J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
Journal of the Physical Society of Japan
Vol. 75, No. 1, January, 2006, 014001
#2006 The Physical Society of Japan
Gauge Theory for Quantum Spin Glasses
Satoshi M ORITA, Yukiyasu O ZEKI1 and Hidetoshi NISHIMORI
Department of Physics, Tokyo Institute of Technology, Oh-okayama, Meguro-ku, Tokyo 152-8551
1
Department of Applied Physics and Chemistry, The University of Electro-Communications,
Chofugaoka, Chofu, Tokyo 182-8585
(Received August 22, 2005; accepted October 21, 2005; published December 26, 2005)
The gauge theory for random spin systems is extended to quantum spin glasses to derive a number of
exact and/or rigorous results. The transverse Ising model and the quantum gauge glass are shown to be
gauge invariant. For these models, an identity is proved that the expectation value of the gauge invariant
operator in the ferromagnetic limit is equal to the one in the classical equilibrium state on the Nishimori
line. As a result, a set of inequalities for the correlation function are proved, which restrict the location of
the ordered phase. It is also proved that there is no long-range order in the two-dimensional quantum
gauge glass in the ground state. The phase diagram for the quantum XY Mattis model is determined.
KEYWORDS: spin glass, gauge theory, correlation function, quantum spin system
DOI: 10.1143/JPSJ.75.014001
1.
Introduction
The problem of spin glasses has been attracting continued
attention.1,2) After the pioneering work by Edwards and
Anderson (EA),3) Sherrington and Kirkpatrick have exactly
solved the mean-field model by assuming the replica
symmetry.4) Parisi has proposed the replica symmetry
breaking solution and established the theoretical picture that
the low-temperature phase is composed of infinitely many
stable states with ultrametric structure.5)
A topic of active investigations in recent years concerns
the properties of short-range systems. Numerical studies
have provided strong evidence for the existence of the spin
glass (SG) transition for both the EA model6–9) and the XY
gauge glass10–13) in three dimensions but against it for the
two-dimensional EA model.8,9) For the two-dimensional
gauge glass, although the long-range SG order has been
denied rigorously,14) it is still possible that the system has a
quasi long-range order in which the SG correlation decays in
a power low. There remains the controversy about the
existence of this order: some numerical studies have
supported the absent of a finite-temperature transition13,15,16)
but some groups argue against such a conclusion.17,18)
Analytical calculations for spin glasses in finite dimensions are difficult because of randomness and frustration.
However, a method using the gauge symmetry of the system
is well-known as a powerful technique.19,20) This method
provides various rigorous results, for instance, the exact
internal energy and an upper bound for the specific heat in
the special region of the phase diagram. Another noteworthy
result is a set of inequalities for the correlation function,
which restrict the topology of the phase diagram. In this
relation, it has been suggested that the phase boundary
between the ferromagnetic (FM) and SG phases is vertical
by modifying the probability distribution.21) These results
are generalized to a wider class of systems including the
usual Ising SG and the Zq and XY gauge glasses.22)
Although the gauge theory provides us with surprising
results, its targets have so far been limited to classical spin
systems. In the present paper, we generalize this theory so
that it applies to quantum spin systems. A difficulty of this
generalization is the fact that we must define the gauge
transformation for spins without violating the commutation
rule. We circumvent this problem by using a rotational
operator on the Hilbert space as the gauge transformation.
This paper consists of six sections. In the next section, we
formulate the gauge transformation in two quantum spin
glasses, the transverse Ising model and the quantum gauge
glass (QGG), and show that these models have gauge
symmetry. In §3, we prove an identity for a gauge invariant
operator. This identity is valid even when the system
parameters depend on time following the time-dependent
Schrödinger equation. In §4, we derive a set of inequalities
for correlation functions and order parameters. These results
restrict the location of the FM phase or the Kosterlitz–
Thouless (KT) phase in the phase diagram. In §5, we extend
these inequalities to the ground state. The resulting inequalities for the order parameters show that the FM order does not
exist at zero-temperature in the two-dimensional QGG. In §6,
we consider the quantum XY Mattis model and determine its
phase diagram. The last section is devoted to summary.
2.
Gauge Transformation for Random Quantum Spin
Systems
2.1 Transverse Ising model
First, let us consider the random-bond Ising model in a
transverse field. The Hamiltonian for this model is written as
X
X
H¼
Ji j iz zj h
ix ;
ð1Þ
i
hi ji
where i is the Pauli matrix at site i. Although we treat spin1=2 systems in this paper, one can straightforwardly
generalize all the results to spin-S systems. There is no
restriction in the spatial dimensionality or lattice structure.
The exchange interaction Ji j is a quenched random variable.
One of the useful probability distributions for Ji j is the
binary distribution
PðJi j Þ ¼ p ðJi j JÞ þ ð1 pÞ ðJi j þ JÞ:
ð2Þ
It is convenient for later arguments to rewrite this distribution as
014001-1
PðJi j Þ ¼
eK p i j
;
2 cosh K p
Kp ¼
1
p
log
;
2
1 p
ð3Þ
J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
J. Phys. Soc. Jpn., Vol. 75, No. 1
S. MORITA et al.
where i j ¼ Ji j =J is the sign of the exchange interaction Ji j .
Another useful distribution is the Gaussian distribution
1
ðJi j J0 Þ2
PðJi j Þ ¼ pffiffiffiffiffiffiffiffiffiffi exp
;
ð4Þ
2J 2
2J 2
2
where J0 and J denote the average and variance, respectively.
For quantum spin systems, the classical gauge transformation, which simultaneously changes the sign of all
components, is not valid because the commutation rule
½ix ; iy ¼ 2iiz is changed to ½ix ; iy ¼ 2iiz . Thus we
define a gauge transformation for spins using a unitary
operator as
Y
U : i ! Gi G1 ;
G¼
Gi ;
8
< 1i
i
Gi ¼
: exp ix
2
i
ði ¼ þ1Þ
ði ¼ 1Þ
ð5Þ
U:
!
ðix ; i iy ; i iz Þ:
ð6Þ
ð7Þ
The transverse-field term in eq. (1) does not change by the
gauge transformation.
The Hamiltonian (1) is clearly invariant under the
successive operations of V and U: ðUVÞH ¼ H. However,
the distribution function of bond configuration is changed,
for the J Ising model, as
PðJi j Þ !
e K p i j i j
:
2 cosh K p
Thus this model can be quantized straightforwardly by
replacing the elements of the above spin vectors by the Pauli
matrices. The Hamiltonian of the QGG is therefore written
explicitly as
Xn
H ¼ J
cos !i j ix xj þ iy yj
hi ji
ð12Þ
o
y x
x y
sin !i j i j i j :
The phase factor !i j 2 ½0; 2Þ is a quenched random variable
whose probability distribution is of cosine type
Pð!i j Þ ¼
A difference of gauge transformations between quantum and
classical systems is the transformation rule of ix .
The gauge transformation for the bond variables fJi j g is
the same as in classical systems, namely
V : Ji j ! Ji j i j :
hi ji
;
where i is a classical gauge variable at site i and takes two
values 1. If i ¼ 1, iy;z ! iy;z and ix ! ix . Equivalently we can write
ðix ; iy ; iz Þ
x
composed of x and y components Si ¼ ð SSyi Þ and rotational
sin i
matrix in the XY plane, RðÞ ¼ ð cos
Þ as
sin cos
X
t
Hcl ¼ J
Si Rð!i j ÞS j :
ð11Þ
ð8Þ
It is important for the following argument that this transformed distribution is proportional to the Boltzmann factor
of a classical system. Similarly, the Gaussian distribution (4)
is changed as
!
Ji2j þ J02
1
J0
p
ffiffiffiffiffiffiffiffiffiffi
PðJi j Þ !
exp 2 Ji j i j : ð9Þ
exp
2J 2
J
2J 2
To simplify the arguments, we focus on the binary
distribution (3) and (8), hereafter. It is straightforward to
apply the same methods to the Gaussian distribution.
2.2 Quantum gauge glass
Next, we consider the quantum gauge glass (QGG).
Similarly to the transverse Ising model, the gauge transformation is defined by the rotation operator.
To properly define the quantum version of gauge glass, let
us first consider the Hamiltonian of the classical gauge glass
(CGG),
X
Hcl ¼ J
cosði j !i j Þ:
ð10Þ
hi ji
eK p cos !i j
;
2I0 ðK p Þ
a periodic Gaussian (Villain) type
rffiffiffiffiffiffiffi 1
Kp X
K p ð!i j 2nÞ2
Pð!i j Þ ¼
exp
2 n¼1
2
ð13Þ
ð14Þ
or a binary type
Pð!i j Þ ¼ p ð!i j Þ þ ð1 pÞð!i j Þ:
ð15Þ
Equation (12) is a special case of the XY model with
Dzyaloshinskii–Moriya interactions. This Hamiltonian is
written as
X
X
H ¼ J
i j
Ji j ð i j Þz ;
ð16Þ
hi ji
hi ji
where the second term is the random Dzyaloshinskii–Moriya
interaction. If we set new parameters,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 Ji j
2
2
~
Ji j ¼ J þ Ji j ; !i j ¼ tan
;
ð17Þ
J
the above Hamiltonian is rewritten as
X n
J~i j cos !i j ix xj þ iy yj
H¼
hi ji
o
sin !i j ix yj iy xj :
ð18Þ
This is equal to eq. (12) except that the interaction depends
on bonds.
In the CGG, the gauge transformation for spins is defined
by the shift of spin variables as i ! i i , where i 2
½0; 2Þ is the gauge variable. Using the same notation as in
eq. (11), this transformation is expressed as
U : Si ! Rð i ÞSi :
ð19Þ
Thus we use this definition of gauge transformation for the
QGG. Using a rotational operator on the Hilbert space, we
define
Y
i i z
U : i ! G i G1 G ¼
exp
i : ð20Þ
2
i
The transformation rule for the transposed vector t i is
defined as
This Hamiltonian can be rewritten using the spin vector
014001-2
U : t i ! t i Rð i Þ ¼ G t i G1 :
ð21Þ
J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
J. Phys. Soc. Jpn., Vol. 75, No. 1
S. MORITA et al.
The gauge transformation of random variables is the same as
in the classical case,
V : !i j ! !i j
i
j:
þ
X
hQiFz ¼
fi j g
ð22Þ
Under the gauge transformation UV, the Hamiltonian is
invariant because
X
t
ðUVÞH ¼ J
i Rð i ÞRð!i j i þ j ÞRð j Þ j
hi ji
ð23Þ
¼ H;
eK p
P
ð29Þ
i j i j
hjQji:
The last part of the right-hand side is rewritten in terms of
cl ðK p Þ as
P
X
eK p i j i j hjQji
fg
¼
X
eK p
P
i j i j
!
ð30Þ
Tr
cl ðK p ÞQ
:
fg
jÞ
:
ð24Þ
This transformed distribution is proportional to the Boltzmann factor for the CGG. If we choose the Gaussian or
binary type, the Boltzmann factor for the Villain or J
model appears, respectively.
3.
X
NB
pÞ
fg
where we used the property of rotation matrices,
Rð ÞRðÞ ¼ Rð þ Þ. The probability distribution (13) is
changed as
eK p cosð!i j i þ
Pð!i j Þ !
2I0 ðK p Þ
1
2N ð2 cosh K
Identity for Gauge Invariant Operators
The gauge symmetry of the Hamiltonian yields a useful
identity for gauge invariant operators. First, let us suppose
that the system was initially in the perfect FM state jFz i in
the transverse Ising model. This state appears in the FM
limit, T ¼ 0, p ¼ 1, h ¼ 0. The gauge transformation
operator G defined in eq. (5) changes this state as
Therefore, we obtain
hQiFz ¼
X
fi j g
P
eK p
P
i j i j
2N ð2 cosh K p ÞNB
Tr
cl ðK p ÞQ:
ð31Þ
Since Tr cl ðK p ÞQ is invariant under the transformation V,
this is identical to the right-hand side of eq. (26)
The above result can be generalized to the case that the
transverse field hðtÞ depends on time, following the classical
example.23) We consider the zero-temperature time evolution following the Schrödinger equation. Using the time
ordered product, the time evolution operator is written as
Zt
0
0
Ut ¼ T exp i Hðt Þ dt :
ð32Þ
0
GjFz i ¼ ji;
ji ¼ ji1 ji2 jiN :
ð25Þ
If i ¼ þ1, jii denotes the state with up spin in the z
direction, and if i ¼ 1, the spin at site i is down.
Using the property of jFz i in (25), we prove the following
identity for a gauge-invariant operator Q which satisfies Q ¼
ðUVÞQ or equivalently VQ ¼ G1 QG,
i
h
hQiFz ¼ hQi cl ðK p Þ ;
ð26Þ
where hi cl ðK p Þ is the expectation value for the classical
equilibrium state on the Nishimori line (NL), that is,
hQi
cl ðK p Þ
cl ðK p Þ
¼ Tr
¼
e
cl ðK p ÞQ;
Kp
Tr e
P
Kp
z z
ij ij i j
P
z z
ij ij i j
ð27Þ
:
To prove the identity (26), we apply the gauge transformation for the configuration of randomness of eq. (7)
appearing on left-hand side of eq. (26). This operation does
not change its value because the transformation V of eq. (7)
only changes the order of the summation over i j . Thus the
left-hand side of eq. (26) is rewritten as
P
X eK p i j i j
hQ iF z ¼
NB hFz jVQjFz i
fi j g ð2 cosh K p Þ
ð28Þ
P
X eK p i j i j
¼
NB hjQji;
fi j g ð2 cosh K p Þ
where we used the assumption that the operator Q is gauge
invariant, VQ ¼ G1 QG. Since the expectation value on the
left-hand side does not depend on , the summation over
and division by 2N yield
Since the time dependence of the Hamiltonian does not
invalidate the gauge symmetry, this operator is also gauge
invariant
ðUVÞUt ¼ GðVUt ÞG1 ¼ Ut :
ð33Þ
Examples of the gauge invariant operator include the
transverse magnetization ix ðtÞ ¼ Uty ix Ut , the autocorrelation function iz ð0Þiz ðtÞ and the interaction term H0 ðtÞ of the
Hamiltonian (the first term on the right-hand side of eq. (1)).
For the QGG, we can prove a similar identity
i
h
hQiFx ¼ hQi cl ðK p Þ :
ð34Þ
Here one should note that cl ðK p Þ is different from the
normal density operator. If we choose the cosine-type
distribution (13), cl ðK p Þ is defined by the Boltzmann factor
for the CGG as
P
K
cosð!i j i þ j Þ
Tr e p i j
j ih j
P
ðK
Þ
¼
;
ð35Þ
cl
p
K
cosð!i j i þ j Þ
Tr e p i j
where j i ¼ GjFx i and Tr stands for integration over i
from 0 to 2. Since the state vector j i does not diagonalize
the Hamiltonian for the quantum gauge glass, cl ðK p Þ is not
equal to the density operator
ðK p Þ ¼
e H
:
Tr e H
ð36Þ
The identities (26) and (34) show that the expectation
value of gauge invariant operator in the FM limit is equal to
the one in the classical equilibrium state on the NL. The
equivalence of the two states has already been proved in the
dynamical gauge theory for classical systems.23,24) The
014001-3
J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
J. Phys. Soc. Jpn., Vol. 75, No. 1
S. MORITA et al.
present results are generalization of these dynamical cases to
quantum systems. Note that the zero-temperature time
evolution for quantum systems is deterministic in contrast
to the stochastic dynamics for the classical SG.
4.
NL
p = pc
T
PM
Tc
Correlation Function and Order Parameter
Using the identities proved in the previous section, we can
derive a class of inequalities for the correlation function.
First, we treat the transverse Ising model. Since the
correlation function is not invariant under the gauge transformation, let us consider the following gauge-invariant
quantity
Q ¼ iz zj hiz zj iK;h ;
K;h
K;h
Here hi j icl
K p is the correlation function for the classical
Ising system with the same configuration fi j g and no
external field. By taking the absolute value of both sides of
this equation, we find
h
D
i
z zE
i j cl ;
ð39Þ
i j
Kp
K;h
where we used the fact that correlation function hiz zj iK;h
does not exceed unity. Similarly, we can prove
D
i
E h
hi j icl
sgn z z
ð40Þ
Kp
i j
K;h
3 2
cl
h
i
6 D i j EK p 7 6 D 1 E
4
5¼4
iz zj
iz zj
K;h
3
7
5 1:
ð41Þ
FM
SG
p
FM
h
Fig. 1. The phase diagram of the transverse Ising model. The paramagnetic (PM), the ferromagnetic (FM) and spin glass (SG) phases meet
at the multicritical point (MCP) in the plane h ¼ 0. The dashed line at
p ¼ pc sets a bound for the existence of the FM phase also for h 6¼ 0.
U
ð42Þ
¼ t Si S j i t Si Rð=2ÞS j :
¼ ðix xj þ iy yj Þ iðix yj iy xj Þ:
ij
¼G
i jG
ij
for the
ð43Þ
The gauge transformation U changes this operator according
to
1
¼ eið
i
jÞ
i j:
ð44Þ
The first factor is the same as the correlation function
appearing in the classical gauge glass except for a minus
sign.
Usually, the correlation function for XY-like systems is
defined as the expectation value of ix xj þ iy yj , which
corresponds to cosði j Þ for classical systems. However,
this expression is not useful for the gauge theory because it
does not separate into gauge variables and spin operators
after transformation by U as
Uðix xj þ iy yj Þ ¼ cosð
eiði j Þ ¼ Si S j i ðSi S j Þz
This motivates us to define a correlation operator
quantum gauge glass as,
hc
PM
K;h
If the probability p of the FM interaction is less than the
critical probability pc at the multicritical point for the
classical Ising system, the right-hand
side of the inequality
(39) vanishes in the limit i j ! 1. Thus the correlation
function for the transverse Ising model on the left-hand side
is also equal to zero. Therefore the region of the FM phase is
restricted to the range p > pc (Fig. 1). Since the transverse
field represents quantum fluctuations, the correlation function for the transverse Ising model should be reduced from
classical system with h ¼ 0, which is the physical origin of
the above-mentioned restriction.
Next, let us consider how to define a correlation function
of the QGG which has convenient properties for the gauge
theory. In the CGG, a useful correlation function is defined
in terms of an exponential function as eiði j Þ . This is
rewritten using the notation of eq. (12) as
ij
PM
ð37Þ
where hiK;h denotes thermal average with temperature
1
¼ J=K under a transverse field h. Substitution of the
above quantity into eq. (26) yields
D
E
D
E
cl
iz zj
¼ i j K p iz zj
:
ð38Þ
2
MCP
i
sinð
i
x x
j Þði j
þ iy yj Þ
x y
j Þði j
iy xj Þ:
In addition, one can easily prove that
E i
hD x x
i j þ iy yj
ij K ¼
ð45Þ
ð46Þ
K
since the QGG is invariant under the following transformation:
ðix ; iy ; iz Þ ! ðiy ; ix ; iz Þ;
!i j ! !i j :
ð47Þ
Thus we choose to discuss the correlation operator (43).
To derive an inequality similar to eq. (39), it is useful to
consider
Q¼
y
i j h i j iK :
ð48Þ
It is easy to prove that this quantity is gauge invariant
because of the property of i j given in eq. (44). Substituting
this quantity into eq. (34), we obtain
i
h ið i j Þ cl
;
ð49Þ
ij K ¼ e
i
j
Kp
K
where hicl
K p stands for the expectation value for the CGG on
the NL. By taking the absolute value of both sides of this
equation, we find
i
h
cl
ð50Þ
i j K eið i j Þ K p :
Similarly, we obtain
014001-4
J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
J. Phys. Soc. Jpn., Vol. 75, No. 1
sgn
2
4
eið
ij K
i
jÞ
S. MORITA et al.
h ið
e
cl
Kp
3
"
i
5¼
i
cl
Kp
;
1:
! ei
i
cl 2
Kp
of inequality (50) is estimated as
i
e
j
i
jÞ
cl 2
Kp
cl 2
Kp
i j ! 1
ðVHÞ G1 jx; !i ¼ x G1 jx; !i:
ð52Þ
ij K
ið
e
ð54Þ
Vjx; !i ¼ G1 jx; !i:
mðK; K p Þ2 mcl ðK p ; K p Þ:
ð55Þ
If the parameter K p is smaller than the critical point K clpc for
the classical system, the right-hand side vanishes. Consequently, the FM phase for the quantum gauge glass lies in
the region satisfying K p > K clpc . This result is consistent with
the intuitive picture that quantum effects reduce long-range
order.
If the spatial dimensionality of the system is equal to the
lower critical dimension, d ¼ 2, there is no long-range order
but quasi long-range order. The Kosterlitz–Thouless (KT)
phase25) exists when both K and K p are sufficiently large.
The ordering tendency of the KT phase is observed by the
correlation length in the paramagnetic (PM) phase as
ji jj=m ðK;K p Þ
;
ð56Þ
ij K e
2
ið i j Þ cl
e
Kp
h
i
cl
cl
¼ eið i j Þ K p eji jj=m ðK p ;K p Þ :
ð57Þ
From the square of the inequality (50), the limit i j !
1 yields
2cl
m ðK p ; K p Þ:
ð58Þ
Thus, the KT phase for the QGG is also restricted to the
region K p > K clpc .
5.
Ground State Property of Quantum Gauge Glass
Next we consider the ground state property of the QGG. It
is necessary to consider the transformation rule for eigenstates of the Hamiltonian. Let us denote an eigenstate of the
Hamiltonian with the eigenvalue x by jx; !i,
Hjx; !i ¼ x jx; !i:
ð59Þ
Since the invariance of the Hamiltonian (23) can be
rewritten as
ð62Þ
This rule is derived straightforwardly, if the state is not
degenerate. When the state is degenerate, the rule is not
unique. However, it is reasonable and has no problem, if we
use this rule as the transformation rule.
Now, we consider the ground sate. Let us denote the
ground state of the Hamiltonian by jg; !i. The above
transformation rule leads to the gauge transformation of the
average of any operator Q in the ground state;
Vhg; !jQjg; !i ¼ hg; !jGðVQÞG1 jg; !i:
where we used the identity mcl ¼ qcl resulting from the
gauge theory on the NL for the CGG. Therefore we obtain
ð61Þ
Note that the effect of the operator V is restricted to the
inside of the brackets ð Þ. The state G1 jx; !i is an
eigenstate of the gauge-transformed Hamiltonian ðVHÞ.
Therefore, one can derive the transformation rule of the
eigenstate
¼ qcl ðK p ; K p Þ2 ¼ mcl ðK p ; K p Þ2 ;
m ðK; K p Þ
ð60Þ
we have an eigen-value equation
Since the lower critical dimension dl is two for continuous
spin systems, for d > 2, we expect a FM phase to exist at
low temperature under small randomness. In this case, a
two-spin correlation function tends to the square of magnetization when the two spins are sufficiently separated,
2
i j ! 1:
ð53Þ
i j K ! mðK; K p Þ ;
The right-hand side
follows:
i
h
ið i j Þ cl 2
e
Kp
GðVHÞG1 ¼ H;
ð51Þ
#
1
ij K
jÞ
ð63Þ
The thermal average h iK in equations derived in the
previous section can be replaced by the ground state
expectation value. For example,
i
h
cl
g; !j i j jg; ! ¼ eið i j Þ K p g; !j i j jg; !
ð64Þ
is derived instead of eq. (49), which provides the inequality
for the order parameters in the ground state,
mð1; K p Þ2 mcl ðK p ; K p Þ
ð65Þ
instead of eq. (55), and
m ð1; K p Þ 2cl
m ðK p ; K p Þ
ð66Þ
instead of eq. (58).
In two dimensions, it has been shown that the FM long
range order exists in the ground state of the pure quantum
XY model (K p ¼ þ1).26) However, in the disordered
regime (K p < þ1), the FM order must disappear since
the FM order in the CGG model, the right-hand side of the
inequality (65), does not exist. The only possibility in this
regime is the existence of the KT phase, which is consistent
with the inequality (66).
6.
Phase Diagram of Quantum Mattis Model
In this section, we introduce and discuss the properties of
the quantum XY Mattis model which has no frustration.
Using the gauge transformation, one can obtain the phase
diagram for this model explicitly. One of the phase
boundaries is determined by the critical point of the pure
quantum system and the other by that of the classical one.
This is an important difference from the classical nonfrustrated systems.27)
Let us locate a quenched random variable !i at each site
and define the phase factor !i j as !i j ¼ ! j !i . The
Hamiltonian for the quantum Mattis model is defined in
terms of the pure quantum XY model, H0 , that is,
Y
i!i z
H ¼ G! H0 G1
;
G
¼
exp
: ð67Þ
!
!
2 i
i
014001-5
J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
J. Phys. Soc. Jpn., Vol. 75, No. 1
S. MORITA et al.
Thus, the ground-state energy is always the same as that of
H0 and the ground state is obtained by operating G! on the
ground state of the pure system. This system has no
frustration in this sense. It is easy to show that the
Hamiltonian for this model is invariant under the gauge
transformation, ðUVÞH ¼ H, where the gauge transformation for the configuration is defined as !i ! !i i .
Using eqs. (44) and (67), we immediately obtain
ij K¼
eið!i ! j Þ
i j 0;K ;
ð68Þ
where the angular brackets on the right-hand side denote the
thermal average with respect to H0 . Note that the correlation
function for the pure system h i j i0;K does not depend the
quenched variable !i . Here, we assume that the distribution
function for the quenched random variable is proportional to
the Boltzmann factor of the pure classical XY model,
P
K
cosð!i ! j Þ
e p ij
Pð!Þ ¼
:
ð69Þ
Z0cl ðK p Þ
Hereafter, the subscript 0 and the superscript cl stand for
pure and classical systems, respectively. Then the configuration average is equal to the thermal average for the pure
classical XY model with coupling K p . Consequently, we find
ið!i ! j Þ cl
:
ð70Þ
ij K ¼ e
0;K p i j 0;K
Similarly, from eq. (68), the spin-glass correlation function
satisfies
2
2
ð71Þ
i j K ¼ i j 0;K :
Taking the limit i j ! 1 of eqs. (70) and (71), we
obtain
mðK; K p Þ ¼ mcl
ð72Þ
0 ðK p Þm0 ðKÞ;
qðK; K p Þ ¼ m0 ðKÞ2 :
ð73Þ
For d > 2, there are three phases, PM, FM and Mattis spinglass (MSG) phases. Figure 2(a) shows the phase diagram.
The phase boundary between PM and other phases is at K ¼
cl
K0c and the one between FM and MSG is at K ¼ K0c
.
If d ¼ 2, the correlation length determines the phase
structure. From eqs. (70) and (71), we find
1
1
1
¼
þ
m ðK; K p Þ cl
ðK
Þ
ðKÞ
p
0
0
a)
ð74Þ
b)
0
0
K
PM
K
PM
K0c
MSG
0
Kp
cl
K0c
K0c
FM
RKT
∞
∞
0
Kp
cl
K0c
UKT
∞
∞
q ðK; K p Þ ¼
0 ðKÞ
;
2
ð75Þ
where q ðK; K p Þ denotes the spin-glass correlation length.
Thus, similarly to the d > 2 case, three phases exist: (i) m <
1 and q < 1: paramagnetic phase (PM), (ii) m ¼ 1 and
q ¼ 1: uniform KT phase (UKT) and (iii) m < 1 and
q ¼ 1: random KT phase (RKT). The phase diagram is
shown in Fig. 2(b).
We note again that the location of the horizontal phase
boundary is determined by the critical point of the quantum
pure system, K0c , and the vertical one comes from that of the
cl
classical pure system, K0c
.
7.
Conclusions
In this paper, we have investigated quantum spin glasses
using the gauge theory. First, we considered the transverse
Ising model and the QGG. To construct the gauge theory, we
defined the gauge transformation by rotational operator on
the Hilbert space. It is essential that interaction of these
models is written in term of one or two components of spin
operators. If a system has Heisenberg-type interactions, we
can not define a gauge transformation which satisfies the
commutation rule.
Using the gauge theory, we obtained mainly two results.
One is the identity for gauge invariant operators. The FM
limit state and the classical equilibrium state on the NL
provide the same expectation value for gauge invariant
operators. We note that this result remains valid when we
introduce time evolution following the Schrödinger equation. This result has already been pointed out for classical
spin glasses with stochastic dynamics.23,24) We have shown
that the same applies to quantum spin systems.
The other result is a set of inequalities for the correlation
function. These inequalities show that the correlation
function for the quantum model never exceeds the classical
counterpart on the NL. The corresponding classical system is
determined by a transformation rule for probability distribution. As a result, the order parameter is smaller than the
square of the classical one. Therefore the FM phase (or the
KT phase) should lie within the corresponding classical one.
This is natural intuitively since quantum effects reduce
ordering tendency, but to prove it rigorously is a different
and quite a non-trivial problem. Moreover, these results are
valid even if the system is in the ground state. Thus FM long
range order vanishes in the two-dimensional QGG although
the ground state of the pure quantum XY model has FM
order.
Next, we determined the phase diagram for the quantum
XY Mattis model. This model is not a real SG because of
lack of frustration. It is interesting, nevertheless, that both
quantum and classical phase transitions occur in a single
system, which may serve as a starting point for investigations of more realistic quantum spin glasses with frustration.
Acknowledgment
Fig. 2. (a) The phase diagram for the quantum XY Mattis model in d > 2.
There are three kinds of phases, the PM, the FM and the SG. (b) The
phase diagram in d ¼ 2. There exist three kinds of phases, the PM, the
uniform KT (UKT) and the random KT (RKT).
This work was supported by the Grant-in-Aid for
Scientific Research on Priority Area ‘‘Statistical-Mechanical
Approach to Probabilistic Information Processing’’ by the
Ministry of Education, Culture, Sports, Science and Technology.
014001-6
J. Phys. Soc. Jpn.
Downloaded from journals.jps.jp by 54.167.13.156 on 09/16/22
J. Phys. Soc. Jpn., Vol. 75, No. 1
S. MORITA et al.
1) K. Binder and A. P. Young: Rev. Mod. Phys. 58 (1986) 801.
2) K. H. Fischer and J. A. Hertz: Spin Glasses (Cambridge University
Press, Cambridge, 1991).
3) S. F. Edwards and P. W. Anderson: J. Phys. F 5 (1975) 965.
4) D. Sherrington and S. Kirkpatrick: Phys. Rev. Lett. 35 (1975) 1792.
5) G. Parisi: Phys. Lett. A 73 (1979) 203; G. Parisi: J. Phys. A 13 (1980);
G. Parisi: J. Phys. A (1980) 13 1101; G. Parisi: J. Phys. A 13 (1980)
1887; G. Parisi: J. Phys. A 13 (1980) L115.
6) R. N. Bhatt and A. P. Young: Phys. Rev. Lett. 54 (1985) 924.
7) A. T. Ogielski and I. Morgenstern: Phys. Rev. Lett. 54 (1985) 928.
8) R. R. P. Singh and S. Chakravarty: Phys. Rev. Lett. 57 (1986) 245.
9) R. N. Bhatt and A. P. Young: Phys. Rev. B 37 (1988) 5606.
10) D. A. Huse and H. S. Seung: Phys. Rev. B 42 (1990) 1059.
11) J. D. Reger, T. A. Tokuyasu, A. P. Young and M. P. A. Fisher: Phys.
Rev. B 44 (1991) 7147.
12) M. Cieplak, J. R. Banavar and A. Khaurana: J. Phys. A 24 (1991)
L145.
13) J. M. Kosterlitz and M. V. Simkin: Phys. Rev. Lett. 79 (1997) 1098.
14)
15)
16)
17)
18)
19)
20)
21)
22)
23)
24)
25)
26)
27)
014001-7
H. Nishimori and H. Kawamura: J. Phys. Soc. Jpn. 62 (1993) 3266.
M. J. P. Gingras: Phys. Rev. B 45 (1992) 7547.
J. D. Reger and A. P. Young: J. Phys. A 26 (1993) L1067.
Y.-H. Lie: Phys. Rev. Lett. 69 (1992) 1819.
M. Y. Choi and S. Y. Park: Phys. Rev. B 60 (1999) 4070.
H. Nishimori: Prog. Theor. Phys. 66 (1981) 1169.
H. Nishimori: Statistical Physics of Spin Glasses and Information
Processing: An Introduction (Oxford University Press, Oxford, 2001).
H. Kitatani: J. Phys. Soc. Jpn. 61 (1992) 4049.
Y. Ozeki and H. Nishimori: J. Phys. A 26 (1993) 3399.
Y. Ozeki: J. Phys. A 28 (1995) 3645; Y. Ozeki: J. Phys.: Condens.
Matter 9 (1997) 11171.
Y. Ozeki: J. Phys. A 36 (2003) 2673.
J. M. Kosterlitz and D. J. Thouless: J. Phys. C 6 (1973) 1181.
T. Kennedy, E. H. Lieb and B. S. Shastry: Phys. Rev. Lett. 61 (1988)
2582.
Y. Ozeki: J. Phys. A 29 (1996) 5805.