be estimated
by noting that in constructing such a domain wall, one can proceed by laying
down a series of “sticks” connecting the nearest neighbors until all s sticks are
used up. For each step there are at most 3 different ways in which the stick can
be laid, and assuming that each move is independent of all previous moves, we
have
Ns ∼ N · 3s . (3.2)
The prefactor N arises from N different points from which a domain wall of a
given shape can originate. Thus,
∞ 2 2 X∞
X s N d
N↓ ∼ N 3s e−2βJs = x xs
s=0
4 16 dx s=0
1 If you like to read a more rigorous argument, refer to Robert B. Griffiths, Phys. Rev.
2
N d 1 N x(1 + x)
= x = . (3.3)
16 dx 1−x 16 (1 − x)3
and the equation of motion for the individual spin Si that follows from it. We
first remind ourselves of the fundamental commutation relations of spin
Combined with the equation of motion for the y- and z-components, we get
dSi X
= −J Sj × Si . (3.8)
dt j∈i
26 CHAPTER 3. FLUCTUATIONS AND EXCITATIONS
d X
δSi = −J mj + δSj × mi + δSi
dt j∈i
X X X
≈ −J mj × mi − J mj × δSi − J δSj × mi (3.9)
j∈i j∈i j∈i
2
ignoring terms of order δS . Once again, the justification is that δS is a small
quantity, and (δS)2 , even smaller. We take the quantization axis ẑ and write
mi = mẑ everywhere. Then2
X
d
δSi = −Jzmẑ × δSi − Jm δSj × ẑ. (3.10)
dt j
d z
The ẑ-component of this equation reads dt δSi = 0, hence the fluctuation in the
2 I apologize for the multiple use of z here: z is the number of neighbors, or the coordination
For a 3-dimensional cubic lattice,P j∈i eik·(rj −ri ) = 2(cos kx + cos ky + cos kz ),
P
and with z = 6, ωk = 2Jm(3 − α cos kα ). For 2D square lattice we will get
ωk = 2Jm(2 − cos kx − cos ky ). This is the desired spin wave energy for a spin
disturbance associated with a wavevector k.
then verify that the funny new expressions on the r.h.s. of the above equations
satisfy all the commutation algebras of ordinary spin operators: [Siα , Sjβ ] =
iδij αβγ Siγ , provided the new operator bi is a canonical boson operator, with
the familiar bosonic commutation relations, [bi , b+
j ] = δij , etc.
How this comes about does not concern us so much here; rather we will
proceed to apply the brand new technique and see if something good comes out.
First rewrite
1
Si · Sj = Siz Sjz + (Si+ Sj− + Si− Sj+ )
2
1 1 1
= (S − ni )(S − nj ) + (2S −ni ) 2 (2S −nj ) 2 (b+ +
i bj + bj bi ). (3.16)
2
3 This section is largely optional on first reading.
28 CHAPTER 3. FLUCTUATIONS AND EXCITATIONS
Well, now the new Hamiltonian is just a quadratic Hamiltonian involving only
a product of two boson operators. Such operators can be diagonalized easily, in
this case by going to the Fourier space:
X
bi = eik·ri bk . (3.19)
k
Substitution of the Fourier expression for bi into Eq. (12.31) immediately gives
X
H= ωk b+
k bk (3.20)
k
where
ωk = JSz(1 − γk )
X
γk = z −1 eik·(rj −ri ) . (3.21)
j∈i
The eigenenergy ωk agrees with the spin-wave spectrum derived earlier, in Eq.
(12.10), using the equation-of-motion theory.
We have derived the same spin wave excitation energy using two completely
disjoint methods. The real power of the HP boson theory lies in the simplicity it
offers in dealing with excitation spectra of complicated magnetic systems, such
as those of an antiferromagnet. The strategy is essentially the same, but the
technique is more advanced. The following segment is for those of you who like
3.2. SPIN WAVES FOR CONTINUOUS SPIN 29
where θi is the angle of the classical ground state spins, θi = π(ix + iy ). This
represents a rotation of spins, with respect to the x-axis, by the angle that just
equals the orientation of the classical ground state spins. The x-component
remains unchanged, hence Six = Six . Furthermore, one can easily verify that
the new operators Siα satisfy all the commutation rules of the original spins, Siz ,
hence they are as good a representation of the local spin operator as the previous
one we used. In terms of Si , we have the antiferromagnetic Hamiltonian written
as
X JX + + X
Si Sj + Si− Sj− − J Siz Sjz .
H=J Si · Sj → (3.23)
ij
2 ij ij
Note that we have two raising (lowering) operators instead of one raising and one
lowering, as was the case before. The ground state of the new Hamiltonian, Eq.
(12.36), is given by the rotation of the antiferromagnetic spins, hSiz i = Seiθi .
But this implies that in the new representation, hSiz i = S everywhere! Since
hSiy i = 0 initially, we have hSiy i = 0, too. So the new Hamiltonian has a
ferromagnetic ground state with all the spins pointing in the +ẑ direction, and
we can make the assumption hSiz i ≈ S, or hni i S after the HP substitution,
for the small-fluctuation analysis.
Under these assumptions the HP-transformed Hamiltonian becomes
X
H ≈ JS (b+ +
i bj + bi bj + ni + nj ). (3.24)
ij
Unlike its ferromagnetic counterpart, Eq. (12.31), this Hamiltonian has a prod-
uct of two annihilation (creation) operators. As before, we take the Fourier
transform and obtain
Xh γk + + i
H = JSz b+
k bk + (bk b−k + bk b−k ) . (3.25)
2
k
Whereas doing the Fourier transform was sufficient to yield excitation en-
ergies in the previous example of ferromagnetic spins, now the Hamiltonian
written in momentum space fails to give eigenenergies. The final trick lies in
working out the equation of motion of the boson operators, which gives
[H, bk ] = −JSz bk + γk b+
−k . (3.26)
Here it’s clear why we could not get the energy from Fourier transform alone.
The dynamics of the operator bk is coupled to that of b+ −k , and the dynamics of
b+
−k is coupled to that of bk by the equation
[H, b+ +
−k ] = JSz b−k + γk bk . (3.27)
Noting that [H, X] = −idX/dt in quantum mechanics, and that for the eigen-
modes we can re-write the time derivative as −iE, where E is the eigenenergy,
30 CHAPTER 3. FLUCTUATIONS AND EXCITATIONS
Each spin is like the arms of a clock, free to rotate within the plane. Indeed, a
spin can be characterized by an angle as (Six , Siy ) = (cos θi , sin θi ) if we normalize
the magnitude to be one. In terms of the angle variable the XY model becomes
X
HXY = −J cos(θi − θj ). (3.30)
ij
There are of course spin wave excitations in this model. Their properties can
be understood using the techniques already explained in the previous section.
Another important class of excitations can be found in the planar model which
are known as the “vortices”. A vortex is a configuration of spins given by
yi − y0
θi = ± arctan . (3.31)
xi − x0
Fv = (J − T ) log(L/a). (3.32)
5 There is a subtlety involved, because long-range order of spins is already lost even before
we consider effects of vortices. Spin wave excitations turn out to be sufficiently violent to
destroy all long range order. So a more precise statement is that proliferation of vortices
above Tc makes the spins from being disordered to even more disordered. This is therefore
quite different situation from the usual kind of transition with the order parameter vanishing
to zero.
6 Some people call it BKT transition, named after another physicist, V. L. Berezinskii, who
worked on the same model a few years earlier than Kosterlitz and Thouless. KT knew about
Berezinkii’s work, but that did not stop them from working out their own, and more complete
theory of the phase transition in the planar model. Thouless himself calls it a “topological
phase transition” in his book.
Chapter 4
XY Model
4.1 Definition
The XY model, as it is commonly known, is given by the classical energy function
X
EXY = −J cos(θi − θj ), (4.1)
ij
where J can be of either sign, and ij spans all the nearest neighbors of the lattice.
The model can be defined in arbitrary dimension, with the two-dimensional case
being the most interesting in view of its critical properties.
The PXY model can be obtained as the planar limit of the spin model,
H = J ij Si · Sj where each spin is represented by the planar component
Si =(cos θi , sin θi ). The signs of J then corresponds to ferromagnetic (J > 0) or
antiferromagnetic (J < 0) spin exchanges, respectively.
The energy function, Eq. (4.1) is invariant under the global rotation of
all spins, θi → θi + α. In addition, it is periodic in each variable θi with
the periodicity of 2π. This latter symmetry reflects the angular nature of the
variable θi . In dealing with the statistical mechanics of the XY model, one often
defines the classical action and the partition function as
X
S = K (1 − cos[θi − θj ])
ij
!
Z 2π Y
Z = dθi e−S . (4.2)
0 i
33
34 CHAPTER 4. XY MODEL
1
hSi · Sj i ∼ hei(θi −θj ) i ∼ η(T )
. (4.3)
rij
Spontaneous symmetry breaking, which would imply hSi ·Sj i → m2 as rij → ∞,
clearly does not occur.
At a sufficiently high temperature, T J, or K 1, one must arrive at
a disordered phase characterized by the exponential decay of the correlation
function
e−rij /ξ
hSi · Sj i ∼ . (4.4)
rij
The behavior of the two correlation functions, Eqs. (4.3) and (4.4), are qualita-
tively different. A phase transition separating the two regimes is expected at a
critical temperature, TKT . The phase diagram of the XY model is qualitatively
described in Fig. 4.1.
In Sec. 4.2, we present the spin-wave theory which describes adequately
the low-temperature behavior of the model. Then in Sec. 4.4 we introduce the
concept of vortex excitation and the vortex-drive phase transition known as the
Kosterlitz-Thouless (KT) transition.
XX 0
h ih 0
i
[θi − θj ]2 = ei(k−k )·ri eik·(ri −rj ) − 1 e−ik ·(ri −rj ) − 1 θk θ−k0 , (4.6)
k k0
4.2. SPIN-WAVE THEORY 35
0
ei(k−k )ri = δkk0 , Eq. (4.5) can be rewritten as
P
Because i
K XX K XX
S= [θi − θj ]2 = (1 − cos[k · rij ]) θk θ−k . (4.7)
4 i j∈i 2 j∈i k
XX 0
h ih 0
i
h(θi − θj )2 i = ei(k−k )·ri eik·(ri −rj ) − 1 e−ik ·(ri −rj ) − 1 hθk θ−k0 i.
k k0
(4.11)
1 1
hθk θ−k i = ,
2K D
P
α=1 (1 − cos[k · eα ])
1 X 1 − cos[k · (ri − rj )]
h(θi − θj )2 i = PD . (4.12)
K α=1 (1 − cos k · eα )
k
P
Contribution to the sum k is dominated by the small-k region, where one can
PD
approximate α=1 (1 − cos k · eα ) ≈ k2 /2,
d2 k 1 − cos[k · (ri − rj )]
Z
2
h(θi − θj )2 i = . (4.13)
K 4π 2 k2
Integrating over all angular orientations of ri − rj with the same magnitude rij
gives
Z
1 dk
h(θi − θj )2 i = [1 − J0 (krij )] . (4.14)
πK k
One can separate the domain of integration into two: 0 < k < 1/rij , and 1/rij <
k < 1/a. The first integration gives a numerical factor which is independent
36 CHAPTER 4. XY MODEL
of rij and can be ignored. For the second integration one can ignore the fast-
oscillating J0 (krij ) and obtain
Z 1/a
1 dk 1 r
ij
h(θi − θj )2 i ≈ ≈ ln ,
πK 1/rij k πK a
1
2πK 2πJ T
i[θi −θj ] a a
he i≈ = . (4.15)
rij rij
This calculation indeed shows that the thermally activated small spin fluctuation
is capable of destroying the long-range order of spins at arbitrary non-zero
temperature. Note, however, the critical exponent η(T ) ≡ T /2πJ controlling
the rate of decay becomes smaller and eventually becomes zero as T → 0.
where the sequence of ij bonds forms a closed loop. Evaluation of the numerator
involves the integration of the function
s
∞ s
X X 1 K X
ei(θI −θJ ) exp K eiθIJ eiθij + e−iθij .
cos θij =
ij s=0
s! 2 ij
(4.18)
By the same reasoning as above, the only non-vanishing contribution arises from
the path that joins the two points I and J. If the separation rIJ is about m,
then the nonzero integral arises from s ≈ m in Eq. (4.18) and there are roughly
m! different ways to connect the two end points, leading to the estimate of the
numerator m
K
NUMERATOR ≈ . (4.19)
2
4.4. VORTEX EXCITATION AND KOSTERLITZ-THOULESS TRANSITION37
The denominator, at each order s of the expansion, has roughly s! different ways
to produce the non-zero combination of bond angles, and gives
∞ s
X K 1
DENOMINATOR ≈ ≈ ≈ 1. (4.20)
s=0
2 1 − K/2
Here ξ = 1/ln(2T /J) defines the correlation length. Higher the temperature,
shorter the correlation length as expected.
−(yi − y0 ) xi − x0
∇θi = ± , , (4.23)
(xi − x0 )2 + (yi − y0 )2 (xi − x0 )2 + (yi − y0 )2
R
and then integrating it ∇θi · dr along a closed path encircling the center
(x0 , y0 ). The quantized value ±2π is independent of the path chosen, as long as
the different paths enclose the same origin. A finite circulation, as it is called,
is the definition of a vortex, and defines the “topological” (path-independent)
property of this peculiar excitation.
Sufficiently far from the core of the vortex where the change of the spin angle
is too swift between nearby spins to allow a continuum description, one can still
apply Eq. (4.5) to evaluate the energy cost of creating a single vortex.
Z Z L
J 2 2 J rdr L
E≈ [∇θ(r)] d r ≈ × 2π = πJ × log . (4.24)
2 2 a r2 a
The lower cutoff a represents the typical size of the core where the continuum
approximation would fail, and L is the linear dimension of the lattice we are
considering. It is a big cost in energy, in principle diverging to infinity as we
try to create it for a larger system size L. Making such an excitation would be
suppressed due to the tiny Boltzmann weight associated with it.
38 CHAPTER 4. XY MODEL
As long as the temperature is lower than Tc ≡ TKT πJ/2, we are better off
without a vortex, because the energetic cost to create one well surpasses the
entropy gain. For T > Tc , however, the situation is reversed, and we gain free
energy by proliferating the system with vortices.
A second type of consideration leads to the same estimate of the critical
temperature. Imagine creating a pair of vortices each having the opposite cir-
culation, or vorticity. For a size much larger than the separation of the vortex
cores all the spins point in the same direction without an energy cost. Then one
can estimate the energy cost of creating a vortex-antivortex pair as 2πJ ln(r/a),
where r now represents the separation of the two cores. Then the mean separa-
tion of the vortices is calculated to be
R∞ 3
2 r dre−2πβJ ln(r/a) πβJ − 1 TKT − T /2
hr i = Ra∞ 3 −2πβJ
= = . (4.26)
a
r dre ln(r/a) πβJ − 2 TKT − T
elaborate theory for the nature of vortex-driven phase transition is worked out
by Kosterlitz and Thouless in 1973 and bears the name of KT transition.
A more thorough analysis of the KT transition involves the use of the real-
space renormalization group analysis, which lies outside the scope of this thesis.
X JX + −
H = −Jz Siz Sjz − (Si Sj + Si− Sj+ ), (4.27)
2
hiji hiji
with Si± = Six ± iSiy . Jz = 0 corresponds to the XY spin model. Now a phase
twist θ is imposed across the x-direction,
J X iθ/L −
H[θ]−H[0] = − (e − 1)Si+ Si+x + c.c.
2 i
J iθ X + − Jθ2 X + −
≈− (Si Si+x − Si− Si+x
+
)+ (Si Si+x + Si− Si+x
+
). (4.29)
2L i 4L2 i
At this point it is already clear that the difference between XY model and XXZ
model disappears and one can use the same definition of the helicity modulus
for both spin systems. Difference in free energy evaluated to second order in θ
gives
1 Michael E. Fisher, Michael N. Barber, and David Jasnow, Phys. Rev. A 8, 1111 (1973).
2 For instance, see N. Schultka, Phys. Rev. B 55, 41 (1997).
40 CHAPTER 4. XY MODEL
Dθe−βH[θ]
R
Z[θ]
=R = e−β(F [θ]−F [0])
Z[0] Dθe−βH[0]
!2
βJθ2 X + − − + β 2 J 2 θ2 X + − − +
=1− h (Si Si+x + Si Si+x )i − h (Si Si+x − Si Si+x ) i,
4L2 i 8L2 i
!2
Jθ2 X − βJ 2 θ2 X + −
F [θ]−F [0] ≈ 2 h (Si+ Si+x + Si− Si+x
+
)i + 2
h (Si Si+x − Si− Si+x
+
) i.
4L i
8L i
(4.30)
Helicity modulus follows from the definition F [θ] − F [0] = (ρs /2)θ2 :
!2
J X J2 X
ρs = 2 h (Six Si+x
x
+Siy Si+x
y
)i− 2
h y
(Six Si+x −Siy Si+x
x
) i. (4.31)
L i
T L i
hM 4 i
UL = 1 − , (4.32)
3hM 2 i2
where M is a order parameter. By using this formula, we can find the transition
temperature for chirality and Ising orders. If you calculate the Binder cumulent
for different lattice size, the line is cross at some point. This crossing point is
the phase transition temperature and this is the same effect to get the critical
temperature at thermodynamic limit.
∞
X 1 2
P [φij ] ∝ e K(cos[ϕij ]−1)
→ e− 2 K[ϕij +2πnij ] (4.34)
nij =−∞
For each link ij, the formula can be used to re-write the probability weight
∞ ∞ Z ∞
− 12 K[ϕij +2πnij ]2 1 2
X X
e = dθij e− 2 K[ϕij +2πθij ] +2πilij θij
. (4.36)
nij =−∞ lij =−∞ −∞
∞
X 1 2
P [ϕij ] ∝ e− 2K lij −ilij ϕij (4.37)
lij =−∞
li,i+x (ϕi − ϕi+x ) + li,i+y (ϕi − ϕi+y ) + li−x,i (ϕi−x − ϕi ) + li−y,i (ϕi−y − ϕi )
X
→ ϕi [li,i+x + li,i+y − li−x,i − li−y,i ] = ϕi [li,i+x + li,i+y + li,i−x + li,i−y ] = ϕi lij .
j∈i
(4.40)
The last manipulation follows from the antisymmetry lij = −lji of the link
field. The coupling term has a geometric interpretation. Think of lij as a
current flowing from i to j. It is also clear from such an interpretation that
lji = −lij . Then the expression that couples to ϕi is the total current flowing
42 CHAPTER 4. XY MODEL
Figure 4.2: Integer-valued link (current) variable lij emanating from site i on a
square lattice. The sum of the current flowing away from i must be zero.
Figure 4.3: Representation of the original and dual lattice sites labeled by lower
case and upper case letters, respectively.
out from i, which is a conserved quantity in the case of conserved charge. This
is in fact the discretized version of divergence ∇ · l.
the P
−i i ϕi ( j∈i lij )
P
Each ϕi appears linearly, e P , in the partition function. Inte-
grating over ϕi yields zero unless j∈i lij = 0, which acts as a constraint on the
possible values of lij . The constraint is solved by writing lij as the difference
lij = hI − hJ (4.41)
of another integer fields hI defined on the dual lattice. In the picture shown in
Fig. 4.3, we will have
li,i+x = hI − hI−y
li,i+y = hI−x − hI
li,i−x = hI−x−y − hI−x
li,i−y = hI−y − hI−x−y . (4.42)
Indeed li,i+x + li,i+y + li,i−x + li,i−y = 0. Initially there were 2N link variables
lij for a lattice of N sites. The constraint being imposed at each site reduces
the number of independent fields to N , which is precisely the number of dual
variables hI . Now the probability weight P [{ϕi }] has lost its meaning because
ϕi ’s have been integrated out. What we have is an equivalent representation of
the partition function Z using the new dual variables,
X − 1 P (h −h )2
Z= e 2K hIJi I J . (4.43)
{hI }
4.7. VILLAIN TRANSFORMATION 43
The new model defined on the dual lattice is called the “height model” because
the greater difference in the heights of the nearby sites results in a larger en-
ergy difference. The high-temperature phase of the XY model corresponds to
the low-temperature phase of the height model and vice versa, because of the
reversed location of K in the Boltzmann factor. The 2D height model is in the
same universality class as the 2D XY model.
We have not yet come to the full vortex representation of the partition
function. To achieve this, invoke once more the Poisson sum formula to get
XZ ∞ 1
P 2 P
Z= dθI e− 2K hIJi (θI −θJ ) +2πi I mI θI . (4.44)
mI −∞
Here mI is another integer field defined at the dual sites I. We will integrate
out θI to obtain the partition functionPsolely in terms of the “vortex variable”
mI . Using the Fourier expansion θI = k θk eik·rI , the quadratic part becomes
1 X 1 X |θk |2
(θI − θJ )2 = . (4.45)
2K 2K Gk
hIJi k
where the Green’s function Gk = (4 − 2 cos kx − 2 cos ky )−1 is used. The linear
part becomes
X X
−2πi mI θI = −πi (θ−k mk + θk m−k ). (4.46)
I k
1 X 1
(θ−k − 2πKim−k Gk ) (θk − 2πKiGk mk ) + 2π 2 Km−k Gk mk . (4.47)
2K Gk
k
d2 k eik·rIJ
Z
GIJ = a2 2 . (4.49)
(2π) 4 − 2 cos(kx a) − 2 cos(ky a)
It is convenient to use the regularized Green’s function G0IJ
!2 !
X X X
Z= exp −2π 2 KG0 mI × exp −2π 2 K mI G0IJ mJ .
{mI } I IJ
(4.51)
Since G0 ∼ (1/2π) ln(L/a), the allowed configurations are those with total vor-
ticity zero. With the help of the result at large distances
44 CHAPTER 4. XY MODEL
1 rIJ 1
G0IJ ∼ −
ln − , (4.52)
2π a 4
the partition function may be rewritten as
0 2
X π K X X r
IJ
Z= exp − m2I + πK mI mJ ln . (4.53)
2 a
{mI } I I6=J
final expression above. Often the fugacity term is defined by y = exp(−π 2 K/2).
This term controls the number of (total) vortices in the system. Equation
(4.53) is the desired action expressed solely in terms of the vortex variables mI .
Equation (4.53) is the start of renormalization group analysis of Kosterlitz.
With K large, the only kind of vortex states that can make a meaningful
contribution to the partition function are those with
P m I = 0, ±1. Going back to
2 2
Eq. (4.44), inserting the fugacity term e−(π K/2) I mI by hand, and restricting
the sum of mI to 0, ±1,
Z ∞ Yh i
1
P 2 2
Z ≈ dθI e− 2K hIJi (θI −θJ ) 1 + e−π K/2
(e2πiθI + e−2πiθI ) .
−∞ I
Z ∞ Yh i
1
P 2 2
− 2K hIJi (θI −θJ )
≈ dθI e 1 + 2e−π K/2
cos(2πθI )
−∞ I
Z ∞
1
P 2
− 2K hIJi (θI −θJ ) +2y cos(2πθI )
≈ dθI e
−∞
Z
1
d2 r[∇θ(r)]2 +2y d2 r cos[2πθ(r)]
R R
≈ Dθ(r) e− 2K . (4.54)
We have arrived at the sine-Gordon model. One can analyze this model using
the Wilson’s renormalization group idea. The existence of the confinement-
deconfinement transition in the sine-Gordon model is consistent with the vortex
binding-unbinding transition in the original XY model or the Villain model. All
analysis point to the existence of the defect(vortex)-mediated phase transition
at a temperature set by Tc = (π/2)J. A simpler, variational analysis of the
confinement-definement transition of the sine-Gordon model can be found in
Nagaosa’s book.
√
In writing the boson operator bi = ρi eiθi , ρi = ρ, one obtains Eq. (4.55)
through 2tρ ≡ J. Thus one finds that αij in Eq. (4.55) reflects the influence of
externally imposed flux. The flux around an elementary plaquette is given by
the directed sum
1 X
ϕ= αij . (4.57)
2π
ij∈plaquette
Now the optional choice of angles θi would be to satisfy θi − θj = αij for all the
bonds. For certain geometries and depending on the value of the flux, this may
not be possible to achieve - hence the “frustration”.
The frustration effect can be introduced purely due to the geometric reason.
To see how this happens, we start with an antiferromagnetic XY model on the
triangular lattice
X X X
E=J cos(θi − θj ) = J cos(θi − θj ). (4.58)
ij i j=i+eα
√ √
The three unit vectors are given by e1 = (1, 0), e2 = (− 21 , 23 ), e3 = (− 12 , − 23 ).
We already know that the ground state of the model is obtained with the spin
configuration
0 4π
θi = k · ri , k = ,0 . (4.59)
3
With this knowledge we can replace θi in Eq. (4.58) by θi + θi0 where θi0 is given
by the above equation. On substitution one finds that Eq. (4.58) becomes
XX XX π
E=J cos[θi −θi+eα −k · eα ] = −J cos[θi −θi+eα − ]. (4.60)
i eα i eα
3
Figure 4.4(a) shows the system which is not frustated. If there is an odd
number of AF bond, the formula of ground state is written by
π π
ϕi − ϕj = (1 − εij ) + τij , (4.63)
2 4
46 CHAPTER 4. XY MODEL
Figure 4.4: (a) Even number of AF bond. (b) and (c) Odd number of AF bond.
where τij has the value of ±1 and is the plaquette variable. The rotation of
spin, which is clockwise or counter-clockwise, is determined by the value of τij .
The chirality of plaquette is defined by this τij . The triangular lattice always
has the chirality because it is fully frustrated. However, the square lattice has
no chirality because it is not frustrated.
Another interesting phenomenon by frustration occurs in vortex. If there is
only one frustrated plaquette in all lattice like Fig. 4.5, it has a half-integer
vortex of ground states. If there is other frustration nearby the original one,
two possible cases for the rotation of vortex exist. First, the rotation of vortex
is equal to original half-integer vortex. In this case, it looks like a large vortex
and hence it costs much energy. Second, the rotation of vortex is opposite to
original half-integer vortex. In this case, they cancel out each other. Therefore,
it has a lower energy so that the system prefer this state.
Chapter 5
Heisenberg Model
X
H = −J Si · Sj , (5.1)
ij
where now the uni-modular spin Si can point in any direction in space. The
ground state of the model is the fully polarized state, Si = S0 . Taking this
arbitrary spin orientation to be the z-direction, a small fluctuation of the spin
vector away from equilibrium is denoted as Si = ni . Then the Heisenberg model
can be re-written as
Z
JX J h i
H= (ni − nj )2 ≈ d2 r (∂x nr )2 + (∂y nr )2 . (5.2)
2 ij 2
Z Z !
1 X
Z = δ(n2 − 1) exp − dd r (∂µ nr )2
g µ
Z Z Z !
1 d
X
2 d 2
= exp − d r (∂µ nr ) − d rλr (nr − 1)
g µ
Z Z Z
d 1 2 d
= exp − d r nr − ∂µ + λr nr + d r λr (5.3)
g
47
48 CHAPTER 5. HEISENBERG MODEL
Z
3
Z = exp dd r λr − Tr ln G . (5.4)
2
The factor 3 reflects the three components of the spin vector. The saddle-point
equation follows from taking λr = λ and minimizing with respect to it,
dd k
Z
3 3 1
1 − TrG−1 = 1 − = 0. (5.5)
2 2 k<Λ (2π) λ + k2 /g
d
Λ
Λ2 Λ2
Z
3 g 1 3g
1= kdk = ln 1 + → λg = . (5.6)
2 2π 0 λg + k 2 8π λg exp(8π/3g) − 1
For any coupling strength g (or temperature T ), there exists a finite solution λ.
In turn, λ gives the mass gap, or the inverse correlation length.
Extending the saddle-point analysis to D = d + 1 dimensions is easy with
the substitution
dd k
Z
1
→
(2π)d k 2 +λg
p
β 2 +λg
Z
d kd X 1
Z d
d k coth 2 k 2
= = . (5.7)
(2π)d iω ωn2 +k 2 +λg (2π)d
p
2
2 k +λg 3g
n
Λ2 Λ2
Z
2 kdk 1 T
≈T 2
= ln 1 + → λg = . (5.8)
3g 2π k +λg 4π λg exp(8π/3gT ) − 1
Z Z !
2 2 1 d
X
2 2
Z = dπdσδ(σ + π − 1) exp − d r [(∂µ π) + (∂µ σ) ] (, 5.9)
g µ
Z Z !
dπ 1 X 2
p
2
Z = √ exp − [(∂µ π) + (∂µ 1 − π 2 ) ]
2 1 − π2 g µ
Z Z Z !
1 X 2
p
2 2 ρ 2
= dπ exp − [(∂µ π) + (∂µ 1 − π ) ] − ln(1 − π ) .
g µ
2
(5.10)
Z Z X Z !
1 2 2
ρ 2
Z = dπ exp − (∂µ π) + (π · ∂µ π) + π . (5.11)
g µ
2
Z X
1
S0 = (∂µ π)2
g µ
Z X Z
1 2 ρ
S1 = (π · ∂µ π) − π2
g µ
2
Z
Z = dπ exp (−S0 − S1 ) . (5.12)
r2 − 4R2
nz = cos θ = ,
r2 + 4R2
4R(x + iy)
nx + iny = eiφ sin θ = 2 . (5.13)
r + 4R2
Here R serves as the “radius” of the Skyrmion. A Skyrmion carries with it a
topological number, obtained as the integral
Z
1
S= dr n · (∂x n × ∂y n). (5.14)
4π
For a singly-quantized Skyrmion given in Eq. (5.13) this number equals +1.
There is a convenient formalism for dealing with rotational excitations of
spins by introducing the complex fields z1 and z2 , and re-write the orientation
vector as
50 CHAPTER 5. HEISENBERG MODEL
1
n · (∂x n × ∂y n) = ∂x Ay − ∂y Ax (5.16)
2
where the “vector potential” Ai (i = x, y) is given by
X
Ai = zα∗ ∂i zα . (5.17)
α
X X X
(∂i n)2 = 4 (∂i zα∗ )(∂i zα ) + 4 zα∗ ∂i zα zβ∗ ∂i zβ . (5.18)
α α β
1 This word is a favorite of professor R. B. Laughlin, the president of KAIST and a Nobel
51
52CHAPTER 6. MONTE CARLO METHODS AND CRITICAL PHENOMENA
m(t) ∼ (−t)β
C(t) ∼ |t|−α
χ(t) ∼ |t|−γ . (6.1)
The three numbers (α, β, γ) (and there are a few others) are the criti-
cal exponents which define the critical properties of a given system. For
instance, the mean-field theory discussed in chapter 2 gives (α, β, γ) =
(0, 1/2, 1). This is an example of the scaling behavior, although we did
not make too big a deal about it when we discussed mean-field theory.
Understanding critical phenomena is reduced to the understanding of (i)
why the scaling behavior arises near Tc , and (ii) how to calculate the crit-
ical exponents.
HA = H0 + VA
HB = H0 + VB . (6.2)
Assuming that you can compute these quantities, the free energy is written
as the logarithm, FA = −T log ZA , FB = −T log ZB , and from the free
energies one can derive all the physical quantities through differentiation.
Hence, the fact that physical observables display identical behavior (when
plotted vs. t) is a statement ultimately about the identity of the partition
functions ZA ∼ ZB . But this is remarkable, given the fact that generally
EA 6= EB !
e−βEC
P
S
P tot,C
X
C
hStot i = −βEC
= Stot,C × pC . (6.4)
Ce C
54CHAPTER 6. MONTE CARLO METHODS AND CRITICAL PHENOMENA
P
In the above, Stot,C = i Si is the total spin for a given configuration C. There
are a total of 2N configurations accessible in the Ising model for L × L lattice.
Each one of 2N allowed configuration contribute to the thermodynamic average,
with a probability weight given by pC = e−βEC /Z, Z = C e−βEC .
P
One way to calculate the average magnetization will be like this. (Of course
no one can do this awful sum exactly, so we will have to put the question to a
computer.)
• Calculate also the energy EC for that configuration and multiply e−βEC
with Stot,C previously calculated.
This process is OK, in principle, except for the astronomical cost in time it
takes to actually do it! Take L = 100. Total number of allowed configurations
are 2100×100 = (210 )1000 ≈ (103 )1000 = 103000 . This is a truly awesome number.
No one, not even the best computers in the world all collaborating on this single
project, can perform this many operations in a reasonable time frame. Clearly,
we need an alternative.
Metropolis and co-workers devised a clever way2 to overcome this difficulty.
Let me first explain the scheme, known as the Metropolis algorithm. It may
be helpful to read the following instruction in comparison with the previous
instructions for evaluating Boltzmann average.
• Take some random, initial configuration, {Si0 }. Pick one spin, at site k,
and change it to some other value, Sk0 → Sk1 . The change in energy for
Ising model as a result is
X X
∆E = −J Sj0 (Sk1 − Sk0 ) = 2J Sj0 Sk0 . (6.5)
j∈k j∈k
• If ∆E is less than zero, you have found a way to lower the energy, therefore
you accept it.
State Calculations for Fast Computing Machines”, Journal of Chemical Physics 21, 1087-1092
(1953).
6.2. MONTE CARLO (MC) FOR ISING SPINS 55
• Move to the next site and play the same game over. Having scanned the
whole lattice this way, we will have arrived at a new configuration given
by {Si1 }.
• Go back to step 1. Repeat the Metropolis run using {Si1 } as an initial
configuration.
For a given Metropolis run, you calculate Stot,MC and square it, then you sum
over all the squared total spins obtained over the Metropolis runs, and that
2
will be the hStot i. We already discussed how to obtain hStot,MC i in Eq. (6.6)
and combining the two numbers you can easily access χ(T ). The same princi-
ple applies for the specific heat calculation, where hEi and hE 2 i are equal to
PM PMmax 2
( Mmax=1 EMC )/Mmax and ( M =1 EMC )/Mmax , respectively.
Actual MC calculation runs on a finite lattice of size L×L. Figure 6.1 is a plot
of the average magnetization as a function of temperature T /J, for progressively
larger system sizes L = 4 through L = 60. The decrease of magnetization with
temperature is more abrupt the larger the system size, and it appears that it
will drop to zero at some sharply defined temperature Tc as L → ∞. But the
thing is that we can never run a simulation for the truly infinite!!
The new rule of the game is finite size scaling. The rough idea is that by
looking at the simulation data for a number of different sizes L, one tries to guess
what will be the correct answer in the truly thermodynamic limit. In fact, one
can do better than just guess. A well-established rule of critical phenomena says
that for a finite-sized system near its critical temperature the physical quantities
must behave as
(α, β, γ) are the critical exponents already introduced in Eq. (6.1) and ν is
another critical exponent defining the correlation length ξ as
The quantities on the r.h.s. of Eq. (6.10) with a subscript s are the scaling
functions which depend on a particular combination L1/ν t only. According to
Eq. (6.10) magnetization obtained in the MC calculation at the same tempera-
ture t but for a different size L will give different results, but with the difference
behaving in the predictable manner. So if we plot, instead of m(L, t), the quan-
tity Lβ/ν m(L, t) vs. L1/ν t rather than t itself, all data should collapse onto a
single curve!!
Figure 6.2 is a plot of the same data as in Fig. 6.1 using the scaled equation.
Indeed, all data points collapse onto more or less a single curve, verifying the
validity of the finite-size scaling hypothesis. Similar plot of the susceptibility for
the 2D Ising model on a square lattice in Fig. 6.3 shows that finite-size scaling
works well for that quantity, too. By fitting the numerical data to a scaling
form one can do a rather nice job of extract the critical exponents3 .
Figure 6.1: Magnetization obtained for 2D square Ising model for varying system
sizes L. (excerpt from Landau’s book)
3 Due to lack of time and also of my own depth in this subject, I’m leaving out a lot of
important details of the critical phenomena in general, and of Monte Carlo in specific. The
in-depth discussion will have to wait another summer school and another expert.
58CHAPTER 6. MONTE CARLO METHODS AND CRITICAL PHENOMENA
Figure 6.2: Same magnetization as in the previous figure plotted in the scaling
form. (excerpt from Landau’s book)
Figure 6.3: Susceptibility for 2D square lattice Ising model, in scaling form
(excerpt from Newman’s book)
Chapter 7
Miscellaneous
with ĥ = (sin θ cos φ, sin θ sin φ, cos θ) denoting the spin orientation. This is
diagonalized by the unitary rotation U + HU with
cos(θ/2) − sin(θ/2)
U= (7.2)
sin(θ/2)eiφ cos(θ/2)eiφ
cos(θ/2)
|1i = , E1 = +1
sin(θ/2)eiφ
− sin(θ/2)
|2i = , E1 = −1. (7.3)
cos(θ/2)eiφ
r
5
Y22 (θ, φ) = 3 sin2 θe2iφ = |2i
96π
r
5
Y21 (θ, φ) = −3 sin θ cos θeiφ = |1i
24π
r
5 3 2 1
Y20 (θ, φ) = cos θ − = |0i
4π 2 2
59
60 CHAPTER 7. MISCELLANEOUS
r
5
Y21 (θ, φ) = 3 sin θ cos θe−iφ = |1i
24π
r
5
Y22 (θ, φ) = 3 sin2 θe−2iφ = |2i. (7.4)
96π
Three t2g orbitals are
i
|xyi = √ (|2i − |2i) ∼ sin2 θ sin φ cos φ ∼ xy,
2
i
|yzi = √ (|1i + |1i) ∼ sin θ cos θ sin φ ∼ yz,
2
1
|zxi = √ (|1i − |1i) ∼ sin θ cos θ cos φ ∼ zx. (7.5)
2
The two eg orbitals are
1
|x2 − y 2 i = √ (|2i + |2i) ∼ sin2 θ(cos2 φ − sin2 φ) ∼ x2 − y 2 ,
2
|3z 2 − r2 i = |0i ∼ 3z 2 − r2 . (7.6)
7.4 p-orbitalogy
The three p-orbitals are classified using the spherical harmonics
r
3
Y11 (θ, φ) = − sin θeiφ = |1i
8π
r
3
Y10 (θ, φ) = cos θ = |0i
4π
r
3
Y11 (θ, φ) = sin θe−iφ = |1i. (7.7)
8π
x, y, z-orbitals are given by
1
|xi = √ (|1i − |1i)
2
i
|yi = √ (|1i + |1i)
2
|zi = |0i. (7.8)
Chapter 8
z1 θ θ
|ni = z = , z1 = e−iφ/2 cos , z2 = eiφ/2 sin . (8.1)
z2 2 2
The angles are those of the classical spin vector n = (sin θ cos φ, sin θ sin φ, cos θ).
The path integral for spin contains the term
i
hṅ|ni = ż† z = φ̇ cos θ. (8.2)
2
For an arbitrary spin length I there is a Berry phase contribution to the action
e−iSB with
Z
SB = I d2 rdt (1 − cos θ) φ̇. (8.3)
R
The term dt(1 − cos θ)φ̇ gives the solid angle subtended by the unit vector
over time t. Equation of motion for spin follows from varying the Berry action
δSB
= I ṅ × n. (8.4)
δn
X
AF
S1D = SB [ni ] (8.5)
i
61
62 CHAPTER 8. PATH INTEGRAL FOR SPINS
X X X δSB
SB [(−1)i Li + Mi ] ≈ SB [(−1)i Li ] + Mi · . (8.6)
δni ni =(−1)i Li
i i i
When the spin direction is reversed, the sign of the Berry phase is also reversed,
SB [(−1)i Li ] = (−1)i SB [Li ]. We get
X X
AF
S1D ≈ (−1)i SB [Li ] + I Mi · L̇i × Li . (8.7)
i i
In the case of the one-dimensional chain one can manipulate the first term in
the Berry phase action a little more. For a chain of length N (N =even) the
first term can be re-arranged as
N/2 N/2
X X δSB IX
(SB [L2i ] − SB [L2i−1 ]) ≈ · (L2i − L2i−1 ) ≈ L̇i × Li · ∂x Li .(8.8)
i=1 i=1
δL 2 i
In the final expression all sites are included in the sum. The continuum limit of
the Berry phase action for 1D AF chain reads
Z Z
AF I
S1D ≈ L · ∂x L × ∂t L + I M · L̇ × L. (8.9)
2
What about the Heisenberg part? By using the same strategy one can re-
write the Heisenberg spin interaction as
X X
(−1)i Li + Mi −(−1)i Li+1 + Mi+1
H = J ni · ni+1 = J
i i
JX X
≈ (Li+1 − Li )2 + J M2i
2 i
Z X Z i
J
→ (∂µ L)2 + J M2 . (8.10)
2 µ
Chapter 9
X
|ψi = φ(n1 , n2 , · · · , nM )|n1 , n2 , · · · nM i,
1≤n1 <n2 <···<nM ≤N
63
64 CHAPTER 9. 1D QUANTUM SPIN CHAIN
M
X M
X
E=J (cos kj − 1), P = kj . (9.4)
j=1 j=1
N
1 X − + −
H⊥ = J (S S + Sn+ Sn+1 )
2 n=1 n n+1
X
Hz = J Snz Sn+1
z
. (9.5)
n
Here |{n0j }i refers to all set of configurations related to the original configuration
|{nj }i by the reversal of the nearest pair of up and down spins. For example,
if a given basis configuration is |3, 6i (spin down position at 3 and 6), the set
of |{n0j }i is |2, 6i, |4, 6i, |3, 5i, |3, 1i. The last configuration arises in the periodic
boundary condition geometry. The total number of configurations {n0j } arising
from {nj } equals the the total number of spin up-down pairs, or the number of
kinks. The action of Hz gives a numerical factor proportional to the number of
kinks,
1
Hz |{nj }i = − J × (number of kinks)|{nj }i. (9.7)
2
The Schrodinger equation becomes (J = 2)
X
Eφ({n}) = (φ({n0 }) − φ({n})) . (9.8)
{n0 }
One-spin down state: The one-spin down state, M = 1, has Eq. (9.8) reduced
to
Eφ(n1 , n2 ) = −4φ(n1 , n2 )
+φ(n1 − 1, n2 ) + φ(n1 + 1, n2 ) + φ(n1 , n2 − 1) + φ(n1 , n2 + 1).
(9.13)
Substituting the ansatz, Eq. (9.15), into the constraint, Eq. (9.14), gives
k1 −k2
A1 eik1 +ik2 + 1 − 2eik1 ei 2 − cos k1 +k
2
2
− = ik1 +ik2 ik
= k2 −k1 . (9.17)
A2 e + 1 − 2e 2
ei 2 − cos k1 +k2
2
The ratio |A1 /A2 | is one, hence one can write A1 = eiθ12 /2 and A2 = eiθ21 /2 ,
θ21 = −θ12 , and show that θ12 in terms of k1 , k2 is obtained from
θ12 sin k1 −k
2
2
cot = . (9.18)
2 cos k1 +k
2
2
− cos k1 −k
2
2
M-spin down state: We return to the general, M -spin down state with the
ansatz solution given by
X M
X
φ(n1 , · · · , ·nM ) = AP exp i kP (j) nj . (9.22)
P j=1
M
X
φ(n1 , · · · , nj − 1, · · · , nM ) + φ(n1 , · · · , nj + 1, · · · , nM )
j=1
− 2φ(n1 , · · · , nM ) = Eφ(n1 , · · · , nM ). (9.23)
The terms which contains two adjacent spins of the same orientation are handled
by the constraints
Inserting the M -particle wave function into this constraint yields the relation
X X X
AP + AP eikP (j) +ikP (j+1) = 2 AP eikP (j+1) ,
P P P
X
AP = AP exp i nl kP (l) + nj kP (j) + nj kP (j+1) . (9.25)
l6=j,j6=j+1
AP + AP (j,j+1) + (AP + AP (j,j+1) )eikP (j) +ikP (j+1) = 2(AP eikP (j+1) + AP (j,j+1) eikP (j) )
θP (1),P (2) + θP (1),P (3) + θP (1),P (4) + θP (2),P (3) + θP (2),P (4) + θP (3),P (4) . (9.28)
9.3. ANTIFERROMAGNETIC CASE 67
The phase factor for AP (1,2) , for instance, is given by 1 and 2 in the above
equation replaced by 2 and 1. An inspection reveals that the net difference
only occurs in replacing θP (1),P (2) by θP (2),P (1) , hence the ratio AP (1),P (2) /AP
is indeed equal to e−iθP (1),P (2) . The same result obtains for interchange of any
other consecutive indices j and j + 1.
Following the earlier practice we impose the periodic boundary condition
φ(n1 , · · · , nM ) = φ(n2 , · · · , nM , n1 + N ) →
X P X
AP e i j kP (j) nj
= AP eikP (1) n2 +···+ikP (M −1) nM +ikP (M ) (n1 +N ) .
P P
(9.29)
One can define a permutation P0 which takes
X
AP eikP P0 (2) n2 +···+ikP P0 (M ) nM +ikP P0 (1) (n1 +N ) =
P
X P X P
AP P −1 ei j kP (j) nj +ikP (1) N
= AP e i j kP (j) nj
. (9.31)
0
P P
i X
AP = exp θP (i),P (j)
2 i<j
i X
AP P −1 = exp θ −1 −1 . (9.33)
0 2 i<j P P0 (i),P P0 (j)
Hubbard Model
H = T +V
X X
tij c+ +
c+
T = jσ ciσ + ciσ cjσ − µ iσ ciσ
hijiσ iσ
X
V = U ni↑ ni↓ (10.1)
i
T represents hopping of electron between a pair of sites hiji with hopping am-
plitude tij = tji . For simplicity we take tij = −t for all the bonds. Unlike
in the previous chapter we include electron spin σ explicitly because, after all,
real electrons
P do carry spin. Diagonalizing T in the absence of the interaction
V leads to kσ k c+ kσ ckσ , where k is the energy of the eigenstate |ki. Thus T
alone describes delocalized electrons that form a band inside the solid. This has
been the essence of the discussion of previous chapter.
Let’s examine the properties of the V -term. First it is completely local, i.e.
it only acts at a given site i: niσ = c+iσ ciσ (no sum on σ) counts the number of
electrons of spin σ at site i. This being an electron model, niσ can only be 0 or
1. When a given site i is occupied by zero electron (ni↑ = ni↓ = 0) or by only
one electron of spin σ (niσ = 1, niσ = 0, ni↑ ni↓ = 0), V has an expectation
value of zero. When the site is doubly occupied, i.e. ni↑ = ni↓ = 1, the site
has an energy increase of U . We may therefore understand U as the Coulomb
repulsion energy of the two electrons. The challenge we must face now is to
understand the properties of, and to solve, the model Hamiltonian (10.1) in the
simultaneous presence of T and V .
69
70 CHAPTER 10. HUBBARD MODEL
state ni E
|0i 0 0
c†i↑ |0i, c†i↓ |0i 1 −µ
c†i↓ c†i↑ |0i 2 U − 2µ
P In the atomic limit tij → 0, the Hamiltonian reduces to the atomic form H =
i Hi , Hi = −µni + U ni↑ ni↓ , that lacks communication between neighboring
sites. Diagonalizing Hi is trivial since each fixed occupation of electron gives an
eigenstate. Eigenstates and their corresponding energies are
Suppose the chemical potential is chosen to satisfy 0 < µ < U2 . Then among
the three possible eigenenergies, only the −µ, corresponding to single occupa-
−1
tion, lies below zero energy. According to the Fermi factor F (E) = eβE + 1
that determines the occupation probability of a given eigenstate, only the ni = 1
state will occur at zero temperature. That means each site of the lattice will
be occupied by one, and only one electron at zero temperature. Since both spin
states give rise to the same energy, each electron occupying the site can be either
spin up or spin down. For N electrons occupying N different lattice sites, this
leads to enormous ground state degeneracy 2 × 2 × · · · × 2 = 2N . In other words,
any state of the spins given by |σ1 σ2 · · · σN i, σi = ±1, is an eigenstate having
the same energy −µN .
Next we raise the question: when we add the hopping part T to the Hamil-
tonian will the eigenstates of V alone remain eigenstates, and will their energies
differ due to the hopping effect?
We will call matrix element hn|H|mi = Vnm , then we have a linear matrix
equation X
Vnm cm = Ecn . (10.4)
m
By diagonalizing the matrix Vnm , the eigenvectors {cmP } are obtained. Then
the eigenstate of the Hamiltonian is expressed as |ψi = m cm |mi, having the
energy E.
We apply precisely this strategy to solve the Hubbard model. And before
doing so, we must first identify the basis states to be used in the diagonalization
10.2. PROJECTION TO TRUNCATED HILBERT SPACE 71
process. For a given site i there are four possible states, listed in Table 10.1,
corresponding to zero, one (of spin up and down), and two electron occupation,
which we will label |0i i, |σi i, and |di i in obvious notation. When there are N
sites, the basis states are given by
where each αi is one of the four possible states. Working out Vnm corresponds
to calculating the overlap of H between two arbitrary basis states (10.5). For
N sites, there are 4N such basis states, and the size of the matrix Vnm is
4N × 4N . Even for N = 10, the dimension of the matrix is 410 × 410 ' 106 × 106 !!
Diagonalizing such a huge matrix is impossible, even with the world’s best
computer available right now. In this chapter we will try to diagonalize H up
to N = 4.
To get a flavor of the steps needed to diagonalize H, we start with the
two-site Hubbard model
X X
c+ +
H = U n1↑ n1↓ + U n2↑ n2↓ − µ (n1σ + n2σ ) − t 2σ c1σ + c1σ c2σ . (10.6)
σ σ
First there are sixteen basis states given as the direct product
We know how to work out the matrix element of T already. Since T conserves
the total spin, nonzero element only occurs between states such as
You should carefully work out the matrix element for each of these.
Next, the interaction term V has a nonzero element if and only if a site is
doubly occupied. Thus
hd1 , d2 |V |d1 , d2 i = U + U = 2U
hd1 , α2 |V |d1 , α2 i = U (α2 6= d2 )
hα1 , d2 |V |α1 , d2 i = U (α1 6= d1 )
hα1 , α2 |V |α1 , α2 i = 0 (α1 6= d1 , α2 6= d2 ) (10.9)
In this way we can work out all the matrix elements among 16 basis states. The
16 × 16 matrix can be diagonalized easily, and will yield 16 energy levels. The
lowest energy and the corresponding eigenvector defines the ground state of the
Hubbard Hamiltonian.
It is often the case that the effective Hamiltonian acting within a restricted
Hilbert space looks quite different from the original Hamiltonian defined in the
larger space. Being able to systematically carry out the transformation to the
restricted space is of practical value and the techniques are presented in this
section using the Hubbard model as the starting point.
We have used Tji above which reduces to Tji = −t in the case of the nearest-
neighbor hopping only. The following derivation is largely taken from Yoshioka,
Girvin, and MacDonald.
Using the fact that the electron occupation plus the hole occupation at any
site for spin σ adds to one, niσ + hiσ = 1, one can split the kinetic term into
three pieces, T = T0 + T1 + T−1 where
Tji hjσ c†jσ ciσ hiσ + Tji njσ c†jσ ciσ niσ .
X X
T0 = (10.11)
ijσ ijσ
T1 increases the number of doubly occupied site by one because it has a non-zero
matrix element if and only if the initial state has the configuration | ↑i ↓j i, or
| ↓i ↑j i. Then T1 acting on the initial state produces |0i dj i or |di 0j i, d implying
the doubly occupied site. Similarly, T−1 decrease the number of doubly occupied
sites by one by acting on the initial state |0i dj i or |di 0j i. T0 does not change
the number of double occupancy because it acts only on the initial state |0i σj i,
|σi 0j i, or |di σj i, |σi dj i, σi =↑, ↓.
Now consider a canonical transformation
0
H = eiS He−iS (10.12)
using iS = (T1 − T−1 )/U . T±1 P satisfies the commutation [V, T1 ] = U T1 ,
[V, T−1 ] = −U T−1 where V = U i ni↑ ni↓ is the Hubbard term. By a straight-
0
forward expansion of H up to second order in S and keeping the leading terms
in O(1/U ) we get
0 1
H = T0 + V + ([T1 , T−1 ] + [T1 , T0 ] + [T0 , T−1 ]) . (10.13)
U
0
When we let the Hamiltonian H act between states with no doubly occupied
sites it is further reduced to
0 1
H = T0 − T−1 T1 (10.14)
U
because the other two terms would change the number of doubly occupied sites.
Write out T−1 T1 as
Tlk Tji hlλ c†lλ ckλ nkλ njσ c†jσ ciσ hiσ .
X
(10.15)
ijklσλ
10.2. PROJECTION TO TRUNCATED HILBERT SPACE 73
Tij Tji hiλ c†iλ cjλ njλ njσ c†jσ ciσ hiσ .
X
T−1 T1 = (10.16)
ijσλ
A short consideration gives that the operators have the following action on
the possible spin states on sites i, j:
X
T−1 T1 = t2ij (| ↑i ↓j i − | ↓i ↑j i) (h↑i ↓j | − h↓i ↑j |) = 2t2ij |sij ihsij |. (10.17)
ij
The symbol sij refers to the singlet state in the ij-bond. For less than half-filled
case it is possible to have an empty site either at i or j, for which T−1 T1 would
give zero. Taking this into account, |sij ihsij | may be re-written
1
|sij ihsij | = ni nj − Si · Sj (10.18)
4
employing the spin operator Si now. In all, we get
0 2 X 2 1
H = T0 + tij Si · Sj − ni nj ,
U ij 4
X hg|T |exihex|T |g 0 i
hg|Heff |g 0 i = , (10.20)
ex
Eg − Eex
By working out the nonzero elements of Eq. (10.20) explicitly as shown in Table
10.2, hg|Heff |g 0 i is found to have the same 2 × 2 matrix elements as those of the
effective Hamiltonian
74 CHAPTER 10. HUBBARD MODEL
| ↑i ↑j i | ↑ i ↓j i | ↓ i ↑j i | ↓i ↓j i
h↑i ↑j | 0 0 0 0
h↑i ↓j | 0 2 −2 0
h↓i ↑j | 0 −2 2 0
h↓i ↓j | 0 0 0 0
4t2 X
1
Heff = Si · Sj − , (10.22)
U 4
hiji
U 2
U ni↑ ni↓ = (n − m2i ). (10.25)
4 i
We will carry out the meanfield decoupling at half-filling, hni i = 1.
U 2 U U U U
ni − m2i → hni ini − hmi imi = ni − si mi .
(10.26)
4 2 2 2 2
The average of the magnetic moment is denoted si . The meanfield Hubbard
Hamiltonian is
X
X U UX
HM F = −t c+ c
jσ iσ + − µ ni − si mi . (10.27)
ijσ
2 i
2 i
10.3. MEAN-FIELD THEORY AND ANTIFERROMAGNETISM 75
X X X X
HM F = −t c+
jσ ciσ − η (−1)i mi = k c+
kσ ckσ − η σc+
k+Qσ ckσ ,
ijσ i kσ kσ
(10.28)
where η = U s/2. The last line follows from Fourier transform. Q implies
the AF wavevector (π, π)(2D) or (π, π, π)(3D). We divide the original Brillouin
zone [−π, π] ⊗ [−π, π] into two: one covers the region bounded by the four lines
kx + ky = π, kx + ky = −π, kx − ky = π, kx − ky = −π, and the other, the
remaining part of the BZ. The first region, shaded in Fig. ??, is the reduced
Brillouin zone (RBZ).
Now one can rewrite the Hamiltonian
X 0 X0
k c+ +
σ c+ +
HM F = kσ ckσ + k+Q ck+Qσ ck+Qσ −η k+Qσ ckσ +ckσ ck+Qσ
kσ kσ
0
X k −ησ ckσ
c+ c+
= kσ k+Qσ . (10.29)
−ησ k+Q ck+Qσ
kσ
r
ση k − k+Q 1
sin 2θk = , cos 2θk = , Ek = η 2 + (k −k+Q )2 . (10.31)
Ek 2Ek 4
In the standard cubic(square)
p lattice k+Q = −k , and thus sin 2θk = ση/Ek ,
cos 2θk = k /Ek , Ek = 2k + η 2 . Using this transformation the Hamiltonian is
brought to a diagonalized form:
0
X
+ +
H= Ek [γ1σ γ1kσ − γ2kσ γ2kσ ], (10.32)
kσ
assume k = −k+Q . We have managed to reduce the Hamiltonian to the diag-
onalized form, and obtain the meanfield energies ±Ek . Furthermore, we have
explicitly derived a relation between the quasiparticle operators (γ) and the orig-
inal, electron operators. The eigenvectors of the matrix act as a “connection”
between the two operators. Since Ek > 0, all the k-states in the RBZ for the
lower energy branch are occupied, and completely empty for the upper branch
at zero temperature, in agreement with the original prescription of half-filling.
The ground state is expressed as
0
Y
+ +
|GSi = γ2k↑ γ2k↓ |0i. (10.33)
k
76 CHAPTER 10. HUBBARD MODEL
+
γ2kσ = uk c+ +
kσ + σvk ck+Qσ ,
s s
1 εk 1 εk
uk = 1− , vk = 1+ . (10.36)
2 Ek 2 Ek
Y
|GSi = uk c+ +
k↑ + vk ck+Q↑ uk c+
k↓ − v +
k k+Q↓ |0i.
c (10.37)
k
The real-space wave function is a product of Slater determinants, D[φ↑ (k, r)]D[φ↓ (k, r)]:
This equation has an exact, T = 0 solution in the case of the constant density
of states D() = 1/(2D),
2D/U
s= , (10.40)
sinh[2D/U ]
The transition temperature TN is given when s = 0, Ek = |k |.
In this section I have given a rather detailed derivation of the gap equation
for the antiferromagnetic ordering. The BCS problem is entirely similar to this
in its mathematical structure. Technically, if you know how to solve the AF
problem, then you are (almost) equally well prepared for the BCS or any other
kind of mean-field problems of condensed matter physics.
The antiferromagnetic ordering at (π, π) introduced a spatial modulation in
an initially translationally invariant Hamiltonian that led to (i) doubling of the
unit cell and halving of the Brillouin zone and (ii) folding of the band into two
subbands and a gap between the bands. Full occupation of the lower band and a
complete absence in the upper band leads to the opening of the gap at half-filling.
10.3. MEAN-FIELD THEORY AND ANTIFERROMAGNETISM 77
U U U U
ni↑ ni↓ → hni ini − hsi i · si = ni − si · mi . (10.41)
2 2 2 2
The average moment mi are taken to lie within the plane, and it is useful to
use the characterization in the complex notation
UX ∗ − Um − Um X +
− (mi si + mi s+
i )=− (sQ + s+ Q) = − (ck+Q↑ ck↓ + c+
k↓ ck+Q↑ ).
2 i 2 2
k
(10.43)
The Brillouin zone of the triangular lattice can be broken up into three pieces
of equal area which are connected to each other by the translation of +Q or
−Q in momentum space. The Hamiltonian within the reduced Brillouin zone
becomes
0
X
k c+ + +
kσ ckσ + k+Q ck+Q,σ ck+Q,σ + k−Q ck−Q,σ ck−Q,σ
kσ
0
X
−η c+ + +
k+Q↑ ck↓ + ck−Q↑ ck+Q↓ + ck↑ ck−Q↓ + h.c.
k
T
c+
k↑ k 0 0 0 0 −η ck↑
c+
k↓
0 k −η 0 0 0 ck↓
0 +
ck+Q↑ 0 −η k+Q 0 0 0 ck+Q↑
X
=
+
ck+Q↓ 0 0 0 k+Q −η 0 ck+Q↓
k +
0 0 −η 0 k−Q 0 ck−Q↑
ck−Q↑
c+ −η 0 0 0 0 k−Q ck−Q↓
k−Q↓
+
T
ck↑
k −η 0 0 0 0 ck↑
c+
−η k−Q 0 0 0 0 ck−Q↓
k−Q↓
0 +
ck+Q↑ 0 0 k+Q −η 0 0 ck+Q↑
X
= +
c 0 0 −η k 0 0 ck↓
k +k↓
0 0 0 0 k−Q −η ck−Q↑
ck−Q↑
c+ 0 0 0 0 −η k+Q ck+Q↓
k+Q↓
(10.44)
78 CHAPTER 10. HUBBARD MODEL
c+
k↑ cos θ1 − sin θ1 0 0 0 0 a1↑
c+
sin θ1 cos θ1 0 0 0 0 a1↓
k−Q↓
c+
0 0 cos θ2 − sin θ2 0 0 a2↑
k+Q↑
=
c+
0 0 sin θ2 cos θ2 0 0 a2↓
+k↓
0 0 0 0 cos θ3 − sin θ3 a3↑
ck−Q↑
c+
k+Q↓
0 0 0 0 sin θ3 cos θ3 a3↓
(10.45)
with
2η 2η 2η
tan 2θ1 = , tan 2θ2 = , tan 2θ3 = .
k−Q − k k − k+Q k+Q − k−Q
(10.46)
The six eigenenergies are given by
1
q
±
E1k = [(k + k−Q ) ± (k − k−Q )2 + 4η 2 ]
2
1
q
±
E2k = [(k+Q + k ) ± (k+Q − k )2 + 4η 2 ]
2
1
q
±
E3k = [(k−Q + k+Q ) ± (k−Q − k+Q )2 + 4η 2 ].
2
√
√ bare dispersion is given by k = −2t[cos kx +cos(kx /2+ 3ky /2)+cos(kx /2−
The
3ky /2)] − µ.
+
Once we fix the chemical potential at zero, µ = 0, one can see that Eαk >0
−
but Eαk < 0 for all α = 1, 2, 3. (See Fig. ??.) That means as long as we
take µ = 0 the three lowest bands are completely filled, and the three highest
completely empty, with the net electron density of one per site, or half-filling.
+ −
The gap magnitude is given by min(Eαk − Eαk ) = 2η.
In the triangular lattice, the antiferromagnetic ordering at Q = (4π/3, 0)
introduces a spatial modulation that led to (i) tripling of the unit cell and 1/3-
ing of the Brillouin zone and (ii) folding of the band into six subbands and a gap
separating the lower three bands from the upper three. Full occupation of the
lower bands and a complete absence in the upper bands leads to the opening of
the gap at half-filling. The reduced Brillouin zone has the shape of a hexagon
with the edge-to-edge separation exactly equalling 4π/3.
The most general fermion hopping between a pair of sites hiji obeying the
time-
ci↑
reversal symmetry can be written down using the spinor notation ψi = ,
ci↓
ˆ
Hij = tψi† [cos θ + i(dˆ · σ) sin θ]ψj = tψi† [eiθd·σ ]ψj . (10.48)
Here dˆ is an arbitrary
3-dimensional
real unit vector. The time-reversal opera-
0 1
tion renders ψi → ψi = iσy ψi and one can easily see that the above
−1 0
expression for Hij is invariant under such a change of ψi plus the conjugation
of idˆ · σ to −idˆ · σ ∗ . Note that σ ∗ = (σx , −σy , σz ). Most generally t, θ and dˆ
will depend on the bond, so we can write
ˆ
Hij = tij ψi† [eiθij dij ·σ ]ψj . (10.49)
To ensure that the Hamiltonian remains Hermitian we will require dˆji = −dˆij .
Let’s consider a specialized case where dˆij is uniform, so that we can take
dˆij = +dˆ for j = i + x̂ and j = i + ŷ, and θij = θ, tij = t everywhere. (I’m
assuming a square lattice.) Then the Hamiltonian can be written in the form
X † ˆ † ˆ † ˆ † ˆ
H= t ψi+x eiθd·σ + ψi−x e−iθd·σ + ψi+y eiθd·σ + ψi−y e−iθd·σ ψi . (10.50)
i
t2ij
Jij Si · Sj , Jij =
. (10.53)
U
Now we must rotate the spins back to the original basis.
(1992).
80 CHAPTER 10. HUBBARD MODEL
to show that
Si · Sj → (cos 2θij )Si · Sj − (sin 2θij )dˆij · (Si × Sj ) + (1 − cos 2θij )(Si · dˆij )(Sj · dˆij ).
(10.57)
The terms generated are superexchange, DM, and Kitaev interactions, respec-
tively.
X tij X
H = − (d+ + + + +
j pj−σ + pjσ ej )(pi−σ di + ei piσ )fjσ fiσ +
+
(ρiσ − µ)fiσ fiσ
mi mj iσ
hijiσ
X X X X
+ U d+
i di − λi (e+
i ei + p+ +
iσ piσ + di di − 1) − ρiσ (p+ +
iσ piσ + di di )
i i σ iσ
(10.61)
+
p
where we have introduced the abbreviation mi = niσ (1 − niσ ), niσ = hfiσ fiσ i.
In the KR theory this factor is expressed in terms of bosons, e.g. niσ →
1 − e+ +
i ei − pi−σ pi−σ , and becomes equal to the present definition only if the
constraints are satisfied exactly. Our scheme has the advantage that it will
simplify the minimization equation
PN considerably. Fermionic part of the Hamil-
tonian is diagonalized as fiσ = m=1 umiσ ψmσ for each species of spin σ. The
number of lattice sites is N .
Minimizing the free energy with respect to d+ +
i , piσ lead to the condition
2
1 X tij
di = P − pi−σ (p+ + +
j−σ dj + ej pjσ )hfiσ fjσ i , (10.62)
σ ρiσ + λi − U jσ
m i mj
1 X tij
piσ = − [ei (p+ + + + + +
j−σ dj + ej pjσ )hfiσ fjσ i + di (dj pjσ + pj−σ ej )hfj−σ fi−σ i] .
λi + ρiσ j
mi mj
(10.63)
Minimization with respect to e+i leads to
X tij
λi ei = − piσ (d+ + +
j pj−σ + pjσ ej )hfjσ fiσ i. (10.64)
jσ
m i mj
and setting them equal to zero. Expressions involving fermions are replaced by their finite-
temperature averages. Although in the actual calculation complex fields such as d+ i will be
treated as a real number, and thus d+ i = di , it is convenient during the derivation to keep the
complex nature of the boson fields.
3 In invoking this relation we are already carrying out the minimization with respect to λ .
i
82 CHAPTER 10. HUBBARD MODEL
+
ρiσ imposes the constraint hfiσ fiσ i = p+ +
iσ piσ + di di . They can be inde-
pendently obtained by diagonalization the fermion Hamiltonian and from Eqs.
+
(10.62)-(10.63). If hfiσ fiσ i turns out to be greater than p+ +
iσ piσ + di di one raises
(by hand) the value of ρiσ in the next diagonalization step. This will lower the
+
expectation value hfiσ fiσ i in the next iteration. Iteration will continue until the
identity is achieved.
Chemical potential is adjusted (also by hand) to satisfy the global constraint
1 X +
hf fiσ i = ν (10.65)
N iσ iσ
X
+ + + +
H = k (zBσ zAσ fBkσ fAkσ + zAσ zBσ fAkσ fBkσ )
kσ
X X
+ +
+ (ρAσ − µ)fAkσ fAkσ + (ρBσ − µ)fBkσ fBkσ
kσ kσ
N X
+ × {U d+ A dA − λ A (e +
A e A + p+ +
Aσ pAσ + dA dA − 1)
2 σ
X
+ +
− ρAσ (pAσ pAσ + dA dA ) + (A → B)}. (10.68)
σ
1
zAσ = p (p+ +
A−σ dA + eA pAσ )
nAσ (1 − nAσ )
1
zBσ = p (p+ +
B−σ dB + eB pBσ ). (10.69)
nBσ (1 − nBσ )
10.6. FOUR-BOSON THEORY 83
r
1 1
Ekσ± = (ρAσ + ρBσ ) − µ ± (ρAσ − ρBσ )2 + 2k |zAσ |2 |zBσ |2 . (10.70)
2 4
Free energy of the system becomes
X N
F = Ekσ± − T ln(1 + eβEkσ± ) + ×
2
kσ±
( )
X X
U d+
A dA − λA (e+
A eA + p+
Aσ pAσ + d+
A dA − 1) − ρAσ (p+
Aσ pAσ + d+
A dA ) + (A → B) .
σ σ
(10.71)
1 2 X ∂Ekσ±
dA = × × F (Ekσ± )
ρA↑ + ρA↓ + λA − U N ∂d+
A
kσ±
1 2 X ∂Ekσ±
dB = × × F (Ekσ± ) . (10.72)
ρB↑ + ρB↓ + λB − U N ∂d+
B
kσ±
∂F/∂p+
A,Bσ = 0 gives
1 2 X ∂Ekσ0 ±
pAσ = × × F (Ekσ0 ± )
ρAσ + λA N 0 kσ ±
∂p+
Aσ
1 2 X ∂Ekσ0 ±
pBσ = × × F (Ekσ0 ± ) . (10.73)
ρBσ + λB N 0 kσ ±
∂p+
Bσ
∂F/∂e+
A,B = 0 gives
2 X ∂Ekσ±
λ A eA = × F (Ekσ± )
N
kσ±
∂e+
A
2 X ∂Ekσ±
λB dB = × F (Ekσ± ) . (10.74)
N
kσ±
∂e+
B
∂F/∂ρA,Bσ = 0 gives
2 X ∂Ekσ±
p+ +
Aσ pAσ + dA dA = F (Ekσ± )
N ∂ρAσ
k±
2 X ∂Ekσ±
p+ +
Bσ pBσ + dB dB = F (Ekσ± ) . (10.75)
N ∂ρBσ
k±
4 Since for each k and σ there is mixing between A and B, one ends up diagonalizing a 2 × 2
+ +
1 1 X 2 |zBσ |2 pA−σ (pA−σ dA + eA pAσ )
dA = × × ±F (Ekσ± ) √k
ρA↑ + ρA↓ + λA − U N D nAσ (1 − nAσ )
kσ±
+ +
1 1 X 2 |zAσ |2 pB−σ (pB−σ dB + eB pBσ )
dB = × × ±F (Ekσ± ) √k .
ρB↑ + ρB↓ + λB − U N D nBσ (1 − nBσ )
kσ±
(10.76)
1 1 X 2
pAσ = × × ±F (Ekσ± ) √k ×
ρAσ + λA N D
k±
" #
2
|zB−σ | dA (pAσ d+
+ eA p+
A A−σ ) |zBσ |2 eA (pAσ e+ +
A + dA pA−σ )
+
nA−σ (1 − nA−σ ) nAσ (1 − nAσ )
1 1 X 2
pBσ = × × ±F (Ekσ± ) √k ×
ρBσ + λB N D
k±
" #
2
|zA−σ | dB (pBσ d+
+ eB p +
B B−σ ) |zAσ |2 eB (pBσ e+ +
B + dB pB−σ )
+ .
nB−σ (1 − nB−σ ) nBσ (1 − nBσ )
(10.77)
1 X ρAσ − ρBσ
p+ +
Aσ pAσ + dA dA = F (Ekσ± ) 1 ± √
N 2 D
k±
1 X ρAσ − ρBσ
p+ p
Bσ Bσ + d+
d
B B = F (Ekσ± ) 1 ∓ √ (10.79)
N 2 D
k±
Strategy for updating the variables are as follows. From Eqs. (10.76) and
+
(10.77) one obtains
P + updated values for dA , dB , pAσ , and pBσ . Using eA eA =
+
1 − dA dA − σ pAσ pAσ and the corresponding relation for B sublattice one
obtains eA , eB . From Eq. (10.78), using known values of eA , eB from above
10.6. FOUR-BOSON THEORY 85
one obtains λA , λB . Finally, Eq. (10.79) determines ρAσ − ρBσ . The even
combination ρAσ + ρBσ leads to the uniform shift of the chemical potential (see
Eq. (10.70)), or it can be absorbed into µ. Taking the difference of the two
equations in Eq. (10.79) gives
p+ + + + + +
Aσ pAσ + dA dA − pBσ pBσ + dB dB = hfAσ fAσ i − hfBσ fBσ i
1 X F (Ekσ± )
= (ρAσ − ρBσ ) × ± √ . (10.81)
N D
k±
The occupation of electrons with spin σ will be different on the two sublattices
provided ρAσ 6= ρBσ , indicating antiferromagnetic component of spin. In prac-
tice I think one should set ρAσ = −ρBσ = ρσ from the outset. This simplifies
the equation to
2 X F (Ekσ± )
p+ + + +
Aσ pAσ + dA dA − pBσ pBσ + dB dB = ρσ × ± √ (10.82)
N D
k±
√
for D = ρ2σ + k |zAσ |2 |zBσ |2 and Ekσ± = −µ ± D.
The ground state can be (i) uniform and paramagnetic, (ii) uniform and
spin-polarized (for nonzero ρσ ) or (iii) possess staggered magnetic moment.
86 CHAPTER 10. HUBBARD MODEL
Chapter 11
Spin Interactions
Now, let’s see what happens when we act with the r.h.s. of Eq. (11.1)
1 1 11 1
2 Si · Sj + | ↑↑i = 2 Siz Sjz + | ↑↑i = 2 + | ↑↑i = | ↑↑i.
4 2 22 4
OK, so both operators produce the same state, | ↑↑i, when acting on | ↑↑i. For
| ↑↓i, one has Pij | ↑↓i = | ↑↓i, because the up- and down-spins simply swap
their positions. Before acting with the r.h.s. of Eq. (11.1) again, first remember
that
1
Si · Sj = Siz Sjz + (Si+ Sj− + Si− Sj+ ). (11.2)
2
87
88 CHAPTER 11. SPIN INTERACTIONS
After collecting the result of action with each of the three operators of Eq. (11.2)
on | ↑↓i we obtain
1 1 1
2 Siz Sjz + Si+ Sj− + Si− Sj+ + | ↑↓i
2 2 4
1 1 1
= 2 − | ↑↓i + 0 + | ↑↓i + | ↑↓i
4 2 4
= | ↑↓i.
Again, we obtain agreement. The case for | ↑↓i and | ↓↓i states proceeds
similarly and need not be reproduced here. Hence we learn that the cherished
Heisenberg exchange model may equally well be written
JX
H= Pij .
2
hiji
In reality, there is no reason to just stop with exchanging two particles at a time.
Why not exchange the spins of three, or even four sites at once. For example,
the 3-site exchange process is depicted in Fig. ??.
The exchange process of this type is denoted Pijk , in obvious generaliza-
tion of the two particle exchange. In turn, the three particle exchange can be
decomposed as a series of two-particle exchanges:
−1
Pijk = Pij Pik = Pjk Pji = Pki Pkj .
In any sensible physical system the two processes - clockwise and anti-
clockwise - occur with same amplitude, and thus the most general 3-site ex-
change process is described by a term in the Hamiltonian
X
−1 1 X
(Pijk + Pijk )= (Pik Pij + Pji Pjk + Pkj Pki + Pij Pik + Pjk Pji + Pki Pkj ).
3
hijki hijki
plus 2-site spin interactions which are readily absorbed into the Heisenberg term.
So, allowing a 3-particle exchange generates a new spin interaction term in the
11.1. HIGHER-ORDER EXCHANGE PROCESSES 89
Hamiltonian of the form shown in Eq. (11.3). The general spin Hamiltonian
involving 2- and 3-particle exchanges can be written down
X
H = J1 Si · Sj
ij
X
+J2 (Si · Sj )(Si · Sk ) + (Sj · Sk )(Sj · Si ) + (Sk · Si )(Sk · Sj ).(11.4)
hijki
There are several kinds of representations for spin operators. Some of them are
exclusively used for S = 1/2, and others are applicable for general S. They also
lead to different ways of obtaining the spin excitations of a given spin interaction
model.
and the equation of motion for the individual spin Si that follows from it. We
first remind ourselves of the fundamental commutation relations of spin
dSix X y X x
= −i[Six , H] = −J Sj Siz − Sjz Siy = −J
Sj × Si . (12.3)
dt j∈i j∈i
Combined with the equation of motion for the y- and z-components, we get
dSi X
= −J Sj × Si . (12.4)
dt j∈i
91
92 CHAPTER 12. SPIN REPRESENTATION AND SPIN EXCITATION
solve the above equation exactly. However we must remind ourselves that it
is the small fluctuation away from the ground state that concerns us, and as
such we can decompose the spin operator as Si = hSi i + δSi = mi + δSi ,
where mi represents the ground state spin average. For a ferromagnetic ground
state, mi = S ẑ. A small fluctuation means |hSi i| |δSi |. In making such
a comparison of magnitudes we must be careful to remember that mi is a
number, but δSi is not. So, it’s not altogether clear what exactly we mean
by the magnitude of δSi . Nevertheless, we can at least make the substitution
Si = mi + δSi as a matter of formal definition, and proceed to plug it into Eq.
(12.11).
d X
δSi = −J mj + δSj × mi + δSi ≈
dt j∈i
X X X
−J mj × mi − J mj × δSi − J δSj × mi (12.5)
j∈i j∈i j∈i
2 P
ignoring terms of order δS . The sum j∈i comprises all the nearest-neighbor
sites j for i. Once again, the justification is that δS is a small quantity, and
(δS)2 , even smaller. We take the quantization axis ẑ and write mi = mẑ
everywhere. Then1
X
d
δSi = −Jzmẑ × δSi − Jm δSj × ẑ. (12.6)
dt j
d
The ẑ-component of this equationreads dt δSzi = 0, hence the fluctuation in the
z
ẑ-component of spin is zero, Si t = m. For the x̂-and ŷ-components we get
d x X
S = JzmSiy − Jm Sjy ,
dt i j∈i
d y X
S = −JzmSix + Jm Sjx . (12.7)
dt i j∈i
1 I apologize for the multiple use of z here: z is the number of neighbors, or the coordination
For a 3-dimensional cubic lattice,P j∈i eik·(rj −ri ) = 2(cos kx + cos ky + cos kz ),
P
and with z = 6, ωk = 2Jm(3 − α cos kα ). For 2D square lattice we will get
ωk = 2Jm(2 − cos kx − cos ky ). This is the desired spin wave energy for a spin
disturbance associated with a wavevector k.
12.1.2 Antiferromagnet
For an antiferromagnet, the equation of motion reads
dSi X
=J Sj × Si . (12.11)
dt j∈i
X X
d
δSi = J mj × δSi + J δSj × mi . (12.12)
dt j∈i j∈i
We take the quantization axis ẑ and write mi = (−1)i mẑ. Then, the whole
square lattice can be divided into two sublattices: i ∈ A if (−1)i = 1, and i ∈ B
if (−1)i = −1. A lattice for which such a division into two sublattices can be
carried out is known as a bipartite lattice.
We write δSi as Si and note that mj for the neighbors j surrounding the
site i is always opposite to mi : mj = −mi .
X
d
Si = zJSi × mi + J Sj × mi . (12.13)
dt j∈i
X
d 0
S = zJS1i × msi −J S1j × msi
dt i j∈i
X
d 1
S = zJS0i × msi +J S0j × msi (12.15)
dt i j∈i
The staggered magnetization msi = (−1)i mi = mẑ is uniform. Since all the
quantities appearing in the above equation is slowly varying, one can write
S0i = S0 eik·ri −iωt , S1i = S1 eik·ri −iωt , and obtain
−iωS0 = mzJ S1 × ẑ − εk S1 × ẑ ,
−iωS1 = mzJ S0 × ẑ + εk S0 × ẑ . (12.16)
94 CHAPTER 12. SPIN REPRESENTATION AND SPIN EXCITATION
where we write zεk = j∈i eik·(rj −ri ) . Using the complex notation S 0 = Sx0 +
P
ωS 0 = mzJ(1 − εk )S 1 ,
1
= mzJ(1 + εk )S 0 .
ωS (12.17)
p
It readily follows that ωk = mzJ 1 − ε2k .
X X
HHDM = −J Si · (Si+x̂ +Si+ŷ ) − K Si ×Si+x̂ · x̂+Si ×Si+ŷ · ŷ
i i
X
=− Si · JSj + KSj ×êji . (12.18)
ij
where we have introduced the unit vector êji extending from site i to site j.
Using the previously derived equation of motion for a ferromagnet, one can
write down the equation of motion for the spiral ferromagnet easily
dSi X
= Si × JSj + KSj × êji . (12.19)
dt j∈i
defines the classical ground state. The spiral spin state that concerns us is given
by mi = meik·ri + c.c. where
1
m = √ (ê1 − iê2 ), ê1 × ê2 = k. (12.21)
2
√
Here k̂ = (1, 1)/ 2 is the propagation vector of the spiral spins. One can show
that
X
Jmj + Kmj × êji = αmi (12.22)
j∈i
√
α = 4J cos Q + 2 2K sin Q. (12.23)
From
√ an independent calculation we know that in two dimensions tan Q =
K/ 2J, so that
12.2. HOLSTEIN-PRIMAKOFF THEORY 95
√ √ p
α = 4J cos Q + 2 2K sin Q = 2 2 K 2 + 2J 2 . (12.24)
Then the linearized equation of motion for the spiral ferromagnet becomes
d X
Si = mi × JSj + KSj × êji + αSi × mi . (12.25)
dt j∈i
d k
S = −αSin − J cos Q Si−x̂n n
+ Si+x̂ n
+ Si−ŷ n
+ Si+ŷ
dt i
Kh
k k k k
− √ cos[Q · ri ] Si+x̂ − Si−x̂ + Si−ŷ − Si+ŷ
2
i
n n n n
+ sin Q Si−x̂ + Si+x̂ + Si−ŷ + Si+ŷ
d n
S = αSik + J Si−x̂
k k
+ Si+x̂ k
+ Si−ŷ k
+ Si+ŷ
dt i
K n n
− √ cos[Q · (ri + x̂)]Si+x̂ − cos[Q · (ri − x̂)]Si−x̂
2
n n
+ cos[Q · (ri − ŷ)]Si−ŷ − cos[Q · (ri + ŷ)]Si+ŷ . (12.27)
12.2.1 Ferromagnet
First, I will introduce the representation (ni = b+
i bi ),
then verify that the new expressions on the r.h.s. of the above equations satisfy
all the commutation algebras of ordinary spin operators: [Siα , Sjβ ] = iδij αβγ Siγ ,
provided the new operator bi is a canonical boson operator, with the familiar
bosonic commutation relations, [bi , b+
j ] = δij , etc.
How this comes about does not concern us so much here; rather we will
proceed to apply the brand new technique and see if something good comes out.
First rewrite
1
Si · Sj = Siz Sjz + (Si+ Sj− + Si− Sj+ )
2
1 1 1
= (S − ni )(S − nj ) + (2S −ni ) 2 (2S −nj ) 2 (b+ +
i bj + bj bi ). (12.29)
2
OK, so it doesn’t appear particular illuminating after the substitution. But
let’s try to appeal to our physical senses to see if some simplifications can occur.
We already know the ground state of the Hamiltonian is given by a ferromagnetic
arrangement of spins, with all the spins pointing in the ẑ-direction. If we take
the average of the operator Siz in the ground state, at zero temperature, we
should get hSiz i = S. The same can be said of the r.h.s. of the last of Eq.
(12.28) because it is just a mathematically equivalent way of writing down the
same quantum-mechanical operator, so we must get hS − ni i = S, or hni i = 0!
This means that the quantum-mechanical ground state |0i of the ferromagnetic
Heisenberg spin model is defined by the condition that h0|ni |0i = 0 for all sites i.
An excitation is just a small wiggle of the spins away from the absolutely frozen
spin configuration that is the ground state, hence we can suppose that hSiz i
deviates only slightly from its ground state value S, or in the language of bosons,
that ni is some small, but non-zero number. In this limiting√ circumstance,
√ we
can simplify the square roots in Eq. (12.28) by taking 2S − ni ≈ 2S, and
writing
1
Si · Sj ≈ (S − ni )(S − nj ) + · 2S(b+ +
i bj + bj bi ).
2
≈ S 2 − S(ni + nj ) + S(b+ +
i bj + bj bi ). (12.30)
Well, now the new Hamiltonian is just a quadratic Hamiltonian involving only
a product of two boson operators. Such operators can be diagonalized easily, in
this case by going to the Fourier space:
X
bi = eik·ri bk . (12.32)
k
Substitution of the Fourier expression for bi into Eq. (12.31) immediately gives
X
H= ω k b+
k bk (12.33)
k
where
ωk = JSz(1 − γk )
X
γk = z −1 eik·(rj −ri ) . (12.34)
j∈i
The eigenenergy ωk agrees with the spin-wave spectrum derived earlier, in Eq.
(12.10), using the equation-of-motion theory.
12.2.2 Antiferromagnet
The strategy for dealing with elementary spin excitations of an antiferromagnet
is essentially the same, but the technique is more advanced.
First of all, we define a new quantum-mechanical operator Siy , Siz related to
y
Si , Siz by the relation
where θi is the angle of the classical ground state spins, θi = π(ix + iy ). This
represents a rotation of spins, with respect to the x-axis, by the angle that just
equals the orientation of the classical ground state spins. The x-component
remains unchanged, hence Six = Six . Furthermore, one can easily verify that
the new operators Siα satisfy all the commutation rules of the original spins, Siz ,
hence they are as good a representation of the local spin operator as the previous
one we used. In terms of Si , we have the antiferromagnetic Hamiltonian written
as
X JX + + X
Si Sj + Si− Sj− − J Siz Sjz .
H=J Si · Sj → (12.36)
ij
2 ij ij
Note that we have two raising (lowering) operators instead of one raising and one
lowering, as was the case before. The ground state of the new Hamiltonian, Eq.
(12.36), is given by the rotation of the antiferromagnetic spins, hSiz i = Seiθi .
But this implies that in the new representation, hSiz i = S everywhere! Since
hSiy i = 0 initially, we have hSiy i = 0, too. So the new Hamiltonian has a
ferromagnetic ground state with all the spins pointing in the +ẑ direction, and
we can make the assumption hSiz i ≈ S, or hni i S after the HP substitution,
for the small-fluctuation analysis.
Under these assumptions the HP-transformed Hamiltonian becomes
X
H ≈ JS (b+ +
i bj + bi bj + ni + nj ). (12.37)
ij
98 CHAPTER 12. SPIN REPRESENTATION AND SPIN EXCITATION
Unlike its ferromagnetic counterpart, Eq. (12.31), this Hamiltonian has a prod-
uct of two annihilation (creation) operators. As before, we take the Fourier
transform and obtain
Xh γk + + i
H = JSz b+ b k + (b b + bk b −k ) . (12.38)
k
2 k −k
k
Whereas doing the Fourier transform was sufficient to yield excitation en-
ergies in the previous example of ferromagnetic spins, now the Hamiltonian
written in momentum space fails to give eigenenergies. The final trick lies in
working out the equation of motion of the boson operators, which gives
[H, bk ] = −JSz bk + γk b+
−k . (12.39)
Here it’s clear why we could not get the energy from Fourier transform alone.
The dynamics of the operator bk is coupled to that of b+
−k , and the dynamics of
+
b−k is coupled to that of bk by the equation
[H, b+ +
−k ] = JSz b−k + γk bk . (12.40)
Noting that [H, X] = −idX/dt in quantum mechanics, and that for the eigen-
modes we can re-write the time derivative as −iE, where E is the eigenenergy,
the coupled oscillator equation we must solve becomes
bk 1 γk bk
Ek = JSz . (12.41)
b+
−k −γk −1 b+
−k
p
Eigenvalues are readily found as Ek = ±JSz 1 − γk2 . Here I state, without
proof, that the negative energies are just artifacts of the theory, and that only
the positive branch matters. Hence, the spin p waves in the antiferromagnetic
background carries the energy given by JSz 1 − γk2 for a wave vector k.
X X
HHDM = −J Sr · (Sr+x̂ +Sr+ŷ ) − K Sr ×Sr+x̂ · x̂+Sr ×Sr+ŷ · ŷ . (12.42)
r r
S+ = S x + iS y = b+
1 b2
S− = S x − iS y = b+2 b1
1 +
Sz = (b b1 − b+2 b2 ). (12.44)
2 1
More concisely one can write S = 12 b+
α σαβ bβ . The spin operators thus defined
obey the canonical spin operator relation
[S α , S β ] = iαβγ S γ . (12.45)
The original spin operator has the additional property S · S = S(S + 1) which
translates into (b+ + + +
1 b1 + b2 b2 )(b1 b1 + b2 b2 + 2)/4 upon the substitution of Eq.
(12.44). To match the original relation, the sum of the two boson occupation
numbers must equal 2S:
b+ +
1 b1 + b2 b2 = 2S. (12.46)
This puts constraint on the the allowed boson Hilbert space of a and b operators.
The spin state with total spin S and Sz = m is represented
ωij is the solid angle subtended by ẑ, Ωi and Ωj . Substituting Eq. (12.48) in
the action b+ i ∂τ bi also gives rise to the Berry phase action first discussed by
Haldane.
Now let’s see how the Heisenberg Hamiltonian looks like in the Schwinger
boson representation. Unlike the Holstein-Primakoff theory, we do not need to
know a priori what the classical ground state looks like. For this reason the
Schwinger boson theory is best applied to cases where the ground state does
not possess broken symmetry.
It is useful to decompose the spin-spin interaction term Si · Sj as
2 Patrick Lee, PRL 63, 680 (1989)
100 CHAPTER 12. SPIN REPRESENTATION AND SPIN EXCITATION
1
S i · S j = S 2 − A+ Aij (12.50)
2 ij
where Aij = b1i b2j − b2i b1j if it is an antiferromagnet, and as
1 + 1 +
Si · Sj = Bij Bij − S(S + 1) = : Bij Bij : −S 2 (12.51)
2 2
where Bij = b+ +
i1 bj1 + bi2 bj2 if it is a ferromagnet. With this substitution the
Heisenberg Hamiltonian, which originally expressed how the nearby spins inter-
act, becomes one for two bosons interacting with each other through the A+ ij Aij
term. Not surprisingly, such interaction Hamiltonian is not exactly solvable.
Instead one relies on an approximation scheme which renders the Hamiltonian
more analytically tractable. We call it the Schwinger boson mean-field theory
(SBMFT) which we describe below.
12.3.1 Ferromagnet
12.3.2 Antiferromagnet
Assume that one of the operators in A+ ij Aij can be replaced by its average,
Qij = hb1i b2j − b2i b1j i. Then we obtain the mean-field Hamiltonian
1X ∗
Qij Aij + Qij A+
HM F = − ij . (12.52)
2 ij
The new Hamiltonian contains only a pair of boson operators, and like all quan-
tum theories involving only a product of two operators, it is solvable.
Although the theory of diagonalization of the mean-field Hamiltonian can be
readily derived, we first introduce a method that will prove useful for treating
a variety of lattice situations. In spirit, this is similar to the rotation of the
classical spin angles to make all the spins aligned ferromagnetically along a
specific direction.
Let’s say the putative classical spin average is given by hSi i = SΩi and
Ωi = (sin θi cos φi , sin θi sin φi , cos θi ). This will in turn impose the condition on
the Schwinger boson average
hb+
i1 bi2 i = Se
iφi
sin θi
−iφi
hb+
i2 bi1 i = Se sin θi
1 +
hb bi1 − b+
i2 bi2 i = S cos θi . (12.53)
2 i1
Introduce the unitary matrix Ui
− sin(θi /2)e−iφi
cos(θi /2)
Ui = (12.54)
sin(θi /2)eiφi cos(θi /2)
which has the relation
Ui+ (Ωi · σ)Ui = σ z . (12.55)
It is now claimed that the rotation of the Schwinger boson spinor ψi = (bi1 bi2 )
by ψi → Ui ψi yields for the new Schwinger boson operators the average
1 +
hb+ +
i1 bi2 i = hbi2 bi1 i = 0, hb bi1 − b+
i2 bi2 i = S. (12.56)
2 i1
12.3. SCHWINGER BOSONS 101
It also implies that in the rotated basis the classical spin average becomes fer-
romagnetic, hSi i = S ẑ.
To make the case more concrete, consider a square-lattice antiferromagnetic
for which the classical spin orientation is hSi i = (−1)i S ẑ. The unitary matrix
(12.54) becomes a unit matrix for A sublattice sites and σx for all the B sub-
lattice sites. Therefore, for B lattice sites, (b1j , b2j ) becomes (−b2j , b1j ). The
pair amplitude Qij , i ∈ A, j ∈ B becomes b1i b1j + b2i b2j . Of course it doesn’t
really matter
P2 what we call the A sublattice, so for every pair of sites ij we have
Aij = m=1 bim bjm .
To carry out the mean-field analysis of the transformed Hamiltonian one
writes, similar to Eq. (12.57),
1X
Qij Aij + Qij A+
HM F = − ij . (12.57)
2 ij
with Aij = bi1 bj1 + bi2 bj2 , Qij = hAij i. We treat the mean-field amplitude as
real and uniform, Qij = Q, and obtain
2
QXX
HM F = − [bim bjm ] + h.c.
2 ij m=1
Q X ik·(rj −ri ) X
= − e bkm b−km + h.c.
4 j∈i
km
X
+ 0 −Qk b1k
= ( b1k b2k ) (12.58)
−Qk 0 b+
2k
k
where Qk = (Q/4) j∈i eik·(rj −ri ) . Inclusion of the Lagrange multiplier i λi (a†i ai +
P P
b+
i bi − 2S) renders the mean-field Hamiltonian
X λ −Qk b1k
HM F = ( b+ b ) . (12.59)
1k 2k −Qk λ b+
2k
k
One can bring the Hamiltonian to its diagonalized form by the rotation
b1k cosh θk sinh θk γ1k
= (12.60)
b+
2k
sinh θk cosh θk +
γ2k
The minus sign in the energy dispersion (12.62) raises concern for what
happens when λ < |Qk |. In fact, this never happens since the moment when
λ equals min[Qk ] is when the bosons begin to condense, and due to the Bose
condensation λ is always fixed to maintain λ = min[Qk ].
102 CHAPTER 12. SPIN REPRESENTATION AND SPIN EXCITATION
In fact, this is identical to the spin wave spectrum obtained using the Holstein-
Primakoff theory, provided we identify Q with JSz, J =exchange energy, S=spin,
z=coordination number. And this is no coincidence. One will find that the same
Schwinger boson calculation on the triangular lattice, at the onset of Bose con-
densation, also has an energy spectrum of the spin waves on the triangular
lattice. Schwinger boson theory has the power to capture the physics of both
the ordered and the disordered phases of the continuous magnet.
To complete the analysis, the self-consistency relation for Q is derived as
1X
Q= (12.64)
4 i
{cσ , c+
σ } = |0ihσ|σih0| + |σih0|0ihσ| = |0ih0| + |σihσ|. (12.65)
Normally this should have been one, but due to the limitations of the Hilbert
space, the electron creation operator can only connect an empty state to a
singly occupied state, but not a singly occupied state to a doubly occupied
state. Besides, the result (12.65) equals 1 − |σihσ| because we have
X
|0ih0| + |σihσ| = 1, (12.66)
σ
also due to the limited Hilbert space imposed by the no double occupancy
constraint. Using a similar idea, we can derive
{cσ , c+
σ} = 1 − cσ+ cσ
{cσ , c+
σ} = c+
σ cσ . (12.67)
P
On the other hand, commutation relation with the density operator n = σ |σihσ|
is the same.
[n, c+ +
σ ] = cσ , [n, cσ ] = −cσ . (12.68)
12.5. SLAVE-FERMION SCHWINGER BOSON 103
cσ = b+ fσ . (12.69)
Quite remarkably, this prescription recovers all the commutations of the re-
stricted Hilbert space, (12.67)-(12.68), provided we impose the constraint
X
b+ b + fσ+ fσ = 1. (12.70)
σ
X
c+ + + + +
σ cσ = bb fσ fσ = (1 + b b)fσ fσ = (2 − fσ+ fσ )fσ+ fσ = fσ+ fσ . (12.71)
σ
Similarly one can show that the spin operator has the expression
1 + 1
S= c σαβ cβ = fα+ σαβ fβ . (12.72)
2 α 2
cσ = f + bσ . (12.73)
Compared to the slave-boson substitution, here the role of boson and fermion is
simply reversed. The advantage is that now one obtains an easy generalization
of the Schwinger boson theory to treat the conducting electrons. Similar to the
slave-boson case, we have
c+ +
σ cσ = bσ bσ
c+ +
α σαβ cβ = bα σαβ bβ (12.74)
t-J Model
The “standard model” for the high-Tc cuprates is the t-J model. A popular
method used in the solution of the t-J model is first to introduce the slave
boson coordinates to rewrite the Hamiltonian, and apply mean-field theory.
Other, more sophisticated techniques exist, e.g. DMRG, exact diagonalization,
QMC, etc, which I do not talk about here. While the mean-field theory is the
least convincing of all the techniques, it is the most versatile and can be applied
to all sorts of problems involving t-J Hamiltonian.
The t-J model is defined by the Hamiltonian
1 Vc X 1
(c†jα ciα + h.c.) + J
X X
H = −t (Si · Sj − ni nj ) + (ni − n̄)(nj − n̄),
4 2 rij
hiji hiji i6=j
(13.1)
with typically the Coulomb term Vc set to zero. Each site i is roughly like an
atomic orbital, whose wavefunction hybridizes with its neighbors’ to produce
itinerancy. Because of the large Coulomb energy of putting two electrons in a
given atomic orbital (a few eV), it is usually believed that double occupation of
orbitals never happens.
A naive diagonalization of the above Hamiltonian would naturally involve
states which have two electrons at a given site, | ↑↓ii , which should cost very
high energy. In the t-J model this energy is treated as infinity, and the diago-
nalization takes place within the restricted Hilbert space consisting of an empty
site, and single occupied sites of either spin orientation. This sort of prob-
lem was confronted sometime in early 80’s by using the so-called slave-boson,
or auxiliary-boson technique. In a nutshell, one writes an electron operator
as a composite of a boson operator bi and another fermion operator fiσ as
ciσ = b†i fiσ . In this new choice of variables, the no-double-occupancy constraint
becomes
b†i bi +
X †
fiσ fiσ = 1 (13.2)
σ
for all i. In loose terms it means that the occupation of a given orbital by a
vacancy (b†i bi ) or by a single spin (fiσ
†
fiσ ) should always add up to one. Double
P †
occupancy is then by definition excluded because σ fiσ fiσ = 1 − b†i bi ≤ 1.
105
106 CHAPTER 13. T-J MODEL
1
(bj b†i fjα
†
X X
H = −t fiα + h.c.) + J (Si · Sj − ni nj )
4
hiji hiji
Vc X 1 †
λi (b†i bi +fiσ
X X †
+ (ni − n̄)(nj − n̄)+ fiσ − 1)−µ( fiσ fiσ −Ne ).
2 rij i iσ
i6=j
(13.4)
† †
2Six = fi↑ fi↓ + fi↓ fi↑
† †
2Siy = −i(fi↑ fi↓ − fi↓ fi↑ )
† †
2Siz = fi↑ fi↑ − fi↓ fi↓ . (13.6)
This fermion representation of spin satisfies all the commutator algebra of the
original spin. By substituting Eq. (6) back into Eq. (5) one completes the
expression of the t-J model in terms of slave-boson coordinates bi and fiσ .
It is horribly difficult to solve a Hamiltonian which contains a quartic inter-
action term. The J-term is quartic in the fermions, and the t-term is quartic
because it contains two bi ’s and two fiσ ’s. One can reduce the problem to the
fermion-only problem by treating the boson fields bi as condensed, and satisfying
the constraint
hb†i bi i = 1 − hfiσ
†
fiσ i. (13.7)
In reality one replaces the boson fields in the Hamiltonian by a complex number
†
whose magnitude is determined from |bi |2 = 1 − hfiσ fiσ i. Because the Hamilto-
nian is invariant under a local phase change bi → eiθi bi , fiσ → eıθi fiσ , one can
always choose bi to be real and positive without loss of generality.
The density-density interaction term can be rewritten as
Vc X 1 2
(b − x)(b2j − x), (13.8)
2 rij i
i6=j
Having removed the boson degrees of freedom, one can now face the quartic
interaction term by mean-field decoupling technique. One groups the quartic
term as a product of two bilinear operators, and replace one of them by its
expectation values. There are three distinct ways to pair up the fields, hence
three different order parameters in the meanfield theory. Keeping track of the
Fock and pairing terms in the mean-field decoupling scheme gives
3 s 1 t
Si · Sj → − (∆ij Pijs + χsij Hij
s
) + (∆ij Pijt + χtij Hij
t
)
8 8
1 1 s s 1 t
ni nj → (∆ P − χsij Hij s
) + (∆ij Pijt − χtij Hij
t
)
4 8 ij ij 8
(13.9)
s/t
Pij = fi1 fj2 ∓ fi2 fj1 ,
s/t † †
Hij = fj1 fi1 ± fj2 fi2 , (13.10)
† 3J X
(∆ij Pij† + χij Hij
†
X
HM F = −t (bj bi fjα fiα + h.c.) − + h.c.)
8
hiji hiji
JX JX
+ Mi σiz − (1 − b2i )(1 − b2j )
4 i 4
hiji
Vc X 1 2 X †
X †
+ (bi − x)(b2j − x) + λi (b2i + fiσ fiσ − 1) − µ fiσ fiσ ,
2 rij i i
i6=j
(13.12)
−χ∗ij − (8t/3J)bi bj
∆ij
Tij = ,
∆∗ij χij + (8t/3J)bi bj
λi + σmi 0
Miσ = , (13.14)
0 −λi + σmi
as
3 X + 1X +
HM F = J ψjσ Tij ψiσ + ψ Miσ ψiσ
8 2 iσ iσ
hijiσ
3 XX + 1X +
= J ψjσ Tij ψiσ + ψ Miσ ψiσ . (13.15)
16 iσ j∈i 2 iσ iσ
For actual calculation it is easier if we re-arrange the ψjσ to appear on the right:
3 XX + 1X +
HM F = J ψiσ Tji ψjσ + ψ Miσ ψiσ ,
16 iσ j∈i 2 iσ iσ
−χij − (8t/3J)bi bj ∆ij
Tji = .
∆∗ij χ∗ij + (8t/3J)bi bj
(13.16)
X uni −v ∗
ψiσ = ni Γnσ ,
vni u∗ni
n
γi1 γi2
Γn1 = + , Γ i2 = + . (13.17)
γi2 −γi1
3 3
− χ0ij unj + ∆ij vnj + λi uni = En uni
8 8
3¯ 3
∆ij unj + χ̄0ij vnj − λi vni = En vni , (13.18)
8 8
we will have
109
∗ ∗
X unj −vnj uni vni
Tji = En
vnj u∗nj vni −u∗ni
j∈i
X X u∗ ∗ ∗
ni vni unj −vnj
Tji = E n σz ,
−vni uni vnj u∗nj
i j∈i
(13.19)
3 X 3X
HM F = En Γ+
nσ σz Γnσ =
+
En γnσ γnσ . (13.20)
16 n 8 nσ
X X
χij = h +
fjσ fiσ i = uni u∗nj F (En ) + vni
∗
vnj F (−En )
σ n
X
∗ ∗
∆ij = hαβ fiα fjβ i = uni vnj F (−En ) − vni unj F (En )
n
X † X
ni = h fiσ fiσ i = |uni |2 F (En ) + |vni |2 F (−En ) .
σ n
(13.21)
P
The sum n runs over both positive and negative energy sets. The usual
∗
identity that (uni , vni ) of energy −E is equal to (−vni , u∗ni ) at energy +E still
holds and allows the self-consistency equations to simplify to
X
χij = 2 uni u∗nj F (En )
n
X
∗
∆ij = −2 vni unj F (En )
n
X
ni = 2 |uni |2 F (En ). (13.22)
n
Now we put the magnetism back in and see how to diagonalize the mean field
Hamiltonian. We try a slightly different Bogoliubov rotation than the previous
case,
X X
∗ + ∗ +
fi1 = (uni γn1 − yni γn2 ), fi2 = (xni γn2 + vni γn1 ),
n n
X uni ∗ ∗
X xni
−yni −vni
ψi1 = ∗ Γn1 , ψi2 = ∗ Γn2 .(13.23)
vni xni yni uni
n n
3 3
− χ0ij unj + ∆ij vnj + (mi + λi )uni = En uni
8 8
3¯ 3
∆ij unj + χ̄0ij vnj + (mi − λi )vni = En vni
8 8
(13.24)
2 4
− χ0ij unj + ∆ij vnj + (mi + λi )uni = En uni
8 8
4¯ 2
∆ij unj + χ̄0ij vnj + (mi − λi )vni = En vni
8 8
(13.25)
with χ0ij ≡ χij + (8t/2)bi bj (sII). Eigenfunctions (xni , yni ) are determined from
the same equation with mi → −mi . Alternatively one can obtain (xni , yni ) of
energy E as (v̄ni , −ūni ) where (uni , vni ) is an eigenstate of energy −E. This
statement holds for arbitrary magnetic states: ferromagnetic, antiferromagnetic,
and what not.
Self-consistency equation after considering the relation between (xni , yni )
and (uni , vni ) becomes Eq. (13.21) without it being reducible to Eq. (13.22).
Additional relation for local magnetic moment is obtained
X † X
Mi = h σfiσ fiσ i = |uni |2 F (En ) − |vni |2 F (−En ) . (13.26)
σ n
3 3
− [χ0x eikx +χ0x e−ikx + χ0y eiky +χ0y e−iky ]uk + [∆x cos kx +∆y cos ky ]vk +(m+λ)uk = Ek uk
8 4
3 3 0 −ikx
[∆x cos kx +∆y cos ky ]uk + [χx e +χx e +χ0y e−iky +χ0y eiky ]vk +(m − λ)vk = Ek vk .
0 ikx
4 8
(13.28)
111
3 0 ikx 3
Ak = [χ e + χ0x e−ikx + χ0y eiky + χ0y e−iky ], Bk = [∆x cos kx + ∆y cos ky ],
8 x 4
(13.29)
X
χx = |uk |2 e−ikx F (Ek ) + |vk |2 eikx F (−Ek )
Ek
X
χy = |uk |2 e−iky F (Ek ) + |vk |2 eiky F (−Ek )
Ek
X
∆x = uk vk∗ e−ikx F (−Ek ) − eikx F (Ek )
Ek
X
∆y = uk vk∗ e−iky F (−Ek ) − eiky F (Ek )
Ek
X
n= |uk |2 F (Ek ) + |vk |2 F (−Ek ) .
Ek
(13.31)
1 λ−Ak 1 λ−Ak 1 Bk
|uk |2 = 1+ , |vk |2 = 1− , uk vk∗ = . (13.32)
2 εk 2 εk 2 εk
1 λ−Ak 1 λ−Ak 1 Bk
2
|uk | = 1− 2
, |vk | = 1+ , uk vk∗ = − . (13.33)
2 εk 2 εk 2 εk
112 CHAPTER 13. T-J MODEL
Chapter 14
H AKLT =
X
P2S (ij), (14.1)
hiji
where P2S projects the spin operators Jij = Si + Sj for nearest-neighbor pair of
adjacent spins hiji onto a subspace of magnitude 2S. There are two constraints
on the spin operators in writing down the AKLT states,
Any pair of spins with a net spin less than 2S is projected out, i.e. P2S (ij)|J <
2Si = 0. For J = 2S, which is the maximal value attainable for the sum of two
spin-S operators, we get P2S (ij)|2Si = |2Si. For S = 1,
2 2
Jij (Jij − 2) 1 2 2
P2 (ij) = = J (J − 2) (14.3)
2·3·4 24 ij ij
2
and since Jij = (Si + Sj )2 = 2S(S + 1) + 2Si · Sj ,
1 2
P2 (ij) = Si · Sj + (Si · Sj )2 + . (14.4)
3 3
Hence the AKLT Hamiltonian for a one-dimensional spin-1 chain is written as
Xh 1 i
H AKLT = Si · Sj + (Si · Sj )2 . (14.5)
3
hiji
113
114 CHAPTER 14. EXACT SPIN HAMILTONIANS
Each hiji bond is covered by a dimer, a spin singlet, made up of two spin-1/2
constituents. The AKLT state has, for each hiji bond, a maximum Jij,z value of
2S−1. Then we must have J of that bond less than 2S, and P2S (ij)|AKLTi = 0.
To see that Jij,z does not exceed 2S−1, we take the square lattice as an example.
From Eq. (14.8), it is clear that a†i b†j − b†i a†j for the hiji bond gives Jij,z = 0.
According to Fig. ??, bonds hi2i, hi3i, and hi4i contribute a†i b†2 , a†i b†3 , and a†i b†4 ,
respectively and bonds hj5i, hj6i, and hj7i contribute a†j b†5 , a†j b†6 , and a†j b†7 ,
respectively. We have up to six up S = 1/2 spins or (a†i )3 (a†j )3 for the hiji bonds.
max
So Jij,z = 3 = 2(z − 1) · 21 = z − 1 = 2S − 1. Since the wavefunction does not
2
possess any J = 2S component for hiji bonds, one must have J2S (ij)|AKLTi =
0.
and the other is a translation of this state by one lattice unit. The proof is as
follows: construct the total spin operator for a triad of adjacent spins,
Ji = Si−1 + Si + Si+1
1 3
Ji2 = J(J + 1), J = , . (14.11)
2 2
14.2. MAJUMDAR-GHOSH STATES 115
Since one is adding up three S = 1/2 operators, the Hilbert space for the triad
of spins must be decomposed into either J = 3/2 or J = 1/2 sectors. The
appropriate projection operator is constructed as
1 3
Ji2 − ≡ P3/2 (i − 1, i, i + 1) . (14.12)
3 4
If {Si−1 , Si , Si+1 } form a J = 1/2 state, then P3/2 (i−1, i, i+1) |J = 1/2i = 0,
but if J = 3/2 then P3/2 P(i − 1, i, i + 1) |J = 3/2i = |J = 3/2i. HMG is a sum
of projectors, HMG = J i P3/2 (i − 1, i, i + 1). Due to the preceding argument
about the allowed values of J, each average hP3/2 (i − 1, i, i + 1)i must be non-
negative, and we can anticipate that the zero-energy state, if it exists, is the
ground state.
Expanding
1 3
P3/2 (i − 1, i, i + 1) = (Si−1 + Si + Si+1 )2 −
3 4
1 3
= + 2(Si−1 · Si + Si · Si+1 + Si−1 · Si+1 )
3 2
1 2
= + (Si−1 · Si + Si · Si+1 + Si−1 · Si+1 ) . (14.13)
2 3
Hence,
N
!
X NJ 4 X 1X
HMG =J P3/2 (i − 1, i, i + 1) = + J Si · Si+1 + Si · Si+2 .
i=1
2 3 i
2 i
(14.14)
Two of the three sites in the |MGi states are already bound into a singlet. The
portion of the wave function featuring the three sites i − 1, i, i + 1 are linear
combinations of
(a†i−1 b+ + † † † + + † +
i − bi−1 ai )ai+1 & (ai−1 bi − bi−1 ai )bi+1 . (14.15)