ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4094
Committed: Tue Apr 1 00:11:42 2014 UTC (10 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 32034 byte(s)
Log Message:
Updates

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
28 \title{Nitrile vibrations as reporters of field-induced phase
29 transitions in 4-cyano-4'-pentylbiphenyl (5CB)}
30 \author{James M. Marr}
31 \author{J. Daniel Gezelter}
32 \email{gezelter@nd.edu}
33 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
34 Department of Chemistry and Biochemistry\\
35 University of Notre Dame\\
36 Notre Dame, Indiana 46556}
37
38 \begin{document}
39
40
41 \begin{tocentry}
42 %\includegraphics[width=9cm]{Elip_3}
43 \includegraphics[width=9cm]{Figure2}
44 \end{tocentry}
45
46 \begin{abstract}
47 4-cyano-4'-pentylbiphenyl (5CB) is a liquid-crystal-forming compound
48 with a terminal nitrile group aligned with the long axis of the
49 molecule. Simulations of condensed-phase 5CB were carried out both
50 with and without applied electric fields to provide an understanding
51 of the Stark shift of the terminal nitrile group. A field-induced
52 isotropic-nematic phase transition was observed in the simulations,
53 and the effects of this transition on the distribution of nitrile
54 frequencies were computed. Classical bond displacement correlation
55 functions exhibit a $\sim~3~\mathrm{cm}^{-1}$ red shift of a
56 portion of the main nitrile peak, and this shift was observed only
57 when the fields were large enough to induce orientational ordering
58 of the bulk phase. Joint spatial-angular distribution functions
59 indicate that phase-induced anti-caging of the nitrile bond is
60 contributing to the change in the nitrile spectrum.
61 \end{abstract}
62
63 \newpage
64
65 \section{Introduction}
66
67 Nitrile groups can serve as very precise electric field reporters via
68 their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
69 triple bond between the nitrogen and the carbon atom is very sensitive
70 to local field changes and has been observed to have a direct impact
71 on the peak position within the spectrum. The Stark shift in the
72 spectrum can be quantified and mapped onto a field that is impinging
73 upon the nitrile bond. The response of nitrile groups to electric
74 fields has now been investigated for a number of small
75 molecules,\cite{Andrews:2000qv} as well as in biochemical settings,
76 where nitrile groups can act as minimally invasive probes of structure
77 and
78 dynamics.\cite{Tucker:2004qq,Webb:2008kn,Lindquist:2009fk,Fafarman:2010dq}
79 The vibrational Stark effect has also been used to study the effects
80 of electric fields on nitrile-containing self-assembled monolayers at
81 metallic interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
82
83
84 Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
85 molecule with a terminal nitrile group, has seen renewed interest as
86 one way to impart order on the surfactant interfaces of
87 nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
88 that can be used to promote particular kinds of
89 self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
90 is a particularly interesting case for studying electric field
91 effects, as 5CB exhibits an isotropic to nematic phase transition that
92 can be triggered by the application of an external field near room
93 temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
94 possiblity that the field-induced changes in the local environment
95 could have dramatic effects on the vibrations of this particular CN
96 bond. Although the infrared spectroscopy of 5CB has been
97 well-investigated, particularly as a measure of the kinetics of the
98 phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
99 seen the detailed theoretical treatment that biologically-relevant
100 small molecules have
101 received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Morales:2009fp,Waegele:2010ve}
102
103 The fundamental characteristic of liquid crystal mesophases is that
104 they maintain some degree of orientational order while translational
105 order is limited or absent. This orientational order produces a
106 complex direction-dependent response to external perturbations like
107 electric fields and mechanical distortions. The anisotropy of the
108 macroscopic phases originates in the anisotropy of the constituent
109 molecules, which typically have highly non-spherical structures with a
110 significant degree of internal rigidity. In nematic phases, rod-like
111 molecules are orientationally ordered with isotropic distributions of
112 molecular centers of mass. For example, 5CB has a solid to nematic
113 phase transition at 18C and a nematic to isotropic transition at
114 35C.\cite{Gray:1973ca}
115
116 In smectic phases, the molecules arrange themselves into layers with
117 their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
118 respect to the layer planes. The behavior of the $S_{A}$ phase can be
119 partially explained with models mainly based on geometric factors and
120 van der Waals interactions. The Gay-Berne potential, in particular,
121 has been widely used in the liquid crystal community to describe this
122 anisotropic phase
123 behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
124 However, these simple models are insufficient to describe liquid
125 crystal phases which exhibit more complex polymorphic nature.
126 Molecules which form $S_{A}$ phases can exhibit a wide variety of
127 subphases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
128 partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
129 ($S_{A_{d}}$), and often have a terminal cyano or nitro group. In
130 particular, lyotropic liquid crystals (those exhibiting liquid crystal
131 phase transitions as a function of water concentration), often have
132 polar head groups or zwitterionic charge separated groups that result
133 in strong dipolar interactions,\cite{Collings:1997rz} and terminal
134 cyano groups (like the one in 5CB) can induce permanent longitudinal
135 dipoles.\cite{Levelut:1981eu} Modeling of the phase behavior of these
136 molecules either requires additional dipolar
137 interactions,\cite{Bose:2012eu} or a unified-atom approach utilizing
138 point charges on the sites that contribute to the dipole
139 moment.\cite{Zhang:2011hh}
140
141 Macroscopic electric fields applied using electrodes on opposing sides
142 of a sample of 5CB have demonstrated the phase change of the molecule
143 as a function of electric field.\cite{Lim:2006xq} These previous
144 studies have shown the nitrile group serves as an excellent indicator
145 of the molecular orientation within the applied field. Lee {\it et
146 al.}~showed a 180 degree change in field direction could be probed
147 with the nitrile peak intensity as it changed along with molecular
148 alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
149
150 While these macroscopic fields work well at indicating the bulk
151 response, the response at a molecular scale has not been studied. With
152 the advent of nano-electrodes and the ability to couple these
153 electrodes to atomic force microscopy, control of electric fields
154 applied across nanometer distances is now possible.\cite{C3AN01651J}
155 In special cases where the macroscopic fields are insufficient to
156 cause an observable Stark effect without dielectric breakdown of the
157 material, small potentials across nanometer-sized gaps may be of
158 sufficient strength. For a gap of 5 nm between a lower electrode
159 having a nanoelectrode placed near it via an atomic force microscope,
160 a potential of 1 V applied across the electrodes is equivalent to a
161 field of 2x10\textsuperscript{8} $\frac{V}{M}$. This field is
162 certainly strong enough to cause the isotropic-nematic phase change
163 and as well as a visible Stark tuning of the nitrile bond. We expect
164 that this would be readily visible experimentally through Raman or IR
165 spectroscopy.
166
167 In the sections that follow, we outline a series of coarse-grained
168 classical molecular dynamics simulations of 5CB that were done in the
169 presence of static electric fields. These simulations were then
170 coupled with both {\it ab intio} calculations of CN-deformations and
171 classical bond-length correlation functions to predict spectral
172 shifts. These predictions made should be easily verifiable with
173 scanning electrochemical microscopy experiments.
174
175 \section{Computational Details}
176 The force-field used to model 5CB was a united-atom model that was
177 parameterized by Guo {\it et al.}\cite{Zhang:2011hh} However, for most
178 of the simulations, each of the phenyl rings was treated as a rigid
179 body to allow for larger time steps and longer simulation times. The
180 geometries of the rigid bodies were taken from equilibrium bond
181 distances and angles. Although the individual phenyl rings were held
182 rigid, bonds, bends, torsions and inversion centers that involved
183 atoms in these substructures (but with connectivity to the rest of the
184 molecule) were still included in the potential and force calculations.
185
186 Periodic simulations cells containing 270 molecules in random
187 orientations were constructed and were locked at experimental
188 densities. Electrostatic interactions were computed using damped
189 shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
190 were equilibrated for 1~ns at a temperature of 300K. Simulations with
191 applied fields were carried out in the microcanonical (NVE) ensemble
192 with an energy corresponding to the average energy from the canonical
193 (NVT) equilibration runs. Typical applied-field equilibration runs
194 were more than 60~ns in length.
195
196 Static electric fields with magnitudes similar to what would be
197 available in an experimental setup were applied to the different
198 simulations. With an assumed electrode seperation of 5 nm and an
199 electrostatic potential that is limited by the voltage required to
200 split water (1.23V), the maximum realistic field that could be applied
201 is $\sim 0.024$ V/\AA. Three field environments were investigated:
202 (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
203 field = 0.024 V/\AA\ .
204
205 After the systems had come to equilibrium under the applied fields,
206 additional simulations were carried out with a flexible (Morse)
207 nitrile bond,
208 \begin{displaymath}
209 V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
210 \label{eq:morse}
211 \end{displaymath}
212 where $r_e= 1.157437$ \AA , $D_e = 212.95 \mathrm{~kcal~} /
213 \mathrm{mol}^{-1}$ and $\beta = 2.67566 $\AA~$^{-1}$. These
214 parameters correspond to a vibrational frequency of $2358
215 \mathrm{~cm}^{-1}$, somewhat higher than the experimental
216 frequency. The flexible nitrile moiety required simulation time steps
217 of 1~fs, so the additional flexibility was introducuced only after the
218 rigid systems had come to equilibrium under the applied fields.
219 Whenever time correlation functions were computed from the flexible
220 simulations, statistically-independent configurations were sampled
221 from the last ns of the induced-field runs. These configurations were
222 then equilibrated with the flexible nitrile moiety for 100 ps, and
223 time correlation functions were computed using data sampled from an
224 additional 200 ps of run time carried out in the microcanonical
225 ensemble.
226
227 \section{Field-induced Nematic Ordering}
228
229 In order to characterize the orientational ordering of the system, the
230 primary quantity of interest is the nematic (orientational) order
231 parameter. This was determined using the tensor
232 \begin{equation}
233 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{u}_{i
234 \alpha} \hat{u}_{i \beta} - \delta_{\alpha \beta} \right)
235 \end{equation}
236 where $\alpha, \beta = x, y, z$, and $\hat{u}_i$ is the molecular
237 end-to-end unit vector for molecule $i$. The nematic order parameter
238 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
239 corresponding eigenvector defines the director axis for the phase.
240 $S$ takes on values close to 1 in highly ordered (smectic A) phases,
241 but falls to much smaller values ($0 \rightarrow 0.3$) for isotropic
242 fluids. Note that the nitrogen and the terminal chain atom were used
243 to define the vectors for each molecule, so the typical order
244 parameters are lower than if one defined a vector using only the rigid
245 core of the molecule. In nematic phases, typical values for $S$ are
246 close to 0.5.
247
248 The field-induced phase transition can be clearly seen over the course
249 of a 60 ns equilibration runs in figure \ref{fig:orderParameter}. All
250 three of the systems started in a random (isotropic) packing, with
251 order parameters near 0.2. Over the course 10 ns, the full field
252 causes an alignment of the molecules (due primarily to the interaction
253 of the nitrile group dipole with the electric field). Once this
254 system began exhibiting nematic ordering, the orientational order
255 parameter became stable for the remaining 150 ns of simulation time.
256 It is possible that the partial-field simulation is meta-stable and
257 given enough time, it would eventually find a nematic-ordered phase,
258 but the partial-field simulation was stable as an isotropic phase for
259 the full duration of a 60 ns simulation. Ellipsoidal renderings of the
260 final configurations of the runs shows that the full-field (0.024
261 V/\AA\ ) experienced a isotropic-nematic phase transition and has
262 ordered with a director axis that is parallel to the direction of the
263 applied field.
264
265 \begin{figure}[H]
266 \includegraphics[width=\linewidth]{Figure1}
267 \caption{Evolution of the orientational order parameters for the
268 no-field, partial field, and full field simulations over the
269 course of 60 ns. Each simulation was started from a
270 statistically-independent isotropic configuration. On the right
271 are ellipsoids representing the final configurations at three
272 different field strengths: zero field (bottom), partial field
273 (middle), and full field (top)}
274 \label{fig:orderParameter}
275 \end{figure}
276
277
278 \section{Sampling the CN bond frequency}
279
280 The vibrational frequency of the nitrile bond in 5CB depends on
281 features of the local solvent environment of the individual molecules
282 as well as the bond's orientation relative to the applied field. The
283 primary quantity of interest for interpreting the condensed phase
284 spectrum of this vibration is the distribution of frequencies
285 exhibited by the 5CB nitrile bond under the different electric fields.
286 There have been a number of elegant techniques for obtaining
287 vibrational lineshapes from classical simulations, including a
288 perturbation theory approach,\cite{Morales:2009fp} the use of an
289 optimized QM/MM approach coupled with the fluctuating frequency
290 approximation,\cite{Lindquist:2008qf} and empirical frequency
291 correlation maps.\cite{Oh:2008fk} Three distinct (and comparatively
292 primitive) methods for mapping classical simulations onto vibrational
293 spectra were brought to bear on the simulations in this work:
294 \begin{enumerate}
295 \item Isolated 5CB molecules and their immediate surroundings were
296 extracted from the simulations. These nitrile bonds were stretched
297 and single-point {\em ab initio} calculations were used to obtain
298 Morse-oscillator fits for the local vibrational motion along that
299 bond.
300 \item A static-field extension of the empirical frequency correlation
301 maps developed by Cho {\it et al.}~\cite{Oh:2008fk} for nitrile
302 moieties in water was attempted.
303 \item Classical bond-length autocorrelation functions were Fourier
304 transformed to directly obtain the vibrational spectrum from
305 molecular dynamics simulations.
306 \end{enumerate}
307
308 \subsection{CN frequencies from isolated clusters}
309 The size of the periodic condensed phase system prevented direct
310 computation of the complete library of nitrile bond frequencies using
311 {\it ab initio} methods. In order to sample the nitrile frequencies
312 present in the condensed-phase, individual molecules were selected
313 randomly to serve as the center of a local (gas phase) cluster. To
314 include steric, electrostatic, and other effects from molecules
315 located near the targeted nitrile group, portions of other molecules
316 nearest to the nitrile group were included in the quantum mechanical
317 calculations. The surrounding solvent molecules were divided into
318 ``body'' (the two phenyl rings and the nitrile bond) and ``tail'' (the
319 alkyl chain). Any molecule which had a body atom within 6~\AA\ of the
320 midpoint of the target nitrile bond had its own molecular body (the
321 4-cyano-biphenyl moiety) included in the configuration. Likewise, the
322 entire alkyl tail was included if any tail atom was within 4~\AA\ of
323 the target nitrile bond. If tail atoms (but no body atoms) were
324 included within these distances, only the tail was included as a
325 capped propane molecule.
326
327 \begin{figure}[H]
328 \includegraphics[width=\linewidth]{Figure2}
329 \caption{Cluster calculations were performed on randomly sampled 5CB
330 molecules (shown in red) from each of the simulations. Surrounding
331 molecular bodies were included if any body atoms were within 6
332 \AA\ of the target nitrile bond, and tails were included if they
333 were within 4 \AA. Included portions of these molecules are shown
334 in green. The CN bond on the target molecule was stretched and
335 compressed, and the resulting single point energies were fit to
336 Morse oscillators to obtain a distribution of frequencies.}
337 \label{fig:cluster}
338 \end{figure}
339
340 Inferred hydrogen atom locations were added to the cluster geometries,
341 and the nitrile bond was stretched from 0.87 to 1.52~\AA\ at
342 increments of 0.05~\AA. This generated 13 configurations per gas phase
343 cluster. Single-point energies were computed using the B3LYP
344 functional~\cite{Becke:1993kq,Lee:1988qf} and the 6-311++G(d,p) basis
345 set. For the cluster configurations that had been generated from
346 molecular dynamics running under applied fields, the density
347 functional calculations had a field of $5 \times 10^{-4}$ atomic units
348 ($E_h / (e a_0)$) applied in the $+z$ direction in order to match the
349 molecular dynamics simulations.
350
351 The energies for the stretched / compressed nitrile bond in each of
352 the clusters were used to fit Morse potentials, and the frequencies
353 were obtained from the $0 \rightarrow 1$ transition for the energy
354 levels for this potential.\cite{Morse:1929xy} To obtain a spectrum,
355 each of the frequencies was convoluted with a Lorentzian lineshape
356 with a width of 1.5 $\mathrm{cm}^{-1}$. Available computing resources
357 limited the sampling to 100 clusters for both the zero-field and full
358 field soectra. Comparisons of the quantum mechanical spectrum to the
359 classical are shown in figure \ref{fig:spectra}.
360
361 \begin{figure}
362 \includegraphics[width=\linewidth]{Figure3}
363 \caption{Spectrum of nitrile frequency shifts for the no-field
364 (black) and the full-field (red) simulations. Upper
365 panel: frequency shifts obtained from {\it ab initio} cluster
366 calculations. Lower panel: classical bond-length autocorrelation
367 spectrum for the flexible nitrile measured relative to the natural
368 frequency for the flexible bond.}
369 \label{fig:spectra}
370 \end{figure}
371
372 \subsection{CN frequencies from potential-frequency maps}
373
374 One approach which has been used to successfully analyze the spectrum
375 of nitrile and thiocyanate probes in aqueous environments was
376 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} This
377 method involves finding a multi-parameter fit that maps between the
378 local electrostatic potential at selected sites surrounding the
379 nitrile bond and the vibrational frequency of that bond obtained from
380 more expensive {\it ab initio} methods. This approach is similar in
381 character to the field-frequency maps developed by the Skinner group
382 for OH stretches in liquid water.\cite{Corcelli:2004ai,Auer:2007dp}
383
384 To use the potential-frequency maps, the local electrostatic
385 potential, $\phi_a$, is computed at 20 sites ($a = 1 \rightarrow 20$)
386 that surround the nitrile bond,
387 \begin{equation}
388 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{j}
389 \frac{q_j}{\left|r_{aj}\right|}.
390 \end{equation}
391 Here $q_j$ is the partial site on atom $j$ (residing on a different
392 molecule) and $r_{aj}$ is the distance between site $a$ and atom $j$.
393 The original map was parameterized in liquid water and comprises a set
394 of parameters, $l_a$, that predict the shift in nitrile peak
395 frequency,
396 \begin{equation}
397 \delta\tilde{\nu} =\sum^{20}_{a=1} l_{a}\phi_{a}.
398 \end{equation}
399
400 The simulations of 5CB were carried out in the presence of
401 externally-applied uniform electric fields. Although uniform fields
402 exert forces on charge sites, they only contribute to the potential if
403 one defines a reference point that can serve as an origin. One simple
404 modification to the potential at each of the probe sites is to use the
405 centroid of the \ce{CN} bond as the origin for that site,
406 \begin{equation}
407 \phi_a^\prime = \phi_a + \frac{1}{4\pi\epsilon_{0}} \vec{E} \cdot
408 \left(\vec{r}_a - \vec{r}_\ce{CN} \right)
409 \end{equation}
410 where $\vec{E}$ is the uniform electric field, $\left( \vec{r}_{a} -
411 \vec{r}_\ce{CN} \right)$ is the displacement between the
412 cooridinates described by Choi {\it et
413 al.}~\cite{Choi:2008cr,Oh:2008fk} and the \ce{CN} bond centroid.
414 $\phi_a^\prime$ then contains an effective potential contributed by
415 the uniform field in addition to the local potential contributions
416 from other molecules.
417
418 The sites $\{\vec{r}_a\}$ and weights $\left\{l_a \right\}$
419 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} are quite
420 symmetric around the \ce{CN} centroid, and even at large uniform field
421 values we observed nearly-complete cancellation of the potenial
422 contributions from the uniform field. In order to utilize the
423 potential-frequency maps for this problem, one would therefore need
424 extensive reparameterization of the maps to include explicit
425 contributions from the external field. This reparameterization is
426 outside the scope of the current work, but would make a useful
427 addition to the potential-frequency map approach.
428
429 We note that in 5CB there does not appear to be a particularly strong
430 correlation between the electric field observed at the nitrile
431 centroid and the calculated vibrational frequency. In
432 Fig. \ref{fig:fieldMap} we show the calculated frequencies plotted
433 against the field magnitude and the parallel and perpendicular
434 components of the field.
435
436 \begin{figure}
437 \includegraphics[width=\linewidth]{Figure7}
438 \caption{The observed cluster frequencies have no apparent
439 correlation with the electric field felt at the centroid of the
440 nitrile bond. Lower panel: vibrational frequencies plotted
441 against the total field magnitude. Middle panel: mapped to the
442 component of the field parallel to the CN bond. Upper panel:
443 mapped to the magnitude of the field perpendicular to the CN
444 bond.}
445 \label{fig:fieldMap}
446 \end{figure}
447
448
449 \subsection{CN frequencies from bond length autocorrelation functions}
450
451 The distribution of nitrile vibrational frequencies can also be found
452 using classical time correlation functions. This was done by
453 replacing the rigid \ce{CN} bond with a flexible Morse oscillator
454 described in Eq. \ref{eq:morse}. Since the systems were perturbed by
455 the addition of a flexible high-frequency bond, they were allowed to
456 re-equilibrate in the canonical (NVT) ensemble for 100 ps with 1 fs
457 timesteps. After equilibration, each configuration was run in the
458 microcanonical (NVE) ensemble for 20 ps. Configurations sampled every
459 fs were then used to compute bond-length autocorrelation functions,
460 \begin{equation}
461 C(t) = \langle \delta r(t) \cdot \delta r(0) ) \rangle
462 \end{equation}
463 %
464 where $\delta r(t) = r(t) - r_0$ is the deviation from the equilibrium
465 bond distance at time $t$. Because the other atomic sites have very
466 small partial charges, this correlation function is an approximation
467 to the dipole autocorrelation function for the molecule, which would
468 be particularly relevant to computing the IR spectrum. Ten
469 statistically-independent correlation functions were obtained by
470 allowing the systems to run 10 ns with rigid \ce{CN} bonds followed by
471 120 ps equilibration and data collection using the flexible \ce{CN}
472 bonds. This process was repeated 10 times, and the total sampling
473 time, from sample preparation to final configurations, exceeded 150 ns
474 for each of the field strengths investigated.
475
476 The correlation functions were filtered using exponential apodization
477 functions,\cite{FILLER:1964yg} $f(t) = e^{-|t|/c}$, with a time
478 constant, $c =$ 3.5 ps, and were Fourier transformed to yield a
479 spectrum,
480 \begin{equation}
481 I(\omega) = \int_{-\infty}^{\infty} C(t) f(t) e^{-i \omega t} dt.
482 \end{equation}
483 The sample-averaged classical nitrile spectrum can be seen in Figure
484 \ref{fig:spectra}. Note that the Morse oscillator parameters listed
485 above yield a natural frequency of 2358 $\mathrm{cm}^{-1}$, somewhat
486 higher than the experimental peak near 2226 $\mathrm{cm}^{-1}$. This
487 shift does not effect the ability to qualitatively compare peaks from
488 the classical and quantum mechanical approaches, so the classical
489 spectra are shown as a shift relative to the natural oscillation of
490 the Morse bond.
491
492
493 The classical approach includes both intramolecular and electrostatic
494 interactions, and so it implicitly couples \ce{CN} vibrations to other
495 vibrations within the molecule as well as to nitrile vibrations on
496 other nearby molecules. The classical frequency spectrum is
497 significantly broader because of this coupling. The {\it
498 ab
499 initio} cluster approach exercises only the targeted nitrile bond,
500 with no additional coupling to other degrees of freedom. As a result
501 the quantum calculations are quite narrowly peaked around the
502 experimental nitrile frequency. Although the spectra are quite noisy,
503 the main effect seen in both the classical and quantum frequency
504 distributions is a moderate shift $\sim 3~\mathrm{cm}^{-1}$ to the
505 red when the full electrostatic field had induced the nematic phase
506 transition.
507
508 \section{Discussion}
509 Our simulations show that the united-atom model can reproduce the
510 field-induced nematic ordering of the 4-cyano-4'-pentylbiphenyl.
511 Because we are simulating a very small electrode separation (5~nm), a
512 voltage drop as low as 1.2~V was sufficient to induce the phase
513 change. This potential is significantly smaller than 100~V that was
514 used with a 5~$\mu$m gap to study the electrochemiluminescence of
515 rubrene in neat 5CB,\cite{Kojima19881789} and suggests that by using
516 electrodes separated by a nanometer-scale gap, it will be relatively
517 straightforward to observe the nitrile Stark shift in 5CB.
518
519 Both the classical correlation function and the isolated cluster
520 approaches to estimating the IR spectrum show that a population of
521 nitrile stretches shift by $\sim~3~\mathrm{cm}^{-1}$ to the red of
522 the unperturbed vibrational line. To understand the origin of this
523 shift, a more complete picture of the spatial ordering around the
524 nitrile bonds is required. We have computed the angle-dependent pair
525 distribution functions,
526 \begin{align}
527 g(r, \cos \omega) = & \frac{1}{\rho N} \left< \sum_{i} \sum_{j}
528 \delta \left(r - r_{ij}\right) \delta\left(\cos \omega_{ij} -
529 \cos \omega\right) \right> \\ \nonumber \\
530 g(r, \cos \theta) = & \frac{1}{\rho N} \left< \sum_{i}
531 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \theta_{i} -
532 \cos \theta \right) \right>
533 \end{align}
534 which provide information about the joint spatial and angular
535 correlations present in the system. The angles $\omega$ and $\theta$
536 are defined by vectors along the CN axis of each nitrile bond (see
537 figure \ref{fig:definition}).
538 \begin{figure}
539 \includegraphics[width=4in]{definition}
540 \caption{Definitions of the angles between two nitrile bonds.}
541 \label{fig:definition}
542 \end{figure}
543
544 The primary structural effect of the field-induced phase transition is
545 apparent in figure \ref{fig:gofromega}. The nematic ordering transfers
546 population from the perpendicular ($\cos\omega\approx 0$) and
547 anti-aligned ($\cos\omega\approx -1$) to the nitrile-alinged peak
548 near $\cos\omega\approx 1$, leaving most other features undisturbed. This
549 change is visible in the simulations as an increased population of
550 aligned nitrile bonds in the first solvation shell.
551
552 \begin{figure}
553 \includegraphics[width=\linewidth]{Figure4}
554 \caption{Contours of the angle-dependent pair distribution functions
555 for nitrile bonds on 5CB in the no field (upper panel) and full
556 field (lower panel) simulations. Dark areas signify regions of
557 enhanced density, while light areas signify depletion relative to
558 the bulk density.}
559 \label{fig:gofromega}
560 \end{figure}
561
562 Although it is certainly possible that the coupling between
563 closely-spaced nitrile pairs is responsible for some of the red-shift,
564 that is not the only structural change that is taking place. The
565 second two-dimensional pair distribution function, $g(r,\cos\theta)$,
566 shows that nematic ordering also transfers population that is directly
567 in line with the nitrile bond (see figure \ref{fig:gofrtheta}) to the
568 sides of the molecule, thereby freeing steric blockage can directly
569 influence the nitrile vibration. This is confirmed by observing the
570 one-dimensional $g(z)$ obtained by following the \ce{C -> N} vector
571 for each nitrile bond and observing the local density ($\rho(z)/\rho$)
572 of other atoms at a distance $z$ along this direction. The full-field
573 simulation shows a significant drop in the first peak of $g(z)$,
574 indicating that the nematic ordering has moved density away from the
575 region that is directly in line with the nitrogen side of the CN bond.
576
577 \begin{figure}
578 \includegraphics[width=\linewidth]{Figure6}
579 \caption{Contours of the angle-dependent pair distribution function,
580 $g(r,\cos \theta)$, for finding any other atom at a distance and
581 angular deviation from the center of a nitrile bond. The top edge
582 of each contour plot corresponds to local density along the
583 direction of the nitrogen in the CN bond, while the bottom is in
584 the direction of the carbon atom. Bottom panel: $g(z)$ data taken
585 by following the \ce{C -> N} vector for each nitrile bond shows
586 that the field-induced phase transition reduces the population of
587 atoms that are directly in line with the nitrogen motion.}
588 \label{fig:gofrtheta}
589 \end{figure}
590
591 We are suggesting an anti-caging mechanism here -- the nematic
592 ordering provides additional space directly inline with the nitrile
593 vibration, and since the oscillator is fairly anharmonic, this freedom
594 provides a fraction of the nitrile bonds with a significant red-shift.
595
596 The cause of this shift does not appear to be related to the alignment
597 of those nitrile bonds with the field, but rather to the change in
598 local steric environment that is brought about by the
599 isotropic-nematic transition. We have compared configurations for many
600 of the cluster that exhibited the lowest frequencies (between 2190 and
601 2215 $\mathrm{cm}^{-1}$) and have observed some similar structural
602 features. The lowest frequencies appear to come from configurations
603 which have nearly-empty pockets directly opposite the nitrogen atom
604 from the nitrile carbon. Because we do not have a particularly large
605 cluster population to interrogate, this is certainly not quantitative
606 confirmation of this effect.
607
608 The prediction of a small red-shift of the nitrile peak in 5CB in
609 response to a field-induced nematic ordering is the primary result of
610 this work, and although the proposed anti-caging mechanism is somewhat
611 speculative, this work provides some impetus for further theory and
612 experiments.
613
614 \section{Acknowledgements}
615 The authors thank Steven Corcelli and Zac Schultz for helpful comments
616 and suggestions. Support for this project was provided by the National
617 Science Foundation under grant CHE-0848243. Computational time was
618 provided by the Center for Research Computing (CRC) at the University
619 of Notre Dame.
620
621 \newpage
622
623 \bibliography{5CB}
624
625 \end{document}