Spin liquids in graphene
Minh-Tien Tran1,2 and Ki-Seok Kim1,3
1
Asia Pacific Center for Theoretical Physics, Pohang, Gyeongbuk 790-784, Republic of Korea
Institute of Physics, Vietnamese Academy of Science and Technology, P.O.Box 429, 10000 Hanoi, Vietnam
3
Department of Physics, Pohang University of Science and Technology, Pohang, Gyeongbuk 790-784, Korea
arXiv:1011.1700v2 [cond-mat.str-el] 24 Jan 2011
2
We reveal that local interactions in graphene allow novel spin liquids between the semi-metal
and antiferromagnetic Mott insulating phases, identified with algebraic spin liquid and Z2 spin
liquid, respectively. We argue that the algebraic spin liquid can be regarded as the two dimensional
realization of one dimensional spin dynamics, where antiferromagnetic correlations show exactly the
same power-law dependence as valence bond correlations. Nature of the Z2 spin liquid turns out
to be d + id′ singlet pairing, but time reversal symmetry is preserved, taking d + id′ in one valley
and d − id′ in the other valley. We propose the quantized thermal valley Hall effect as an essential
feature of this gapped spin liquid state. Quantum phase transitions among the semi-metal, algebraic
spin liquid, and Z2 spin liquid are shown to be continuous while the transition from the Z2 spin
liquid to the antiferromagnetic Mott insulator turns out to be the first order. We emphasize that
both algebraic spin liquid and d ± id′ Z2 spin liquid can be verified by the quantum Monte Carlo
simulation, showing the enhanced symmetry in the algebraic spin liquid and the quantized thermal
valley Hall effect in the Z2 spin liquid.
PACS numbers:
I.
INTRODUCTION
Interplay between the topological band structure and
interaction drives one direction of research in modern
condensed matter physics,1,2 where emergence of Dirac
fermions is at the heart of the interplay. The original example is the system of one dimensional interacting electrons, where interactions become enhanced at low energies, combined with one dimensionality, and electron
fractionalization results, giving rise to a new state of
matter, dubbed as the Tomonaga-Luttinger liquid.3 An
interesting aspect is that such fractionalized excitations
as spinons and holons are identified with topological excitations, carrying fermion quantum numbers associated
with the topological structure of the Dirac theory.4
A recent study based on the quantum Monte Carlo
simulation5 has argued that essentially the same phenomenon as electron fractionalization in the TomonagaLuttinger liquid may happen in two dimensions when local interactions are introduced in the graphene structure.
This study claimed existence of a paramagnetic Mott insulator with a spin gap between the semi-metal and antiferromagnetic Mott insulating phases. Immediately, the
nature of the spin gapped Mott state has been suggested
to be an s-wave spin-singlet pairing order between next
nearest neighbor spins,6–8 thus identified with a Z2 spin
liquid. We point out other scenarios9,10 for the nature of
the spin liquid state.
In the present study we revisit this problem, the nature of possible spin liquids in the Hubbard model on
the graphene structure. An important point of our study
is to solve the Hubbard model directly beyond recent
studies,6–8 where an additional energy scale was introduced to describe the spin-singlet pairing order. The
SU(2) slave-rotor representation, invented by one of the
authors,11 is at the heart of the methodology, where ex-
change correlations via virtual processes are naturally
caught to allow spin singlet-pairing. One may regard
the SU(2) slave-rotor theory of the Hubbard model as an
analogue of the SU(2) slave-boson theory12 for the t-J
model.
We find two kinds of spin liquids, identified with an
algebraic spin liquid and a Z2 spin liquid, respectively,
between the semi-metal and antiferromagnetic phases.
We argue that the algebraic spin liquid13,14 can be regarded as the two dimensional realization of one dimensional spin dynamics, where antiferromagnetic correlations show exactly the same power-law dependence as valence bond correlations.15,16 Increasing interactions, pairing correlations between nearest neighbor sites become
enhanced. As a result, the algebraic spin liquid is shown
to turn into a gapped spin liquid state, where the spin gap
results from d + id′ singlet pairing, believed to originate
from the interplay between the topological structure and
interaction. Actually, this pairing symmetry solution has
been argued to be stable, based on an effective model in
the weak coupling approach.17,18 However, we argue that
time reversal symmetry is preserved, taking the d + id′
singlet pairing to one valley while the d − id′ pairing to
another. This assignment turns out to be essential in order to have a fully gapped spectrum because the d + id′
singlet pairing order parameter in one valley vanishes in
the other valley, allowing the gapless Dirac spectrum. We
propose the quantized thermal valley Hall effect19,20 for
the fingerprint of this gapped Z2 spin liquid. We discuss the nature of quantum phase transitions beyond the
mean-field analysis.
We would like to emphasize that appearance of both
algebraic spin liquid and d ± id′ Z2 spin liquid can be
verified by the quantum Monte Carlo simulation in principle. The fingerprint of the algebraic spin liquid is an
enhanced symmetry, giving rise to the same power-law
2
dependence between antiferromagnetic and valence bond
correlations. The hallmark of the d ± id′ Z2 spin liquid
is the quantized thermal valley Hall effect, as mentioned
above. We hope that the present study motivates quantum Monte Carlo simulation researchers to calculate such
correlation functions.
The present paper is organized as follows. In Sec. II
we present the SU(2) slave-rotor theory of the Hubbard
model. General mean field equations are also obtained.
The mean field analysis of possible phase transitions is
presented in Sec. III. In Sec. IV summary and discussion
are presented.
II.
SU(2) SLAVE-ROTOR THEORY OF THE
HUBBARD MODEL
A.
Formulation
i
ciσ (c†iσ )
is the annihilation (creation) operator
where
for an electron at site i with spin σ. t is the hopping
integral, and U is the on-site Coulomb interaction, where
niσ = c†iσ ciσ represents the density of electrons with spin
σ.
Introducing the Nambu-spinor representation
ci↑
,
ψi =
c†i↓
and performing the Hubbard-Stratonovich transformation for the pairing, density (singlet) and magnetic
(triplet) interaction channels, we obtain an effective Lagrangian
X †
X †
ψi σz ψj + H.c.
ψi (∂τ 1 − µσz )ψi − t
L=
i
hiji
X
†
†
†
I
−i
[ΦR
i (ψi σx ψi ) + Φi (ψi σy ψi ) + ϕi (ψi σz ψi )]
i
3 X R 2
+
[(Φi ) + (ΦIi )2 + (ϕi )2 ]
2U κc i
1 X 2 X
mi (ψi† ψi ).
mi −
+
2U κs i
i
ψi = Zi† Fi ,
Here, ziσ is a boson operator, satisfying the unimodular
†
†
(rotor) constraint, zi↑
zi↑ + zi↓
zi↓ = 1.
The key point of the slave-rotor representation21 is
to extract out collective charge dynamics explicitly from
correlated electrons. Such charge fluctuations are identified with zero sound modes in the case of short range interactions while plasmon modes in the case of long range
interactions. Actually, one can check that the dispersion
of the rotor variable (zi↑ ) is exactly the same as that of
such collective charge excitations.
In the slave-rotor theory the Mott transition is described by gapping of rotor excitations.21 Until now, the
Mott transition has not been achieved successfully, based
on the diagrammatic approach starting from the Fermi
liquid theory. In this respect the slave-rotor theory can
be regarded as an effective field theory, not bad for the
Mott transition.
Resorting to the SU(2) slave rotor representation in
Eq. (3), we rewrite the effective Lagrangian Eq. (2) as
follows
L = L0 + LF + LZ ,
X
†
Tr[Xij Yij† + Yij Xij
]+
L0 = t
hiji
(2)
LF =
R(I)
Φi
and ϕi are associated with pairingHere,
fluctuation and density-excitation potentials, introduced
to decouple the charge channel. mi is an effective magnetic field, which decouples the spin channel. κc and κs
are introduced for decoupling between singlet and triplet
interactions in respect that we do not know how they become renormalized at low energies. One may regard these
(3)
fi↑
is a fermion operator in the Nambu
†
fi↓
representation, and Zi is an SU(2) matrix
!
†
zi↑ −zi↓
Zi =
.
(4)
†
zi↓ zi↑
where Fi =
We start from the Hubbard model on the honeycomb
lattice
X
X †
ni↑ ni↓ ,
(1)
ciσ cjσ + H.c. + U
H = −t
hijiσ
two decoupling coefficients as phenomenological parameters to overcome the mean-field approximation. Several
examples for decoupling are shown in appendix A. We
emphasize that both the semi-metal to algebraic spin liquid and the algebraic spin liquid to Z2 spin liquid phase
transitions are shown not to depend on such phenomenological parameters, where both the critical points are
determined with the combination between U/t and κc .
Only the Z2 spin liquid to antiferromagnetic transition
turns out to depend on such parameters. We should be
careful in determining this phase transition, comparing
various cases (appendix A) with each other.
The SU(2) slave-rotor representation11 means to write
down an electron field as a composite field in terms of a
charge-neutral spinon field and a spinless holon field
X
i
−t
LZ =
Fi† (∂τ 1 − iΩi ·σ)Fi
(5)
1 X 2
mi , (6)
2U κs i
X
X †
mi (Fi† Fi ), (7)
(Fi Xij Fj + H.c.) −
i
hiji
X
3
Tr[Ωi ·σ−iZi ∂τ Zi† + iµZi σz Zi† ]2
4U κc i
X
Tr[Zi σ z Zj† Yij† + H.c.].
(8)
−t
hiji
3
a2
a
d1
d3
d2
a1
FIG. 1: Graphene. A distance between two next nearest
neighbor sites is chosen as the length unit. a1 and a2 are
primitive translation vectors. δ1 , δ2 , and δ3 are three nearest
neighbor bonds.
It is not difficult to see equivalence between the SU(2)
slave-rotor effective Lagrangian and the decoupled Hubbard model [Eq. (2)]. Integrating over field variables of
Xij and Yij , and shifting Ωi ·σ as
Ωi ·σ+iZi ∂τ Zi† − iµZi σz Zi† ,
I
where Ωi = (ΦR
i , Φi , ϕi ) is the pseudospin potential field,
we recover Eq. (2) exactly with an introduction of an
electron field Zi† Fi → ψi . This procedure is well described in the previous study.11 An important feature in
the SU(2) slave-rotor description is appearance of pairing correlations between nearest neighbor electrons, given
by off diagonal hopping in Xij which results from on-site
pairing (virtual) fluctuations, captured by the off diago-
F =−
nal variable zi↓ of the SU(2) matrix field Zi . We note that
the diagonal rotor field zi↑ corresponds to the zero sound
mode, giving rise to the Mott transition via gapping of
their fluctuations. The additional boson rotor variable
zi↓ allows us to catch super-exchange correlations in the
Mott transition. But, the appearance of pairing correlations does not necessarily lead to superconductivity because their global coherence, described by condensation
of SU(2) matrix holons, is not guaranteed. The similar
situation happens in the SU(2) slave-boson theory12 of
the t-J model.
B.
Mean-field ansatz
We perform the mean-field analysis, taking the following ansatz
wδ vδ∗
Xij =
· σz ,
(9)
vδ −wδ∗
w
eδ veδ∗
Yij =
· σz ,
(10)
veδ −w
eδ∗
where δ denotes the bond between the nearest neighbor
sites. In the honeycomb lattice there are three nearest neighbor bonds. We choose wδ = wγδ , vδ = vζδ ,
w
eδ = wγ
e δ , and veδ = veζδ , where γδ and ζδ are symmetric
factors for the hopping parameter w (w)
e and the pairing
order parameter v (e
v ). The choice for γδ and ζδ depends
on the symmetry of the considered phase. For example,
the s-wave pairing symmetry is given by ζδ = (1, 1, 1), the
dx2 −y2 -wave symmetry ζδ = (− 21 , − 21 , 1), and the dxy wave symmetry ζδ = (− 21 , − 21 , 0). For the magnetic order parameter mi we choose an antiferromagnetic ansatz
mi = (−1)i m.
Particle-hole symmetry at half filling results in µ +
I
iϕi = 0 while pairing potentials of ΦR
i and Φi vanish in
the mean-field level. Then, we obtain a general expression for the free energy
1X
1X
log[(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| + m)2 ] −
log[(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| − m)2 ]
β
β
k,iω
k,iω
X
2X
3
N
(iν)2 + λ)2 − 4t2 w
e2 |γ(k)|2 − 4t2 e
v 2 |ζ(k)|2 ] + 4tN
m2 − N λ, (11)
+
log[(−
(ww|γ
e δ |2 + ve
v |ζδ |2 ) +
β
4κc U
2κs U
δ
k,iν
P
where γ(k) = δ γδ exp(irδ · k) is thePenergy dispersion for spinons and holons, and ζ(k) = δ ζδ exp(irδ · k) is
associated with the pairing potential.
δ is performed in the unit cell. λ is a Lagrange multiplier field, introduced
to keep the slave-rotor constraint. N is the total number of sites.
Minimizing the free energy, we obtain fully self-consistent equations for order parameters
X
2tw X
|γ(k)|2
|γ(k)|2
2
w
e
|γδ | = −
, (12)
+
4N β
(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| + m)2
(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| − m)2
P
δ
ve
X
δ
k,iω
|ζδ |2 = −
(tv|ζ(k)| − m)|ζ(k)|
2 X
(tv|ζ(k)| + m)|ζ(k)|
,
+
4N β
(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| + m)2
(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| − m)2
k,iω
(13)
4
w
X
|γδ |2 =
X
|ζδ |2 =
δ
v
δ
m=−
|γ(k)|2
4tw
eX
,
3
2
2
N
(− 4κc U (iν) + λ) − 4t2 w
e2 |γ(k)|2 − 4t2 ve2 |ζ(k)|2
k
(14)
|ζ(k)|2
4te
vX
,
3
N
(− 4κc U (iν)2 + λ)2 − 4t2 w
e2 |γ(k)|2 − 4t2 ve2 |ζ(k)|2
(15)
k
2κs U X
tv|ζ(k)| + m
tv|ζ(k)| − m
2κs U X
+
,
2
2
2
2
2
Nβ
(iω) − t w |γ(k)| − (tv|ζ(k)| + m)
Nβ
(iω)2 − t2 w2 |γ(k)|2 − (tv|ζ(k)| − m)2
k,iω
(16)
k,iω
1=
(− 4κ3c U (iν)2 + λ)
4 X
.
Nβ
e2 |γ(k)|2 − 4t2 ve2 |ζ(k)|2
(− 4κ3c U (iν)2 + λ)2 − 4t2 w
k,iν
(17)
In this study our objective is to reveal the phase structure of the Hubbard model on the honeycomb lattice. It
is convenient to take the zero temperature limit. Performing the Matsubara frequency summation, we obtain selfconsistent mean-field equations at zero temperature
w
e
ve
X
δ
w
|ζδ |2 =
X
δ
v
X
δ
X
w X |γ(k)|2
w X |γ(k)|2
+
,
8N/2
D(k,m) 8N/2
D(k, −m)
(18)
1 X (v|ζ(k)| − m
1 X (v|ζ(k)| + m
t )|ζ(k)|
t )|ζ(k)|
+
,
8N/2
D(k,m)
8N/2
D(k, −m)
(19)
δ
|γδ |2 =
k
k
|γδ | =
r
|ζδ |2 =
r
2
k
m=
k
κb U w
e X |γ(k)|2
3 2N/2
E(k)
k
κb U ve X |ζ(k)|2
3 2N/2
E(k)
k
1
p
−p
λ − 2tE(k)
λ + 2tE(k)
1
1
p
−p
λ − 2tE(k)
λ + 2tE(k)
!
!
,
(20)
,
(21)
κs U X v|ζ(k)| + m
κs U X v|ζ(k)| − m
t
t
−
,
2N/2
D(k,m)
2N/2
D(k, −m)
k
1=
1
r
(22)
k
κb U 1 X
3 N/2
k
1
1
p
+p
λ − 2tE(k)
λ + 2tE(k)
!
,
(23)
count spatially uniform hopping
where
p
e2 |γ(k)|2 + e
v 2 |ζ(k)|2 ,
E(k) = w
r
m
D(k, m) = w2 |γ(k)|2 + (v|ζ(k)| + )2
t
(24)
(25)
are holon and spinon energy spectra in the presence of
pairing and antiferromagnetism, respectively.
Considering symmetry, it is natural to take into ac-
√
1
3
|γ(k)|2 = 3 + 2 cos(ky ) + 4 cos( ky ) cos(
kx ). (26)
2
2
On the other hand, the s-wave pairing potential is not
allowed due to repulsive interactions. Counting the lattice symmetry of the honeycomb structure, the next candidate will be dx2 −y2 or dxy for nearest neighbor singlet
pairing18 . We introduce a general combination of dx2 −y2 and dxy -wave pairing for the pairing term ζ(k)
|ζ(k)|2 = | cos(θ)ζx2 −y2 (k) + i sin(θ)ζxy (k)|2 ,
(27)
5
where θ is a combination factor, and ζx2 −y2 (ζxy ) is the
dx2 −y2 (dxy ) -wave symmetry function
ζx2 −y2 (kx , ky ) = e
kx
−i √
3
ζxy (kx , ky ) = ie
i 2k√x3
−e
i 2k√x3
sin(
cos(
ky
)
2
ky
).
2
For θ = ±π/3 this pairing symmetry becomes d ± id′ .
We also consider the d + d′ -wave pairing symmetry
|ζ(k)|2 = | cos(θ)ζx2 −y2 (k) + sin(θ)|ζxy (k)|2 ,
(28)
but this pairing order turns out to be not a solution of the
mean-field equations. If one tunes κc and κs parameters,
he can make this pairing symmetry a solution. However,
this solution does not give the lowest free energy, compared with the d + id′ pairing solution, consistent with
earlier studies.17,18
One may criticize the ansatz for uniform hopping in
this paper because such an assumption excludes possible
dimerized phases a priori. Actually, the J1 − J2 Heisenberg model
X
X
Sk · Sl
S i · S j + J2
H = J1
hiji
hhklii
has shown several types of dimerized phases when the
ratio of J2 /J1 is beyond a certain critical value8,22 , approximately given by J2 /J1 ≈ 0.2 ∼ 0.3. Here, the first
term represents the exchange interaction between nearest neighbor spins, and the second expresses that between
next nearest neighbor ones. This model can be derived
from the Hubbard model, resorting to the degenerate perturbation theory in the t/U → 0 limit23 , where each parameter is given by8
J1 = 4t
t 3 o
nt
,
−4
U
U
J2 = 4t
t 3
U
antiferromagnetic order disappears, and a paramagnetic
Mott insulating state results, identified with a certain
type of Z2 spin liquids. Such a spin-gapped state turns
out to evolve into a dimerized phase with either translational or rotational symmetry breaking near J2 /J1 ≈
0.2 ∼ 0.3.8,22 It was reported that the spin liquid state
turns into a dimerized phase with three-fold degeneracy
around J2 /J1 ≈ 0.3, where it breaks the C3 symmetry but preserving the translational symmetry.8 On the
other hand, the plaquette order was claimed to appear
near J2 /J1 ≈ 0.2 before the dimerized phase, breaking
the translational symmetry only.22 An important point
is that if we translate the critical J2 /J1 value in terms of
U/t of the Hubbard model, J2 /J1 ≈ 0.3 corresponds to
U/t ≈ 2.7 and J2 /J1 ≈ 0.2, U/t ≈ 3.0. Comparing these
critical values with the critical value for the semi-metal
to spin liquid transition in the quantum Monte Carlo
simulation5 , the Mott critical value given by U/t ≈ 3.5
turns out to be larger than those for dimerized phases.
This means that the semi-metal phase will appear before
reaching such dimerized phases in the Hubbard model
owing to charge fluctuations, not introduced in the J1 −J2
Heisenberg model. In other words, the J1 − J2 model
seems to be an effective low energy model only in the
limit of U/t → ∞ while physics of such a model will be
different from that of the Hubbard model in the small
U/t case.
However, it should be pointed out that these critical
values cannot be guaranteed. Thus, we cannot exclude
the possibility of dimerization near the Mott criticality of
the Hubbard model completely. In addition, introduction
of the next nearest neighbor hopping t′ will favor dimerization. In this respect it will be the best interpretation
that the spin liquid physics may appear at finite temperatures at least, actually observed from the quantum
Monte Carlo simulation.
III.
up to the fourth order process. Then, the J2 /J1 ratio
can be expressed in terms of U/t as follows
J2
1
=
.
J1
(U/t)2 − 4
It was argued that higher order terms such as third
neighbor and ring exchange terms may be ignored because third neighbor exchange terms are not frustrating, just renormalizing the J1 term effectively, while the
ring exchange term is expected to be small.8 However,
the role of the ring exchange term has been also studied
carefully.24,25
An antiferromagnetic phase has been reported in
J2 /J1 < (J2 /J1 )AF ≈ 0.08.8,22 This corresponds to
(U/t)AF ≈ 4.3, consistent with the result of the quantum Monte Carlo simulation.5 Increasing frustration, the
A.
SADDLE-POINT ANALYSIS
From semi-metal to algebraic spin liquid
The semi-metal phase is described by condensation of
holons hzσ i 6= 0 with v = ve = m = 0. Considering the
symmetry factor γδ = (1, 1, 1), the condensation occurs
when the effective chemical potential given by the Lagrange multiplier field λ touches the maximum point of
the holon dispersion, i.e., λc = 2tw
ec max |γ(k)| = 6tw
ec .
These collective charge excitations become gapped when
λ > λc , and a Mott insulating state appears, characterized by hzσ i = 0 with v = e
v = m = 0.
Taking λ = λc with v = ve = m = 0, we can determine
the quantum critical point from the following mean-field
equations
6
wc
X
δ
|γδ |2 =
1=
r
r
w
ec
X
δ
|γδ |2 =
k
κc U c 1 X
6tw
ec N/2
k
k
3−γ(k)
(29)
k
κc U c 1 X
|γ(k)|
6tw
ec 2N/2
1
1
p
−p
3 − γ(k)
3 + γ(k)
1
1
p
+p
3 − γ(k)
3 + γ(k)
Inserting w
ec from Eq. (29) into Eq. (31), one obtains
the critical value for the interaction strength
P
1
|γ(k)|
N/2
κc U c
3
k
=
2
t
2P
P
1
√ 1
+√ 1
|γδ |2 N/2
δ
1 X
|γ(k)|,
4N/2
!
.
!
,
(30)
(31)
and the gapped spin liquid state because such a simulation should be performed at finite temperatures. But,
calculations for correlation functions need not be done
at zero temperature. It is sufficient to show equivalent
correlation behaviors in the quantum critical region at
finite temperatures.
3+γ(k)
= 0.312.
(32)
It is interesting to notice that the resulting paramagnetic Mott insulator has all kinds of lattice symmetries. In particular, spin dynamics is described by gapless
spinons. An effective field theory for spinon dynamics
was proposed to be an SU(2) gauge theory with Dirac
fermions.13
It is not at all straightforward to understand dynamics
of such gapless spinons due to complexity of the SU(2)
gauge theory. It has been shown that an interacting stable fixed point arises in the large-Nf limit,13 where Nf
is the number of fermion flavors. Such a conformal invariant fixed point was also shown to appear in the U(1)
gauge theory with gapless Dirac fermions.26 An interesting property of the stable fixed point is that the symmetry of the original microscopic model, here the Hubbard model, is enhanced, associated with special transformation properties of Dirac spinors.14,15 As a result,
spin-spin correlations at an antiferromagnetic wave vector have exactly the same power-law dependence as valence bond-valance bond correlations, which means that
the scaling dimension of the staggered spin operator is
the same as that of the valence bond operator.16 This
situation is completely unusual because scaling dimensions of these two operators cannot be the same in the
level of the microscopic model.
It is clear that one direct way to verify the algebraic
spin liquid state is to observe the symmetry enhancement
at low energies. If the staggered-spin correlation function
turns out to display the same power-law behavior as the
valence-bond correlation function, this will be an undisputable evidence for the algebraic spin liquid phase between the semi-metal phase and gapped spin liquid state.
In the recent quantum Monte Carlo simulation data there
seems to be uncertainty between the semi-metal phase
B.
From algebraic spin liquid to Z2 spin liquid
Increasing κctU more from the semi-metal to algebraic
spin liquid critical point κctUc , we find another paramagnetic Mott insulating phase, characterized by v 6= 0
and e
v 6= 0 with the d ± id′ pairing symmetry. Recall Eq. (27) and Eq. (28) for pairing symmetries
that we checked explicitly. The algebraic spin liquid
(hzσ i = v = ve = m = 0) to gapped spin liquid (v 6= 0
and e
v 6= 0 with hzσ i = m = 0) critical point is found
with an ansatz of v = ve = 0 but v/e
v ≡ ϑ 6= 0. The
FIG. 2: (Color online) Spinon spectra ±D(k) given by Eq.
(25) with w = 1, v = 0.1 and m = 0. Left figure: If we
consider only the d + id′ pairing symmetry, spinon excitations
are gapped only at the K point, but they remain gapless at
the other K ′ point. Right figure: If we consider d + id′ in one
valley (K) and d − id′ in the other valley (K ′ ), spinon excitations become fully gapped. In this figure we assign d + id′
to the upper band and d − id′ to the lower band, respectively,
in order to realize the d ± id′ pairing symmetry.
7
mean-field equations to determine this critical point are
given in appendix B. The system of equations is solved
numerically, explicitly shown in appendix B.
We find that the free energy reaches the lowest value
for the d ± id′ pairing symmetry. Actually, we checked
self-consistency for various values of the angle parameter
θ in Eqs. (27) and (28), and found θ = ±π/3 in Eq. (27),
corresponding to d ± id′ . In Fig. 2 we plot the spinon
spectrum D(k) given by Eq. (25) for the d ± id′ pairing
symmetry. It shows that if only the d + id′ pairing order parameter is taken into account, the energy spectrum
opens a gap at one Brillouin zone edge (for instance, the
K point), but it still keeps the Dirac cone at the other
inequivalent Brillouin zone edge (the K ′ point). On the
other hand, if we consider only the d − id′ pairing symmetry, we see that the K point remains gapless while
only the K ′ point becomes gapped. This demonstration
motivates us to assign the d + id′ pairing symmetry to
one valley (K) and the d − id′ to the other (K ′ ), making the spinon spectrum fully gapped. Of course, this
fully gapped state is energetically more favorable than
the gapless state. In addition, this proposal resolves the
problem of time reversal symmetry breaking at the same
time. The edge state from the d+id′ pairing in one valley
is cancelled by that from the d − id′ pairing in the other
valley, preserving time reversal symmetry. One may regard this cancellation of such edge states as anomaly cancellation due to fermion doubling in condensed matter
physics.
The critical value turns out to be κc Uv /t = 0.315. Note
that Uv > Uc . This intermediate phase between the semimetal and gapped spin liquid is the algebraic spin liquid
with an enhanced symmetry, as discussed in the previous
subsection. We would like to emphasize that this region
of Uc < U < Uv is not wide at zero temperature. But,
the quantum critical region at finite temperatures will
not be so narrow, and it will not be so difficult to verify the algebraic spin liquid, considering staggered-spin
correlations and valence-bond correlations.
This pairing state can be verified by the quantized
thermal valley Hall effect.19,20 The spinon number is not
conserved due to particle-particle pairing, thus the charge
Hall conductivity is not useful. On the other hand, both
spin and energy (thermal) Hall coefficients are important
probes. But, the spin Hall conductivity vanishes due to
the different assignment between two valleys. The thermal valley Hall effect should be observed in this state,
regarded as the fingerprint of our Z2 spin liquid phase.
This may be verified27 by either quantum Monte Carlo
simulation28 or exact diagonalization29.
It should be noted that our time reversal symmetry
preserving Z2 spin liquid state is beyond the classification
scheme based on the projective symmetry group because
their possible Z2 spin liquids in the projective symmetry
group are constrained with complete time reversal symmetric pairing.6,7 In other words, d ± id′ singlet pairing
orders are excluded from the first although these pairing
orders are not only found but also argued to be stable in
recent studies.17,18
C.
From Z2 spin liquid to antiferromagnetic Mott
insulator
Our last subject is to investigate the quantum phase
transition from the Z2 spin liquid to the antiferromagnetic Mott insulator. Here, we should take into account
two order parameters such as the d ± id′ pairing and
antiferromagnetic ones. Generically, we expect four possibilities. The first candidate is coexistence between such
two orders, where the two critical lines cross each other.
As a result, we have two critical points inside each phase.
The second possibility is the multi-critical point, where
the two critical points meet at one point. The third situation will be the first order transition between them.
The last corresponds to an intermediate state without
any ordering, where the two critical points do not meet.
First of all, we exclude the last possibility because this
phase is nothing but the algebraic spin liquid and there
is no reason for this reentrant behavior.
We start to examine the possibility of coexistence. The
antiferromagnetic critical point inside the Z2 spin liquid
phase can be determined by m = 0 while v and ve are
finite, thus determined self-consistently. The mean field
equations for this quantum critical point are given in appendix C-1. The strategy of solving the system of equations is how to reduce the number of self-consistent equations. Detailed calculations are provided in appendix
C-1. As a result, we obtain two self-consistent equations for two unknown variables. These equations can be
solved numerically. For the first (κc = 1, κs = 1) and
third (κc = 3/2, κs = 1/2) decomposition schemes in
appendix A, we could show that there are no mean field
solutions at the transition point. On the other hand,
we find Um /t = 0.360 in the case of the d + id′ pairing
symmetry for the second decomposition scheme (κc = 1,
κs = 1/2).
The other quantum critical point is the Z2 spin liquid
critical point inside the antiferromagnetic phase. It can
be found when v = ve = 0 but m is finite, determined
self-consistently. The mean field equations for this quantum critical point are given in appendix C-2. Solving the
mean field equations self consistently, we could not find
any solution. On the other hand, if the direct phase transition from the antiferromagnetic Mott insulator to the
semi-metal is concerned, we find the critical point occurs
at Um /t = 0.330 for the second decomposition (κc = 1,
κs = 1/2).
Our analysis for the quantum phase transition from
the Z2 spin liquid to the antiferromagnetic Mott insulator shows that the nature of this transition depends
on our phenomenological parameters of κc and κs . We
could find the antiferromagnetic quantum critical point
inside the Z2 spin liquid state for particular values of
κc and κs while we could not obtain the Z2 spin liquid
quantum critical point inside the antiferromagnetic Mott
8
insulating phase. We could not find the multi-critical
point solution, either. As mentioned before, it is difficult
to expect the algebraic spin liquid solution between the
Z2 spin liquid and antiferromagnetic phases. Actually,
we could find only one solution for the Z2 spin liquid
to algebraic spin liquid transition, given by the previous
subsection. The remaining possibility is the first order
transition between the Z2 spin liquid and antiferromagnetic Mott insulator. We believe that the first order transition is the generic case for the phase transition between
these two phases. The formal procedure will be to integrate over spinons and holons and to obtain an effective
Landau-Ginzburg-Wilson free energy functional for both
d ± id′ spin singlet pairing and antiferromagnetic order
parameters. Based on the effective field theory, we can
perform the renormalization group analysis and find the
nature of the phase transition. This study is beyond the
scope of the present study.
One may ask the possibility of the Landau-GinzburgWilson forbidden continuous transition between the Z2
spin liquid with the d ± id′ pairing symmetry and the antiferromagnetic Mott insulator. Classification of LandauGinzburg-Wilson forbidden continuous transitions in two
spatial dimensions has been performed in Ref. 30. Investigating the classification table carefully, we can find that
this transition does not belong to any cases. The main
reason is that the singlet pairing order parameter cannot
be symmetrically equivalent to the antiferromagnetic order parameter. The classification scheme reveals that
the Néel order parameter can form a hyper-vector with a
triplet pairing order parameter. In this respect we are allowed to exclude the possibility of the Landau-GinzburgWilson forbidden continuous transition between the Z2
spin liquid and the antiferromagnetic Mott insulator.
IV.
DISCUSSION AND SUMMARY
In this paper we investigated the phase structure of the
Hubbard model on the honeycomb lattice. Physics of one
dimensional interacting electrons is our reference. As well
known, even if we start from weak interactions, they become enhanced at low energies, destabilizing the Fermi
liquid state. In one dimension such quantum corrections
can be summed exactly, resorting to the Ward identity.31
The resulting electron Green’s function shows two kinds
SM
ASL
GSL
AFM
X
Uc
Uv
Um
U/t
FIG. 3: Schematic phase diagram. Abbreviations: SM is the
semi-metal phase, ASL is the algebraic spin liquid, GSL is the
gapped spin liquid, and AFM is antiferromagnetism. The SMASL and ASL-GSL quantum phase transitions belong to the
second order while the GSL-AFM quantum phase transition
is the first order.
of branch cuts, corresponding to collective charge and
spin excitations. In this diagrammatic approach it is
difficult to see the nature of such fractionalized excitations. But, the bosonization approach is helpful at low
energies, revealing that spinons and holons are identified with topological solitons such as domain walls.3 One
can interpret this phenomenon in another respect that
topological solitons acquire fermion quantum numbers
via fermion zero modes, regarded as realization of quantum anomaly.4 We believe that the spin-charge separation in one dimensional interacting electrons results from
not only just interaction effects but also hidden topological properties of Dirac fermions. Then, the next natural
question is whether we can find this physics in higher
dimensions.
The graphene structure is an ideal system for realization of Dirac fermions. The first observation in this Dirac
fermion system is that the vanishing density of states
needs a finite value of the interaction strength U for an
antiferromagnetic order to be achieved. Then, the question is whether we can find intermediate phases between
the semi-metal and antiferromagnetic Mott insulator, allowing fractionalized excitations as one dimensional interacting electrons. Indeed, we could find two kinds of
paramagnetic Mott insulating phases, which show fractionalized excitations.
The algebraic spin liquid appears from the semi-metal
state via the Higgs transition, gapping of charge fluctuations. Although it is not clear how the topological
nature of Dirac fermions is introduced to result in such a
spin liquid state, spinon excitations in the algebraic spin
liquid can be identified with topological excitations corresponding to meron (half skyrmion) excitations.32 The
underlying mechanism is that the symmetry of the original microscopic model is enhanced at low energies, allowing a topological term to assign a fermion quantum number to such a topological excitation. The algebraic spin
liquid turns out to have an O(5) symmetry in the physical case, where antiferromagentic correlations exhibit the
same power-law dependence for distance as valence-bond
correlations.14–16 It was pointed out that the corresponding effective field theory would be given by an O(5) WessZumino-Witten theory,16 identifying spinons with such
topological excitations. Comparing the algebraic spin liquid with the Tomonaga-Luttinger liquid, there is one to
one correspondence between them except that charge excitations are critical in the Tomonaga-Luttinger liquid.
Actually, spin dynamics in one dimension is governed by
the O(4) Wess-Zumino-Witten theory,3 describing critical dynamics of spinons.
Because the stability of the algebraic spin liquid is not
guaranteed beyond the large-Nf limit, we proposed how
the quantum Monte Carlo simulation can prove the existence of such a phase. As discussed before, the symmetry
enhancement can be verified, calculating both antiferromagnetic and valence-bond correlations at finite temperatures. If such correlations turn out to have the same
scaling behavior, we have the algebraic spin liquid phase
9
just beside the semi-metal state.
When interactions are increased more, pairing correlations between nearest neighbor sites become enhanced
in the singlet channel, destabilizing the algebraic spin
liquid. As a result, spinon excitations are gapped due
to their pairing orders. An interesting point is that the
nature of this gapped spin liquid state is given by the
d + id′ singlet pairing order, which breaks time reversal
symmetry. We would like to emphasize that time reversal symmetric combinations based on the d-wave pairing symmetry turn out to give higher energies than the
d + id′ pairing order. We suspect that this time reversal
symmetry breaking may be related with the Berry phase
effect of the momentum space.33 One way to verify this
statement is to check how the d + id′ pairing symmetry is changed, increasing the chemical potential from
the Dirac point, where the Berry phase effect becomes
weaken. Unfortunately, the quantum Monte Carlo simulation claimed that there is no time reversal symmetry
breaking in the gapped spin liquid state. This inconsistency was resolved, taking d − id′ pairing to another
valley. As a result, the edge state from the d + id′ pairing
is cancelled by that from the d − id′ one. In addition to
this time reversal symmetry, our proposal for the pairing
order parameter turns out to be essential in order to have
a fully gapped spectrum of spinon excitations, thus energetically more favorable than the case of only the d + id′
pairing, where spin excitations remain gapless. We suggested an experimental signature, that is, the quantized
thermal valley Hall effect as the fingerprint of this gapped
spin liquid.
Finally, we investigated the quantum phase transition
from the Z2 spin liquid to the antiferromagnetic Mott
insulator. We concluded that the first order transition
will take place generically. We argue that this first order
transition is involved with two symmetrically unrelated
order parameters, displaying different discrete symmetry
properties, here time reversal symmetry. We claim that
the Landau-Ginzburg-Wilson forbidden continuous transition will not appear, based on the existing classification
scheme in the two dimensional Dirac theory on the honeycomb lattice.30
We would like to point out that the SU(2) slave-rotor
theory seems to overestimate quantum fluctuations. If
one sets κc as the order of 1, the critical strength of the
Mott transition is the order of 10−1 for the critical value,
compared with that from the quantum Monte Carlo simulation. This overestimation originates from strong band
renormalization for spinons and holons, given by effective
hopping integrals, Xij and Yij . Qualitatively the same
situation also happens in the U(1) slave-rotor theory13
while the SU(2) slave-rotor theory seems to overestimate
quantum fluctuations more. We believe that this aspect
should be investigated more sincerely.
Recently, the role of the spin-orbit interaction in the
Hubbard model on the honeycomb lattice has been studied both extensively and intensively, where one purpose
is to reveal the interplay between the topological band
structure given by the spin-orbit coupling and strong
correlation effect. Novel exotic phases have been suggested in this Kane-Mele-Hubbard model, some of which
are quantum spin Hall effect in a transition metal oxide such as Na2 IrO3 34 , a spin liquid state with a topological band structure35 , and the chiral spin liquid state
with the anyon nature of excitations36 . These interesting proposals will be verified based on ”exact” numerical
calculations37 .
Acknowledgments
We would like to thank T. Takimoto for useful discussions. We were supported by the National Research
Foundation of Korea (NRF) grant funded by the Korea
government (MEST) (No. 2010-0074542). M.-T. was
also supported by the National Foundation for Science
and Technology Development (NAFOSTED) of Vietnam.
Appendix A: Decoupling scheme
We discuss several decoupling schemes. The first example is
UX †
U X †
2HU =
(ψi σx ψi )2 +
(ψi σy ψi )2
6 i
6 i
U XX
UXX
(
(
niσ − 1)2 +
niσ − 1)
+
6 i σ
6 i σ
UX †
U X
−
(σciσ ciσ )2 +
niσ .
(A1)
2 iσ
2 iσ
Formally, this magnetic decoupling does not correspond
to the conventional Hartree-Fock analysis for antiferromagnetism because the interaction strength is twice
larger the standard mean field value.
The second possible decoupling is
UX †
U X †
2HU =
(ψi σx ψi )2 +
(ψi σy ψi )2
6 i
6 i
U XX
UXX
(
(
niσ − 1)2 +
niσ − 1)
+
6 i σ
6 i σ
X †
UX X †
+
[(
ciσ ciσ )2 − (
σciσ ciσ )2 ]. (A2)
4 i
σ
σ
This decoupling recovers the standard mean field theory
for
antiferromagnetism, but the coefficient of the term
PP
5
U , not equal to the coefficient of the
( niσ − 1)2 is 12
σ
i
P †
term (ψi σx ψi )2 .
i
The third possible decoupling is
U X †
U X †
(ψi σx ψi )2 +
(ψi σy ψi )2
2HU =
4 i
4 i
X †
U X X †
+
[(
ciσ ciσ )2 − (
σciσ ciσ )2 ]. (A3)
4 i
σ
σ
10
The third decoupling scheme seems natural, but we in-
troduce phenomenological parameters κc and κs .
Appendix B: The algebraic spin liquid to Z2 spin liquid critical point
The mean field equations for the algebraic spin liquid to Z2 spin liquid critical point are given by
w
ev
X
δ
1 X
|γ(k)|,
4N/2
|γδ |2 =
(B1)
k
1 1 X |ζ(k)|2
1 X
|ζδ |2 =
,
ϑv
4wv N/2
|γ(k)|
δ
X
wv
δ
ϑv
X
δ
2
|γδ | =
2
|ζδ | =
r
1=
r
1
κc U v 1 X
1
−q
,
|γ(k)| q
6tw
ec 2N/2
ev − γ(k)
ev + γ(k)
k
λ
λ
X |ζ(k)|2
1
κc U v
6tw
ev 2w
ev N/2
γ(k)
k
r
(B2)
k
κc U v 1 X
6tw
ev N/2
k
1
1
q
−q
ev − γ(k)
ev + γ(k)
λ
λ
1
1
q
+q
ev − γ(k)
ev + γ(k)
λ
λ
,
,
(B3)
(B4)
(B5)
ev is redefined.
where λv = 2tw
ev λ
The strategy for the critical interaction strength is to solve Eq. (B5) with w
ev from Eq. (B1). The point is how to
e
find λv from other equations. From Eqs. (B3) and (B5) we obtain
P
1
1
1
√
√
|γ(k)|
−
2N/2
ev −γ(k)
ev +γ(k)
λ
λ
k
.
wv =
(B6)
P
P
1
√ 1
+√ 1
|γδ |2 N/2
δ
ev −γ(k)
λ
k
ev +γ(k)
λ
Inserting Eq. (B6) into Eq. (B2), we get
2
ϑv = P
δ
P
|ζδ |
δ
1
2
|γδ | N/2
1
N/2
2
|γ(k)|
1 P √
N/2
P |ζ(k)|2
k
|γ(k)|
ϑv
X
δ
|ζδ |2 =
P |ζ(k)|2
k
k
γ(k)
1
2w
ev N/2
P
k
P
δ
1
N/2
2
2
|ζδ |
P |ζ(k)|2 =
k
|γ(k)|
P
2
2
|γδ |
P
1
|γ(k)|
N/2
δ
k
1
ev −γ(k)
λ
√
1
ev −γ(k)
λ
√
Equations (B7) and (B8) with w
ev from Eq. (B1) give
1
N/2
1
N/2
√ 1
ev −γ(k)
λ
1
ev −γ(k)
λ
k
From Eqs. (B4) and (B5) we obtain
1
N/2
P
P |ζ(k)|2
k
P
k
γ(k)
√ 1
ev +γ(k)
λ
−
1
λv +γ(k)
+ √e
1
λv +γ(k)
− √e
.
1
ev −γ(k)
λ
− √e
1
λv −γ(k)
− √e
√
|γ(k)| √e
.
+ √e
1
λv +γ(k)
(B7)
(B8)
1
λv +γ(k)
1
λv +γ(k)
.
(B9)
11
ev . Once we find λ
ev , we can obtain the critical value from Eq. (B5) together with Eq.
This equation determines λ
(B1), given by
κc U v
=
t
1
3 N/2
2
P
δ
|γδ
|2
1
N/2
P
k
P
k
|γ(k)|
√ 1
ev −γ(k)
λ
+
√ 1
ev +γ(k)
λ
2 .
(B10)
Appendix C: To analyze the quantum phase transition from the Z2 spin liquid to the antiferromagnetic Mott
insulator
1.
To find the antiferromagnetic quantum critical point inside the Z2 spin liquid state
The mean field equations for this quantum critical point are given by
w
em
vem
X
wm
δ
vm
X
δ
2
|γδ | =
2
|ζδ | =
|γδ |2 =
X
|ζδ |2 =
δ
δ
r
r
X
1 X
wm |γ(k)|2
p
,
2 |γ(k)|2 + v 2 |ζ(k)|2
4N/2
wm
m
vm |ζ(k)|2
1 X
p
,
2 |γ(k)|2 + v 2 |ζ(k)|2
4N/2
wm
m
k
em X |γ(k)|2
κc U m w
3 2N/2
E(k)
k
κb Um vem X |ζ(k)|2
3 2N/2
E(k)
k
1
1
p
p
,
−
λm − 2tE(k)
λm + 2tE(k)
1
1
p
,
−p
λm − 2tE(k)
λm + 2tE(k)
1
κs U m 1 X
p
,
2
2
2 |ζ(k)|2
t N/2
wm |γ(k)| + vm
1=
(C1)
k
(C2)
(C3)
(C4)
(C5)
k
1=
r
κc U m 1 X
3 N/2
k
1
1
p
.
+p
λm − 2tE(k)
λm + 2tE(k)
Introducing xm = vm /wm and x
em = vem /w
em , we rewrite the above equations as
w
em
wm
X
δ
|γδ |
2
=
r
vem
X
|γδ |2 =
X
|ζδ |2 =
δ
δ
|γ(k)|2
1 X
p
,
4N/2
|γ(k)|2 + x2m |ζ(k)|2
(C6)
(C7)
k
xm |ζ(k)|2
1 X
p
,
4N/2
|γ(k)|2 + x2m |ζ(k)|2
(C8)
k
κc U m 1 X
|γ(k)|2
p
6tw
em 2N/2
|γ(k)|2 + x
e2m |ζ(k)|2
k
1
1
,
−q
× q
p
p
em − |γ(k)|2 + x
em + |γ(k)|2 + x
λ
λ
e2m |ζ(k)|2
e2m |ζ(k)|2
(C9)
12
vm
X
δ
|ζδ |2 =
r
em X
|ζ(k)|2
κc U m x
p
6tw
em 2N/2
|γ(k)|2 + x
e2m |ζ(k)|2
k
1
1
q
,
−
× q
p
p
em − |γ(k)|2 + x
em + |γ(k)|2 + x
e2m |ζ(k)|2
e2m |ζ(k)|2
λ
λ
κs U m 1 X
1
p
,
twm N/2
|γ(k)|2 + x2m |ζ(k)|2
1=
(C10)
(C11)
k
1=
r
κc U m 1 X
6tw
em N/2
k
1
1
q
+q
p
p
em − |γ(k)|2 + x
em + |γ(k)|2 + x
λ
λ
e2m |ζ(k)|2
e2m |ζ(k)|2
em is redefined. From Eqs. (C7) and (C8) we get
where λm = 2tw
em λ
P
δ
x
em P
δ
|ζδ |2
|γδ |2
1
N/2
= xm
1
N/2
Similarly, Equations (C9) and (C10) give
P
δ
xm P
δ
|ζδ |2
|γδ
|2
1
N/2
=x
em
1
N/2
P
√
|ζ(k)|2
|γ(k)|2 +e
x2m |ζ(k)|2
P
√
|γ(k)|2
|γ(k)|2 +e
x2m |ζ(k)|2
P
√
k
k
P
k
√
|ζ(k)|2
|γ(k)|2 +x2m |ζ(k)|2
k
√
|γ(k)|2
|γ(k)|2 +x2m |ζ(k)|2
P
,
(C12)
.
(C13)
q
√ 1
em − |γ(k)|2 +e
λ
x2m |ζ(k)|2
−
q
√ 1
em + |γ(k)|2 +e
λ
x2m |ζ(k)|2
q
√ 1
2 |ζ(k)|2
em − |γ(k)|2 +e
λ
xm
−
q
√ 1
2 |ζ(k)|2
em + |γ(k)|2 +e
λ
xm
q
√ 1
em − |γ(k)|2 +e
λ
x2m |ζ(k)|2
−
q
.
(C14)
From Eqs. (C9) and (C12) we obtain
wm
X
δ
|γδ |2 =
1
2
1
N/2
k
|γ(k)|2
|γ(k)|2 +e
x2m |ζ(k)|2
1
N/2
P
k
q
√ 1
em − |γ(k)|2 +e
λ
x2m |ζ(k)|2
+
q
em +
λ
√
em +
λ
√
1
|γ(k)|2 +e
x2m |ζ(k)|2
1
|γ(k)|2 +e
x2m |ζ(k)|2
.
Taking both sides of Eq. (C9) to the square power with w
em from Eq. (C7), we obtain
"
#2
P
|γ(k)|2
1
1
1
q
√
−q
√
√
N/2
em − |γ(k)|2 +e
em + |γ(k)|2 +e
|γ(k)|2 +e
x2m |ζ(k)|2
X
λ
λ
x2m |ζ(k)|2
x2m |ζ(k)|2
k
κc U m
2
wm
.
|γδ | =
P
|γ(k)|2
1
6twm
√
δ
N/2
k
|γ(k)| +xm |ζ(k)|
k
λm −
(C17)
|γ(k)| +xm |ζ(k)|
Equations (C15), (C16), and (C17) give us
1
κc 1 X
|γ(k)|2
1
p
q
q
−
p
p
3κs N/2
|γ(k)|2 + x
e2m |ζ(k)|2
em − |γ(k)|2 + x
em + |γ(k)|2 + x
k
e2m |ζ(k)|2
e2m |ζ(k)|2
λ
λ
2
P
P
|γ(k)|
1
1
1
√
√
N/2
2 +x2 |ζ(k)|2 N/2
2 +x2 |ζ(k)|2
|γ(k)|
|γ(k)|
m
m
k
.
k
=
P
1
1
1
q
q
+
√
√
N/2
2
2
2
2
2
2
e
e
k
(C16)
|γ(k)|2 +x2m |ζ(k)|2
Inserting Um /wm from Eq. (C11) into this equation, we obtain
!#2
"
|γ(k)|2
1
1
1 P √
q
q
−
√
√
N/2
|γ(k)|2 +e
x2m |ζ(k)|2
em − |γ(k)|2 +e
em + |γ(k)|2 +e
X
λ
x2m |ζ(k)|2
λ
x2m |ζ(k)|2
k
κc
2
.
wm
|γδ | =
P
P
2
|γ(k)|
1
1
1
6κs
√
√
δ
N/2
2
2
2
2
2 N/2
2
k
(C15)
|γ(k)| +e
xm |ζ(k)|
λm +
|γ(k)| +e
xm |ζ(k)|
(C18)
13
Inserting
P
δ
x
em = xm P
δ
|γδ |2
1
N/2
|ζδ |2
1
N/2
P
|ζ(k)|2
|γ(k)|2 +x2m |ζ(k)|2
k
√
k
√
P
(C19)
|γ(k)|2
|γ(k)|2 +x2m |ζ(k)|2
from Eq. (C13) into Eqs. (C14) and (C18), we obtain two self-consistent equations for two unknown variables, xm
em . These equations can be solved numerically. We fix xm first, and solve these two equations for λ
em . Then, we
and λ
e
obtain two functions, λm of xm . When two lines of these functions intersect, we obtain the solution of such equations.
em are determined, the critical value of Um is also found from Eq. (C11).
Once xm and λ
2.
To find the Z2 spin liquid quantum critical point inside the antiferromagnetic Mott insulator
The mean field equations at this critical point are given by
X
1 X
|γ(k)|2
q
w
ea
|γδ |2 =
,
4N/2
|γ(k)|2 + ( ma )2
δ
k
wa
δ
ϑa
X
δ
twa
X
1
|ζ(k)|2
1 X
q
|ζδ |2 =
,
ϑa
4wa N/2
|γ(k)|2 + ( ma )2
δ
X
(C20)
2
|γδ | =
2
|ζδ | =
r
r
k
twa
κc U a 1 X
1
1
q
q
−
,
|γ(k)|
6tw
ea 2N/2
ea − γ(k)
ea + γ(k)
k
λ
λ
X |ζ(k)|2
1
κc U a
6tw
ea 2w
ea N/2
γ(k)
k
1=
r
1
1
q
−q
ea − γ(k)
ea + γ(k)
λ
λ
κs U a 1 X
1
q
,
twa N/2
2
|γ(k)| + ( ma )2
k
1=
(C21)
κc U a 1 X
6tw
ea N/2
k
,
(C22)
(C23)
(C24)
twa
1
1
q
+q
,
ea − γ(k)
ea + γ(k)
λ
λ
(C25)
ea is redefined. We solve these equations basically in the same way as the previous case. First,
where λa = 2tw
ea λ
we reduce the system of six equations into two equations of two unknown variables, and solve the two equations
numerically.
From Eqs. (C22) and (C25) we get
P
1
1
1
√
√
|γ(k)|
−
N/2
ea −γ(k)
ea +γ(k)
λ
λ
1
k
.
wa = P
(C26)
P
2 |γδ |2
1
1
1
√
√
+
δ
N/2
ea −γ(k)
λ
k
ea +γ(k)
λ
Equations (C23) and (C25) together with w
ea from Eq. (C20) give
P |ζ(k)|2
1
1
1
√
√
− e
N/2
γ(k)
ea −γ(k)
λ
λa +γ(k)
1
k
P
ϑa =
2
P
2w
ea |ζδ |
1
1
1
√
√
+
δ
N/2
P
2 |γδ |
= Pδ
|ζδ |2
δ
1
N/2
2
1
N/2
P
k
√
P |ζ(k)|2
k
γ(k)
ea −γ(k)
λ
k
1
ea −γ(k)
λ
√
|γ(k)|2
1
ma 2 N/2
)
|γ(k)|2 +( tw
a
P
k
ea +γ(k)
λ
1
λa +γ(k)
− √e
1
ea −γ(k)
λ
√
1
λa +γ(k)
+ √e
.
(C27)
14
From Eqs. (C21) and (C26) we obtain
P
1
N/2
2
2 |ζδ |
ϑa = Pδ
|γδ |2
δ
1
N/2
Equations (C27) and (C28) leads to
P
δ
P
δ
|γδ |2
|ζδ
|2
2
1
N/2
2 =
1
N/2
P
k
P
k
k
|γ(k)|
√ 1
ea −γ(k)
λ
2
1
√ |ζ(k)|
ma 2 N/2
)
|γ(k)|2 +( tw
a
|γ(k)|
P |ζ(k)|2
k
P
γ(k)
1
ea −γ(k)
λ
√
P
k
−
√ 1
ea −γ(k)
λ
1
λa +γ(k)
− √e
√ 1
ea +γ(k)
λ
√ 1
ea −γ(k)
λ
−
P
√
√ 1
ea +γ(k)
λ
1
N/2
1
N/2
√ 1
ea +γ(k)
λ
+
P
k
P
k
.
|γ(k)|2
ma 2
)
|γ(k)|2 +( tw
a
√
(C28)
.
(C29)
|ζ(k)|2
ma 2
)
|γ(k)|2 +( tw
a
√
From Eqs. (C24) and (C25) we obtain
6κs w
ea
1=
κc wa
1
N/2
1
N/2
P
k
k
1
ma 2
)
|γ(k)|2 +( tw
a
1
ea −γ(k)
λ
√
1
λa +γ(k)
+ √e
2 .
(C30)
Inserting wa and w
ea from Eqs. (C20) and (C26) into this equation, one obtains
3κs
1=
κc
1
N/2
1
N/2
P
k
P
k
|γ(k)|2
ma 2
)
|γ(k)|2 +( tw
a
|γ(k)| √e
1
λa −γ(k)
1
λa +γ(k)
− √e
Equations (C29) and (C31) are the last two equations
ea and ma /twa . We fix ma /twa first, and
determining λ
ea numerically. Then, we
solve the two equations for λ
1
2
3
4
5
1
N/2
√
X.-L. Qi and S.-C. Zhang, arXiv:1008.2026 (unpublished);
M. Z. Hasan and C. L. Kane, arXiv:1002.3895 (unpublished); P. G. Silvestrov and E. G. Mishchenko,
arXiv:0912.4658 (unpublished); S. Murakami, Prog. Theor.
Phys. Suppl. 176, 279 (2008); M. Koenig, H. Buhmann, L.
W. Molenkamp, T. L. Hughes, C.-X. Liu, X.-L. Qi, and S.C. Zhang, J. Phys. Soc. Jpn. 77, 031007 (2008).
E. R. Mucciolo and C. H. Lewenkopf, J. Phys.: Condens.
Matter 22, 273201 (2010); S. Das Sarma, S. Adam, E.
H. Hwang, and E. Rossi, arXiv:1003.4731 (unpublished);
D. S. L. Abergel, V. Apalkov, J. Berashevich, K. Ziegler,
and T. Chakraborty, Advances in Physics, 59 (4), 261
(2010); A. H. Castro Neto, F. Guinea, N. M. R. Peres, K.
S. Novoselov, and A. K. Geim, Rev. Mod. Phys. 81, 109
(2009).
A. O. Gogolin, A. A. Nersesyan, and A. M. Tsvelik,
Bosonization and Strongly Correlated Systems (Cambridge
University Press, New York, 2004).
R. Rajaraman, Solitons and Instantons(Elsevier Science,
New York, 2003).
Z. Y. Meng, T. C. Lang, S. Wessel, F. F. Assaad, and A.
Muramatsu, Nature 464, 847 (2010).
1
N/2
P
k
P
k
√
√
1
ma 2
)
|γ(k)|2 +( tw
a
1
ea −γ(k)
λ
1
λa +γ(k)
+ √e
.
(C31)
ea of ma /twa . When two lines of
obtain two functions, λ
e
λa and ma /twa intersect, we find the solution of these
equations.
6
7
8
9
10
11
12
13
14
15
16
17
18
F. Wang, Phys. Rev. B 82, 024419 (2010).
Y.-M. Lu and Y. Ran, arXiv:1005.4229 (unpublished); Y.M. Lu and Y. Ran, arXiv:1007.3266 (unpublished).
B. K. Clark, D. A. Abanin, and S. L. Sondhi,
arXiv:1010.3011 (unpublished).
G. Wang, M. O. Goerbig, B. Gremaud, and C. Miniatura,
arXiv:1006.4456 (unpublished).
A. Vaezi and X.-G. Wen, arXiv:1010.5744 (unpublished).
Ki-Seok Kim, Phys. Rev. Lett. 97, 136402 (2006); Ki-Seok
Kim, Phys. Rev. B 75, 245105 (2007); Ki-Seok Kim and
Mun Dae Kim, Phys. Rev. B. 81, 075121 (2010).
P. A. Lee, N. Nagaosa, and X.-G. Wen, Rev. Mod. Phys.
78, 17 (2006).
M. Hermele, Phys. Rev. B 76, 035125 (2007), and references therein.
M. Hermele, T. Senthil, and M. P. A. Fisher, Phys. Rev.
B 72, 104404 (2005).
Y. Ran and X.-G. Wen, arXiv:cond-mat/0609620v3 (unpublished).
A. Tanaka and X. Hu, Phys. Rev. Lett. 95, 036402 (2005).
C. Honerkamp, Phys. Rev. Lett. 100, 146404 (2008).
A. M. Black-Schaffer and S. Doniach, Phys. Rev. B 75,
15
19
20
21
22
23
24
25
26
27
134512 (2007).
O. Vafek, A. Melikyan, and Z. Tesanovic, Phys. Rev. B 64,
224508 (2001).
D. Xiao, W. Yao, and Q. Niu, Phys. Rev. Lett. 99, 236809
(2007); W. Yao, D. Xiao, and Q. Niu, Phys. Rev. B 77,
235406 (2008).
S. Florens and A. Georges, Phys. Rev. B 70, 035114 (2004).
H. Mosadeq, F. Shahbazi, and S. A. Jafari,
arXiv:1007.0127 (unpublished).
A. H. MacDonald, S. M. Girvin, and D. Yoshioka, Phys.
Rev. B 41, 2565 (1990).
R. Kumar, D. Kumar, and B. Kumar, Phys. Rev. B 80,
214428 (2009).
A. Banerjee, K. Damle, and A. Paramekanti,
arXiv:1012.4546 (unpublished).
M. Hermele, T. Senthil, M. P. A. Fisher, P. A. Lee, N.
Nagaosa, and X.-G. Wen, Phys. Rev. B 70, 214437 (2004).
Unfortunately, we could not find any references on numerically ”exact” calculations for thermal Hall effect. Instead,
it has been performed to evaluate electrical Hall effect
based on exact simulations28,29 .
28
29
30
31
32
33
34
35
36
37
F. F. Assaad and M. Imada, Phys. Rev. Lett. 74, 3868
(1995).
J. O. Haerter and B. S. Shastry, Phys. Rev. B 77, 045127
(2008).
S. Ryu, C. Mudry, C.-Y. Hou, and C. Chamon, Phys. Rev.
B 80, 205319 (2009).
D. L. Maslov, arXiv:cond-mat/0506035 (unpublished).
T. Senthil, A. Vishwanath, L. Balents, S. Sachdev, and
M. P. A. Fisher, Science 303, 1490 (2004); T. Senthil, L.
Balents, S. Sachdev, A. Vishwanath, and M. P.A. Fisher,
Phys. Rev. B 70, 144407 (2004).
S. Matsuura and S. Ryu, arXiv:1007.2200 (unpublished).
A. Shitade, H. Katsura, J. Kunes, X.-L. Qi, S.-C. Zhang,
and N. Nagaosa, Phys. Rev. Lett. 102, 256403 (2009).
S. Rachel and K. L. Hur, Phys. Rev. B 82, 075106 (2010).
J. He, S.-P. Kou, Y. Liang, and S. Feng, arXiv:1012.0620
(unpublished).
M. Hohenadler, T. C. Lang, F. F. Assaad, arXiv:1011.5063
(unpublished).