You are on page 1of 4

VOLUME 82, NUMBER 18 PHYSICAL REVIEW LETTERS 3 MAY 1999

Elementary Excitations of the Symmetric Spin-Orbital Model: The XY Limit


Frédéric Mila,1 Beat Frischmuth,2 Andreas Deppeler,2 and Matthias Troyer3
1
Laboratoire de Physique Quantique, Université Paul Sabatier, 118 route de Narbonne,
F-31062 Toulouse Cedex, France
2
Institute of Theoretical Physics, ETH Hönggerberg, CH-8093 Zürich, Switzerland
3
Institute for Solid State Physics, University of Tokyo, Roppongi 7-22-1, Tokyo, 106, Japan
(Received 4 August 1998)
The elementary excitations of the 1D, symmetric, spin-orbital model are studied in two anisotropic
cases: the pure XY and the dimerized XXZ models. While they preserve the symmetry between spin
and orbital degrees of freedom, these models allow for a picture of the low-lying excitations. In
the pure XY case, a phase separation takes place between two phases with free-fermion-like gapless
excitations. In the dimerized case the low-energy effective Hamiltonian reduces to the 1D Ising model
with gapped excitations. All the elementary excitations involve simultaneous flips of the spin and
orbital degrees of freedom: a clear indication of the breakdown of the traditional mean-field theory.
[S0031-9007(99)08987-5]

PACS numbers: 75.10. – b, 71.27. + a, 75.30.Et

The impact of orbital degeneracy on the low-energy complexity comes from the large local degeneracy: For a
properties of Mott-Hubbard insulators is currently attract- single bond, the ground state is sixfold degenerate (spin
ing a lot of attention following the progress in synthesiz- singlet 3 any of the three orbital triplets or any of the spin
ing and studying materials with these characteristics [1]. triplets 3 orbital singlet). It is thus interesting to study the
It was already pointed out a long time ago by Kugel and XXZ version of the model defined by
Khomskii that such systems should have low-lying orbital X µ y y l

excitations in addition to spin excitations [2]. More re- H ­ Ji 2sSix Si11
x
1 Si Si11 1 lSiz Si11 z
d1
i 2
cently, it has been suggested that the interplay between µ ∂
both degrees of freedom can have more dramatic conse- y y l
3 2stix ti11
x
1 ti ti11 1 ltiz ti11 z
d1 . (2)
quences. For instance, under suitable conditions the or- 2
bital degeneracy can enhance quantum fluctuations in the In that case, the degeneracy is lifted within the triplet sector
spin degrees of freedom and lead to gapped spin excita- as soon as l , 1, and the ground state of a given bond is
tions even in the 3D case [3]. Another interesting situation only twofold degenerate (spin singlet 3 orbital triplet with
is the SU(4) symmetric case [4,5] where the system can- z
ttot ­ 0 or spin triplet with Stot z
­ 0 3 orbital singlet).
not choose locally between the configurations (spin singlet The essential ingredient, namely, the symmetry between
3 orbital triplet) and (spin triplet 3 orbital singlet). Then spin and orbital degrees of freedom, is preserved, but
the mean-field approach that decouples spin and orbital de- the Hilbert space of the low-lying sector is considerably
grees of freedom on each bond cannot be a good starting reduced.
point in that case since it violates basic SU(4) relationships In the following, we will concentrate on two versions
between correlation functions on a given bond, as empha- of this model for which a transparent picture of the low-
sized in Ref. [6]. As a consequence, the traditional picture lying excitations can be obtained.
of relatively independent spin and orbital excitations must The pure XY model.—It corresponds to the previous
be abandoned. A clear picture of the low-lying excitations Hamiltonian [Eq. (2)] with l ­ 0 and Ji ­ J for all
in such a case is still lacking though. bonds, which can be written more compactly as
In this Letter, we concentrate on the symmetric case. X
The basic model is the SU(4) symmetric Hamiltonian given H ­ J sSi1 Si11 2
1 Si2 Si11
1
d sti1 ti11
2
1 ti2 ti11
1
d , (3)
i
by
Xµ 1
∂µ
1
∂ or in expanded form,
X
H ­J 2S$ i ? S$ i11 1 2t$ i ? t$ i11 1 , (1) H ­ J sSi1 ti1 Si11
2 2
ti11 1 Si2 ti2 Si11
1 1
ti11 d
i 2 2
i
X
where S$ i and t$ i are spin-1y2 operators corresponding to 1 J sSi1 ti2 Si11 d.
2 1
ti11 1 Si2 ti1 Si11
1 2
ti11 (4)
spin and orbital degrees of freedom, respectively. This i
model has already been studied rather extensively by Analyzing this Hamiltonian in the product basis ≠i jhli ,
1 1
several methods [4–9]. In particular, it is known from where jhli ­ jSiz ­ 6 2 , tiz ­ 6 2 l and denoting
the Bethe ansatz solution that there are three branches of
ja6li ­ jSiz ­ 61y2, tiz ­ 61y2l ,
low-energy excitations [7]. The physical interpretation of (5)
these branches is not straightforward though. The essential jb6li ­ jSiz ­ 61y2, tiz ­ 71y2l ,

0031-9007y99y82(18)y3697(4)$15.00 © 1999 The American Physical Society 3697


VOLUME 82, NUMBER 18 PHYSICAL REVIEW LETTERS 3 MAY 1999

one can easily show that all the matrix elements of


H between states hja6li j and hjb6lj j vanish. As a
consequence, the eigenstates can be classified according to
the sequence of domains of phase A (with parallel orbital
and spin states) and phase B (with antiparallel states) that
include only states of type ja6li and jb6li , respectively,
and Eq. (4) can be written as
X 1
H ­ 2J sai1 ai11
2
1 ai2 ai11
1
d
i 2
X 1
1 2J sbi1 bi11
2
1 bi2 bi11
1
d, (6)
i 2
where ai6 ­ Si6 ti6 (bi6 ­ Si6 ti7 ) are the raising and
lowering spin-1y2 operators corresponding to ja6li
(jb6li ). Now the Hamiltonian within a phase can be
mapped onto spinless fermions with a Jordan-Wigner FIG. 1. Temperature dependence of the entropy of the SU(4)
symmetric spin-orbital model of Eq. (1) (dashed line, taken
transformation, and the ground state energy of a domain from Ref. [6]) and of the pure XY spin-orbital model of
of length L is given by ≥ ¥ Eq. (3) (solid line) for chains of length L ­ 200 with periodic
µ ∂ sLy2dp
boundary conditions in two different temperature scales. For
sLy2 1 1dp sin 2sL11d
EsLd ­ 22J cos ≥ ¥ . (7) comparison, the analytical result for the entropy of the XY
2sL 1 1d p
sin 2sL11d
model with coupling 2J and infinite length is also shown
(dotted line). The statistical errors are always less than 0.0006.
It is easy to check that EsLd , EsL1 d 1 · · · 1 EsLn d,
where L1 , . . . , Ln are integers such that L1 1 · · · 1 Ln ­
L. So the model of Eq. (6) undergoes a phase separation, hand, coincides with the linear behavior of the XY model
and the ground state is twofold degenerate with a single only up to a crossover temperature T p ø 0.05J. Above
domain of phase A or B. T p a sharp increase of the entropy takes place (see inset of
The phase separation has some drastic consequences for Fig. 1). This increase comes from the additional entropy
the thermodynamics, and the low-temperature properties contribution of the domain walls which are more and more
of the pure XY spin-orbital model turn out to be signifi- frequent with increasing temperature. This can be seen
cantly different from those of the SU(4) symmetric model in Fig. 2 where the average density of domain walls p
[Eq. (1)]. In Fig. 1, we show the entropies s per site as a (average number of domain walls per site) is depicted as a
function of temperature for both models, the SU(4) sym- function of temperature. Up to T p , the spin-orbital model
metric model [Eq. (1)] and the pure XY case [Eq. (3)]. is in one of the phases A or B and p ­ 0 within the
They have been calculated numerically for chains of statistical errors of the Monte Carlo simulation. Above
length L ­ 200 with periodic boundary conditions, using T p the number of phase sectors increases very sharply
the continuous time quantum Monte Carlo loop algorithm with increasing T .
[10,11]. Using the relation cy ­ T ≠Sy≠T , the entropy This behavior can be easily understood in terms of the
was determined by integrating the measured internal ener- following, approximate free energy per site:
gies UsT d: Z 1yT fsT , pd ­ pEDW 2 TsDW sT , pd 2 TsD sT d . (9)
SsT d ­ Ss`d 1 UsT dyT 2 Us1ybd db . (8)
0 The various quantities entering this expression are as
At a very low temperature, both entropies show a linear follows: (i) p, the concentration of domain walls, to be
behavior (see inset of Fig. 1), indicating the presence determined by minimizing the free energy; (ii) EDW , the
of gapless excitations. But the slope for the SU(4) is energy of a domain wall. This is the energy required to
much larger than that in the pure XY spin-orbital case. split a finite chain of length L into two chains of length
In fact, in the first case the slope is 3 times bigger L 2 L1 and L1 , i.e., EsL 2 L1 d 1 EsL1 d 2 EsLd, where
than that of a single SU(2) antiferromagnetic Heisenberg EsLd is defined in Eq. (7). EDW is a priori a function of
chain [6], while the slope of the entropy at low T in L and L1 . It turns out that, for large enough L, EDW .
the pure XY spin-orbital model is equal to that of the 0.36J regardless of L and L1 except for L1 , L0 . 20.
XY model with coupling 2J, as expected from Eq. (6). Since L0 does not depend on L, this difference will play
A further difference is visible as the temperature is no role in the thermodynamic limit at low temperatures,
increased. The entropy of the SU(4) symmetric spin- and one can safely assume EDW ­ 0.36J; (iii) sDW sT , pd,
orbital model remains approximately linear also in the the entropy of the domain walls. For small p, it is
intermediate temperature range (up to T ø 0.2J). The given by sDW sT , pd ­ 2p ln p; (iv) sD sT d, the entropy
entropy of the pure XY spin-orbital model, on the other contribution of the domains. When p is small, finite-size

3698
VOLUME 82, NUMBER 18 PHYSICAL REVIEW LETTERS 3 MAY 1999

again in very good agreement with 2EDW . Note that


the temperature T0 below which finite-size effects due
to periodic boundary conditions start to influence the
thermodynamics is given by exps2EDW yT0 d ­ 1yL, i.e.,
T0 . EDW y ln L. Since it vanishes when the system size
goes to infinity, the free energy of Eq. (9) is expected to
be valid down to zero temperature in the thermodynamic
limit.
To summarize this section, the very low tempera-
ture excitations correspond to simultaneous flips of spin
and orbital degrees of freedom within one domain, and
domain-wall excitations corresponding to collective ex-
citations involving spins and orbital degrees of freedom
play an important role above a crossover temperature
T p . 0.05J.
FIG. 2. Temperature dependence of the average density p of The dimerized XXZ model.—It corresponds to the
domain walls in the pure XY spin-orbital model. The error Hamiltonian of Eq. (2) with Ji ­ J if i is even, and
bars are smaller than the symbols. The inset shows lnp as Ji ­ aJ if i is odd. We wish to study that model in the
function of the inverse temperature bJ as well as fits of the limit a ø 1. Let us start by introducing some notations.
form p ­ exps2E0 yTd and p ­ sL 2 1d exps2E1 yT d in the The Hilbert space of a given bond is spanned by the
intermediate and very low temperature ranges (for details, see
text). 16 states jSSl, jSTi l (i ­ 21, 0, 1), jTi Sl (i ­ 21, 0, 1),
and jTi Tj l (i, j ­ 21, 0, 1), where the first (second) letter
refers to the spins (orbitals), while jSl and jTi l are pthe
effects are negligible, and sD sT d is equal to the entropy of usual singlet and triplets given by p jSl ­ sj"#l 2 j#"ldy 2,
the XY model with coupling 2J, i.e., spy6d sT yJd at low jT1 l ­ j""l, jT0 l ­ sj"#l 1 j#"ldy 2, and jT21 l ­ j##l. In
temperature. the pure Heisenberg case, the six states jSTi l and jTi Sl
As for the 1D-Ising model, minimizing with re- (i ­ 21, 0, 1) are degenerate ground states. However this
spect to the density of domain walls p leads to p ­ degeneracy is partially lifted if l , 1 in Eq.(2), and the
exps2EDW yT d and sDW sT d ­ sEDW yT d exps2EDW yT d. ground state of a given bond is only twofold degenerate
The total entropy, which is the sum of sDW sT d and (jST0 l and jT0 Sl). If a ­ 0, the ground state of Eq.(2)
sD sT d, is then dominated at low temperature by sD sT d . is then 2Ly2 -fold degenerate, where L is the number of
spy6d sT yJd, while the domain-wall contribution takes sites, since each dimer si, i 1 1d, i even, can be in any
over at higher temperature. To be more quantitative, of the two states jST0 li or jT0 Sli . Let us study how
let us define the temperature T1 where both contribu- this degeneracy is lifted when a is switched on. Since
tions are equal. It is given by sEDW yT1 d exps2EDW y we have a two-level system on each dimer si, i 1 1d,
T1 d ­ spy6d sT1 yJd, which leads to T1 ­ 0.074J. This i even, we can define a pseudo-spin-1y2 operator s$ i
is in very good agreement with the numerical results of that acts on this dimer with the identification jST0 l ; j#l
Fig. 1. and jT0 Sl ; j"l. An effective Hamiltonian can then be
The prediction for the density of domain walls is also in derived using a standard many-body perturbation theory.
very good agreement with the numerical simulations (see The result depends on l. If l . 0, then the perturbation
inset of Fig. 2, dashed line) for not too low temperatures: is lifted to first order in a, and the effective Hamiltonian
A fit with an exponential law p ­ exps2E0 yT d for 8 # reads
b # 12 gives E0 ­ 0.36s1d, in very good agreement with X µ ∂
1
the domain-wall energy EDW . For very low temperatures, l.0
Heff ­ al2 six si12
x
1 , (10)
namely, for temperatures where the average number of i-even 4
domain walls is of order 1 or smaller, and for finite
while if l ­ 0 one has to go to second order perturbation
systems, the above picture cannot work because the
theory to lift the degeneracy, and the effective Hamilton-
numerical simulations were performed using periodic
ian reads
boundary conditions, and domain walls can be created
only by pairs with a minimum energy 2EDW . Neglecting X µ 1

configurations with more than one pair of domain walls, Heff ­ 2a
l­0 2 x x
si si12 1 . (11)
i-even 4
one can show that the concentration is expected to
behave such as p . sL 2 1d exps22EDW yT d. A fit of Several remarks can be made about these results: First,
the very low temperature numerical data could indeed the effective Hamiltonian is always an Ising model
be performed with the law p ­ sL 2 1d exps2E1 yT d for to the first non-vanishing order in perturbation theory.
15 # b # 25 (see inset of Fig. 2) with E1 ­ 0.72s2d, Second, the effective Ising model is antiferromagnetic if

3699
VOLUME 82, NUMBER 18 PHYSICAL REVIEW LETTERS 3 MAY 1999

l . 0 and ferromagnetic if l ­ 0. These models are, S a t b which can be seen as generators of the SU(4)
of course, related by a simple transformation, but we algebra (see Ref. [5]), and it is likely that at least part
expect to have a transition line between these cases in of the low-lying modes of the SU(4) symmetric model
the (a, l) plane along which the effective Hamiltonian will be predominantly built on these generators and will
presumably takes a more complicated form. Finally, and retain the mixed character observed here. Besides, the
more importantly, we obtain an Ising model in terms of fact that the correlation functions kS$ i ? S$ i11 l, kt$ i ? t$ i11 l,
six ­ ssi1 1 si2 dy2, not siz . So the eigenstates must and ksS$ i ? S$ i11 d st$ i ? t$ i11 dl are all negative on a given
x
be written in termsp of the eigenstates ofp si , namely, bond appears in the XXZ case as a direct consequence of
sjST0 l 1 jT0 Sldy 2 and sjST0Q l 2 jT0 Sldy 2. They are the local degeneracy between the states (spin singlet 3
thus of the general form 22Ly4 i-even sjST0 li 6 jT0 Sli d. orbital triplet) and (spin triplet 3 orbital singlet). So the
Let us now briefly discuss the low-energy properties. picture that emerges is that the symmetry between spin
Quite generally, we expect to have a twofold degenerate and orbital degrees of freedom has dramatic consequences
ground state, and a gapped excitation spectrum. More on the low-lying excitations: The system is not able to
specifically, the ground states are given by choose between spin or orbital singlets or triplets, and
Y the excitations are an intricate mixture of spin and orbital
jGS6l ­ 22Ly4 sjST0 li 6 jT0 Sli d , (12) degrees of freedom.
i-even
To complete the picture, it will be useful to study the
in the ferromagnetic case, and by XXZ model in the whole parameter range 0 # a, l # 1.
Y The main issues are as follows: (i) the evolution of the
jGS6l ­ 22Ly4 sjST0 li 6 s21diy2 jT0 Sli d , (13)
i-even
spectrum along the line sa ­ 1, l ­ 0d ! sa ­ 1, l ­
1d joining the XY case and the model of Eq. (1); (ii)
in the antiferromagnetic case. The first p excitations are the number of low-lying modes, and, in particular, the
obtained by replacing p sjST 0 l 2 jT 0 Sldy 2 (respectively,
p presence of a gap as a function of a and l; (iii) the nature
sjST0 l 1 jT0 Sldy 2) byp sjST0 l 1 jT0 Sldy 2 (respec- of the effective model in the limit l ­ 1, a ø 1. Work
tively, sjST0 l 2 jT0 Sldy 2) in one of these ground is in progress along these lines.
states with energy a 2 y2 in the ferromagnetic case and We acknowledge very useful discussions with T. M.
al2 y2 in the antiferromagnetic case. So there is indeed Rice and F.-C. Zhang. One of us (B. F.) is also grateful
a gap in the spectrum. More importantly, it is clearly for financial support from the Swiss Nationalfonds. The
impossible to separate spin p and orbital degrees of freedom
p calculations were performed on the Intel Paragon at the
since sjST0 l 1 jT0 Sldy 2 and sjST0 l 2 jT0 Sldy 2 are ETH Zürich.
not eigenstates of sS$ i 1 S$ i11 d2 or of st$ i 1 t$ i11 d2 , and
the excitations are neither spin excitations nor orbital
excitations. They are transitions between resonating
valence-bond states that intimately mix spin and orbital
degrees of freedom.
It is also interesting to note that the correlation func- [1] W. Bao, C. Broholm, G. Aeppli, P. Dai, J. M. Honig, and
tions on a strong bond (i even) kS$ i ? S$ i11 l, kt$ i ? t$ i11 l, P. Metcalf, Phys. Rev. Lett. 78, 507 (1997); C. Broholm,
and ksS$ i ? S$ i11 d st$ i ? t$ i11 dl are all negative, as in the G. Aeppli, S.-H. Lee, W. Bao, and J. F. DiTusa, J. Appl.
Phys. 79, 5023 (1996).
SU(4) symmetric case, which excludes mean-field the-
[2] K. I. Kugel and D. I. Khomskii, Zh. Eksp. Teor. Fiz. 64,
ory as a good starting point for the same reasons (see 1429 (1973) [Sov. Phys. JETP 37, 725 (1973)]; Usp. Fiz.
Ref. [6]). Nauk. 136, 621 (1982) [Sov. Phys. Usp. 25, 231 (1982)].
Coming back to the original problem of the nature of [3] L. F. Feiner, A. M. Oleś, and J. Zaanen, Phys. Rev. Lett.
the excitations in the SU(4) symmetric model, let us put 78, 2799 (1997).
our results in perspective. In both cases studied above, [4] D. P. Arovas and A. Auerbach, Phys. Rev. B 52, 10 114
exact results have been obtained, and the low-energy (1995).
excitations are neither spin nor orbital excitations, but [5] Y. Q. Li, M. Ma, D. N. Shi, and F. C. Zhang, Phys. Rev.
involve both spin and orbital degrees of freedom on an Lett. 81, 3527 (1998).
equal footing. This is reminiscent of the classical version [6] B. Frischmuth, F. Mila, and M. Troyer, Phys. Rev. Lett.
of the Hamiltonian of Eq. (1) for which new ground 82, 835 (1999).
[7] B. Sutherland, Phys. Rev. B 12, 3795 (1975).
states can be generated from a given ground state by
[8] I. Affleck, Nucl. Phys. B265, 409 (1986).
flipping simultaneously the spin and the orbital at a given [9] Y. Yamashita, N. Shibata, and K. Ueda, Phys. Rev. B 58,
site. This is also a clear indication of the breakdown of 9114 (1998).
mean-field theory. It strongly suggests that the model [10] H. G. Evertz, G. Lana, and M. Marcu, Phys. Rev. Lett. 70,
of Eq. (1) also possesses such low-lying excitations. In 875 (1993).
particular, the operators ai6 ­ Si6 ti6 and bi6 ­ Si6 ti7 [11] B. B. Beard and U. J. Wiese, Phys. Rev. Lett. 77, 5130
of the XY case are linear combinations of the operators (1996).

3700

You might also like