4
arXiv:0801.0791v2 [cond-mat.stat-mech] 9 Jan 2008 Variational wavefunction study of the triangular lattice supersolid Arnab Sen 1 , Prasenjit Dutt 1 , Kedar Damle 1,2 , and R. Moessner 3,4 1 Department of Theoretical Physics, Tata Institute of Fundamental Research, Mumbai 400005, India 2 Physics Department, Indian Institute of Technology Bombay, Mumbai 400076, India 3 Rudolf Peierls Centre for Theoretical Physics, Oxford University, Oxford OX1 3NP, UK 4 Max-Planck-Institut f¨ ur Physik komplexer Systeme, 01189 Dresden, Germany (Dated: February 2, 2008) We present a variational wavefunction which explains the behaviour of the supersolid state formed by hard-core bosons on the triangular lattice. The wavefunction is a linear superposition of only and all configurations minimising the repulsion between the bosons (which it thus implements as a hard constraint). Its properties can be evaluated exactly—in particular, the variational minimisation of the energy yields (i) the surprising and initially controversial spontaneous density deviation from half-filling (ii) a quantitatively accurate estimate of the corresponding density wave (solid) order parameter. PACS numbers: 75.10.Jm 05.30.Jp 71.27.+a There has been sustained interest in using ultracold atoms confined in optical lattice potentials to realize strongly-correlated systems of interest to condensed mat- ter physics [1]. The recently discussed possibility that Helium has a supersolid phase[2] leads, in this context, to a natural question: Can the lattice analog of this, namely a superfluid phase that simultaneously breaks lat- tice translation symmetry, be seen in such optical lattice experiments? Although other examples are known [2], perhaps the best candidate for such a lattice supersolid persisting over a wide range of parameters (such as chemical potential and interaction energy) is that observed numerically in several recent Quantum Monte-Carlo (QMC) studies of a two-dimensional system of strongly interacting bosons in a triangular lattice potential [3, 4, 5], and confirmed in subsequent follow-up work [6]. Following earlier work [7], these QMC studies considered the model Hamiltonian: H = t ij(b i b j + b j b i )+ V ij(n i 1/2) (n j 1/2)(1) where b i (b i ) is the boson creation (annihilation) opera- tor, t represents the strength of the boson hopping am- plitude, V is the strength of the nearest neighbour re- pulsion, and the bosons are restricted to be in the hard- core limit (n i =0, 1) by a strong onsite repulsive term not written down explicitly (here, we have only displayed the Hamiltonian for the value of the chemical potential μ at which the system has particle-hole symmetry). Strik- ingly, in this particle-hole symmetric, hard-core case, an extended supersolid state was observed numerically for all V/t 8.9, and seen to possess both a non-zero su- perfluid stiffness, and density wave (‘solid’) order at the three sublattice ( 3 × 3 ordering) wavevector Q (the state is stable to changes in μ as well). There are two other genuinely surprisingly features of this phase. The first [4] concerns the nature of the solid order. Plausible mean-field theory arguments [5] predict that the density wave order in the supersolid at μ =0 should involve a ‘(+ 0)’ type three sub-lattice pattern of density with ρ a =1/2, ρ b =1/2+ δ, ρ c =1/2 δ, while in reality the system prefers a different ‘(+ −−)’ pattern with the same ordering wavevector: ρ a =1/2+ δ 1 , ρ b = ρ c =1/2 δ 2 , with δ 1,2 > 0 (Fig 1). General symmetry arguments [4, 5] predict that the total density of the system should also exhibit a spontaneous deviation from half-filling in a ‘(+ −−)’ state. The second surprise is that although there is such a deviation, it is extremely small in ‘natural’ units (i.e compared to the strength of the density wave order itself), and therefore extremely difficult to see numerically [4, 5]. In order to understand this large V/t supersolid phase in the hardcore limit, it is necessary to take into account this strong repulsion as a hard constraint. Here we con- sider an analytically tractable variational wavefunction that takes into account this hard constraint from the start, and uses a variational parameter to tune the am- plitudes of different minimum repulsion energy configura- tions. The energetically optimal wavefunction accounts for both qualitative (‘(+ −−)’ order) and quantitative (size of the order parameter) aspects of the supersolid phase in the limit of strong nearest neighbour repulsion. Our treatment thus represents a remarkable instance in which a non-trivial constraint on the lattice scale imposed by a dominant term in the Hamiltonian can be taken into account in an exact manner in a variational calculation— for instance, the celebrated problem of accounting for the single-occupancy constraint imposed (by strong coulomb repulsion) on holes in the Cu-O planes of high-T c super- conductors has resisted analytical treatment thus far. Our starting point is the observation that the super- solid ground state at large V/t must have a wavefunction that lies entirely in the subspace of minimum interac- tion energy configurations. More precisely, in the classi- cal (t = 0) limit, the ground states are diagonal in the particle-number basis, and form an extensively degener-

Arnab Sen et al- Variational wavefunction study of the triangular lattice supersolid

  • Upload
    kuiasm

  • View
    20

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Arnab Sen et al- Variational wavefunction study of the triangular lattice supersolid

arX

iv:0

801.

0791

v2 [

cond

-mat

.sta

t-m

ech]

9 J

an 2

008

Variational wavefunction study of the triangular lattice supersolid

Arnab Sen1, Prasenjit Dutt1, Kedar Damle1,2, and R. Moessner3,4

1Department of Theoretical Physics, Tata Institute of Fundamental Research, Mumbai 400005, India2 Physics Department, Indian Institute of Technology Bombay, Mumbai 400076, India

3 Rudolf Peierls Centre for Theoretical Physics, Oxford University, Oxford OX1 3NP, UK4 Max-Planck-Institut fur Physik komplexer Systeme, 01189 Dresden, Germany

(Dated: February 2, 2008)

We present a variational wavefunction which explains the behaviour of the supersolid state formedby hard-core bosons on the triangular lattice. The wavefunction is a linear superposition of only and

all configurations minimising the repulsion between the bosons (which it thus implements as a hardconstraint). Its properties can be evaluated exactly—in particular, the variational minimisation ofthe energy yields (i) the surprising and initially controversial spontaneous density deviation fromhalf-filling (ii) a quantitatively accurate estimate of the corresponding density wave (solid) orderparameter.

PACS numbers: 75.10.Jm 05.30.Jp 71.27.+a

There has been sustained interest in using ultracoldatoms confined in optical lattice potentials to realizestrongly-correlated systems of interest to condensed mat-ter physics [1]. The recently discussed possibility thatHelium has a supersolid phase[2] leads, in this context,to a natural question: Can the lattice analog of this,namely a superfluid phase that simultaneously breaks lat-tice translation symmetry, be seen in such optical latticeexperiments?

Although other examples are known [2], perhaps thebest candidate for such a lattice supersolid persisting overa wide range of parameters (such as chemical potentialand interaction energy) is that observed numerically inseveral recent Quantum Monte-Carlo (QMC) studies ofa two-dimensional system of strongly interacting bosonsin a triangular lattice potential [3, 4, 5], and confirmed insubsequent follow-up work [6]. Following earlier work [7],these QMC studies considered the model Hamiltonian:

H = −t∑

〈ij〉

(b†ibj + b†jbi) + V∑

〈ij〉

(ni − 1/2) (nj − 1/2)(1)

where b†i (bi) is the boson creation (annihilation) opera-tor, t represents the strength of the boson hopping am-plitude, V is the strength of the nearest neighbour re-pulsion, and the bosons are restricted to be in the hard-core limit (ni = 0, 1) by a strong onsite repulsive termnot written down explicitly (here, we have only displayedthe Hamiltonian for the value of the chemical potential µat which the system has particle-hole symmetry). Strik-ingly, in this particle-hole symmetric, hard-core case, anextended supersolid state was observed numerically forall V/t ≥ 8.9, and seen to possess both a non-zero su-perfluid stiffness, and density wave (‘solid’) order at the

three sublattice (√

3 ×√

3 ordering) wavevector ~Q (thestate is stable to changes in µ as well).

There are two other genuinely surprisingly features ofthis phase. The first [4] concerns the nature of the solidorder. Plausible mean-field theory arguments [5] predict

that the density wave order in the supersolid at µ = 0should involve a ‘(+ − 0)’ type three sub-lattice patternof density with ρa = 1/2, ρb = 1/2 + δ, ρc = 1/2 − δ,while in reality the system prefers a different ‘(+ − −)’pattern with the same ordering wavevector: ρa = 1/2 +δ1, ρb = ρc = 1/2 − δ2, with δ1,2 > 0 (Fig 1). Generalsymmetry arguments [4, 5] predict that the total densityof the system should also exhibit a spontaneous deviationfrom half-filling in a ‘(+−−)’ state. The second surpriseis that although there is such a deviation, it is extremely

small in ‘natural’ units (i.e compared to the strength ofthe density wave order itself), and therefore extremelydifficult to see numerically [4, 5].

In order to understand this large V/t supersolid phasein the hardcore limit, it is necessary to take into accountthis strong repulsion as a hard constraint. Here we con-sider an analytically tractable variational wavefunctionthat takes into account this hard constraint from thestart, and uses a variational parameter to tune the am-plitudes of different minimum repulsion energy configura-tions. The energetically optimal wavefunction accountsfor both qualitative (‘(+ − −)’ order) and quantitative(size of the order parameter) aspects of the supersolidphase in the limit of strong nearest neighbour repulsion.Our treatment thus represents a remarkable instance inwhich a non-trivial constraint on the lattice scale imposedby a dominant term in the Hamiltonian can be taken intoaccount in an exact manner in a variational calculation—for instance, the celebrated problem of accounting for thesingle-occupancy constraint imposed (by strong coulombrepulsion) on holes in the Cu-O planes of high-Tc super-conductors has resisted analytical treatment thus far.

Our starting point is the observation that the super-solid ground state at large V/t must have a wavefunctionthat lies entirely in the subspace of minimum interac-tion energy configurations. More precisely, in the classi-cal (t = 0) limit, the ground states are diagonal in theparticle-number basis, and form an extensively degener-

Page 2: Arnab Sen et al- Variational wavefunction study of the triangular lattice supersolid

2

C

B

A

CC A BBA

B

1/2

1/2 −δ1/2 + δ

C

CBACBA

B

C

A

B

C

1/2 −δ 2

1/2 +δ1

n=1

n=0

Dimer

(a)T

TT

0

12

(b)

FIG. 1: (color online). a) The two possible supersolid states

at the three-sublattice wavevector ~Q. Dark (light) links rep-resent higher (lower) bond kinetic energy, and sites are colorcoded to reflect mean density. b) Flippable single and doublehexagons in the dimer representation, and the correspondinglocal density configuration.

ate set of states corresponding to all minimum repulsionenergy configurations of particles. To leading order int/V , the slow dynamics induced by the kinetic energyterm then leads to an effective Hamiltonian H acting inwithin this manifold of states:

Heff = −t∑

〈ij〉

Pg(b†i bj + b†jbi)Pg (2)

where Pg is the projection operator to the minimum re-pulsion energy subspace.

As H maps to a system of S = 1/2 spins (Szi = ni−1/2

where ni is the boson number at site i) with frustrated

antiferromagnetic exchange Jz = V between the z com-ponents of neighbouring spins, ferromagnetic exchangeJ⊥ = 2t between their x and y components, and magneticfield Bz ∝ µ, the ground states in this t = 0 limit may beidentified with minimally frustrated states of the classicalIsing antiferromagnet on the triangular lattice [8]. As iswell-known, these may be conveniently characterized interms of dimer coverings of the dual honeycomb lattice(with a dimer placed on the dual link perpendicular to ev-ery frustrated bond of a given spin configuration). In thislanguage, Heff is simply a quantum dimer model witha ring-exchange term that acts on each pair of adjacenthexagons on the dual honeycomb lattice (correspondingto each bond 〈ij〉 of the triangular lattice):

Heff = −t∑

〈ij〉

(| 〈ij〉〉〈 〈ij〉| + h.c.) (3)

Thus, the supersolid behaviour at large V/t should beunderstood in terms of the ground state of this quantumdimer model. As was noted in earlier literature [4, 5],a trial state obtained from an equal amplitude superpo-sition of all minimum interaction energy configurations(which maps, apart from a global particle-hole transfor-mation, to a uniform superposition of dimer covers of the

1

3

56

42

x

y

^

^

1

z

b)

C 1

C2

1C

C2

(a)(

FIG. 2: (color online). a) Dimer ensemble with a three-sublattice pattern of fugacities which breaks lattice transla-tion symmetry at wavevector ~Q. b) The two types of flippabledouble-hexagons, with two flippable configurations each—note that light (dark) links correspond to fugacities 1 (z).

honeycomb lattice) already provides a substantial kineticenergy gain, while minimizing the inter-particle repulsionby construction.

Since 〈Ψ0|b†i |Ψ0〉 is clearly proportional to the non-

zero probability that hexagon i is flippable in the clas-sical dimer model, this trial state immediately providesa rationale for the persistence of off-diagonal long-rangeorder and superfluidity in the large V/t limit. On theother hand, this trial state is unable to fully account forthe density wave order of the true supersolid state, asdensity correlators in |Ψ0〉 map on to spin correlations ofthe T = 0 classical Ising antiferromagnet on the trian-gular lattice which does not support genuine long-rangeorder [8], but instead displays power-law order at thethree-sublattice wavevector [9].

Here, we focus instead on a variational state obtainedfrom a dimer model with two types of links, with differentdimer fugacities 1 and z ≥ 0, defined on the honeycomblattice in a pattern (Fig 2) which breaks lattice symmetry

at wavevector ~Q for all z 6= 1:

|Ψvar(z)〉 =∑

Cd

PCd(z)/2 (|n+(Cd)〉 + |n−(Cd)〉) (4)

where PCd(z) is the probability of a dimer configuration

Cd in this dimer model, and the two particle-hole conju-gate configurations |n±(Cd)〉 corresponding to Cd occurwith the same amplitude in this nodeless wavefunction(this follows from the Perron-Frobenius theorem, as wasnoted earlier in a different context[10]).

Clearly, |Ψ(z)〉 is characterized by three-sublattice den-sity wave order of the (+−−) type (see Fig 1) for z < 1while for z > 1, it displays density wave order of the(+ − 0) type. Furthermore, as the state is constructedfrom a coherent superposition of Fock states with a con-siderably wide distribution of average density, |Ψ(z)〉 isexpected to also possess off-diagonal long range order as-sociated with superfluidity, at least for z close to z = 1.

In order to perform an unbiased variationalstudy, we need to locate minima of the varia-

Page 3: Arnab Sen et al- Variational wavefunction study of the triangular lattice supersolid

3

-1.380

-1.374

-1.368

-1.362

-1.356

0.8 0.85 0.9 0.95 1 1.05 1.1 1.15 1.2

E(z

) (x

10)

z

MathematicaL = 72L = 96

L = 120

FIG. 3: (color online). Variational energy per site in units of2t as a function of z.

tional estimate of energy per site (in units of 2t):E(z) = 〈Ψvar(z)|Heff |Ψvar(z)〉/2tL2. To calculate

E(z), we note that 〈Ψvar(z)|Pgb†ibjPg|Ψvar(z)〉 =

P ( 〈ij〉)P ( 〈ij〉), where P ( 〈ij〉) and P ( 〈ij〉)are the probabilities that the double hexagon 〈ij〉 is inthe flippable configurations depicted in their respectivearguments. This allows us to write

E(z) = −(

2√

P (1C)P (1C) +√

P (2C)P (2C)

)

(5)

where P (1C) and P (1C) (P (2C) and P (2C)) are therespective probabilities that type 1 (type 2) double-hexagons are in flippable configurations C and C (seeFig 2). Thus, the variational energy is lowered if theexpected number of flippable double-hexagons is large.Since the number of flippable double hexagons must de-crease rapidly to zero in the z → 0 as well as the z → ∞limit as the dimers freeze into a perfect columnar or pla-quette state in these limits, it is immediately clear thatone or more variational minima of E(z) must occur ator in the vicinity of the translationally invariant pointz = 1.

Calculation of the probabilities P is greatly facilitatedby the well-known formulation of dimer models on planargraphs in terms of Grassmann variables [11, 12, 13]: Forthe case at hand, this can be obtained by using the ‘arrowconvention’ displayed in Fig 2 to define an antisymmetricmatrix M , with Mij = +µ〈ij〉 (where the fugacity asso-ciated with link 〈ij〉) if an arrow points from point i to jand Mij = −µ〈ij〉 if the arrow goes from j to i (Mij = 0if i,j are not nearest neighbours, µ〈ij〉 equals z or 1 asshown in Fig 2). The dimer partition function Z is thenobtained as Z = |Pf [M ]| =

[Dψ] exp(S)∣

∣, where Pfdenotes the Pfaffian and S =

i<j Mijψiψj is the ac-tion for Grassmann variables ψi defined on sites of thehoneycomb lattice.

To proceed further, we represent the honeycomb net

0

0.02

0.04

0.06

0.08

0.1

0.12

-0.025 -0.015 -0.005 0.005 0.015 0.025

P(

δρ )

δρ = ρ -1/2

z=0.925z=1.075

FIG. 4: (color online). Distribution of the total density atthe global variational minimum at z ≈ 0.925, as well as atthe competing local minimum at z ≈ 1.07. Data shown is forL = 96.

as an underlying triangular Bravais lattice with Romanletter coordinates (representing centers of those hexagonswhose links all have fugacity z) that is decorated by sixbasis points (corresponding to vertices of such hexagons)labeled by Greek letters (Fig 2 ), and write the action

as S = 1

2

~x,α

~y,β Mα,β~x,~y ψα,~xψβ,~y. Transforming to

Fourier space, we obtain S = 1

2

~k,α,β Mα,β~k

ψα,~kψβ,−~k,

with the 6 × 6 matrix Mα,β~k

is given as

M(~k) =

(

0 R(~k)

−R†(~k) 0

)

where 0 is the 3× 3 null matrix and R can be written as

R(~k) =

z −e−iky −z−z −z eikx

−e−ikx+iky z −z

Next, we note that this Grassmann formulation pro-vides a simple prescription for the probability that fivedimers occupy alternate links l1 . . . l5 on the perimeter ofdouble hexagon made up of points i = 1, 2 . . . 10:

P ({l1 . . . l5}) =

(

5∏

m=1

µlm

)

· |〈ψ1ψ2ψ3 · ·ψ10〉| (6)

where µlm are the fugacities of the 5 links occupied bydimers. This allows us to write E(z) as

E(z) = −2z7

2 |〈ψ1ψ2ψ3 · ·ψ10〉1| − z3|〈ψ1ψ2ψ3 · ·ψ10〉2|(7)

where the subscripts on the 10-point correlators referto the type of double-hexagon on which they are cal-culated (see Fig 2).

Page 4: Arnab Sen et al- Variational wavefunction study of the triangular lattice supersolid

4

The 10 point correlators involved in this expressioncan be computed in the free field action S using Wick’stheorem and knowledge of the two point correlators:

〈ψα,~xψβ,~y〉 =

d2k

4π2exp(−i~k · (~x− ~y))(−M

−1(−~k))α,β(8)

where the integration is over the Brillouin zonekx, kyǫ(−π, π]. As 〈ψiψj〉 = 0 if i and j belong to thesame sublattice on the honeycomb lattice, one needs tokeep track of ‘only’ 5! = 120 contractions in evaluatingeither of the two 10-point correlators. Keeping track ofthese contractions and performing the required k integra-tions using MATHEMATICA, we obtain the variationalenergy E(z) shown in Fig 3. As is clear from this figure,E(z) has two minima, a global minimum at z ≈ 0.9250,and local minimum at z ≈ 1.0750 with energy only veryslightly higher (by about ≈ 0.047%) than its value at theglobal minimum.

Thus, our variational calculation yields a (+−−) typethree sublattice ordered supersolid state in the large V/tlimit, and suggests that the energy of the (+ − 0) su-persolid at the same wavevector is only slightly higherthan that of the (+ − −) supersolid. In order to obtainan independent verification of this result, we have alsoused the method of Ref [14] to simulate the correspondingdimer model and numerically calculate E(z)—the resultsof these calculations are seen to match precisely with theresults obtained by Grassmann techniques (Fig 3).

This simulation also allows us to measure directly thesolid order parameter ψ = ma + mbe

2πi/3 + mce4πi/3,

where ma,mb,mc are the number densities on the threesublattices of the triangular lattice. We obtain |ψ|2 =0.03885 for the global minimum at z = 0.9250, whichagrees within 10% with the results from QMC [5] ex-trapolated to V/t → ∞. ( the value in the competing(+ − 0) state is |ψ|2 = 0.03697 at the slightly higher en-ergy local minimum at z = 1.0750). In addition, we alsomeasure the histogram of the total density, from whichone may calculate the ground state expectation values ofall powers of the density in the supersolid state. Whilethe local minimum at slightly higher energy has no spon-taneous deviation of density from half-filling, we find thatthe density histogram at the global minimum shows acharacteristic two-peak structure, reflecting a very small(∼ 2%) spontaneous deviation from half-filling charac-teristic of the (+ −−) supersolid (Fig 4).

To obtain further insight into the smallness of thisspontaneous density deviation from half-filling, we notethat reducing z slightly from z = 1 introduces a fieldcoupled to the Fourier component of the particle-densityat wavevector ~Q. The effect of this field can be under-stood by using the well-known [15] height model for-mulation of z = 1 dimer model in terms of an actionSh = κ

d2x(∇h)2 for a ‘height’ field h(x) in terms ofwhich one may write the Fourier component of the par-ticle density at wavevector ~Q as ρ~Q ∼ exp(iπh/3), and

at wavevector 0 as ρtot ∼ exp(iπh).

For z < 1 but close to z = 1, the response of thedensity at these two wavevectors can be obtained by cal-culating the corresponding susceptibilities, and one findsthat the response at wavevector ~Q is much larger than atwavevector 0, thereby providing a rationalization for thesmallness of this symmetry allowed density deviation (incomparison with the size of the order parameter).

Thus, our variational approach provides a coherent pic-ture of the triangular lattice supersolid at large V/t: Ki-netic energy effects are seen to select a (+−−) supersolidstate with a three-sublattice density wave order parame-ter comparable in magnitude to the mean density itself.In addition, the competing (+ − 0) supersolid is seen tobe very close in energy to this variational ground state,consistent with numerical evidence and analytical argu-ments [16] that the order parameter phase (which distin-guishes between these two states) is only weakly pinnedin the supersolid state. Furthermore, the spontaneous de-viation of the density from 1/2 is seen to be a very smallfraction (∼ 2%) of the mean density, consistent with thesurprisingly small value obtained in earlier QMC studies.

Acknowledgements: We thank D. Dhar for very in-sightful discussions and S. L. Sondhi for collaborationon closely related work. We acknowledge computationalresources at TIFR and Oxford and support from DSTSR/S2/RJN-25/2006 (KD), the John Fell OUP ResearchFund–Oxford-India network in Theoretical Physical Sci-ences (AS), and a British Council RXP grant (RM).

[1] I. Bloch, J. Dalibard, and W. Zwerger, arXiv:0704.3011(to appear in Rev. Mod. Phys.).

[2] http://online.kitp.ucsb.edu/online/smatter m06[3] S. Wessel and M. Troyer, Phys. Rev. Lett. 95, 127205

(2005).[4] D. Heidarian and K. Damle, Phys. Rev. Lett. 95, 127206

(2005).[5] R. Melko et al., Phys. Rev. Lett. 95, 127207 (2005).[6] M. Bonninsegni and N. Prokof’ev, Phys. Rev. Lett. 95,

237204 (2005).[7] G. Murthy, D. Arovas, and A. Auerbach, Phys. Rev. B

56, 3104 (1997).[8] G. H. Wannier, Phys. Rev. 79, 357 (1950).[9] J. Stephenson, J. Math. Phys. 11, 413 (1970).

[10] R. Moessner and S. L. Sondhi, Phys. Rev. B 63, 224401(2001).

[11] P. W. Kasteleyn, J. Math. Phys. 4, 287 (1963).[12] S. Samuel, J. Math. Phys. 21, 2806 (1980).[13] R. Moessner and S. L. Sondhi, Phys. Rev. B 62, 14122

(2000).[14] A. W. Sandvik and R. Moessner, Phys. Rev. B 73, 144504

(2006).[15] B. Nienhuis, H. J. Hilhorst, and H. J. W. Blote, J. Phys.

A 17, 3559 (1984).[16] A. A. Burkov and L. Balents, Phys. Rev. B 72, 134502

(2005).