ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4048
Committed: Tue Feb 25 18:52:18 2014 UTC (10 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 29823 byte(s)
Log Message:
Edits

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3
4 \usepackage{caption}
5 \usepackage{float}
6 \usepackage{geometry}
7 \usepackage{natbib}
8 \usepackage{setspace}
9 \usepackage{xkeyval}
10 \usepackage{amsmath}
11 \usepackage{amssymb}
12 \usepackage{times}
13 \usepackage{mathptm}
14 \usepackage{setspace}
15 %\usepackage{endfloat}
16 \usepackage{tabularx}
17 \usepackage{longtable}
18 \usepackage{graphicx}
19 \usepackage{multirow}
20 \usepackage{multicol}
21 \usepackage{achemso}
22 \usepackage{subcaption}
23 \usepackage[colorinlistoftodos]{todonotes}
24 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25 % \usepackage[square, comma, sort&compress]{natbib}
26 \usepackage{url}
27 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29 9.0in \textwidth 6.5in \brokenpenalty=10000
30
31 % double space list of tables and figures
32 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33 \setlength{\abovecaptionskip}{20 pt}
34 \setlength{\belowcaptionskip}{30 pt}
35
36 % \bibpunct{}{}{,}{s}{}{;}
37
38 %\citestyle{nature}
39 % \bibliographystyle{achemso}
40
41
42 \title{Nitrile vibrations as reporters of field-induced phase
43 transitions in 4-cyano-4'-pentylbiphenyl (5CB)}
44 \author{James M. Marr}
45 \author{J. Daniel Gezelter}
46 \email{gezelter@nd.edu}
47 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
48 Department of Chemistry and Biochemistry\\
49 University of Notre Dame\\
50 Notre Dame, Indiana 46556}
51
52 \date{\today}
53
54 \begin{document}
55
56 \maketitle
57
58 \begin{doublespace}
59
60 \begin{abstract}
61 4-cyano-4'-pentylbiphenyl (5CB) is a liquid-crystal-forming compound
62 with a terminal nitrile group aligned with the long axis of the
63 molecule. Simulations of condensed-phase 5CB were carried out both
64 with and without applied electric fields to provide an understanding
65 of the the Stark shift of the terminal nitrile group. A
66 field-induced isotropic-nematic phase transition was observed in the
67 simulations, and the effects of this transition on the distribution
68 of nitrile frequencies were computed. Classical bond displacement
69 correlation functions exhibit a $\sim 40 \mathrm{~cm}^{-1}$ red
70 shift of a portion of the main nitrile peak, and this shift was
71 observed only when the fields were large enough to induce
72 orientational ordering of the bulk phase. Our simulations appear to
73 indicate that phase-induced changes to the local surroundings are a
74 larger contribution to the change in the nitrile spectrum than
75 direct field contributions.
76 \end{abstract}
77
78 \newpage
79
80 \section{Introduction}
81
82 Nitrile groups can serve as very precise electric field reporters via
83 their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
84 triple bond between the nitrogen and the carbon atom is very sensitive
85 to local field changes and has been observed to have a direct impact
86 on the peak position within the spectrum. The Stark shift in the
87 spectrum can be quantified and mapped onto a field that is impinging
88 upon the nitrile bond. The response of nitrile groups to electric
89 fields has now been investigated for a number of small
90 molecules,\cite{Andrews:2000qv} as well as in biochemical settings,
91 where nitrile groups can act as minimally invasive probes of structure
92 and
93 dynamics.\cite{Tucker:2004qq,Webb:2008kn,Lindquist:2009fk,Fafarman:2010dq}
94 The vibrational Stark effect has also been used to study the effects
95 of electric fields on nitrile-containing self-assembled monolayers at
96 metallic interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
97
98
99 Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
100 molecule with a terminal nitrile group, has seen renewed interest as
101 one way to impart order on the surfactant interfaces of
102 nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
103 that can be used to promote particular kinds of
104 self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
105 is a particularly interesting case for studying electric field
106 effects, as 5CB exhibits an isotropic to nematic phase transition that
107 can be triggered by the application of an external field near room
108 temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
109 possiblity that the field-induced changes in the local environment
110 could have dramatic effects on the vibrations of this particular CN
111 bond. Although the infrared spectroscopy of 5CB has been
112 well-investigated, particularly as a measure of the kinetics of the
113 phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
114 seen the detailed theoretical treatment that biologically-relevant
115 small molecules have
116 received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Morales:2009fp,Waegele:2010ve}
117
118 The fundamental characteristic of liquid crystal mesophases is that
119 they maintain some degree of orientational order while translational
120 order is limited or absent. This orientational order produces a
121 complex direction-dependent response to external perturbations like
122 electric fields and mechanical distortions. The anisotropy of the
123 macroscopic phases originates in the anisotropy of the constituent
124 molecules, which typically have highly non-spherical structures with a
125 significant degree of internal rigidity. In nematic phases, rod-like
126 molecules are orientationally ordered with isotropic distributions of
127 molecular centers of mass. For example, 5CB has a solid to nematic
128 phase transition at 18C and a nematic to isotropic transition at
129 35C.\cite{Gray:1973ca}
130
131 In smectic phases, the molecules arrange themselves into layers with
132 their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
133 respect to the layer planes. The behavior of the $S_{A}$ phase can be
134 partially explained with models mainly based on geometric factors and
135 van der Waals interactions. The Gay-Berne potential, in particular,
136 has been widely used in the liquid crystal community to describe this
137 anisotropic phase
138 behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
139 However, these simple models are insufficient to describe liquid
140 crystal phases which exhibit more complex polymorphic nature.
141 Molecules which form $S_{A}$ phases can exhibit a wide variety of
142 subphases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
143 partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
144 ($S_{A_{d}}$), and often have a terminal cyano or nitro group. In
145 particular, lyotropic liquid crystals (those exhibiting liquid crystal
146 phase transitions as a function of water concentration), often have
147 polar head groups or zwitterionic charge separated groups that result
148 in strong dipolar interactions,\cite{Collings:1997rz} and terminal
149 cyano groups (like the one in 5CB) can induce permanent longitudinal
150 dipoles.\cite{Levelut:1981eu} Modeling of the phase behavior of these
151 molecules either requires additional dipolar
152 interactions,\cite{Bose:2012eu} or a unified-atom approach utilizing
153 point charges on the sites that contribute to the dipole
154 moment.\cite{Zhang:2011hh}
155
156 Macroscopic electric fields applied using electrodes on opposing sides
157 of a sample of 5CB have demonstrated the phase change of the molecule
158 as a function of electric field.\cite{Lim:2006xq} These previous
159 studies have shown the nitrile group serves as an excellent indicator
160 of the molecular orientation within the applied field. Lee {\it et
161 al.}~showed a 180 degree change in field direction could be probed
162 with the nitrile peak intensity as it changed along with molecular
163 alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
164
165 While these macroscopic fields work well at indicating the bulk
166 response, the response at a molecular scale has not been studied. With
167 the advent of nano-electrodes and the ability to couple these
168 electrodes to atomic force microscopy, control of electric fields
169 applied across nanometer distances is now possible.\cite{citation1} In
170 special cases where the macroscopic fields are insufficient to cause
171 an observable Stark effect without dielectric breakdown of the
172 material, small potentials across nanometer-sized gaps may be of
173 sufficient strength. For a gap of 5 nm between a lower electrode
174 having a nanoelectrode placed near it via an atomic force microscope,
175 a potential of 1 V applied across the electrodes is equivalent to a
176 field of 2x10\textsuperscript{8} $\frac{V}{M}$. This field is
177 certainly strong enough to cause the isotropic-nematic phase change
178 and as well as a visible Stark tuning of the nitrile bond. We expect
179 that this would be readily visible experimentally through Raman or IR
180 spectroscopy.
181
182 In the sections that follow, we outline a series of coarse-grained
183 classical molecular dynamics simulations of 5CB that were done in the
184 presence of static electric fields. These simulations were then
185 coupled with both {\it ab intio} calculations of CN-deformations and
186 classical bond-length correlation functions to predict spectral
187 shifts. These predictions made should be easily varifiable with
188 scanning electrochemical microscopy experiments.
189
190 \section{Computational Details}
191 The force field used for 5CB was a united-atom model that was
192 parameterized by Guo {\it et al.}\cite{Zhang:2011hh} However, for most
193 of the simulations, each of the phenyl rings was treated as a rigid
194 body to allow for larger time steps and very long simulation times.
195 The geometries of the rigid bodies were taken from equilibrium bond
196 distances and angles. Although the individual phenyl rings were held
197 rigid, bonds, bends, torsions and inversion centers that involved
198 atoms in these substructures (but with connectivity to the rest of the
199 molecule) were still included in the potential and force calculations.
200
201 Periodic simulations cells containing 270 molecules in random
202 orientations were constructed and were locked at experimental
203 densities. Electrostatic interactions were computed using damped
204 shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
205 were equilibrated for 1~ns at a temperature of 300K. Simulations with
206 applied fields were carried out in the microcanonical (NVE) ensemble
207 with an energy corresponding to the average energy from the canonical
208 (NVT) equilibration runs. Typical applied-field equilibration runs
209 were more than 60ns in length.
210
211 Static electric fields with magnitudes similar to what would be
212 available in an experimental setup were applied to the different
213 simulations. With an assumed electrode seperation of 5 nm and an
214 electrostatic potential that is limited by the voltage required to
215 split water (1.23V), the maximum realistic field that could be applied
216 is $\sim 0.024$ V/\AA. Three field environments were investigated:
217 (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
218 field = 0.024 V/\AA\ .
219
220 After the systems had come to equilibrium under the applied fields,
221 additional simulations were carried out with a flexible (Morse)
222 nitrile bond,
223 \begin{displaymath}
224 V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
225 \label{eq:morse}
226 \end{displaymath}
227 where $r_e= 1.157437$ \AA , $D_e = 212.95 \mathrm{~kcal~} /
228 \mathrm{mol}^{-1}$ and $\beta = 2.67566 $\AA~$^{-1}$. These
229 parameters correspond to a vibrational frequency of $2358
230 \mathrm{~cm}^{-1}$, somewhat higher than the experimental
231 frequency. The flexible nitrile moiety required simulation time steps
232 of 1~fs, so the additional flexibility was introducuced only after the
233 rigid systems had come to equilibrium under the applied fields.
234 Whenever time correlation functions were computed from the flexible
235 simulations, statistically-independent configurations were sampled
236 from the last ns of the induced-field runs. These configurations were
237 then equilibrated with the flexible nitrile moiety for 100 ps, and
238 time correlation functions were computed using data sampled from an
239 additional 200 ps of run time carried out in the microcanonical
240 ensemble.
241
242 \section{Field-induced Nematic Ordering}
243
244 In order to characterize the orientational ordering of the system, the
245 primary quantity of interest is the nematic (orientational) order
246 parameter. This was determined using the tensor
247 \begin{equation}
248 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{u}_{i
249 \alpha} \hat{u}_{i \beta} - \delta_{\alpha \beta} \right)
250 \end{equation}
251 where $\alpha, \beta = x, y, z$, and $\hat{u}_i$ is the molecular
252 end-to-end unit vector for molecule $i$. The nematic order parameter
253 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
254 corresponding eigenvector defines the director axis for the phase.
255 $S$ takes on values close to 1 in highly ordered (smectic A) phases,
256 but falls to much smaller values ($0 \rightarrow 0.3$) for isotropic
257 fluids. Note that the nitrogen and the terminal chain atom were used
258 to define the vectors for each molecule, so the typical order
259 parameters are lower than if one defined a vector using only the rigid
260 core of the molecule. In nematic phases, typical values for $S$ are
261 close to 0.5.
262
263 The field-induced phase transition can be clearly seen over the course
264 of a 60 ns equilibration runs in figure \ref{fig:orderParameter}. All
265 three of the systems started in a random (isotropic) packing, with
266 order parameters near 0.2. Over the course 10 ns, the full field
267 causes an alignment of the molecules (due primarily to the interaction
268 of the nitrile group dipole with the electric field). Once this
269 system began exhibiting nematic ordering, the orientational order
270 parameter became stable for the remaining 150 ns of simulation time.
271 It is possible that the partial-field simulation is meta-stable and
272 given enough time, it would eventually find a nematic-ordered phase,
273 but the partial-field simulation was stable as an isotropic phase for
274 the full duration of a 60 ns simulation. Ellipsoidal renderings of the
275 final configurations of the runs shows that the full-field (0.024
276 V/\AA\ ) experienced a isotropic-nematic phase transition and has
277 ordered with a director axis that is parallel to the direction of the
278 applied field.
279
280 \begin{figure}[H]
281 \includegraphics[width=\linewidth]{Figure1}
282 \caption{Evolution of the orientational order parameters for the
283 no-field, partial field, and full field simulations over the
284 course of 60 ns. Each simulation was started from a
285 statistically-independent isotropic configuration. On the right
286 are ellipsoids representing the final configurations at three
287 different field strengths: zero field (bottom), partial field
288 (middle), and full field (top)}
289 \label{fig:orderParameter}
290 \end{figure}
291
292
293 \section{Sampling the CN bond frequency}
294
295 The vibrational frequency of the nitrile bond in 5CB depends on
296 features of the local solvent environment of the individual molecules
297 as well as the bond's orientation relative to the applied field. The
298 primary quantity of interest for interpreting the condensed phase
299 spectrum of this vibration is the distribution of frequencies
300 exhibited by the 5CB nitrile bond under the different electric fields.
301 There have been a number of elegant techniques for obtaining
302 vibrational lineshapes from classical simulations, including a
303 perturbation theory approach,\cite{Morales:2009fp} the use of an
304 optimized QM/MM approach coupled with the fluctuating frequency
305 approximation,\cite{Lindquist:2008qf} and empirical frequency
306 correlation maps.\cite{Oh:2008fk} Three distinct (and comparatively
307 primitive) methods for mapping classical simulations onto vibrational
308 spectra were brought to bear on the simulations in this work:
309 \begin{enumerate}
310 \item Isolated 5CB molecules and their immediate surroundings were
311 extracted from the simulations. These nitrile bonds were stretched
312 and single-point {\em ab initio} calculations were used to obtain
313 Morse-oscillator fits for the local vibrational motion along that
314 bond.
315 \item A static-field extension of the empirical frequency correlation
316 maps developed by Cho {\it et al.}~\cite{Oh:2008fk} for nitrile
317 moieties in water was attempted.
318 \item Classical bond-length autocorrelation functions were Fourier
319 transformed to directly obtain the vibrational spectrum from
320 molecular dynamics simulations.
321 \end{enumerate}
322
323 \subsection{CN frequencies from isolated clusters}
324 The size of the periodic condensed phase system prevented direct
325 computation of the complete library of nitrile bond frequencies using
326 {\it ab initio} methods. In order to sample the nitrile frequencies
327 present in the condensed-phase, individual molecules were selected
328 randomly to serve as the center of a local (gas phase) cluster. To
329 include steric, electrostatic, and other effects from molecules
330 located near the targeted nitrile group, portions of other molecules
331 nearest to the nitrile group were included in the quantum mechanical
332 calculations. The surrounding solvent molecules were divided into
333 ``body'' (the two phenyl rings and the nitrile bond) and ``tail'' (the
334 alkyl chain). Any molecule which had a body atom within 6~\AA\ of the
335 midpoint of the target nitrile bond had its own molecular body (the
336 4-cyano-biphenyl moiety) included in the configuration. Likewise, the
337 entire alkyl tail was included if any tail atom was within 4~\AA\ of
338 the target nitrile bond. If tail atoms (but no body atoms) were
339 included within these distances, only the tail was included as a
340 capped propane molecule.
341
342 \begin{figure}[H]
343 \includegraphics[width=\linewidth]{Figure2}
344 \caption{Cluster calculations were performed on randomly sampled 5CB
345 molecules (shown in red) from each of the simulations. Surrounding
346 molecular bodies were included if any body atoms were within 6
347 \AA\ of the target nitrile bond, and tails were included if they
348 were within 4 \AA. Included portions of these molecules are shown
349 in green. The CN bond on the target molecule was stretched and
350 compressed, and the resulting single point energies were fit to
351 Morse oscillators to obtain a distribution of frequencies.}
352 \label{fig:cluster}
353 \end{figure}
354
355 Inferred hydrogen atom locations were added to the cluster geometries,
356 and the nitrile bond was stretched from 0.87 to 1.52~\AA\ at
357 increments of 0.05~\AA. This generated 13 configurations per gas phase
358 cluster. Single-point energies were computed using the B3LYP
359 functional~\cite{Becke:1993kq,Lee:1988qf} and the 6-311++G(d,p) basis
360 set. For the cluster configurations that had been generated from
361 molecular dynamics running under applied fields, the density
362 functional calculations had a field of $5 \times 10^{-4}$ atomic units
363 ($E_h / (e a_0)$) applied in the $+z$ direction in order to match the
364 molecular dynamics simulations.
365
366 The energies for the stretched / compressed nitrile bond in each of
367 the clusters were used to fit Morse potentials, and the frequencies
368 were obtained from the $0 \rightarrow 1$ transition for the energy
369 levels for this potential.\cite{Morse:1929xy} To obtain a spectrum,
370 each of the frequencies was convoluted with a Lorentzian lineshape
371 with a width of 1.5 $\mathrm{cm}^{-1}$. Available computing resources
372 limited the sampling to 67 clusters for the zero-field spectrum, and
373 59 for the full field. Comparisons of the quantum mechanical spectrum
374 to the classical are shown in figure \ref{fig:spectrum}.
375
376 \subsection{CN frequencies from potential-frequency maps}
377
378 One approach which has been used to successfully analyze the spectrum
379 of nitrile and thiocyanate probes in aqueous environments was
380 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} This
381 method involves finding a multi-parameter fit that maps between the
382 local electrostatic potential at selected sites surrounding the
383 nitrile bond and the vibrational frequency of that bond obtained from
384 more expensive {\it ab initio} methods. This approach is similar in
385 character to the field-frequency maps developed by the Skinner group
386 for OH stretches in liquid water.\cite{Corcelli:2004ai,Auer:2007dp}
387
388 To use the potential-frequency maps, the local electrostatic
389 potential, $\phi_a$, is computed at 20 sites ($a = 1 \rightarrow 20$)
390 that surround the nitrile bond,
391 \begin{equation}
392 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{j}
393 \frac{q_j}{\left|r_{aj}\right|}.
394 \end{equation}
395 Here $q_j$ is the partial site on atom $j$ (residing on a different
396 molecule) and $r_{aj}$ is the distance between site $a$ and atom $j$.
397 The original map was parameterized in liquid water and comprises a set
398 of parameters, $l_a$, that predict the shift in nitrile peak
399 frequency,
400 \begin{equation}
401 \delta\tilde{\nu} =\sum^{20}_{a=1} l_{a}\phi_{a}.
402 \end{equation}
403
404 The simulations of 5CB were carried out in the presence of
405 externally-applied uniform electric fields. Although uniform fields
406 exert forces on charge sites, they only contribute to the potential if
407 one defines a reference point that can serve as an origin. One simple
408 modification to the potential at each of the probe sites is to use the
409 centroid of the \ce{CN} bond as the origin for that site,
410 \begin{equation}
411 \phi_a^\prime = \phi_a + \frac{1}{4\pi\epsilon_{0}} \vec{E} \cdot
412 \left(\vec{r}_a - \vec{r}_\ce{CN} \right)
413 \end{equation}
414 where $\vec{E}$ is the uniform electric field, $\left( \vec{r}_{a} -
415 \vec{r}_\ce{CN} \right)$ is the displacement between the
416 cooridinates described by Choi {\it et
417 al.}~\cite{Choi:2008cr,Oh:2008fk} and the \ce{CN} bond centroid.
418 $\phi_a^\prime$ then contains an effective potential contributed by
419 the uniform field in addition to the local potential contributions
420 from other molecules.
421
422 The sites $\{\vec{r}_a\}$ and weights $\left\{l_a \right\}$
423 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} are quite
424 symmetric around the \ce{CN} centroid, and even at large uniform field
425 values we observed nearly-complete cancellation of the potenial
426 contributions from the uniform field. In order to utilize the
427 potential-frequency maps for this problem, one would therefore need
428 extensive reparameterization of the maps to include explicit
429 contributions from the external field. This reparameterization is
430 outside the scope of the current work, but would make a useful
431 addition to the potential-frequency map approach.
432
433 \subsection{CN frequencies from bond length autocorrelation functions}
434
435 The distribution of nitrile vibrational frequencies can also be found
436 using classical time correlation functions. This was done by
437 replacing the rigid \ce{CN} bond with a flexible Morse oscillator
438 described in Eq. \ref{eq:morse}. Since the systems were perturbed by
439 the addition of a flexible high-frequency bond, they were allowed to
440 re-equilibrate in the canonical (NVT) ensemble for 100 ps with 1 fs
441 timesteps. After equilibration, each configuration was run in the
442 microcanonical (NVE) ensemble for 20 ps. Configurations sampled every
443 fs were then used to compute bond-length autocorrelation functions,
444 \begin{equation}
445 C(t) = \langle \delta r(t) \cdot \delta r(0) ) \rangle
446 \end{equation}
447 %
448 where $\delta r(t) = r(t) - r_0$ is the deviation from the equilibrium
449 bond distance at time $t$. Because the other atomic sites have very
450 small partial charges, this correlation function is an approximation
451 to the dipole autocorrelation function for the molecule, which would
452 be particularly relevant to computing the IR spectrum. Ten
453 statistically-independent correlation functions were obtained by
454 allowing the systems to run 10 ns with rigid \ce{CN} bonds followed by
455 120 ps equilibration and data collection using the flexible \ce{CN}
456 bonds. This process was repeated 10 times, and the total sampling
457 time, from sample preparation to final configurations, exceeded 150 ns
458 for each of the field strengths investigated.
459
460 The correlation functions were filtered using exponential apodization
461 functions,\cite{FILLER:1964yg} $f(t) = e^{-|t|/c}$, with a time
462 constant, $c =$ 3.5 ps, and were Fourier transformed to yield a
463 spectrum,
464 \begin{equation}
465 I(\omega) = \int_{-\infty}^{\infty} C(t) f(t) e^{-i \omega t} dt.
466 \end{equation}
467 The sample-averaged classical nitrile spectrum can be seen in Figure
468 \ref{fig:spectra}. Note that the Morse oscillator parameters listed
469 above yield a natural frequency of 2358 $\mathrm{cm}^{-1}$, somewhat
470 higher than the experimental peak near 2226 $\mathrm{cm}^{-1}$. This
471 shift does not effect the ability to qualitatively compare peaks from
472 the classical and quantum mechanical approaches, so the classical
473 spectra are shown as a shift relative to the natural oscillation of
474 the Morse bond.
475
476 \begin{figure}
477 \includegraphics[width=3.25in]{Convolved}
478 \includegraphics[width=3.25in]{2Spectra}
479 \caption{Quantum mechanical nitrile spectrum for the no-field simulation
480 (black) and the full field simulation (red). The lower panel
481 shows the corresponding classical bond-length autocorrelation
482 spectrum for the flexible nitrile measured relative to the natural
483 frequency for the flexible bond.}
484 \label{fig:spectra}
485 \end{figure}
486
487 Note that due to electrostatic interactions, the classical approach
488 implicitly couples \ce{CN} vibrations to the same vibrational mode on
489 other nearby molecules. This coupling is not handled in the {\it ab
490 initio} cluster approach.
491
492 \section{Discussion}
493
494 Our simulations show that the united-atom model can reproduce the
495 field-induced nematic ordering of the 4-cyano-4'-pentylbiphenyl.
496 Because we are simulating what is in effect a small electrode
497 separation (5nm), a voltage drop as low as 1.2 V was sufficient to
498 induce the phase change. This potential is significantly lower than
499 the 500V that is known to cause dielectric breakdown in 5CB.\cite{XXX}
500
501 Both the classical correlation function and the isolated cluster
502 approaches to estimating the field-induced changes to the IR spectrum
503 show an increase in the population of nitrile stretches that appear at
504 a shift of $\sim 40 \mathrm{cm}^{-1}$ to the red of the unperturbed
505 vibrational line. The cause of this shift does not appear to be
506 related to the alignment of those nitrile bonds with the field, but
507 rather to the change in local environment that is brought about by the
508 isotropic-nematic transition.
509
510 The angle-dependent pair distribution functions,
511 \begin{align}
512 g(r, \cos \omega) = & \frac{1}{\rho N} \left< \sum_{i}
513 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \omega_{ij} -
514 \cos \omega\right) \right> \\ \nonumber \\
515 g(r, \cos \theta) = & \frac{1}{\rho N} \left< \sum_{i}
516 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \theta_{i} -
517 \cos \theta \right) \right>
518 \end{align}
519 provide information about the joint spatial and angular correlations
520 in the system. The angles $\omega$ and $\theta$ are defined by vectors
521 along the CN axis of each nitrile bond (see figure
522 \ref{fig:definition}).
523
524 \begin{figure}
525 \includegraphics[width=\linewidth]{definition}
526 \caption{Definitions of the angles between two nitrile bonds.}
527 \label{fig:definition}
528 \end{figure}
529
530 In figure \ref{fig:gofromega}, one of the structural effects of the
531 field-induced phase transition is clear. The nematic ordering
532 transfers population from the perpendicular or unaligned region in the
533 center of the plot to the nitrile-alinged peak near $\cos\omega =
534 1$. Most other features are undisturbed. The major change visible is
535 the increased population of aligned nitrile bonds in the first
536 solvation shells.
537
538 \begin{figure}
539 \includegraphics[width=\linewidth]{Figure4}
540 \caption{Contours of the angle-dependent pair distribution functions
541 for nitrile bonds on 5CB in the zero-field (upper panel) and full
542 field (lower panel) simulations. Dark areas signify regions of
543 enhanced density, while light areas signify depletion relative to
544 the bulk density.}
545 \label{fig:gofromega}
546 \end{figure}
547
548 Although it is possible that the coupling between closely-spaced
549 nitrile pairs is responsible for some of the red-shift, that is not
550 the complete picture. The other two dimensional pair distribution
551 function, $g(r,\cos\theta)$, shows that nematic ordering also
552 transfers population that is directly in line with the nitrile bond
553 (see figure \ref{fig:gofrtheta}) to the sides of the molecule, thereby
554 freeing steric blockage that directly blocks the nitrile vibratio
555 \begin{figure}
556
557 \includegraphics[width=\linewidth]{Figure5}
558 \caption{Contours of the angle-dependent pair distribution function,
559 $g(r,\cos \theta)$, for finding any atom at a distance and angular
560 deviation from the nitrile bond centroid. The right side of each
561 plot corresponds to local density directly the direction of
562 nitrile bond. Increased density at $\cos\theta = 1$ corresponds
563 to steric hindrance of the nitrile bond.}
564 \label{fig:gofromega}
565 \end{figure}
566
567 .At the same time, the system exhibits an increase in aligned
568 and anti-a
569
570
571
572
573
574
575
576 While this makes the application of nitrile Stark effects in
577 simulations without water harder, these data show
578 that it is not a deal breaker. The classically calculated nitrile
579 spectrum shows changes in the spectra that will be easily seen through
580 experimental routes. It indicates a shifted peak lower in energy
581 should arise. This peak is a few wavenumbers from the leading edge of
582 the larger peak and almost 75 wavenumbers from the center. This
583 seperation between the two peaks means experimental results will show
584 an easily resolved peak.
585
586 The Gaussian derived spectra do indicate an applied field
587 and subsiquent phase change does cause a narrowing of freuency
588 distrobution. With narrowing, it would indicate an increased
589 homogeneous distrobution of the local field near the nitrile.
590
591
592
593 \section{Conclusions}
594 Field dependent changes
595
596 \section{Acknowledgements}
597 The authors thank Steven Corcelli for helpful comments and
598 suggestions. Support for this project was provided by the National
599 Science Foundation under grant CHE-0848243. Computational time was
600 provided by the Center for Research Computing (CRC) at the University
601 of Notre Dame.
602
603 \newpage
604
605 \bibliography{5CB}
606
607 \end{doublespace}
608 \end{document}