ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/gb-bilayer/bilayer.tex
Revision: 2683
Committed: Mon Apr 3 15:39:57 2006 UTC (18 years, 5 months ago) by tim
Content type: application/x-tex
File size: 27781 byte(s)
Log Message:
Imported using TkCVS

File Contents

# User Rev Content
1 tim 2683 %\documentclass[aps,preprint,showpacs,groupedaddress]{revtex4}
2     \documentclass[aps,floats,twocolumn,showpacs,groupedaddress]{revtex4}
3     \usepackage{epsf}
4     \usepackage{graphicx}
5     \usepackage{times}
6    
7     \begin{document}
8     \bibliographystyle{apsrev}
9    
10     \title{Bilayer phase and antiphase formation in polar liquid crystals}
11    
12     \author{J. Saha\footnote{Current Address: Department of Physics, Visva-Bharati
13     University, Santiniketan - 731235, West Bengal, India} and
14     J. Daniel Gezelter\footnote{e-mail: gezelter@nd.edu}}
15    
16     \affiliation{Department of Chemistry and Biochemistry, University of
17     Notre Dame, Notre Dame, Indiana 46556}
18    
19     \date{\today}
20    
21     \begin{abstract}
22     We present results of a series of Molecular Dynamics simulations on
23     the molecular organization of systems of ellipsoidal Gay-Berne
24     molecules containing two fixed dipole moments. The effects of relative
25     dipolar orientations and positions on the generation of bilayer
26     lamellar phase, antiphases and monolayer phases has been studied. We
27     report on the structural features of the phases formed by molecules of
28     both liquid crystalline and biological interest.
29     \end{abstract}
30    
31     \pacs{61.30}
32    
33     \keywords{Gay-Berne molecule, two dipoles, molecular dynamics, bilayer
34     phase, antiphase}
35    
36     \maketitle
37    
38     \newpage
39     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40     %%%%%%% BODY OF TEXT
41     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
42    
43     \section{Introduction}
44     \label{sec:intro}
45    
46     Long range orientational order is the most fundamental property of
47     liquid crystal mesophases, while positional order is limited or
48     absent. This orientational anisotropy of the macroscopic phases
49     originates in the anisotropy of the constituent molecules. For the
50     existence of orientational ordering, these molecules typically have
51     highly non-spherical structure with some degree of rigidity. Liquid
52     crystalline compounds typically possess cylindrically symmetric rod or
53     disc-like rigid core structures and usually have flexible substituents
54     associated with these central regions. In nematic phases, rod-like
55     molecules are orientationally ordered with isotropic distributions of
56     molecular centers of mass. In smectic phases, the molecules arrange
57     themselves into layers with their long (symmetry) axis normal
58     ($S_{A}$) or tilted ($S_{C}$) with respect to the layer planes. The
59     layers themselves are two dimensional liquids.
60    
61     \subsection{Previous Experimental Work}
62    
63     Experimental and theoretical studies of smectic liquid crystals
64     include a range of polymorphic variations on the basic smectic
65     organization. The behaviour of the $S_{A}$ phase can be explained with
66     theoretical models mainly based on geometric factors and van der Waals
67     interactions. However, these simple models are insufficient to
68     describe liquid crystal phases which exhibit more complex polymorphic
69     nature. X-ray diffraction studies have shown that the ratio between
70     lamellar spacing ($s$) and molecular length ($l$) can take
71     significantly different values.\cite{Leadbetter77,Gray84} Typical
72     $S_{A}$ phases have $s/l$ ratios on the order of $0.8$ , whereas for
73     some compounds e.g. 4-alkyl-4'-cyanobiphenyls the $s/l$ ratio is on
74     the order of $1.4$. Experiments show that depending on dipole
75     delocalization within the molecules, $s$ can take values ranging from
76     the length of a single molecule to twice the molecular
77     length~\cite{Leadbetter77,Hardouin80}. Extensive experimental studies
78     reveal that compounds of the $S_{A}$ type which show a variety of
79     phases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
80     partial bilayers ($S_{\tilde A}$) and interdigitated bilayers
81     ($S_{A_{d}}$), usually have a terminal cyano or nitro group. These
82     classes of liquid crystal and in particular, lyotropic liquid crystals
83     (those exhibiting liquid crystal phase transition as a function of
84     water concentration) which form surfactant and lipid membranes, often
85     have polar head groups or zwitterionic charge separated groups that
86     result in strong dipolar interactions.\cite{Collings97} Apart from the
87     uniaxial $S_{A}$ phases mentioned above, compounds having permanent
88     dipole moments can also give rise to reentrant smectic, biaxial
89     ($S_{C}$) and ferroelectric phases. Because of their versatile
90     polymorphic nature, these liquid crystalline materials have important
91     technological applications in addition to their immense relevance to
92     biological systems.\cite{Collings97}
93    
94     Experimental studies by Levelut {\it et
95     al.}~\cite{Levelut81a,Levelut81b} revealed that terminal cyano or
96     nitro groups usually induce permanent longitudinal dipole moments on
97     the molecules. Strong lateral dipole $(C=O)$ and terminal transverse
98     dipoles can also effect the phase behaviour considerably. Many
99     liquid-crystal forming molecules of biological interest
100     (e.g. phospholipids) could be modelled more accurately with {\em
101     transverse} or angled dipole moments at the terminus of the molecule.
102     In particular, the dipole moment of phosphatidylcholine (PC) head
103     groups are typically oriented perpendicular to the molecular axis,
104     while the dipole of phosphatidylethanolamine (PE) head groups are
105     tilted relative to the molecular axis. Moreover, there is strong
106     indication from the molecular structure of liquid crystalline
107     molecules that form these phases, that in adddition to the terminal
108     dipole, it is advantageous to have the presence of aromatic $\pi$
109     bonds and/or other dipoles near the core which can be easily polarized
110     by the strong electron withdrawing properties of the terminal group,
111     resulting in the various $S_{A}$
112     phases.~\cite{Gray84,Jeu83,Levelut81a,Levelut81b}
113    
114     \subsection{Previous theoretical work}
115    
116     Theoretical models of polar smectics using generic molecular field
117     descriptions were studied by Photinos, Saupe and
118     others.~\cite{Photinos76,Vanakaras98} Prost coupled two different
119     order parameters (the density wave and dipolar ordering of molecules
120     along the director) and using a mean field approch showed that
121     competition between them was responsible for polar liquid crystal
122     behaviour.~\cite{Prost84,Prost80} Comparing some polar liquid crystal
123     compounds, de~Jeu inferred that variation of dipole correlation with
124     molecular structure could affect the phase behaviour of polar liquid
125     crystals.~\cite{Jeu83} A number of other analytical approaches to
126     these phases have been presented in the
127     literature,~\cite{Meyer76,Dowell85,Indekeu86,Baus89,Netz92} but most
128     have been too simplified to confirm any more than the qualitative
129     behavior of these phases. Therefore, it seems that a molecular-scale
130     simulation approach will be required for a more complete understanding
131     of bilayer and antiphases.
132    
133     Levesque {\em et al.} presented a hard rod model which exhibited
134     monolayers and further indicated that the very symmetric nature of the
135     potential was responsible for their failure to generate true
136     bilayer.~\cite{Levesque93} Reproduction of small domains of bilayers
137     was reported for transverse dipoles but these were not always present
138     in their simulations.
139    
140     The Gay-Berne potential has seen widespread use in the liquid crystal
141     community to describe this anisotropic phase
142     behavior.~\cite{Gay81,Berne72,Kushick76,Luckhurst90,Perram96} It is an
143     appropriate model for simulation of these systems because fairly rigid
144     liquid crystal-forming molecules maintain their rod-like or disc-like
145     shapes, which produce interactions that favor local alignment. In its
146     original form, the Gay-Berne potential was a computationally efficient
147     {\em single} site model for the interactions of rigid ellipsoidal
148     molecules.~\cite{Gay81} It can be thought of as a modification of the
149     Gaussian overlap model originally described by Berne and
150     Pechukas.~\cite{Berne72} The potential is constructed in the familiar
151     form of the Lennard-Jones function using orientation-dependent
152     $\sigma$ and $\epsilon$ parameters. The functional form for the
153     potential is given in section 2. Luckhurst has given a particularly
154     good explanation of the choice of the Gay-Berne parameters $\mu$ and
155     $\nu$ for modeling non-polar liquid crystal molecules.
156    
157     Although there have been a large number of studies of the phase
158     behaviour of the {\em non-polar} Gay-Berne potential using Monte Carlo
159     and Molecular Dynamics techniques,~\cite{Zannonibook2000} there has
160     been comparatively little work done on an important class of liquid
161     crystal forming molecules which have {\em dipolar} interactions.
162     There have been some preliminary Monte Carlo studies of the Gay-Berne
163     potential with fixed longitudinal dipoles (i.e. pointed along the
164     principal axis of rotation).~\cite{Berardi96,Satoh96} There have also
165     been some recent molecular dynamics simulations on polar Gay-Berne
166     models by Pasterny {\em et al.}~\cite{Pasterny2000} Zannoni's group
167     has studied the phase behavior of Gay-Berne ellipsoids with
168     longitudinal and transverse dipoles both at the midpoint and terminus
169     of the molecule.~\cite{Berardi99} Their exhaustive simulation with a
170     model potential comprising both the attractive-repulsive G-B
171     interaction and a single dipolar interaction exhibited partial striped
172     bilayer structures.
173    
174     MISSING: AYTON AND VOTH WORK ON GB with terminal LJ spheres.
175    
176     Simulations of small domains of local bilayers for nCB {WHAT IS NCB}
177     have been reported recently.~\cite{Fukunaga2004}
178    
179     Since none of these simulations, which considered molecules with
180     single dipoles and studied the effects of either the terminal or the
181     central dipoles seperately, were able to reproduce perfect bilayer
182     arrangement the molecular origin of these liquid crystal phases is not
183     very well understood. There is, of course, a vast literature of
184     all-atom, or coarse-grained simulation models for lipid bilayers, but
185     typical system sizes with these models allow only relatively small
186     patches of single or double bilayers to be studied. In this work, we
187     present a model that is simple enough to allow us to probe the
188     equilibrated phase behavior at a number of different conditions, while
189     still maintaining enough molecular-scale realism to be useful as a
190     predictive tool.
191    
192     To mimic the terminal dipolar interaction coupled with polar cores we
193     considered systems comprising Gay-Berne particles with an embedded
194     terminal dipole and another weaker central dipole. Performing a series
195     of molecular dynamics simulations, we studied the structural
196     properties of these phases in systems of prolate ellipsoidal particles
197     each having relatively two dipole moments oriented perpendicularly
198     with respect to their respective molecular symmetry axes. In this
199     paper we report the generation of bilayer, monolayer and wavy
200     antiphase structures. To our knowledge, the present simulation work
201     is the first of its kind which could generate the bilayer liquid
202     crystalline phase successfully along with other important
203     experimentally-observed phases.
204    
205     \section{Model}
206     \label{sec:model}
207    
208     In this work, rod-like polar molecules are modelled as prolate
209     ellipsoidal Gay-Berne (GB) particles. The GB interaction potential
210     used to mimic the apolar characteristics of liquid crystal molecules
211     takes the familiar form of Lennard-Jones function with orientation and
212     position dependent range ($\sigma$) and well depth ($\epsilon$)
213     parameters. It can can be expressed as,
214     \begin{equation}
215     \begin{array}{ll}
216     V^{GB}_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}) =
217     4 \epsilon({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}})
218     & \left[ \left(
219     \frac{\sigma_{o}}{r - \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
220     \hat{r}})+\sigma_{o}} \right)^{12} \right. \\ \\
221     & - \left. \left(
222     \frac{\sigma_{o}}{r - \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
223     \hat{r}})+\sigma_{o}}
224     \right)^{6}
225     \right], \\
226     \end{array}
227     \label{eq:gb}
228     \end {equation}
229     where ${\bf \hat{u}_{i},\hat{u}_{j}}$ are unit vectors specifying the
230     orientation of two molecules $i$ and $j$ separated by intermolecular
231     vector ${\bf r}$. ${\bf \hat{r}}$ is the unit vector along the
232     intermolecular vector.
233    
234     The functional form for $\sigma$ is given by
235     \begin {equation}
236     \begin{array}{ll}
237     \sigma ({\bf \hat{u}_{i}, \hat{u}_{j},\hat{r}}) = \sigma_{0}
238     & \left[ 1- \frac {\chi}{2} \left( \frac{({\bf \hat{r}}.{\bf \hat{u}_{i}}+{\bf \hat{r}}.
239     {\bf \hat {u}_{j}})^2}{1+\chi
240     ({\bf \hat{u}_{i}}.{\bf \hat{u}_{j}})} \right. \right. \\ \\
241     & \left.\left. + \frac {({\bf \hat{r}}.{\bf \hat{u}_{i
242     }}-{\bf \hat{r}}.
243     {\bf \hat{u}_{j}})^2}{1-\chi ({\bf \hat{u}_{i}}.{\bf \hat{u}_{j}})} \right)
244     \right ]^{-1/2}
245     \end{array}
246     \end {equation}
247    
248     The aspect ratio of the particles is governed by shape anisotropy
249     parameter
250     \begin {equation}
251     \begin{array}{rcl}
252     \chi & = & \frac
253     {(\sigma_{e}/\sigma_{s})^2-1}{(\sigma_{e}/\sigma_{s})^2+1} \\ \\
254     \end{array}
255     \label{eq:chi}
256     \end {equation}
257     Here, the subscript $s$ indicates the {\it side-by-side} configuration
258     where $\sigma$ has its minimum value, $\sigma_{s}$, and where the
259     potential well is $\epsilon_{s}$ deep. The subscript $e$ refers to
260     the {\it end-to-end} configuration where $\sigma$ has its maximum
261     value, $\sigma_{e}$, and where the well depth, $\epsilon_{e}$ is
262     somewhat smaller than in the side-by-side configuration. For prolate
263     ellipsoids, we have
264     \begin{equation}
265     \begin{array}{rcl}
266     \sigma_{s} & < & \sigma_{e} \\
267     \epsilon_{s} & > & \epsilon_{e}
268     \end{array}
269     \end{equation}
270     where, $\sigma_{e}$ is the measure of the length and $\sigma_{s}$ is
271     the breadth of a molecule. The shape anisotropy parameter $\chi$ has
272     a functional dependence on the length to breadth ratio (i.e. the
273     aspect ratio of the particles.).
274    
275     The functional form of well depth is
276     \begin {equation}
277     \epsilon({\bf \hat{u}_{i},\hat{u}_{j},\hat{r}}) = \epsilon_{0}
278     \epsilon^{\nu}({\bf \hat{u}_{i}.\hat{u}_{j}})
279     \epsilon^{\prime\mu}({\bf \hat{u}_{i},\hat{u}_{j},\hat{r}})
280     \end {equation}
281     where $\epsilon_{0}$ is a constant term and
282     \begin {equation}
283     \epsilon ({\bf \hat{u}_{i},\hat{u}_{j}}) =
284     [1-\chi^{2}({\bf \hat{u}_{i}.\hat{u}_{j}})^{2}]^{-1/2}
285     \end {equation}
286     and
287     \begin {equation}
288     \epsilon^{\prime} ({\bf \hat{u}_{i}, \hat{u}_{j},\hat{r}})
289     = 1- \frac {\chi^{'}}{2} \left
290     ( \frac {({\bf \hat{r}}.{\bf \hat{u}_{i}}+{\bf\hat{r}}.
291     {\bf \hat{u}_{j}})^2}{1+\chi^{\prime}
292     ({\bf \hat{u}_{i}}.{\bf \hat{u}_{j}})} +\frac {({\bf\hat{ r}}.{\bf\hat{
293     u}_{i}}-{\bf \hat{r}}.
294     {\bf \hat{u}_{j}})^2}{1-\chi^{\prime} ({\bf \hat{u}_{i}}.{\bf
295     \hat{u}_{j}})} \right)
296     \end {equation}
297     where the well depth anisotropy parameter $\chi\prime$ can be expressed as
298     \begin {equation}
299     \chi^{\prime} = \frac {1-(\epsilon_{e}/\epsilon_{s})^{1/\mu}}{1+(\epsilon_{e}/
300     \epsilon_{s})^{1/\mu}}.
301     \end {equation}
302    
303    
304    
305     As the molecules have two dipoles, for each pair of them (4 pairs for
306     two interacting molecules), there should be an electrostatic
307     interaction term of the form
308    
309     \begin {equation}
310     U_{dd} = \frac { \mu^{*}_{id} \mu^{*}_{jd}}{r_{d}^{3}}
311     \left [({\bf \hat{u}_{id}.\hat{u}_{jd}}) - 3 ({\bf \hat{u}_{id}. \hat{r}_{d}})(
312     {\bf \hat{u}_{jd}. \hat{r}_{d}}) \right ]
313     \end {equation}
314    
315     where reduced dipole moment
316    
317     \begin{equation}
318     \mu_{d}^{*} = \frac {\mu_{d}^{2}}
319     {(4 \pi {\it \epsilon} \epsilon_{s} \sigma_{s}^{3})^{1/2}}
320     \end{equation}
321    
322     $\it {\epsilon}$ is the permitivity of the free space and $r_{d}$ is
323     the unit vector along the vector joining the two dipoles. $\bf \hat{u}_{id}$
324     and
325     $\bf \hat{u}_{jd}$ are the direction of the unit vectors along the direction of
326     the dipoles situated on molecules i and j respectively.
327    
328    
329     So the total interaction potential for a pair of polar molecules is
330     the sum of attractive-repulsive term and dipole-dople interaction term,
331     which can be expressed as
332    
333     \begin {equation}
334     U_{ij} = U_{GB} + (U_{dd})_{1st~ dipole} + (U_{dd})_{2nd~ dipole}
335     \end {equation}
336    
337     To simulate systems of dipolar Gay-Berne
338     particles with different relative dipolar orientations and positions, we
339     used this model potential.
340    
341     \begin{figure}
342     \begin{center}
343     \epsfxsize=3in
344     \epsfbox{system_sketch.eps}
345     \end{center}
346     \caption{The molecular models studied in this work. All are prolate
347     Gay-Berne ellipsoids which have point dipoles embedded centrally or
348     terminally within the molecular bodies. System A is a molecular-scale
349     model for XXXX, System B is a model of YYYY, and System C could be
350     considered a model for phosphatidylcholine (PC) lipids. Details on
351     the dipolar locations and strengths are given in the text.}
352     \label{fig:gbdp}
353     \end{figure}
354    
355    
356     \section{Computational Methodology}
357     We performed a series of extensive Molecular Dynamics (MD) simulations to
358     study
359     the phase behaviour of a family of polar liquid crystals.
360    
361     In each simulation, rod-like polar
362     molecules have been represented by polar ellipsoidal
363     Gay-Berne (GB) particles. The four parameters characterizing G-B
364     potential were taken as $\mu = 1,~ \nu = 2, ~\epsilon_{e}/\epsilon_{s}
365     = 1/5$ and $\sigma_{e}/\sigma_{s} = 3$. The components of the
366     scaled moment of inertia $(I^{*} = I/m \sigma_{s}^{2})$ along
367     the major and minor axes were $I_{z}^{*} = 0.2$ and $I_{\perp}^{*}
368     = 1.0$. We used the
369     reduced dipole
370     moments $ \mu^{*} = \mu/(4
371     \pi \epsilon_{fs} \sigma_{0}^{3})^{1/2}= 1.0$ for terminal dipole and
372     $ \mu^{*} = \mu/(4
373     \pi \epsilon_{fs} \sigma_{0}^{3})^{1/2}= 0.5$ for second dipole,
374     where $\epsilon_{fs}$
375     was the permitivitty of free space. For all simulations the position of the
376     terminal dipole
377     has been kept
378     at a fixed distance $d^{*} = d/\sigma_{s} = 1.0 $ from the
379     centre of mass on the molecular symmetry axis. The second dipole
380     takes $d^{*} = d/\sigma_{s} = 0.0 $
381     i.e. it is on the centre of mass. To investigate
382     the molecular organization behaviour due to different dipolar
383     orientation with respect to the symmetry axis, we selected dipolar
384     angle $\alpha_{d} = 0$ to model terminal outward longitudinal
385     dipole and $\alpha_{d} = \pi/2$ to model transverse outward dipole where
386     the second
387     dipole takes relative anti
388     antiparallel orientation with respect to the first. System of molecules
389     having a single transverse terminal dipole has also been studied. We ran
390     a series of
391     simulations to investigate the effect of dipoles on molecular organization.
392    
393     In each of the simulations 864 molecules were confined in a cubic box with
394     periodic boundary conditions. The run started from a density
395     $\rho^{*} = \rho \sigma_{0}^{3}$ = 0.01 with nonpolar molecules
396     loacted on the sites of FCC lattice and having parallel
397     orientation. This structure was not a stable structure at this density
398     and it was melted at a reduced temperature $T^{*} = k_{B}T/
399     \epsilon_{0} = 4.0$ .
400     We used this isotropic
401     configuration which was both orientationally and translationally
402     disordered, as the initial configuration for each simulation. The dipoles
403     were also switched on from this point.
404     Initial translational and angular velocities were assigned
405     from the gaussian distribution of velocities.
406    
407     To get the ordered structure for each system of particular dipolar
408     angles we increased the density from
409     $\rho^{*} = 0.01$ to $\rho_{*} = 0.3$ with an increament size
410     of 0.002 upto $\rho^{*} = 0.1$ and 0.01 for the rest at some higher
411     temperature. Temperature was
412     then lowered in finer steps to avoid ending up with disordered glass phase
413     and thus to help the molecules set with more order.
414     For each system this process required altogether $5 \times 10^{6}$ MC cycles
415     for equilibration.
416    
417     The torques and forces were calculated using
418     velocity verlet algorithm. The time step size $\delta t^{*} =
419     \delta t/(m \sigma_{0}^{2} / \epsilon_{0})^{1/2}$ was set at 0.0012 during
420     the process. The orientations of molecules were described by quaternions
421     instead of Eulerian angles to get the singularity-free orientational
422     equations of motion.
423    
424     The interaction potential was truncated at a cut-off radius
425     $r_{c} = 3.8 \sigma_{0}$. The long range dipole-dipole interaction potential
426     and torque were handled by the application of reaction field method
427     ~\cite{Allen87}.
428    
429     To investigate the phase structure of the model liquid
430     crystal family we calculated
431     the orientational order parameter, correlation functions.
432     To identify a particular phase we took configurational snapshots
433     at the onset of each layered phase.
434    
435     The orientational order parameter for uniaxial phase was calculated
436     from the largest eigen value obtained by diagonalization of the order
437     parameter tensor
438    
439     \begin{equation}
440     \begin{array}{lr}
441     Q_{\alpha \beta} = \frac{1}{2 N} \sum(3 e_{i \alpha} e_{i \beta}
442     - \delta_{\alpha \beta}) & \alpha, \beta = x,y,z \\
443     \end{array}
444     \end{equation}
445    
446     where $e_{i \alpha}$ was the $\alpha$ th component of the unit vector
447     $e_{i}$ along the symmetry axis of the i th molecule. Corresponding
448     eigenvector gave the director which defines the average direction
449     of molecular alignment.
450    
451     The density correlation along the director is $g(z) = < \delta
452     (z-z_{ij})>_{ij} / \pi R^{2} \rho $, where $z_{ij} = r_{ij} cos
453     \beta_{r_{ij}}$ was measured in the director frame and $R$ is the
454     radius of the cylindrical sampling region.
455    
456    
457     \section{Results and Conclusion}
458     \label{sec:results and conclusion}
459    
460     Analysis of the simulation results shows that relative dipolar orientation
461     angle of the molecules can give rise to rich polymorphism
462     of polar mesophases.
463    
464     The correlation function g(z) shows layering along perpendicular
465     direction to the plane for a system of G-B molecules with two
466     transverse outward pointing dipoles in fig. \ref{fig:1}. Both the
467     correlation plot and the snapshot (fig. \ref{fig:4}) of their
468     organization indicate a bilayer phase. Snapshot for larger system of
469     1372 molecules also confirms bilayer structure (Fig. \ref{fig:7}).
470     Fig. \ref{fig:2} shows g(z) for a system of molecules having two
471     antiparallel longitudinal dipoles and the snapshot of their
472     organization shows a monolayer phase
473     (Fig. \ref{fig:5}). Fig. \ref{fig:3} gives g(z) for a system of G-B
474     molecules with single transverse outward pointing dipole and
475     fig. \ref{fig:6} gives the snapshot. Their organization is like a wavy
476     antiphase (stripe domain). Fig. \ref{fig:8} gives the snapshot for
477     1372 molecules with single transverse dipole near the end of the
478     molecule.
479    
480     \begin{figure}
481     \begin{center}
482     \epsfxsize=3in
483     \epsfbox{fig1.ps}
484     \end{center}
485     \caption { Density projection of molecular centres (solid) and terminal dipoles (broken) with respect to the director g(z)
486     for a system of G-B molecules with two transverse
487     outward pointing dipoles, the first dipole having $d^{*}=1.0$,
488     $\mu^{*}=1.0$ and the second dipole having $d^{*}=0.0$,
489     $\mu^{*}=0.5$}
490     \label{fig:1}
491     \end{figure}
492    
493    
494     \begin{figure}
495     \begin{center}
496     \epsfxsize=3in
497     \epsfbox{fig2.ps}
498     \end{center}
499     \caption { Density projection of molecular centres (solid) and terminal
500     dipoles (broken) with respect to the director
501     g(z) for a system of G-B molecules with two antiparallel
502     longitudinal dipoles, the first outward pointing dipole having $d^{*}=1.0$,
503     $\mu^{*}=1.0$ and the second dipole having $d^{*}=0.0$,
504     $\mu^{*}=0.5$}
505     \label{fig:2}
506     \end{figure}
507    
508     \begin{figure}
509     \begin{center}
510     \epsfxsize=3in
511     \epsfbox{fig3.ps}
512     \end{center}
513     \caption {Density projection of molecular centres (solid) and terminal
514     dipoles (broken) with respect to the director g(z)
515     for a system of G-B molecules with single transverse
516     outward pointing dipole, having $d^{*}=1.0$,
517     $\mu^{*}=1.0$}
518     \label{fig:3}
519     \end{figure}
520    
521     \begin{figure}
522     \centering
523     \epsfxsize=2.5in
524     \epsfbox{fig4.eps}
525     \caption{Typical configuration for a system of 864 G-B molecules
526     with two transverse dipoles, the first dipole having $d^{*}=1.0$,
527     $\mu^{*}=1.0$ and the second dipole having $d^{*}=0.0$,
528     $\mu^{*}=0.5$. The white caps indicate the location of the terminal
529     dipole, while the orientation of the dipoles is indicated by the
530     blue/gold coloring.}
531     \label{fig:4}
532     \end{figure}
533    
534     \begin{figure}
535     \begin{center}
536     \epsfxsize=3in
537     \epsfbox{fig5.ps}
538     \end{center}
539     \caption {Snapshot of molecular configuration for a system of 864 G-B molecules with
540     two antiparallel longitudinal dipoles, the first outward pointing dipole
541     having $d^{*}=1.0$, $\mu^{*}=1.0$ and the second dipole having $d^{*}=0.0$,
542     $\mu^{*}=0.5$ (fine lines are molecular symmetry axes and small thick lines
543     show terminal dipolar direction, central dipoles are not shown).}
544     \label{fig:5}
545     \end{figure}
546    
547    
548     \begin{figure}
549     \begin{center}
550     \epsfxsize=3in
551     \epsfbox{fig6.ps}
552     \end{center}
553     \caption {Snapshot of molecular configuration for a system of 864 G-B molecules with
554     single transverse outward pointing dipole, having $d^{*}=1.0$, $\mu^{*}=1.0$
555     (fine lines are molecular symmetry axes and small thick lines
556     show terminal dipolar direction).}
557     \label{fig:6}
558     \end{figure}
559    
560     \begin{figure}
561     \begin{center}
562     \epsfxsize=3in
563     \epsfbox{fig7.ps}
564     \end{center}
565     \caption {Snapshot of molecular configuration for a system of 1372 G-B molecules
566     with two transverse outward pointing dipoles, the first dipole having
567     $d^{*}=1.0$, $\mu^{*}=1.0$ and the second dipole having $d^{*}=0.0$,
568     $\mu^{*}=0.5$(fine lines are molecular symmetry axes and small thick lines
569     show terminal dipolar direction, central dipoles are not shown).}
570     \label{fig:7}
571     \end{figure}
572    
573     \begin{figure}
574     \begin{center}
575     \epsfxsize=3in
576     \epsfbox{fig8.ps}
577     \end{center}
578     \caption {Snapshot of molecular configuration for a system of 1372 G-B molecules with
579     single transverse outward pointing dipole, having $d^{*}=1.0$, $\mu^{*}=1.0$
580     (fine lines are molecular symmetry axes and small thick lines
581     show terminal dipolar direction).}
582     \label{fig:8}
583     \end{figure}
584    
585     Starting from an isotropic configuaration of polar Gay-Berne molecules,
586     we could successfully simulate perfect bilayer, antiphase and monolayer
587     structure. To break the up-down symmetry i.e. the nonequivalence of
588     directions ${\bf \hat {n}}$ and ${ -\bf \hat{n}}$, the molecules should have permanent
589     electric or magnetic dipoles. Longitudinal electric dipole interaction could
590     not form
591     polar nematic phase as orientationally disordered phase with larger entropy
592     is stabler than polarly ordered phase. In fact, stronger central dipole moment
593     opposes polar nematic ordering more effectively in case of rod-like
594     molecules. However, polar ordering like bilayer $A_{2}$, interdigitated
595     $A_{d}$, and wavy $\tilde A$ in smectic layers can be achieved, where adjacent
596     layers with opposite polarities makes bulk phase a-polar. More so, lyotropic
597     liquid crystals and bilayer bio-membranes can have polar layers. These
598     arrangements appear to get favours with the shifting of longitudinal dipole
599     moment to the molecular terminus, so that they can have
600     anti-ferroelectric
601     dipolar arrangement giving rise to local (within the sublayer) breaking of
602     up-down symmetry along the director. Transverse polarity breaks two-fold
603     rotational symmetry, which favours more in-plane polar order. However, the
604     molecular origin of these phases requires something more which are apparent
605     from the earlier simulation results. We have shown that to get perfect bilayer
606     structure in a G-B system, alongwith transverse terminal dipole, another
607     central dipole (or
608     a polarizable core) is required so that polar head and a-polar tail
609     of Gay-Berne molecules go to opposite directions within a bilayer. This
610     gives some kind of clipping interactions which forbid the molecular
611     tail go in other way.
612     Moreover,
613     we could simulate other varieties of polar smectic phases e.g. monolayer
614     $A_{1}$,
615     antiphase $\tilde A$
616     successfully.
617     Apart from guiding chemical synthesization of ferroelectric,
618     antiferroelectric liquid crystals for technological applications, the present
619     study will be of scientific interest in understanding molecular level
620     interactions of lyotropic liquid crystals as well as nature-designed
621     bio-membranes.
622    
623     \begin{acknowledgments}
624     Support for this project was provided by the National Science
625     Foundation under grant CHE-0134881. Computation time was provided by
626     the Notre Dame High Performance Computing Cluster and the Notre Dame
627     Bunch-of-Boxes (B.o.B) computer cluster (NSF grant DMR-0079647).
628     \end{acknowledgments}
629    
630    
631     \bibliography{bilayer}
632    
633     \pagebreak
634    
635     \end {document}
636    
637    
638    
639