ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/LiquidCrystal.tex
Revision: 2867
Committed: Sat Jun 17 18:43:58 2006 UTC (18 years, 1 month ago) by tim
Content type: application/x-tex
File size: 21342 byte(s)
Log Message:
finishing temperature control of Langevin

File Contents

# Content
1 \chapter{\label{chapt:liquidcrystal}LIQUID CRYSTAL}
2
3 \section{\label{liquidCrystalSection:introduction}Introduction}
4
5 Long range orientational order is one of the most fundamental
6 properties of liquid crystal mesophases. This orientational
7 anisotropy of the macroscopic phases originates in the shape
8 anisotropy of the constituent molecules. Among these anisotropy
9 mesogens, rod-like (calamitic) and disk-like molecules have been
10 exploited in great detail in the last two decades\cite{Huh2004}.
11 Typically, these mesogens consist of a rigid aromatic core and one
12 or more attached aliphatic chains. For short chain molecules, only
13 nematic phases, in which positional order is limited or absent, can
14 be observed, because the entropy of mixing different parts of the
15 mesogens is paramount to the dispersion interaction. In contrast,
16 formation of the one dimension lamellar sematic phase in rod-like
17 molecules with sufficiently long aliphatic chains has been reported,
18 as well as the segregation phenomena in disk-like molecules.
19
20 Recently, the banana-shaped or bent-core liquid crystal have became
21 one of the most active research areas in mesogenic materials and
22 supramolecular chemistry\cite{Niori1996, Link1997, Pelzl1999}.
23 Unlike rods and disks, the polarity and biaxiality of the
24 banana-shaped molecules allow the molecules organize into a variety
25 of novel liquid crystalline phases which show interesting material
26 properties. Of particular interest is the spontaneous formation of
27 macroscopic chiral layers from achiral banana-shaped molecules,
28 where polar molecule orientational ordering is shown within the
29 layer plane as well as the tilted arrangement of the molecules
30 relative to the polar axis. As a consequence of supramolecular
31 chirality, the spontaneous polarization arises in ferroelectric (FE)
32 and antiferroelectic (AF) switching of smectic liquid crystal
33 phases, demonstrating some promising applications in second-order
34 nonlinear optical devices. The most widely investigated mesophase
35 formed by banana-shaped moleculed is the $\text{B}_2$ phase, which
36 is also referred to as $\text{SmCP}$\cite{Link1997}. Of the most
37 important discover in this tilt lamellar phase is the four distinct
38 packing arrangements (two conglomerates and two macroscopic
39 racemates), which depend on the tilt direction and the polar
40 direction of the molecule in adjacent layer (see
41 Fig.~\ref{LCFig:SMCP}).
42
43 \begin{figure}
44 \centering
45 \includegraphics[width=\linewidth]{smcp.eps}
46 \caption[]
47 {}
48 \label{LCFig:SMCP}
49 \end{figure}
50
51 Many liquid crystal synthesis experiments suggest that the
52 occurrence of polarity and chirality strongly relies on the
53 molecular structure and intermolecular interaction\cite{Reddy2006}.
54 From a theoretical point of view, it is of fundamental interest to
55 study the structural properties of liquid crystal phases formed by
56 banana-shaped molecules and understand their connection to the
57 molecular structure, especially with respect to the spontaneous
58 achiral symmetry breaking. As a complementary tool to experiment,
59 computer simulation can provide unique insight into molecular
60 ordering and phase behavior, and hence improve the development of
61 new experiments and theories. In the last two decades, all-atom
62 models have been adopted to investigate the structural properties of
63 smectic arrangements\cite{Cook2000, Lansac2001}, as well as other
64 bulk properties, such as rotational viscosity and flexoelectric
65 coefficients\cite{Cheung2002, Cheung2004}. However, due to the
66 limitation of time scale required for phase transition and the
67 length scale required for representing bulk behavior,
68 models\cite{Perram1985, Gay1981}, which are based on the observation
69 that liquid crystal order is exhibited by a range of non-molecular
70 bodies with high shape anisotropies, became the dominant models in
71 the field of liquid crystal phase behavior. Previous simulation
72 studies using hard spherocylinder dimer model\cite{Camp1999} produce
73 nematic phases, while hard rod simulation studies identified a
74 Landau point\cite{Bates2005}, at which the isotropic phase undergoes
75 a direct transition to the biaxial nematic, as well as some possible
76 liquid crystal phases\cite{Lansac2003}. Other anisotropic models
77 using Gay-Berne(GB) potential, which produce interactions that favor
78 local alignment, give the evidence of the novel packing arrangements
79 of bent-core molecules\cite{Memmer2002,Orlandi2006}.
80
81 Experimental studies by Levelut {\it et al.}~\cite{Levelut1981}
82 revealed that terminal cyano or nitro groups usually induce
83 permanent longitudinal dipole moments, which affect the phase
84 behavior considerably. A series of theoretical studies also drawn
85 equivalent conclusions. Monte Carlo studies of the GB potential with
86 fixed longitudinal dipoles (i.e. pointed along the principal axis of
87 rotation) were shown to enhance smectic phase
88 stability~\cite{Berardi1996,Satoh1996}. Molecular simulation of GB
89 ellipsoids with transverse dipoles at the terminus of the molecule
90 also demonstrated that partial striped bilayer structures were
91 developed from the smectic phase ~\cite{Berardi1996}. More
92 significant effects have been shown by including multiple
93 electrostatic moments. Adding longitudinal point quadrupole moments
94 to rod-shaped GB mesogens, Withers \textit{et al} induced tilted
95 smectic behaviour in the molecular system~\cite{Withers2003}. Thus,
96 it is clear that many liquid-crystal forming molecules, specially,
97 bent-core molecules, could be modeled more accurately by
98 incorporating electrostatic interaction.
99
100 In this chapter, we consider system consisting of banana-shaped
101 molecule represented by three rigid GB particles with one or two
102 point dipoles at different location. Performing a series of
103 molecular dynamics simulations, we explore the structural properties
104 of tilted smectic phases as well as the effect of electrostatic
105 interactions.
106
107 \section{\label{liquidCrystalSection:model}Model}
108
109 A typical banana-shaped molecule consists of a rigid aromatic
110 central bent unit with several rod-like wings which are held
111 together by some linking units and terminal chains (see
112 Fig.~\ref{LCFig:BananaMolecule}). In this work, each banana-shaped
113 mesogen has been modeled as a rigid body consisting of three
114 equivalent prolate ellipsoidal GB particles. The GB interaction
115 potential used to mimic the apolar characteristics of liquid crystal
116 molecules takes the familiar form of Lennard-Jones function with
117 orientation and position dependent range ($\sigma$) and well depth
118 ($\epsilon$) parameters. The potential between a pair of three-site
119 banana-shaped molecules $a$ and $b$ is given by
120 \begin{equation}
121 V_{ab}^{GB} = \sum\limits_{i \in a,j \in b} {V_{ij}^{GB} }.
122 \end{equation}
123 Every site-site interaction can can be expressed as,
124 \begin{equation}
125 V_{ij}^{GB} = 4\epsilon (\hat u_i ,\hat u_j ,\hat r_{ij} )\left[
126 {\left( {\frac{{\sigma _0 }}{{r_{ij} - \sigma (\hat u_i ,\hat u_j
127 ,\hat r_{ij} )}}} \right)^{12} - \left( {\frac{{\sigma _0
128 }}{{r_{ij} - \sigma (\hat u_i ,\hat u_j ,\hat r_{ij} )}}} \right)^6
129 } \right] \label{LCEquation:gb}
130 \end{equation}
131 where $\hat u_i,\hat u_j$ are unit vectors specifying the
132 orientation of two molecules $i$ and $j$ separated by intermolecular
133 vector $r_{ij}$. $\hat r_{ij}$ is the unit vector along the
134 intermolecular vector. A schematic diagram of the orientation
135 vectors is shown in Fig.\ref{LCFigure:GBScheme}. The functional form
136 for $\sigma$ is given by
137 \begin{equation}
138 \sigma (\hat u_i ,\hat u_i ,\hat r_{ij} ) = \sigma _0 \left[ {1 -
139 \frac{\chi }{2}\left( {\frac{{(\hat r_{ij} \cdot \hat u_i + \hat
140 r_{ij} \cdot \hat u_j )^2 }}{{1 + \chi \hat u_i \cdot \hat u_j }}
141 + \frac{{(\hat r_{ij} \cdot \hat u_i - \hat r_{ij} \cdot \hat u_j
142 )^2 }}{{1 - \chi \hat u_i \cdot \hat u_j }}} \right)} \right]^{ -
143 \frac{1}{2}},
144 \end{equation}
145 where the aspect ratio of the particles is governed by shape
146 anisotropy parameter
147 \begin{equation}
148 \chi = \frac{{(\sigma _e /\sigma _s )^2 - 1}}{{(\sigma _e /\sigma
149 _s )^2 + 1}}.
150 \label{LCEquation:chi}
151 \end{equation}
152 Here, $\sigma_ s$ and $\sigma_{e}$ refer to the side-by-side breadth
153 and the end-to-end length of the ellipsoid, respectively. The well
154 depth parameters takes the form
155 \begin{equation}
156 \epsilon (\hat u_i ,\hat u_j ,\hat r_{ij} ) = \epsilon _0 \epsilon
157 ^v (\hat u_i ,\hat u_j )\epsilon '^\mu (\hat u_i ,\hat u_j ,\hat
158 r_{ij} )
159 \end{equation}
160 where $\epsilon_{0}$ is a constant term and
161 \begin{equation}
162 \epsilon (\hat u_i ,\hat u_j ) = \frac{1}{{\sqrt {1 - \chi ^2 (\hat
163 u_i \cdot \hat u_j )^2 } }}
164 \end{equation}
165 and
166 \begin{equation}
167 \epsilon '(\hat u_i ,\hat u_j ,\hat r_{ij} ) = 1 - \frac{{\chi
168 '}}{2}\left[ {\frac{{(\hat r_{ij} \cdot \hat u_i + \hat r_{ij}
169 \cdot \hat u_j )^2 }}{{1 + \chi '\hat u_i \cdot \hat u_j }} +
170 \frac{{(\hat r_{ij} \cdot \hat u_i - \hat r_{ij} \cdot \hat u_j
171 )^2 }}{{1 - \chi '\hat u_i \cdot \hat u_j }}} \right]
172 \end{equation}
173 where the well depth anisotropy parameter $\chi '$ depends on the
174 ratio between \textit{end-to-end} well depth $\epsilon _e$ and
175 \textit{side-by-side} well depth $\epsilon_s$,
176 \begin{equation}
177 \chi ' = \frac{{1 - (\epsilon _e /\epsilon _s )^{1/\mu} }}{{1 +
178 (\epsilon _e /\epsilon _s )^{1/\mu} }}.
179 \end{equation}
180
181 \begin{figure}
182 \centering
183 \includegraphics[width=\linewidth]{banana.eps}
184 \caption[]{} \label{LCFig:BananaMolecule}
185 \end{figure}
186
187 %\begin{figure}
188 %\centering
189 %\includegraphics[width=\linewidth]{bananGB.eps}
190 %\caption[]{} \label{LCFigure:BananaGB}
191 %\end{figure}
192
193 \begin{figure}
194 \centering
195 \includegraphics[width=\linewidth]{gb_scheme.eps}
196 \caption[]{Schematic diagram showing definitions of the orientation
197 vectors for a pair of Gay-Berne molecules}
198 \label{LCFigure:GBScheme}
199 \end{figure}
200
201 To account for the permanent dipolar interactions, there should be
202 an electrostatic interaction term of the form
203 \begin{equation}
204 V_{ab}^{dp} = \sum\limits_{i \in a,j \in b} {\frac{1}{{4\pi
205 \epsilon _{fs} }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{r_{ij}^3 }}
206 - \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
207 r_{ij} } \right)}}{{r_{ij}^5 }}} \right]}
208 \end{equation}
209 where $\epsilon _{fs}$ is the permittivity of free space.
210
211 \section{Computational Methodology}
212
213 A series of molecular dynamics simulations were perform to study the
214 phase behavior of banana shaped liquid crystals.
215
216 In each simulation, rod-like polar molecules have been represented
217 by polar ellipsoidal Gay-Berne (GB) particles. The four parameters
218 characterizing G-B potential were taken as $\mu = 1,~ \nu = 2,
219 ~\epsilon_{e}/\epsilon_{s}
220 = 1/5$ and $\sigma_{e}/\sigma_{s} = 3$. The components of the
221 scaled moment of inertia $(I^{*} = I/m \sigma_{s}^{2})$ along the
222 major and minor axes were $I_{z}^{*} = 0.2$ and $I_{\perp}^{*} =
223 1.0$. We used the reduced dipole moments $ \mu^{*} = \mu/(4 \pi
224 \epsilon_{fs} \sigma_{0}^{3})^{1/2}= 1.0$ for terminal dipole and
225 $ \mu^{*} = \mu/(4
226 \pi \epsilon_{fs} \sigma_{0}^{3})^{1/2}= 0.5$ for second dipole,
227 where $\epsilon_{fs}$ was the permitivitty of free space. For all
228 simulations the position of the terminal dipole
229 has been kept
230 at a fixed distance $d^{*} = d/\sigma_{s} = 1.0 $ from the
231 centre of mass on the molecular symmetry axis. The second dipole
232 takes $d^{*} = d/\sigma_{s} = 0.0 $ i.e. it is on the centre of
233 mass. To investigate the molecular organization behaviour due to
234 different dipolar orientation with respect to the symmetry axis, we
235 selected dipolar angle $\alpha_{d} = 0$ to model terminal outward
236 longitudinal dipole and $\alpha_{d} = \pi/2$ to model transverse
237 outward dipole where the second dipole takes relative anti
238 antiparallel orientation with respect to the first. System of
239 molecules having a single transverse terminal dipole has also been
240 studied. We ran a series of simulations to investigate the effect of
241 dipoles on molecular organization.
242
243 In each of the simulations 864 molecules were confined in a cubic
244 box with periodic boundary conditions. The run started from a
245 density $\rho^{*} = \rho \sigma_{0}^{3}$ = 0.01 with nonpolar
246 molecules loacted on the sites of FCC lattice and having parallel
247 orientation. This structure was not a stable structure at this
248 density and it was melted at a reduced temperature $T^{*} = k_{B}T/
249 \epsilon_{0} = 4.0$ . We used this isotropic configuration which was
250 both orientationally and translationally disordered, as the initial
251 configuration for each simulation. The dipoles were also switched on
252 from this point. Initial translational and angular velocities were
253 assigned from the gaussian distribution of velocities.
254
255 To get the ordered structure for each system of particular dipolar
256 angles we increased the density from $\rho^{*} = 0.01$ to $\rho_{*}
257 = 0.3$ with an increament size of 0.002 upto $\rho^{*} = 0.1$ and
258 0.01 for the rest at some higher temperature. Temperature was then
259 lowered in finer steps to avoid ending up with disordered glass
260 phase and thus to help the molecules set with more order. For each
261 system this process required altogether $5 \times 10^{6}$ MC cycles
262 for equilibration.
263
264 The torques and forces were calculated using velocity verlet
265 algorithm. The time step size $\delta t^{*} = \delta t/(m
266 \sigma_{0}^{2} / \epsilon_{0})^{1/2}$ was set at 0.0012 during the
267 process. The orientations of molecules were described by quaternions
268 instead of Eulerian angles to get the singularity-free orientational
269 equations of motion.
270
271 The interaction potential was truncated at a cut-off radius $r_{c} =
272 3.8 \sigma_{0}$. The long range dipole-dipole interaction potential
273 and torque were handled by the application of reaction field method
274 ~\cite{Allen87}.
275
276 To investigate the phase structure of the model liquid crystal
277 family we calculated the orientational order parameter, correlation
278 functions. To identify a particular phase we took configurational
279 snapshots at the onset of each layered phase.
280
281 The orientational order parameter for uniaxial phase was calculated
282 from the largest eigen value obtained by diagonalization of the
283 order parameter tensor
284
285 \begin{equation}
286 \begin{array}{lr}
287 Q_{\alpha \beta} = \frac{1}{2 N} \sum(3 e_{i \alpha} e_{i \beta}
288 - \delta_{\alpha \beta}) & \alpha, \beta = x,y,z \\
289 \end{array}
290 \end{equation}
291
292 where $e_{i \alpha}$ was the $\alpha$ th component of the unit
293 vector $e_{i}$ along the symmetry axis of the i th molecule.
294 Corresponding eigenvector gave the director which defines the
295 average direction of molecular alignment.
296
297 The density correlation along the director is $g(z) = < \delta
298 (z-z_{ij})>_{ij} / \pi R^{2} \rho $, where $z_{ij} = r_{ij} cos
299 \beta_{r_{ij}}$ was measured in the director frame and $R$ is the
300 radius of the cylindrical sampling region.
301
302
303 \section{Results and Conclusion}
304 \label{sec:results and conclusion}
305
306 Analysis of the simulation results shows that relative dipolar
307 orientation angle of the molecules can give rise to rich
308 polymorphism of polar mesophases.
309
310 The correlation function g(z) shows layering along perpendicular
311 direction to the plane for a system of G-B molecules with two
312 transverse outward pointing dipoles in fig. \ref{fig:1}. Both the
313 correlation plot and the snapshot (fig. \ref{fig:4}) of their
314 organization indicate a bilayer phase. Snapshot for larger system of
315 1372 molecules also confirms bilayer structure (Fig. \ref{fig:7}).
316 Fig. \ref{fig:2} shows g(z) for a system of molecules having two
317 antiparallel longitudinal dipoles and the snapshot of their
318 organization shows a monolayer phase (Fig. \ref{fig:5}). Fig.
319 \ref{fig:3} gives g(z) for a system of G-B molecules with single
320 transverse outward pointing dipole and fig. \ref{fig:6} gives the
321 snapshot. Their organization is like a wavy antiphase (stripe
322 domain). Fig. \ref{fig:8} gives the snapshot for 1372 molecules
323 with single transverse dipole near the end of the molecule.
324
325 \begin{figure}
326 \begin{center}
327 \epsfxsize=3in \epsfbox{fig1.ps}
328 \end{center}
329 \caption { Density projection of molecular centres (solid) and
330 terminal dipoles (broken) with respect to the director g(z) for a
331 system of G-B molecules with two transverse outward pointing
332 dipoles, the first dipole having $d^{*}=1.0$, $\mu^{*}=1.0$ and the
333 second dipole having $d^{*}=0.0$, $\mu^{*}=0.5$} \label{fig:1}
334 \end{figure}
335
336
337 \begin{figure}
338 \begin{center}
339 \epsfxsize=3in \epsfbox{fig2.ps}
340 \end{center}
341 \caption { Density projection of molecular centres (solid) and
342 terminal dipoles (broken) with respect to the director g(z) for a
343 system of G-B molecules with two antiparallel longitudinal dipoles,
344 the first outward pointing dipole having $d^{*}=1.0$, $\mu^{*}=1.0$
345 and the second dipole having $d^{*}=0.0$, $\mu^{*}=0.5$}
346 \label{fig:2}
347 \end{figure}
348
349 \begin{figure}
350 \begin{center}
351 \epsfxsize=3in \epsfbox{fig3.ps}
352 \end{center}
353 \caption {Density projection of molecular centres (solid) and
354 terminal
355 dipoles (broken) with respect to the director g(z)
356 for a system of G-B molecules with single transverse outward
357 pointing dipole, having $d^{*}=1.0$, $\mu^{*}=1.0$} \label{fig:3}
358 \end{figure}
359
360 \begin{figure}
361 \centering \epsfxsize=2.5in \epsfbox{fig4.eps} \caption{Typical
362 configuration for a system of 864 G-B molecules with two transverse
363 dipoles, the first dipole having $d^{*}=1.0$, $\mu^{*}=1.0$ and the
364 second dipole having $d^{*}=0.0$, $\mu^{*}=0.5$. The white caps
365 indicate the location of the terminal dipole, while the orientation
366 of the dipoles is indicated by the blue/gold coloring.}
367 \label{fig:4}
368 \end{figure}
369
370 \begin{figure}
371 \begin{center}
372 \epsfxsize=3in \epsfbox{fig5.ps}
373 \end{center}
374 \caption {Snapshot of molecular configuration for a system of 864
375 G-B molecules with two antiparallel longitudinal dipoles, the first
376 outward pointing dipole
377 having $d^{*}=1.0$, $\mu^{*}=1.0$ and the second dipole having $d^{*}=0.0$,
378 $\mu^{*}=0.5$ (fine lines are molecular symmetry axes and small
379 thick lines show terminal dipolar direction, central dipoles are not
380 shown).} \label{fig:5}
381 \end{figure}
382
383
384 \begin{figure}
385 \begin{center}
386 \epsfxsize=3in \epsfbox{fig6.ps}
387 \end{center}
388 \caption {Snapshot of molecular configuration for a system of 864
389 G-B molecules with single transverse outward pointing dipole, having
390 $d^{*}=1.0$, $\mu^{*}=1.0$ (fine lines are molecular symmetry axes
391 and small thick lines show terminal dipolar direction).}
392 \label{fig:6}
393 \end{figure}
394
395 \begin{figure}
396 \begin{center}
397 \epsfxsize=3in \epsfbox{fig7.ps}
398 \end{center}
399 \caption {Snapshot of molecular configuration for a system of 1372
400 G-B molecules with two transverse outward pointing dipoles, the
401 first dipole having $d^{*}=1.0$, $\mu^{*}=1.0$ and the second dipole
402 having $d^{*}=0.0$, $\mu^{*}=0.5$(fine lines are molecular symmetry
403 axes and small thick lines show terminal dipolar direction, central
404 dipoles are not shown).} \label{fig:7}
405 \end{figure}
406
407 \begin{figure}
408 \begin{center}
409 \epsfxsize=3in \epsfbox{fig8.ps}
410 \end{center}
411 \caption {Snapshot of molecular configuration for a system of 1372
412 G-B molecules with single transverse outward pointing dipole, having
413 $d^{*}=1.0$, $\mu^{*}=1.0$ (fine lines are molecular symmetry axes
414 and small thick lines show terminal dipolar direction).}
415 \label{fig:8}
416 \end{figure}
417
418 Starting from an isotropic configuaration of polar Gay-Berne
419 molecules, we could successfully simulate perfect bilayer, antiphase
420 and monolayer structure. To break the up-down symmetry i.e. the
421 nonequivalence of directions ${\bf \hat {n}}$ and ${ -\bf \hat{n}}$,
422 the molecules should have permanent electric or magnetic dipoles.
423 Longitudinal electric dipole interaction could not form polar
424 nematic phase as orientationally disordered phase with larger
425 entropy is stabler than polarly ordered phase. In fact, stronger
426 central dipole moment opposes polar nematic ordering more
427 effectively in case of rod-like molecules. However, polar ordering
428 like bilayer $A_{2}$, interdigitated $A_{d}$, and wavy $\tilde A$ in
429 smectic layers can be achieved, where adjacent layers with opposite
430 polarities makes bulk phase a-polar. More so, lyotropic liquid
431 crystals and bilayer bio-membranes can have polar layers. These
432 arrangements appear to get favours with the shifting of longitudinal
433 dipole moment to the molecular terminus, so that they can have
434 anti-ferroelectric dipolar arrangement giving rise to local (within
435 the sublayer) breaking of up-down symmetry along the director.
436 Transverse polarity breaks two-fold rotational symmetry, which
437 favours more in-plane polar order. However, the molecular origin of
438 these phases requires something more which are apparent from the
439 earlier simulation results. We have shown that to get perfect
440 bilayer structure in a G-B system, alongwith transverse terminal
441 dipole, another central dipole (or a polarizable core) is required
442 so that polar head and a-polar tail of Gay-Berne molecules go to
443 opposite directions within a bilayer. This gives some kind of
444 clipping interactions which forbid the molecular tail go in other
445 way. Moreover, we could simulate other varieties of polar smectic
446 phases e.g. monolayer $A_{1}$, antiphase $\tilde A$ successfully.
447 Apart from guiding chemical synthesization of ferroelectric,
448 antiferroelectric liquid crystals for technological applications,
449 the present study will be of scientific interest in understanding
450 molecular level interactions of lyotropic liquid crystals as well as
451 nature-designed bio-membranes.
452
453 \section{\label{liquidCrystalSection:methods}Methods}
454
455 \section{\label{liquidCrystalSection:resultDiscussion}Results and Discussion}
456
457 \section{Conclusion}