• Aucun résultat trouvé

Layer fluctuations and hexatic order in liquid crystals

N/A
N/A
Protected

Academic year: 2021

Partager "Layer fluctuations and hexatic order in liquid crystals"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00210819

https://hal.archives-ouvertes.fr/jpa-00210819

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.

Layer fluctuations and hexatic order in liquid crystals

Jonathan V. Selinger

To cite this version:

Jonathan V. Selinger. Layer fluctuations and hexatic order in liquid crystals. Journal de Physique,

1988, 49 (8), pp.1387-1396. �10.1051/jphys:019880049080138700�. �jpa-00210819�

(2)

E 1387

Layer fluctuations and hexatic order in liquid crystals

Jonathan V. Selinger

Department of Physics, Harvard University, Cambridge, MA 02138, U.S.A.

(Requ le 14 octobre 1987, accepté sous forme définitive le 25 avril 1988)

Résumé.

2014

Dans les phases smectique-A et hexatique-B les couches ne sont pas planes mais présentent des

fluctuations de grande amplitude. Afin de déterminer l’effet de ces fluctuations sur la transition smectique-A- hexatique-B nous construisons un hamiltonien de Ginzburg-Landau qui dépend du paramètre d’ordre hexatique 03C8 et des fluctuations u. Le couplage des champs 03C8 et u est dû à la frustration d’origine géométrique

introduite par la courbure des couches. Nous intégrons les fluctuations u pour obtenir un hamiltonien effectif

en fonction de 03C8. Cette intégration produit une renormalisation finie des coefficients de l’hamiltonien effectif.

Par ce mécanisme, la fluctuation des couches peut transformer la transition smectique-A-hexatique-B en une

transition du premier ordre si les constantes de rigidité hexatiques sont assez grandes. Cet effet est peut-être apparenté aux anomalies de la chaleur spécifique observ6es expérimentalement à cette transition.

Abstract.

2014

In the smectic-A and hexatic-B phases, the layers are not planar but undergo large fluctuations. In order to find the effect of these fluctuations on the smectic-A-hexatic-B transition, we construct a Ginzburg-

Landau Hamiltonian in terms of the hexatic order parameter 03C8 and the layer fluctuations u. The fields 03C8 and u are coupled because of the geometrical frustration introduced by the curvature of the layers. We integrate out the u field to obtain an effective Hamiltonian in terms of 03C8 alone. This integration leads to a finite

renormalization of the coefficients in the effective Hamiltonian. By this mechanism, layer fluctuations may drive the smectic-A-hexatic-B transition to be first-order if the hexatic stiffness constants are sufficiently large.

This effect may be related to observed anomalies in the specific heat at that transition.

J. Phys. France 49 (1988) 1387-1396 AOÙT 1988,

Classification

Physics Abstracts

64.70M - 61.30

-

64.60C

1. Introduction.

Three-dimensional (3D) layered liquid crystal sys- tems can exhibit several phases with different types of in-plane 2D order. The least-ordered layered phase is the smectic-A phase, which has short-range

bond-orientational and positional order within each layer. At lower temperature the system can form the hexatic-B phase, which has long-range bond-

orientational order but only short-range positional

order within each layer [1-3]. At even lower tempera-

ture the system freezes into a crystalline phase, with long-range bond-orientational and positional order.

In the bulk smectic-A and hexatic-B phases, the

molecules do not lie in static, planar layers ; rather,

the layers are constantly fluctuating. In any layered system without long-range in-plane positional order,

the magnitude of the Landau-Peierls [4, 5] fluctu-

ations of the layers diverges logarithmically with the

system size [6, 7]. The smectic-A-hexatic-B ordering

transition therefore takes place in a system of

strongly-fluctuating layers. It is interesting to con-

sider the effects of these strong fluctuations on the smectic-A-hexatic-B transition.

In a recent paper Nelson and Peliti [8] derived the elastic Hamiltonian of a single curved hexatic mem-

brane. They showed that the Hamiltonian must

involve the covariant derivative of the bond-angle

field 0 (x ) to account for the frustration due to the membrane curvature. In this paper we extend their

theory to the stacked system of curved layers in a

smectic-A or hexatic-B liquid crystal. We thereby

obtain a coupling between the coarse-grained hexatic

order parameter .p (x) = e6 i 8 (x» and the layer

fluctuations u (x). This Hamiltonian involves a vector

potential term which is similar to the coupling

between the superconductor order parameter and the electromagnetic field in the normal-supercon-

ductor transition, and to the coupling between the

smectic order parameter and nematic director fluctu- ations in the nematic-smectic transition. Halperin, Lubensky, and Ma [9, 10] have studied how fluctu-

ations in the electromagnetic field or the nematic director field affect the latter two transitions. By

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:019880049080138700

(3)

integrating out these fluctuations, they show that

these transitions are driven weakly first-order. In this paper we perform the analogous calculation for the smectic-A-hexatic-B transition : we integrate out

the u field to obtain an effective Hamiltonian in terms of qi alone. In this case, integration of the

u field leads only to a finite renormalization of the coefficients of the terms in the effective Hamiltonian.

Coupling with layer fluctuations may or may not

cause the smectic-A-hexatic-B transition to be first-

order, depending on the size of the coefficients. In

particular, the transition will be driven first-order if the hexatic stiffness constants are sufficiently large.

This coupling between hexatic order and layer

fluctuations may help to explain observed anomalies

in the specific heat at the smectic-A-hexatic-B tran-

sition. Experiments have studied this transition in several pure and binary liquid crystal systems [11- 17]. In some the transition was found to have a large specific heat exponent a = 0.48 to 0.67. In others the transition showed some thermal hysteresis and

was judged to be weakly first-order. The large value

of a is close to the value a = 0.5 that one would expect from a nearby tricritical point. For systems with herringbone packing order, Bruinsma and Aep- pli [18] have attributed the tricritical point to coupl- ing between the hexatic and herringbone order

parameters. For systems without herringbone order, Aharony et al. [19] have argued that a tricritical

point should exist on the smectic-A-hexatic-B line

near the smectic-hexatic-crystal triple point because

of coupling between the hexatic and crystalline order parameters. The coupling between hexatic order and

layer fluctuations discussed in this paper provides

an additional mechanism for the transition to be driven first-order. If the tricritical point hypothesis is

correct, then this coupling should move the tricritical

point farther from the triple point than one would expect from the argument of Aharony et al.

In principle the theory in this paper should also be relevant to the smectic-A-smectic-C transition. That transition is described by the complex order par- ameter 0 = w e where w is the angle of the tilt away from the layer normal and 0 is the azimuthal angle [20]. By the argument of Nelson and Peliti, the Hamiltonian should involve covariant derivatives of 03A6 , and the analysis of this paper should go through exactly as in the hexatic case. However, experiments

have shown that the smectic-A-smectic-C transition is well described by a simple mean field theory in

03A6 with an unusually large sixth-order term [21-23].

To the best of our knowledge, no experiments have

seen a weakly first-order transition or an anomalous

peak in the specific heat that could be due to

coupling between the tilt order parameter and layer

fluctuations. This coupling must be too weak to

drive the smectic-A-smectic-C transition first-order, probably because smectic-C stiffness constants are

typically much weaker than the corresponding hexa-

tic stiffness constants [24].

The plan of this paper is as follows. In section 2 we

extend the Nelson-Peliti theory of a single hexatic

membrane to a 3D stacked hexatic liquid crystal and

obtain expressions for the intralayer and interlayer coupling. In section 3 we integrate out layer fluctu-

ations to obtain an effective Hamiltonian for the hexatic order parameter t/1, and we show that the

coupling with layer fluctuations can cause the smec-

tic-A-hexatic-B transition to be first-order if the hexatic stiffness constants are large enough.

2. Coupling between hexatic order and layer fluctu-

ations.

In any layered liquid crystal system, each molecule is surrounded by (on average) six nearest neighbours

in the same layer. In the hexatic-B phase, there are long-range correlations in the orientations of the

nearest-neighbour bonds. To describe this phase,

the Hamiltonian must contain a term that tends to

align the bond orientations at different points in the system. To construct such a term, one must define the local bond angle field, and one must find a

method for comparing the bond orientations at different points. If the layers are planar, then one

can measure the angles of all the bonds relative to a

single fixed axis (e.g. the x-axis) in the layer planes,

and thereby obtain a bond-angle field 0 (x) defined

modulo 2 7r/6. One can compare the bond orien- tations at x and x’ by simply taking the difference

0 (x’) - 0 (x). The hexatic elastic Hamiltonian is therefore

where z is the coordinate normal to the layers. In general the elastic coefficients KA and KN for

variations within and normal to the layers may be different.

If the layers are not planar, then it is much more

difficult to construct the hexatic elastic Hamiltonian.

The problem is illustrated in figure 1. At any given point on a layer, the six nearest-neighbour bonds lie

in the local tangent plane to the layer, not in the xy-

plane. Because the tangent plane varies from point

to point, one cannot measure the bond angles in a single plane relative to a single fixed axis, and one

must find another way to compare the bond orien- tations at different points. Nelson and Peliti [8] have

studied this problem for a single 2D hexatic mem-

brane. They use parallel transport to compare the bond vectors at different points on the same mem- brane, and they arrive at a hexatic elastic Hamilto- nian involving the covariant derivative of the bond-

angle field. This parallel-transport method is not

adequate for a 3D stacked hexatic system because

(4)

1389

Fig. 1.

-

Comparison of the bond orientations in the tangent planes at x and x’. The vector v lies in both tangent planes.

one cannot parallel transport bond vectors from one layer to another. In this section we develop a general

method for comparing the bond orientations at

different points in the same or different layers, and

we use it to construct the elastic Hamiltonian of a

stacked hexatic system. In the appendix we show

that our method is equivalent to parallel transport in the case of a single hexatic membrane.

Consider a layered liquid crystal system with the layers specified by

where n is an integer, d is the equilibrium interlayer spacing, and u (x) is the layer displacement field. At

any point on any of the layers, there is a normal

vector and a tangent plane. The normal vector is

and a natural choice of basis for the tangent plane is

This basis is not orthonormal because

In this paper we will use the convention that Greek indices run over the values 1 and 2, and Latin indices

run over the values 1, 2, and 3, unless otherwise noted. We can define an orthonormal basis for the tangent plane by

where c(x) and è(x) are chosen so that

Consistency of (2.6a) and (2.6b) implies

and (2.7) implies

It is straightforward to solve (2.8) and (2.9) to obtain

exact expressions for c(x) and c (x ), valid in any coordinate system. At a later stage of this calcu-

lation, it will be convenient to work in a coordinate system in which z is the average direction normal to the layers and Vu is small. At lowest order in

Vu,

This approximation for c(x) and c (x ) gives e03B1 to

second order in Vu.

If we neglect disclinations, then each molecule in each layer has six nearest-neighbour bonds in the six tangent directions given by

The field 0 (x) specifies the bond angles in the local tangent plane relative to the locally-defined axis ê1. Now consider the bond configurations at two nearby points x and x’, which may be in the same or different layers. The geometry is shown in figure 1.

We cannot compare the bond configurations by simply subtracting 0 (x’) - 0 (x), because those ang- les are defined in different tangent planes relative to

different axes. To compare them, we must use a

common reference direction in both tangent planes.

To find a reference direction, note that the two tangent planes have a common tangent vector v, which is unique up to scalar multiplication. The

vector v can be expressed as

in the orthonormal bases at x and x’. We can use v as

a benchmark for comparing the bond orientations at

(5)

x and x’. The angle 0 (x) - 0 gives the bond

orientation relative to v in the tangent plane at x,

and 0 (x’) - .0’ gives the same at x’. If

0 (x) - 0 = 0 (x’ ) - 0’, then we can map the bond

configuration at x onto the bond configuration at

x’ by a single rotation about v, and the bond

configurations are as closely aligned as they can be.

We hypothesize that there is an energy penalty for disalignment of

for d0 = O(x’)-O(x) and d0 =0’-O both

small. Note that 0 is not just a function of x ; rather, cp and -0’ are both derived from v, which is derived from the intersection of the tangent planes at x and

x’. In the appendix we show that comparison of the

bond orientations with respect to v is equivalent to comparison by parallel transport if x and x’ are in the

same layer.

Because of the So term, the hexatic elastic Hamiltonian couples the bond-angle field 0 (x) and

the layer fluctuations u (x ). To see this coupling explicitly, we must derive expressions for v and 5 0 in terms of u. We can let v = n x n’, where n

and n’ are the normal vectors at x and x’ respectively.

In terms of the natural bases for the tangent planes

at x and x’ we have

where ððiU = aju (x’) - aju (x) and Ea{3 is the anti-

symmetric symbol with E12 =1. In the limit of small ðx = x’ - x, 5 0 is determined by

At this point we can greatly simplify the calculations

by choosing a coordinate system in which z is the average direction normal to the layers and Vu is

small. If we combine equations (2.15), (2.14), (2.6)

and (2.10), and expand all quantities to lowest order in Vu, we obtain

For small 6x, we can write

where

or in standard 3D notation

A is a vector potential for the bond-angle field in this system. Because 0 is not just a function of x, A is not

necessarily the gradient of a scalar function.

We have argued that the energy penalty for

variations in 0 (x) between x and x’ should be

proportional to [ (V 0 + A ) . SX]2 . The coefficient may be different for intralayer and interlayer vari-

ations. We can therefore use the Hamiltonian

We have shown that A is of order (Vu )2. If we

assume that VO is also of order (ou )2 (we will check

this assumption self-consistently in the next section),

then we can neglect the difference between n and z to obtain the Hamiltonian

valid to order (Vu )4. If we rescale distances in the z-

direction by a factor of eo = .J KN/ KA, the Hamil-

tonian simplifies yet further to

where KA has been implicitly rescaled.

So far we have considered systems that are well inside the hexatic phase and hence have well-estab- lished hexatic order. In these systems the coarse-

grained hexatic order parameter t/J (x) = e 6 i 0 (x))

has constant amplitude. At higher temperatures, close to the smectic-A-hexatic-B transition, we must

consider variations in the amplitude as well as the phase of 03C8. In that case the Hamiltonian becomes

The constants KA and KA are related by

KA = KÁ 1 t/J 2 at low temperature.

(6)

1391

3. Integration of layer fluctuations.

We have derived a geometrical coupling (2.22)

between the hexatic order parameter w and a vector potential A due to layer fluctuations. This coupling is analogous to the coupling between the superconduc-

tor order parameter 4/ and the electromagnetic

vector potential A in the normal-superconducting

transition. It is also analogous to a coupling term in

the nematic-smectic-A transition, where w corres- ponds to the amplitude and phase of the smectic

density wave and A to the nematic director field.

Halperin, Lubensky, and Ma [9] [10] have shown

that the critical properties of the normal-supercon- ducting and nematic-smectic-A transitions are ident- ical : in each case, fluctuations in the A field drive the transition from w = 0 to qi :0 0 to be weakly first-order, at least in 4 - E dimensions. Although

the smectic-A-hexatic-B transition is similar to the other two transitions, it is not exactly equivalent to

them because in this transition the vector potential A

is really a composite of derivatives of the u field. It is therefore interesting to ask what effect fluctuations in A have on the smectic-A-hexatic-B transition. In this section we answer that question by integrating

out fluctuations in A to obtain an effective Hamilto- nian in terms of the hexatic order parameter 03C8 alone.

As a first step in this calculation, we must construct the full Hamiltonian for the hexatic order parameter and layer fluctuations. In addition to the coupling

term (2.22) developed in the last section, the Hamil-

tonian must contain (a) a Ginzburg-Landau expan- sion in It/J 2 and (b) the energy of layer fluctuations

u (x ). The Ginzburg-Landau expansion can easily be

written as

Grinstein and Pelcovits [25] have shown that the lowest-order rotationally-invariant form for the en-

ergy of layer fluctuations is

where distances in the x, y, and z-directions are measured in the same units. If we work in a

coordinate system in which z is the average direction normal to the layers, then the cubic and quartic

terms in Vu will be small. Grinstein and Pelcovits have shown that these terms give only a logarithmic

renormalization of the elastic constants B and

Kl at long wavelengths. We will therefore neglect

these terms for the purpose of assessing the signifi-

cance of the coupling between qi and u. Further-

more, we can neglect the a2U terms in comparison

with the a,u term at long wavelength. If we now

rescale distances in the z-direction by a factor of

= J KN/ KA, and implicitly rescale all coef- ficients in (3.1) and (3.2) appropriately, we obtain

the total Hamiltonian

For any specific realization of the layer fluctu-

ations u (x), the state qi (x) = constant does not

minimize the Hamiltonian. This fact is not surpris- ing : if the layers are curved, there is no reason to expect the state in which all bond angles are aligned

with the locally-defined axis ê1 to be the ground

state. Suppose the amplitude of hexatic order

141 (x) I is constant in space. To find the ground state

of the bond-angle field, we write the vector potential

as

where P j = 6ij - °iOj/V2 is the 3D transverse pro-

jection operator. The ground state occurs when the phase of qi (x) satisfies 0 (x) + eo L (x ) = constant.

We therefore make a gauge transformation [10] by defining

Variations in 0 (x ), the phase of IV (x), give the

fluctuations in the bond-angle field away from the

ground state for any specific realization of u (x). In

terms of 03C8 (x), the Hamiltonian becomes

The transverse part of A gives an extra energy that cannot be reduced by any configuration of the bond-

angle field. It gives a geometrical frustration in the hexatic phase.

Because the state 11 (x) = constant minimizes the

Hamiltonian, we use mean field theory in 1/1’ (x) and neglect fluctuations in 0(x). In this mean field theory VO is of order A, which is of order

(Vu )2. In physical terms, this mean field theory

assumes that the correlation length for the amplitude

of the hexatic order parameter is much greater than

the correlation length for layer fluctuations. In the

(7)

analogous superconductor problem, this assumption corresponds to the assumption that the coherence length 6 is much greater than the London pen- etration depth A, or that the superconductor is type I. We will discuss the validity of this approxi-

mation below.

To obtain an effective Hamiltonian as a function

of W alone, we must integrate out the fluctuations in

u by writing

With the neglect of spatial fluctuations in 1/1’ (x), equation (3.7) implies

where fi is the system volume. We must now calculate the expectation value (f d3x PT A 2 as a function

of 1/1’. In contrast with the superconductor problem, we cannot evaluate this expectation value exactly

because of the quartic term in u that is implicit in 1 p T A 12. Rather, we must use perturbation theory in

Kg e)[ W [ 03C8. Let Ho represent the first two terms in equation (3.6), and write the coupling term as

r

_

.1 f

The expectation value is then

where (. ) o denotes the average with respect to the Gaussian weighting function e - f3Ho.

We can express equation (3.10) as a graphical perturbation series. Let the vertex in figure 2 represent

Fig. 2.

-

Vertex corresponding to the operator

Each slash over a leg represents a factor of momentum, and the dashed line indicates how to contract the momenta with the transverse projection operators. At zero-th order in KA e 0 21 T 2, the only non-vanishing

contributions to

(f d3x I p T A 12) V, are the graphs in figure 3, which represent

(8)

1393

Fig. 3.

-

Contribution to

( d’x I pTA 12 ) w at zero-th order in Kg e) [ W [.2

This term is IR convergent in 3D, and it can be made finite by the use of a UV cutoff. When inserted into

equation (3.8), this term gives a finite correction to the coefficient a in Heff (03C8).

The contribution to

( f d 3X I pT A12 ) W at first order in KÁ eJI1Ji’ 12 is

which is explicitly negative. This contribution is

represented by the graphs with the loop structures

shown in figure 4. Because of the transverse projec-

tion operator, none of these graphs factor into

independent integrals. By power counting, all of

these graphs are IR convergent in 3D. When inserted into equation (3.8), this contribution to

(f d3x I pT A 12) gives a finite, negative correction

to the coefficient b in Heff (AP ).

Fig. 4.

-

Loop structures of the contributions to

f d’x I pT A 12) at first order in K, A e21 0 lp 2.

In the normal-superconducting and nematic-smec- tic-A problems, integration of fluctuations in the vector potential generates a non-analytic term in Heff( 1/1’). This term goes as - IT14-, in 4 - e

dimensions, or as + I T I I In 11/1’ I in 4 dimensions.

This non-analytic term is what forces those tran-

sitions to be weakly first-order. If such a term were

present in Heff (T) in the smectic-A-hexatic-B tran-

sition, then a2 H eft/ a ( 11/1’ 12)2 would diverge at

1/1’ = 0, and there would be an infinite, negative

correction to the coefficient b. Because this correc-

tion is only finite, no non-analytic term of this form is present, and Heft ( 1/1’) can be written as

where the subscript R indicates renormalization of a and b. The Halperin-Lubensky-Ma mechanism

therefore does not force the smectic-A-hexatic-B transition to be first-order.

As an aside, we could also calculate how the

coupling term in the Hamiltonian (3.6) renormalizes the elastic constants B and Kl in the hexatic phase, just as Grinstein and Pelcovits [25] calculated how the (ozu) (Vu )2 and (VU)4 terms renormalize B and

Kl. In the case of a single 2D hexatic membrane, Nelson and Peliti [8] have shown that coupling with

hexatic order gives a logarithmically divergent, wavevector-dependent correction to the rigidity

K. In a 3D stacked hexatic, however, the corrections to B and Kl are convergent. Coupling with hexatic

order in 3D gives only a finite renormalization of these elastic constants.

Although the coupling between hexatic order and layer fluctuations does not generate divergent correc-

tions to b, B, or Ki, it may still have important

effects on the smectic-A-hexatic-B transition. If the hexatic stiffness constants KA and KN are sufficiently large, then the coupling may drive bR negative. In

that case, Heff(T) must contain a positive term of

(9)

3CR I T I ’. for thermodynamic stability. We would

then get a first-order transition from the smectic-A

phase to the hexatic-B phase. As usual we assume

that aR OC (T - TO), where To is constant, and

bR and CR are independent of temperature. In mean field theory in 1/1’, the first-order transition occurs at

and the value of the hexatic order parameter just

below the transition is

We can use the Ginzburg criterion [26] to deter-

mine whether fluctuations in 1/1’ invalidate mean

field theory. According to the Ginzburg criterion,

fluctuations have a significant effect only inside a

critical region I aR I - aGe If the hexatic stiffness constants are sufficiently large that bR 0 and

3 bR/16 CR > aG, then mean field theory is self-

consistent. In that case the second-order transition at aR = 0 is preempted by the first-order transition, and aR = 0 is only the limit of metastability of the

smectic-A phase. However, if 3 bR/16 CR aG then

mean field theory is not self-consistent.

It is difficult to make a quantitative estimate of how large the hexatic stiffness constants must be to induce a first-order smectic-A-hexatic-B transition for two reasons. First, all the integrals that contribute

to the renormalization of b are quite UV divergent.

Their numerical values are therefore extremely

sensitive to the choice of the UV cutoffs on

qz and q, . Second, we cannot determine the renor-

malized coefficient bR unless we know the bare coefficient b. This bare coefficient is unobservable.

However, we can make some general comparisons

between different systems. As mentioned in the introduction, the theory in this paper should apply to

the smectic-A-smectic-C transition as well as to the smectic-A-hexatic-B transition. By the argument of section 2, the hexatic-B and smectic-C phases should

have the same coupling between orientational order and layer fluctuations, where the bond-angle field in

the hexatic-B phase corresponds to the azimuthal

angle of the tilt in the smectic-C phase. Because of packing constraints, we would expect the hexatic-B stiffness constants to be larger than the correspond- ing stiffness constants in the smectic-C phase. Recent experiments on thin liquid crystal films confirm that hexatic stiffness constants are much greater than smectic-C stiffness constants [24]. For that reason,

the coupling in this paper is more likely to affect the

smectic-A-hexatic-B transition than to affect the smectic-A-smectic-C transition.

It is interesting to compare the results of our calculations with the observed anomalies in the smectic-A-hexatic-B transition. As noted in sec-

tion 1, this transition is observed to be second-order with a specific-heat exponent a = 0.48 to 0.67 in

some systems, and weakly first-order in other sys- tems. The large value of a in the former case could arise from the influence of a nearby tricritical point,

which Bruinsma and Aeppli attribute to coupling

between hexatic and herringbone order parameters and Aharony et al. to coupling between hexatic and

crystalline order parameters. This tricritical point

must exist in the temperature-concentration phase diagram of the 650BC-40.8 binary system, which shows a second-order transition at low concentration and a first-order transition at higher concentration of 40.8 [13]. However, no tricritical point has been

seen in the temperature-chain length phase diagram

of the homologous series nmOBC, which always

shows a second-order smectic-A-hexatic-B transition with a large value of a [17]. If that system has a tricritical point, it must be unobservably close to the smectic-hexatic-crystal triple point. It is surprising

but possible that such a tricritical point could affect

a for a wide range of chain length. Coupling

between hexatic order and layer fluctuations pro- vides an additional reason for the smectic-A-hexatic- B transition to be driven first-order in a part of the phase diagram where the hexatic stiffness constants are large. It therefore provides an additional

mechanism to generate a tricritical point, which

could account for a large value of a.

Acknowledgments.

I am grateful to D. R. Nelson for suggesting this problem and for many helpful discussions about it,

and to R. G. Petschek for drawing my attention to

errors in a previous version of this paper. This work

was supported by the National Science Foundation

through Grant No. DMR 85-14638 and through the

Harvard Materials Research Laboratory.

Appendix.

In section 2 we develop a general method for

comparing the bond orientations at two nearby points x and x’ in the same or different layers. If the points are in the same layer, then we could also compare the bond orientations by the more standard

method of parallel transport: we could parallel- transport the bond vectors ba along a geodesic from

the tangent plane at x to the tangent plane at

x’ and then compare them with the bond vectors at x’. In the limit of x’ - x -+ 0, tlfis method gives the

covariant derivative of the bond-vector field in the

given layer. Nelson and Peliti [8] use this covariant

derivative to construct the elastic Hamiltonian of a

(10)

1395

single hexatic membrane. In this appendix we show

that comparing bond orientations by the method of section 2 gives the same result as comparing bond

orientations by parallel transport if x and x’ are in

the same layer. Our Hamiltonian (2.21) is therefore

a generalization of the Hamiltonian of Nelson and Peliti.

To see that the two methods of comparing bond

orientations are equivalent, consider a single layer in

the Monge parametrization x(o-) = (o- , 0’

u (o,- 1, Q 2)). The natural basis for the tangent plane

at any point is ta = a a x, where a,, =- d/do-a. In the Monge parametrization, the metric tensor is

and the connection coefficients are

As in section 2, let v be a vector that is in the tangent planes at both x and x’. By the arguments of section 2, it can be expressed as v = v " ta = v’" ta,

where

in the tangent planes at both x and x’.

Now consider the two methods of comparing bond

orientations geometrically. In the method of sec-

tion 2, we use the common tangent vector v as a reference direction to compare the bond angles in

the tangent planes at x and x’. In the parallel- transport method, we draw a geodesic from x to

x’ and use the geodesic as a reference direction in the two tangent planes. Those methods are consist-

ent if the angle between v and the geodesic at x equals the angle between v and the geodesic at

x’. Suppose that the geodesic has tangent vector G at

x = x(o-) and tangent vector G’ at x’ = x (cr + Su).

We then obtain

For small 8 cr , the angle between v and the geodesic

varies only as (8 u )2. Comparison with respect to v

and comparison by parallel transport are therefore, equivalent ways of comparing bond orientations at

nearby points.

We can also show explicitly that the hexatic Hamiltonian of Nelson and Peliti is consistent with the Hamiltonian developed in section 2. We define a

local orthonormal basis ê(IA-)( U ) = e (IA-) ta for the

tangent plane at (1, and let b = cos 0 e (1) + sin 6ê(2) } be any of the six bond-vector fields. If we compare

b at nearby points by parallel transport, we obtain the elastic Hamiltonian

where d2s is the element of area and Vab13 is the

covariant derivative of b. Nelson and Peliti show that this Hamiltonian reduces to

where Aa is defined by

They calculate Eaf3 ðaAf3 and hence the transverse

component of the 2D vector potential A for a general basis ê(JL )(0’). We can calculate both compo- nents of A with a specific choice of ê (JL ) (0’ ). Specializ-

ing (2.6a) to a single 2D layer, we choose e(JL) = SJLa - c(O’) ðJLu ðau. In a coordinate system in which z is the average direction normal to the layers

and au is small, we have c -- 2 . We then obtain

explicitly

and hence

at lowest order in au. The vector potential (A.9) is a

2D special case of the vector potential (2.18). The

3D hexatic Hamiltonian derived in section 2 is therefore a generalization of the Hamiltonian (A.5)

derived from covariant differentiation.

(11)

References

[1] HALPERIN, B. I. and NELSON, D. R., Phys. Rev.

Lett. 41 (1978) 121 ; 41 (1978) 519.

[2] BIRGENEAU, R. J. and LITSTER, J. D., J. Phys.

France Lett. 39 (1978) L-399.

[3] PINDAK, R., MONCTON, D. E., DAVEY, S. C. and GOODBY, J. W., Phys. Rev. Lett. 46 (1981)

1135.

[4] PEIERLS, R. E., Helv. Phys. Acta Suppl. II 7 (1934)

81.

[5] LANDAU, L. D., Phys. Z. Sowjet. 11 (1937) 545,

JETP 7 (1937) 627, Collected Papers of L. D.

Landau, Ed. D. ter Haar (Gordon and Breach,

New York) 1965, p. 209 ;

see also LANDAU, L. D. and LIFSHITZ, E. M., Statisti- cal Physics part 1 (Pergamon, London) 1980, p. 434.

[6] CAILLÉ, A., C. R. Hebd. Séan. Acad. Sci. Ser. B 274

(1972) 891.

[7] ALS-NIELSEN, J., LITSTER, J. D., BIRGENEAU, R. J., KAPLAN, M., SAFINYA, C. R., LINDEGAARD- ANDERSEN, A. and MATHIESEN, S., Phys. Rev.

B 22 (1980) 312.

[8] NELSON, D. R. and PELITI, L., J. Phys. France 48 (1987) 1085 ; 49 (1988) 139 ;

see also DAVID, F., GUITTER, E. and PELITI, L., J.

Phys. France 48 (1987) 2059.

[9] HALPERIN, B. I., LUBENSKY, T. C. and MA, S. K., Phys. Rev. Lett. 32 (1974) 292 ;

see also CHEN, J. H., LUBENSKY, T. C. and NELSON, D. R., Phys. Rev. B 17 (1978) 4274.

[10] HALPERIN, B. I. and LUBENSKY, T. C., Solid State Commun. 14 (1974) 997.

[11] HUANG, C. C., VINER, J. M., PINDAK, R. and GOOD- BY, J. W., Phys. Rev. Lett. 46 (1981) 1289.

[12] ROSENBLATT, C. and Ho, J. T., Phys. Rev. A 26 (1982) 2293.

[13] VINER, J. M., LAMEY, D., HUANG, C. C., PINDAK, R. and GOODBY, J. W., Phys. Rev. A 28 (1983)

2433.

[14] PITCHFORD, T., NOUNESIS, G., DUMRONGRATTANA, S., VINER, J. M., HUANG, C. C. and GOODBY,

J. W., Phys. Rev. A 32 (1985) 1938.

[15] MAHMOOD, R., LEWIS, M., BIGGERS, R., SUREN- DRANATH, V., JOHNSON, D. and NEUBERT, M. E., Phys. Rev. A 33 (1986) 519.

[16] HUANG, C. C., NOUNESIS, G. and GUILLON, D., Phys. Rev. A 33 (1986) 2602.

[17] NOUNESIS, G., GEER, R., LIU, H. Y., HUANG, C. C.

and GOODBY, J. W., preprint.

[18] BRUINSMA, R. and AEPPLI, G., Phys. Rev. Lett. 48

(1982) 1625.

[19] AHARONY, A., BIRGENEAU, R. J., BROCK, J. D.

and LITSTER, J. D., Phys. Rev. Lett. 57 (1986)

1012.

[20] DE GENNES, P. G., The Physics of Liquid Crystals (Oxford) 1974.

[21] HUANG, C. C. and VINER, J. M., Phys. Rev. A 25 (1982) 3385.

[22] BIRGENEAU, R. J., GARLAND, C. W., KORTAN, A. R., LITSTER, J. D., MEICHLE, M., OCKO, B. M., ROSENBLATT, C., YU, L. J. and GooD- BY, J., Phys. Rev. A 27 (1983) 1251.

[23] HUANG, C. C. and LIEN, S. C., Phys. Rev. A 31 (1985) 2621.

[24] DIERKER, S. B. and PINDAK, R., Phys. Rev. Lett. 59

(1987) 1002.

[25] GRINSTEIN, G. and PELCOVITS, R. A., Phys. Rev. A

26 (1982) 915.

[26] GINZBURG, V. L., Fiz. Tverd. Tela 2 (1960) 2031,

Sov. Phys. Solid State 2 (1961) 1824 ;

see also MA, S. K., Modern Theory of Critical

Phenomena (Benjamin/Cummings, Reading,

Mass.) 1976.

Références

Documents relatifs

other hexatic B phases have been identified on the basis of qualitative features associated with the X-ray powder pattern [14], as we are doing here, 650BC and 460BC

amplitude modes in the biaxial phase mix the components of biaxiality and birefringence of the order parameter, the critical (quasi-biaxial) mode presenting

By analogy with the Sn-i,C phase, we can assume that this second network which lies in a plane perpendicular to the layers originates in the periodic modulation of

2014 We study the interplay between crystalline (or hexatic) order and thermal fluctuations in membranes with vanishing surface tension.. If the connectivity of the

2014 The long distance behaviour of a model of fluid membranes with orientational (hexatic) order and.. small surface tension is

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

two-dimensional order (uncorrelated layers) the sam- ple mosaic combined with the two-dimensional fea- tures of the scattering, which consists of rods perpen- dicular to

apply when the membrane is embedded in a space of high enough dimension (a simple estimate yields d>4), it might turn out that the behavior of hexatic membranes in the real