• Aucun résultat trouvé

On quantum spin chains and liquid crystal films

N/A
N/A
Protected

Academic year: 2021

Partager "On quantum spin chains and liquid crystal films"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00210731

https://hal.archives-ouvertes.fr/jpa-00210731

Submitted on 1 Jan 1988

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

On quantum spin chains and liquid crystal films

T.J. Sluckin, Timothy Ziman

To cite this version:

T.J. Sluckin, Timothy Ziman. On quantum spin chains and liquid crystal films. Journal de Physique,

1988, 49 (4), pp.567-576. �10.1051/jphys:01988004904056700�. �jpa-00210731�

(2)

E 567

On quantum spin chains and liquid crystal films

T. J. Sluckin (1, and Timothy Ziman (1, 2)

(1) Institut Laue Langevin, B.P. 156X, 38042 Grenoble Cedex, France

(2) Department of Physics and Astronomy, Rutgers University, P.O. Box 849, Piscataway, New Jersey 08855- 0849, U.S.A.

(Requ le 16 juillet 1987, accepté sous forme définitive le 14 dgcembre 1987)

Résumé.

2014

Nous avons étudié un modèle bidimensionnel de spin, introduit par Korshunov d’une part et par Lee et Grinstein d’autre part, décrivant l’apparition de l’ordre dans des molécules soumises à des interactions

nématiques et ferromagnétiques antagonistes. Utilisant un formalisme de matrices de transfert nous avons

établi une correspondance avec un modèle quantique unidimensionnel de spin 2. Celui-ci a été diagonalisé

pour des chaînes de longueur délimitée. Nous trouvons alors un comportement cohérent avec le diagramme de phase prédit : une phase planaire à basse température séparée d’une phase désordonnée de haute température

par une ligne de transitions du type Kosterlitz-Thouless, et séparée d’une phase nématique à basse température par une ligne de transitions du type Ising. On s’attend à ce que la transition nématique- paramagnétique soit du type Kosterlitz-Thouless. Ici, les résultats sont moins clairs, mais cependant nous

discutons des topologies cohérentes avec nos résultats.

Abstract.

2014

A two dimensional spin model introduced by Korshunov and by Lee and Grinstein, which describes ordering of molecules with competing nematic and ferromagnetic interactions, has been studied.

Using the transfer matrix formalism, we have derived an equivalent spin 2 quantum model in one spatial

dimension. This has been diagonalised for chains of finite length. We find behaviour consistent with the

predicted phase diagram ; a low temperature planar phase separated from a disordered highs temperature phase by a line of transitions of the Kosterlitz-Thouless type, and from a low temperature nematic phase by a

line of Ising transitions. The nematic to paramagnetic transition has also been predicted to be Kosterlitz- Thouless-like. Here the numerical evidence is less clear ; however we discuss topologies consistent with our

results.

J. Phys. France 49 (1988) 567-576 AVRIL 1988,

Classification

Physics Abstracts

75.10H - 75.10J

-

05.70J

-

61.30

1. Introduction.

Many of the most interesting problems in statistical mechanics are concerned with competing ordering

mechanisms. The resolution of the competition

often leads to an interesting low temperature phase diagram. In this paper we study such a problem, the

« mixed » model introduced independently by Kor-

shunov [1] and Lee and Grinstein [2]. This is a two-

dimensional lattice model, with X-Y spins at each

lattice point. The spins at neighbouring lattice points

(*) Permanent address : Department of Mathematics, University of Southampton, Southampton S095NH, Great-Britain.

(**) (Lady Davis fellow) Department of Chemistry,

Technion-Israel Institute of Technology, 32000 Haifa, Israel.

JOURNAL DE PHYSIQUE. - T. 49, 4, AVRIL 1988

are coupled, firstly by a ferromagnetic interaction

favouring parallel orientation and secondly by a

« nematic » interaction favouring either parallel or antiparallel orientation. These two tendencies com-

pete once we include entropy and it has been

predicted that this gives rise to a variety of low temperature phases : we summarise these predictions

at a later stage in the article. This Hamiltonian may describe the physics of liquid crystal films, in which

the molecules composing the liquid are ferroelectric and therefore have a polarity. The Hamiltonian we

shall discuss does correctly include the short range effective potential implied by such a weak polarity although if the ferroelectricity were sufficiently de- veloped it would be essential to include the additional long-range electromagnetic interactions that lead to domain formation.

The formal analogy between the partition function

in statistical mechanics and the Lagrangian in the

37

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01988004904056700

(3)

path integral formulation in quantum mechanics is

by now well-known [3]. It allows a one-dimensional

problem in statistical mechanics to be mapped onto

the quantum mechanics of a single particle, and by extension, maps d-dimensional statistical mechanics

problems onto (d -1 ) dimensional quantum prob-

lems. This permits mathematical techniques and experience developed for understanding quantum mechanics to be brought to bear in statistical mech- anics and vice versa.

In this paper we exploit this analogy to transform

the mixed model into an equivalent one-dimensional chain of quantum spins. This quantum spin chain is

similar to others that have been studied in the past

[3-6], and we can profit from this experience. We perform an exact diagonalisation of the Hamiltonian of finite periodic chains of n spins with n between 4 and 10, and extrapolate to infer properties of infinite

chains.

It turns out that to provide a simple description of

the nematic phase the minimum quantum spin chain

has spin 2 ; in the past intensive studies have been carried out on spin 1 , 1 and 3 , each of which has

2 ’ 2

rather different properties [5, 6]. This chain has

nematic as well as ferromagnetic interactions and as

such is of some intrinsic interest. Indeed Andreev and Grischuk have postulated that solid He3 is a spin

nematic [7] and Papanicolau has studied the ground

state properties of spin 1 nematics [8].

This paper is organised in the following way. In section 2 we discuss in more detail the mixed model.

Then in section 3 we describe how we derive the

equivalent quantum spin problem, and explain the

limitations to the approach. In section 4 we solve the spin problem using a superposition approximation as

an Ansatz to the ground state wave function. This

approximation has a theoretical status more or less

equivalent to mean field theory and we shall refer to it as such. In section 5 we discuss the results obtained from exact diagonalisation of finite length chains.

Finally in section 6 we draw some general con-

clusions.

2. The mixed model.

First we describe the mixed model studied by

Korshunov [1] and Lee and Grinstein [2] ; then we

discuss some related models. The Hamiltonian of the mixed model is

where Jl, J2 . 0 and the angle Oi is defined at each site i of a two-dimensional square lattice, and the pairs [ij] are nearest neighbours. This Hamiltonian

can also be rewritten in terms of the spins cri

=

(cos 0 j, sin 0 i ) in which case it contains terms

quartic as well as quadratic ins ;

It is convenient to define a variable a by

A point in the phase diagram is now defined in terms of the variables (T

=

T/J, a ). T is the scaled

temperature and we can without ambiguity drop the

tilde. The degree of « nematicity » is expressed by

a ; a

=

0 corresponds to the pure XY model, and

a =1 corresponds to a purely nematic version of the XY model.

The authors of references 1 and 2 used a number of techniques to define the phase diagram. The

a

=

0 line is already well-known [4] ; there is a low

temperature phase with algebraic order defined by

At T = 0, TJ 1 (T) = 0 implying true long range order. As T is increased q 1 increases until at the Kosterlitz-Thouless temperature q - 1 4 beyond

which order decays exponentially with distance rij. A quantity of interest in describing the phase diagram as a whole is TJ2(T), defined by

However so long as a

=

0 the Gaussian result

’TJ 2 = 4 ’TJ 1 should hold.

The a =1, nematic line is also easily understood by making the transformation I/J i = 2 P i, which provides a mapping to the a

=

0 line. In the low temperature phase exp i (oi - Oj) decays ex onen-

tially but ’TJ 2 is well-defined and less than 4. For

tially but r 2 is well-defmed and less than 4 For intermediate a we shall show that finite-size studies

can be used to find the phase boundaries. For

a :s:: 1 we expect two transitions as first ’q 2 and then

,q 1 become finite, with decreasing temperature. The

latter transition is expected to be Ising like in that above it (gii) is well defined, implying Pi) is

defined only mod (-fr), whereas below it Pi) is

defined mod (2 7T ). In between there will be a

multicritical point where the two transitions join and

near it some or all of the phase transition lines may be of first order. The hypothetical phase diagram is

shown in figure 1.

It is the equivalent quantum version of this model

on which we concentrate our attention. It is never-

theless of some interest to mention some related

models, all of which in some way pit magnetic

interactions against the effect of nematic liquid

crystallinity. A simple generalization is to allow

(4)

569

Fig. 1.

-

Expected phase diagram of the Hamiltonian (1)

of Korshunov, Lee and Grinstein.

J1 or J2 to be negative [9]. If J2 0 the ground state

becomes multiply degenerate over a certain range of

a and there are a wide variety of ordered phases.

One might also consider the usual mixed model on a

three dimensional cubic lattice. There have been recent predictions [10] that although the transition to a (now long-ranged) ordered phase is continuous for

a close to 1 and 0, in a limited range of a there is a

first order transition. Alternatively one might con-

sider a Hamiltonian of the form (2) but suppose that the ( cr ) are now Heisenberg spins ; this corresponds

to competition between Heisenberg and classical nematic order. Then in two dimensions, using the spin wave arguments of Polyakov [11] there are no

low temperature ordered phases. However in three dimensions there is a small a low T ferromagnetic phase separated from a high a low T nematic phase by an Ising transition, which is itself separated from

the high T disordered phase by a continuous tran- sition. Arguments based on Landau theory suggest that there will be a tricritical point beyond which the

transition is first order. A number of authors, indeed, have studied a magnetic model with multis-

pin interactions [12] and found such first order transitions.

3. The equivalent quantum spin system.

In this section we follow the method of Stoeckly and Scalapino [3] who discussed the mapping between

the planar model, a continuum version of the X-Y

model, and an equivalent quantum system.

The strategy is as follows. We first write down the free energy appropriate to a continuum model with the same order parameters as the mixed model.

This, we believe, will have a phase diagram topologi- cally equivalent to the mixed model. We then shall take an anisotropic limit in which the y-direction is

discretized. The partition function is then equivalent

to the path integral formulation in quantum mech- anics if the x-direction is identified with imaginary

time (it). This enables us to identify the generating quantum mechanical Hamiltonian, which now applies to a chain of sites. Finally we carry out some further approximations and truncations which leave

an effective spin chain.

A system described by the Hamiltonian has two relevant order parameters. These are a vector mag- netisation m, where

where a is a Euclidean subscript, and a traceless

nematic tensor Q where

It is convenient to introduce complex order par- ameters

The angles qi and § give the orientation of the local

magnetic and nematic order parameters respectively,

and the numbers I m I and I Q I give their respective magnitudes.

We suppose the system anisotropic and write

down the Landau free energy in terms of the order parameters m and Q :

where A1 and A2 are the purely magnetic and

nematic free energies respectively :

(there is no cubic term because there is no third order invariant in two dimensions), and A3 is the coupling term

We now suppose the y direction to be discretized, so

that the free energy now refers to chains in the x-

direction, coupled to their neighbours, and a distance

8y apart. The gradient terms in the y direction now

become

(5)

where the i subscripts are labels referring to the (one-dimensional) chain array.

The quantities a,

=

ai (T - Tern), a2

=

a2 (T - Ten) identify the mean field transition temperatures Tcm, Ten to magnetic and nematic states respectively.

At low temperatures there are four order parameters

I m I, qi, I Q I and ç. However from equation (10) it

can be seen that fluctuations in I m I , I Q I and

5 = ç - 2 qi cost free energy locally and may be

ignored ; the important fluctuations, corresponding

to what would be Goldstone modes if there were true long range order, are those in qi, and on those

we concentrate. The varying terms in the free energy functional are now given by

The mapping to the quantum system is effected by observing that the partition function

and its derivatives can be found via the properties of

the transfer matrix defined in the x direction.

Making the mapping x = it (h

=

1 ) the transfer matrix can be written as a quantum-mechanical

Hamiltonian of variables gii i

where T = /3 and Boltzmann’s constant is taken to be 1. This is now the quantum mechanical Hamilto- nian of a chain of sites i. We investigate the ground

state properties of the Hamiltonian, as a function of the parameters D and {3, where

and identification is made between the parameters in equations (14) and (15). Now D is the temperature- like variable and K2 and Kl tend to induce nematic order and magnetism respectively.

We can identify on-site and inter-site terms in the Hamiltonian : the operator = - I a has

a(pi eigenstates [ pi)

=

exp ipgii, we define operators

subject to the commutation relations

In terms of these operators the Hamiltonian is now

and we observe that bt , bi- are raising and lowering

operators with respect to the levels Ipi) though

without the Clebsch-Gordon coefficients which occur

in angular momentum ; bi is analogous to the z- component of the spin.

This Hamiltonian without the K2 term has been

much studied [3-6] ; the fundamental features of the

ground states can be delineated by restricting the

basis set to p

=

0, ± 1 and now p

=

± 2 states. Thus within our approximation

That the analogy is good, can be seen by comparing

the limiting cases of K2

=

0 and p restricted to 0,

± 1 and Ki = 0 and p restricted to 0, + 1, ± 2. In the

latter case the p

=

± 1 states are not coupled to the ground state even indirectly, and may be ignored.

Defining operators ci - - (bt-)2, and ci = 1/2 bi,

the Hamiltonian for the limiting case can be rewrit-

ten

The properties of the operators {cj in this problem

are now identified to those of the operators

(bi) for K2 = 0 and p restricted to 0, ± 1. We

conclude that the minimal basis set for understanding ground states in the Kl

=

0 case is the p = 0 ± 2 set.

To study the competition if Kl =1= 0 the p = ± 1 must

also be included.

Before studying the Hamiltonian (18) some com-

ments on the transformation from the original equation (1) are in order. In reference [3], there is a

discussion of the same transformation for the simple

X-Y model which explains why fluctuations in

I m I can be ignored in that case. In our problem I Q I fluctuations are analogous in being amplitude

fluctuations. There remain fluctuations in the two

phase variables ç and qi. They are coupled with two

inequivalent but equally likely minima corresponding

(6)

571

to, respectively, )

=

2.p and § = 2 (gi + w ) [0 -- 6 -- 2 7T ]. We ignore coupling between these

minima ; in the quantum language this occurs by tunnelling and could be calculated by the WKB

method. This tunnelling corresponds, in the original

model to the « strings » joining spin 1/2 vortices of

opposite parity predicted to exist in the low T nematic phases. These strings cost energy in the low T planar phase (and can presumably be ignored in

the classical to quantum transformation for that

reason) but are soft in the nematic phase and perhaps should be considered explicitly.

We shall seek in subsequent sections an analogue

to the low T nematic phase in the quantum model.

This phase will possess spatial correlation functions

corresponding to the existence of « soft » strings in

the time direction. As the correlations are Lorentz invariant this implies time correlation functions of the same form. Thus although the transformation breaks the symmetry of the two lattice directions this symmetry is restored for long wavelengths. The string tension in the space direction, though initially finite, is renormalised to zero. We return to this

point briefly in the conclusions.

4. Mean field theory.

It is in general difficult to find the ground state of

many body quantum systems. Even in one dimension only special Hamiltonians may be integrated by the

Bethe Ansatz and those that can be are not necess-

arily representative. The first approximation for a qualitative understanding is the superposition ap-

proximation. Quantum fluctuations are simplified

rather as mean field theory simplifies thermal fluctu- ations in classical statistical mechanics. In low dimen- sional systems this may be dangerous ; nevertheless with proper care the results may be useful.

We use a superposition approximation to construct

the ground state of the Hamiltonian (18) following

closely the argument of Solyom and Ziman [5]. The

form of the ground state does not depend on the

overall magnitude of JCQ. There are two dimension-

.

less parameters which change the form of the ground

state. Defining

where

the parameters are D (henceforth we drop the tilde)

and y, which is analogous to a in section 2.

The ground state Ansatz

where

where

and 80, 81, 82 are real quantities.

This state is invariant under time reversal as

appropriate for planar or singlet ground states. The

energy per spin is found using the variational prin- ciple

and yields

Minimization of E with respect to the variational parameters allows the following possibilities :

(1) So = 1, 81= 8 2 = 0. This is singlet ground

state, corresponding to a disordered phase of the equivalent thermodynamic system :

and

This state obtains for large D, which favours the

10 > as against the I -t 1 ) , ± 2 ) states.

(2) 8 0, 81 1 and 52 =1= 0; 0 1 arbitrary but

0 1 - (J 2

=

0 (mod 7T). This corresponds to a state

(7)

with planar and nematic order, the equality of 0 1 and 9 2 indicates that the orientation of the planar

and nematic order parameters is the same, although

the nematic order parameter can reverse in direction without change n the state. The state is degenerate

in the x-y plane since 01 is arbitrary. The planar

order parameter (see Eqs. (8)) is

and the nematic order parameter is

This state obtains for low values of y and D.

(3) 8 0, 5 2 =A 0, 5 1 0 . 9 2 = 0 arbitrary. This is a

nematic state with m = 0 and Q # 0 and obtains for

low D and high y. The full phase diagram, which can

be obtained by a Landau-like analysis of equation (26) is shown if figure 2. We note the identity of the a

=

0 and a 7r lines, and that for

2

any a - 2 , for sufficiently low D the planar state is

favoured. The phase boundaries are of continuous transitions.

Fig. 2.

-

Mean field phase diagram for the Hamiltonian

Eq. (18).

The superposition approximation implies long-

range order which, for phases of continuous broken symmetry, must be destroyed by quantum fluctua-

tions. We discuss the numerical diagonalization of

finite systems in the next section ; this allows calcu- lation of the details of the algebraic order that may

replace true long range order.

5. Analysis of finite size systems.

We consider chains of spins with periodic boundary

conditions. A chain of N spins is a system of 5N states that can be diagonalised exactly. In practice

exploiting symmetries reduces the size of the mat- rices to be diagonalised. We extrapolate properties

of finite systems as N increases in order to predict

the properties of the infinite system. In our calcu- lations 4 , N -- 10. Information is sought as to the

nature of the ground state as D and y are varied. It is necessary to identify the phases and properties thereof, locate the phase transitions between differ-

ent phases, and estimate the critical exponents and

related properties associated with the phase tran-

sitions.

The eigenstates of the model can be classified by a

few quantum numbers [5]. These are M = ¿ hi, the

total spin in the z-direction, the momentum k and

p, which is ± 1 according to whether the state is even or odd under left-right reflection. For M

=

0 there is also the time reversal operator cr.

The phases can be defined from the degeneracy of

the ground state in the thermodynamic limit. As in the mean field approximation three phases are

discerned.

a) The singlet phase, in which the ground state is a non-degenrate M

=

0, k

=

0, cr = + 1, p

=

+ 1 state, for large D predominantly the configuration (0, 0, 0, ... 0) .

b) The (normal) planar phase, again as for the spin 1 planar ferromagnet [5] there are for k

=

0,

-

IN- degenerate states for an N-spin system ; these include levels with M

=

0, ± 1, ± 2, etc.

c) The nematic planar phase, which is similar, but

the degeneracy has been partially lifted. The degen-

erate states have M

=

0, ± 2, ± 4, etc.

We concentrate on two quantities, AEi = Eo(M = 1) - Eo and AE2 = EO(M = 2) - . Eo, where Eo is the ground state and the brackets

denote orthogonal subspaces of the quantum num- bers cited. We recall that in general (to leading order), whereas in the singlet ground state AYE [] ]

remains finite and in a fully ordered state AE [N ] -- exp - AN, in an algebraically ordered phase

AE[N] - N-1.

In fact it is more convenient to examine the

Callan-Symanzik {3 -function

where X

=

D or y is a coupling constant. Roomany

and Wyld introduced [13] a finite system approxim-

ant for {3 i

where i

(8)

573

We also recall that, in the singlet state, AE is the quantum analogue of §-1 , where 6 is the correlation

length. Close to the phase transition to an ordered

x - -1

p hase at Xc, AE- X - Xc ) and jS = I X 1-1

.

phaseatXc,IlE-IX-Xclvand13= xc 1

In figure 3 we plot j3i against tan y for D

=

0.

Along this line 16 2

=

0 to within the accuracy of the

plot (± 0.02). The results, extrapolated to N

=

oo,

are consistent with the conclusion that j31 = 0 for

y l’ c = tan-l (1.3) with v = 1.25 ± 0.25. This

identifies, as expected, the usual planar phase in the region y yc, and identifies these phases as having algebraic order. For the exponent v i along the line

D

=

0 we find v = 1.0 ::!:: 0.25 consistent with the

Ising value v 1=1.

Fig. 3.

-

The beta function fJfN’ (1’) along D = 0.

N’

=

N + 2, N

=

6 dot-dash, N = 8 dash, N = 10 full line.

In the algebraically ordered phases the exponents

1 l and ’TI2 are of interest. These are calculated

following a method originally postulated by Luck [14] and shown by Cardy [15] to be justified by

conformal invariance. This was applied to quantum spin chains by Schulz and Ziman [6] and indepen- dently to a quantum problem of different symmetry by von Gehlen et al. [16].

where vi is the velocity of sound corresponding to

the relevant Goldstone mode. This can be calculated from the energy difference AEi (N, k ) between the

ground state which has k

=

0 and the lowest excited state of wave number k

=

2 ’TT / N yielding

vi = N âEi(N, k)/2 ’TT.

We identify the phase boundary along the (well- understood) planar line ( y

=

0), using evidence

from the behaviour of f3 i and qi. These results are

shown in figure 4. We identify the phase boundary

between the planar disordered phases from the point

at which 1 4,ql = 1. n i

=

4 772

=

For D--0.27 dif-

ferent size systems give stable estimates of qi i and the relation TJ 2 = 4 TJ 1 is satisfied. At D = 0.27 ,q 1 - 0.23. Different measures of q i suggest the

Fig. 4.

-

fl£’’(y ) and "f1m/m2, m

=

1, 2 along y

=

0.

Curve 1: (31’ curve 2: 132 (scale on the right) ; N

=

6,

N=8. Curve a : q i (N = 8 ) ; b: q i (N = 6 ) ; c:

"f12/4 (N = 8); d: T,,/4(N=6).

phase transition occurs at D

=

0.28 ± 0.02. Beyond

this point application of equation (31) to systems of different sizes gives increasingly inconsistent results, and the relation 11 2

=

4 11 is no longer good. All this

is consistent with the high D phase no longer being

critical in which case formula (31) no longer applies.

The behaviour of Qi 1 and 13 2 is consistent with

13i=O for D Dc and /3i- (D-Dc)’+a with

Dc

=

0.25 ± 0.01, and a

=

0.5 ± 0.1; the Koster- litz-Thouless result would be cr = 0.5. The various different criteria for identifying the phase boundary give close but not identical results.

We show in figure 5 phase boundaries over the whole range of D and 0, constructed using the

criteria described above. The general form of the

phase diagram is unambiguous, and is as predicted

from the mean fied picture, although fluctuations

cause the ordering transition to occur at lower D. A number of features are particularly noteworthy.

First we comment on the low D behaviour close to y

=

7T /2. Mean field theory predicts that this will be

a planar phase, with nematic behaviour only occur- ring for high D. The finite size analysis shows that

Fig. 5.

-

Phase diagram for the Hamiltonian Eq. (18).

Full line : limit of planar ordered phase from

,B = 0 (N

=

6). Dashed line : limit of nematic phase

from IB2=0 (N = 6 ).

(9)

for y > tan-1 (1.3 ) at D

=

0 planar order is des-

troyed and only nematic order remains. Secondly,

there is a strong suspicion of reentrant behaviour in

the planar-nematic phase boundary ; in the inter-

mediate y regime the low D phase appears to be the nematic phase, and as D increases one passes

successively through planar and disordered phases.

This is somewhat contrary to intuition and mean

field theory, for which the (more ordered) planar phase occurs at lower D than the (less ordered)

nematic phase. Finally, close to what is presumably

the multicritical point of coexistence of the dis- ordered, planar and nematic phases there appears to be anomalous behaviour. The phase boundary be-

tween the nematic and the nematic and disordered

phases, as determined by the calculated 132, appar-

ently occurs in a region which is, according to the

calculated 13 1, unambiguously in the planar ordered regime. However as earlier discussed one believes in

general that planar order enforces nematic order. To cast further light on this apparent contradiction, we

show in figure 6 some results for 8 1 and {32 as

functions of D in this region of the phase diagram.

For D

=

0.65 and D

=

0.7 the results on 132 are

consistent with a low D nematic ordered phase, although the exact position of the transition is somewhat ambiguous. The 61 I results are more

difficult to interpret. However the N dependence suggests that for large N limiting behaviour in which for D D,l f31 -- (Del - D), for Del D Dc2 /3

=

0, and for D > De2 81 - (D - De2)3/2 would be

consistent with the available data. The apparent

Fig. 6.

-

i3 m (D) for (i) y

=

0.65 7T/2 (ii) y

=

0.7 7r/2.

Curves (a) N = 4, N’ = 6, m = 1 ; (b) N = 6, N’ = 8,

m = 1 ; (c) N

=

4, N’ = 6, m

=

2 ; (d) N

=

6, N’ = 8,

m = 2.

inconsistencies in results are a result of the proximity

to the multi-critical point, and the consequent slow convergence of 8 NN’.

6. Conclusions.

We have demonstrated that the classical mixed XY model may be studied in the critical phase by mapping it onto a quantum spin chain of spin 2.

While it is appealing to relate properties of liquid crystals to those of quantum magnets, one might ask

is there any real practical advantage in doing this ?

In fact there are several. As is frequently the case

when analogies are found new approximations are suggested : here we have displayed a mean-field theory in the spin language which is distinct from the molecular field theory of the classical picture. Fur-

thermore these are numerical consequences : while it is possible, and indeed has been found useful, to study the statistical mechanics of the model by

Monte Carlo on the model as defined directly, the mapping onto a spin chain suggests different numeri- cal procedures that may give supplementary infor-

mation. By restricting the expansion to the lowest

five levels, the phase space of a single spin is reduced

from an infinite to finite dimension, with no change

in the universal scaling properties. Within this re-

stricted space of fluctuations we are able to diagonal-

ise the transfer matrix exactly for up to 10 spins. This corresponds to finding the exact partition function

for periodic strips infinitely long in one direction and

up to 10 wide in the other. Furthermore this geometry is a particularly propitious one, for confor- mal invariance [15, 17] provides one with a detailed scaling theory for the finite size corrections. Nat-

urally this could have been done without reference

to the spin 2 truncation but reformulation in this

language allows us more direct comparison with

other parallel work. The details of solving the

transfer matrix for such a system would involve

solving a complicated integral equation rather than a

matrix problem that is easily manageable. There are

consequences of the operator algebra that have not yet been fully exploited in the work reported here

that may be expected to increase the accuracy of the results in the future. With the procedures followed

to date we have not unambiguously defined the topology close to the multicritical point. For some

classes of spin chains it was found to be possible to identify multicritical points to various spin Hamilto-

nians which are exceptionally integrable via the

Bethe Ansatz [6, 18], it is to be hoped that the same

program will clarify the multicriticality here. It is probably true that until this point is fully understood

that numerical results, both of the kind presented

and those found by Monte Carlo studies of the

original statistical mechanical model will defy clear

interpretation.

(10)

575

Having made the translation to the spin language

one way to understand the critical phases is by

fermion representation of the spin operators, as

proposed by Timonen and Luther [19] for the spin

one chain. Calculation of correlation function func- tions by such methods checks the equivalence of the spin chain and the original lattice formulation for

long wavelengths. For example, while use of the

transfer matrix breaks the symmetry between the

two lattice directions of the original model, in the massless phases this symmetry is restored. Numeri-

cally this is confirmed by the fact that both phases

have finite spin wave velocities (although the veloci- ties are calculated with respect to excitations of different symmetry). The fermion representation,

valid in the perturbative regime, gives critical corre-

lations that are explicitly Lorentz invariant. In other words the correlations could be made isotropic simply by rescaling the time direction with the calculated velocity.

Although the details of such a continuum theory

remain to worked out, we can already make some

remarks useful for comparison to other results. For the spin-one chain with large negative anisotropy D

there is a transition, discussed by one of us and

Schulz [6] from the normal planar phase (XY 1) to

one with effective spin 1 (XY 2). 2 The transition

between these two massless phases, like that from

ferromagnetic to nematic in the present study, is

believed to be Ising-like and the mechanism must be similar : in that case the (S+)2 operator remains

massless while the S+ acquires a gap. This occurs because only the single power includes a contribution from the disorder parameter of the Ising part of the Hamiltonian. This leads to exponential decay in

correlation functions S+ S- but not (S+ )2 (S- )2.

The Ising ordering, or the disordering of its dual, is analogous to the vanishing of the string tension in

the classical language of Lee and Grinstein. There

are important differences that limit the resemblance to the present situation, however. The XY 2 phase

has the correlations of an effective spin 1 2 Hamilto-

nian and is unstable to formation of an ordered

antiferromagnetic phase at a value of 11 2 = 1, i.e. the

same value taken at the boundary of the normal

planar phase. Therefore the transition to the antifer-

romagnetic state can be considered to be driven by

the same vortex operators that disorder the planar

state. In contrast, the nematic phase has the corre-

lations of an integer system and is unstable for

’q 2- 1 This means that in the continuum limit of 4

this model there must be a different operator driving

the transition. This half-vortex operator can be identified from the comparison to the a = 1 line and

the transformation of section 2 from operators b to

c. It is distinct from the vortex operator that becomes relevant in the ferromagnetic-paramagnetic boundary, just as geometrically the half-integer

vortices are distinct from the integer vortices.

Acknowledgements.

We would like to thank John Chalker for useful conversations and communicating unpublished re-

sults. We are grateful to the Lady Davis Foundation for supporting T.J.S. at the Technion. We thank Ph.

Nozieres and the Theory college of the ILL for

facilitating this collaboration. Computer programs

are based on those developed previously with H. J.

Schulz to whom we are most grateful. We thank J.

Solyom for helpful discussions. This work was sup-

ported in part by the National Science Foundation under Grant No. DMR-85-20190 at Rutgers and by a grant from the New Jersey State Commission on

Science and Technology supercomputer fund.

References

[1] KORSHUNOV, S. E., Pis’ma Zh. E. T. F. 41 (1985) 216,

Sov. Physics JETP Letts 41 (1985) 263.

[2] LEE, D. H. and GRINSTEIN, G., Phys. Rev. Lett. 55

(1985) 541.

[3] STOECKLY, B. and SCALAPINO, D. J., Phys. Rev. B

11 (1975) 205.

HERTZ, J., Phys. Rev. B 14 (1976) 1165.

LUTHER, A. and SCALAPINO, D. J., Phys. Rev. B 16 (1977) 1153.

DEN NIJS, M., Physica 111A (1982) 273.

[4] KOSTERLITZ, J. M. and THOULESS, D. J., J. Phys. C

5 (1972) L124 ; J. Phys. C 6 (1973) 1181.

[5] BOTET, R. and JULLIEN, R., Phys. Rev. B 27 (1983)

613.

SÓLYOM, J. and ZIMAN, T. A. L., Phys. Rev. B 30 (1984) 3980.

[6] SCHULZ, H. J. and ZIMAN, T. A. L., Phys. Rev. B 33 (1986) 6545.

SCHULZ, H. J., Phys. Rev. B 34 (1986) 6372.

ZIMAN, T. A. L. and SCHULZ, H. J., Phys. Rev. Lett.

59 (1987) 140.

[7] ANDREEV, A. F. and GRISCHUK, I. A., Zh. E. T. F. 87

(1984) 467, Sov. Phys. JETP 60 (1984) 267.

[8] PAPANICOLAOU, N., Phys. Lett. 116A (1986) 89.

[9] LEE, D. H., GRINSTEIN, G. and TONER, J., Phys.

Rev. Lett. 56 (1986) 2318.

[10] JANKE, W. and KLEINERT, H., Phys. Rev. Lett. 57

(1986) 279.

(11)

[11] POLYAKOV, A. M., Phys. Lett. B 59 (1975) 79.

[12] MUNRO, R. G., J. Phys. C 9 (1976) 2611.

MUNRO, R. G. and GIRARDEAU, M. J., J. Magn.

Magn. Mater. 2 (1976) 319.

MARITAN, A., STELLA, A. and VANDERZANDE, C., Phys. Rev. B 29 (1984) 519.

NOSKOVA, L. M., Fiz. Tver. Tela 25 (1983) 2471, Sov. Phys. Solid State 25 (1983) 1418.

[13] ROOMANY, H. H. and WYLD, H. W., Phys. Rev. D

21 (1980) 3341.

[14] LUCK, J. M., J. Phys. A 15 (1982) L169.

[15] CARDY, J., J. Phys. A 17 (1982) L385.

[16] GEHLEN, G. V., RITTENBERG, V., RUEGG, H., J.

Phys. A 19 (1985) 107.

[17] BELAVIN, A. A., POLYAKOV, A. M. and ZAMOLOD- CHIKOV, A. B., Nucl. Phys. B 241 (1984) 333.

CARDY, J. L., J. Phys. A 19 (1986) L1093.

CARDY, J. L., Nucl. Phys. B 270 (1984) 186.

[18] AFFLECK, I., Nucl. Phys. B 265 (1986) 409.

AFFLECK, I. and HALDANE, F. D. M., Phys. Rev. B

36 (1987) 5291.

[19] TIMONEN, J. and LUTHER, A., J. Phys. C 18 (1985)

1439 ; see also SCHULZ, H. J., reference [6].

Références

Documents relatifs

Constraints on spin observables coming from discrete symmetries such as P, C, T and identical particles may be divided in two types: 1) classical ones, which insure the invariance

Abstract.- The Monte Carlo method has been used to calculate the moments of the relaxation function for the classical planar spin chain.. In this way extended results for the

The class containing just isotropic Heisenberg and biquadratic exchange is particularly interesting be- cause as well as the model studied by Haldane, it con- tains

- The ground state of quantum spin chains with two spin-112 operators per site is determined from finite chain calculations and compared to predictions from the continuum limit..

(1): Nouara hocine les intellectuelle algerienes mouvance et anamophone education algerien mythe mouvace et anamorphone eddition dahlab et ENAG alger 2005 p33... 27

Ces 4 souches ont été isolées chez 3 patients tous porteurs de sonde de cystostomie depuis 5 jours chez un sujet porteur d'un cancer de la vessie, depuis 40 jours chez un sujet

Afin de pouvoir appréhender le contexte territorial dans lequel prend place l’ensemble des activités agricoles, artisanales et commerciales, nous nous attachons

Régulation épigénétique de l’expression de FOXL2 et voies activées en aval de ce gène dans la gonade.. Thèse de doctorat de