ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4056
Committed: Thu Feb 27 19:42:03 2014 UTC (10 years, 6 months ago) by jmarr
Content type: application/x-tex
File size: 30940 byte(s)
Log Message:
citation1 fixed 

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 Stark shift of the terminal nitrile group. A field-induced
66 isotropic-nematic phase transition was observed in the simulations,
67 and the effects of this transition on the distribution of nitrile
68 frequencies were computed. Classical bond displacement correlation
69 functions exhibit a $\sim 40 \mathrm{~cm}^{-1}$ red shift of a
70 portion of the main nitrile peak, and this shift was observed only
71 when the fields were large enough to induce orientational ordering
72 of the bulk phase. Joint spatial-angular distribution functions
73 indicate that phase-induced anti-caging of the nitrile bond is
74 contributing to the change in the nitrile spectrum.
75 \end{abstract}
76
77 \newpage
78
79 \section{Introduction}
80
81 Nitrile groups can serve as very precise electric field reporters via
82 their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
83 triple bond between the nitrogen and the carbon atom is very sensitive
84 to local field changes and has been observed to have a direct impact
85 on the peak position within the spectrum. The Stark shift in the
86 spectrum can be quantified and mapped onto a field that is impinging
87 upon the nitrile bond. The response of nitrile groups to electric
88 fields has now been investigated for a number of small
89 molecules,\cite{Andrews:2000qv} as well as in biochemical settings,
90 where nitrile groups can act as minimally invasive probes of structure
91 and
92 dynamics.\cite{Tucker:2004qq,Webb:2008kn,Lindquist:2009fk,Fafarman:2010dq}
93 The vibrational Stark effect has also been used to study the effects
94 of electric fields on nitrile-containing self-assembled monolayers at
95 metallic interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
96
97
98 Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
99 molecule with a terminal nitrile group, has seen renewed interest as
100 one way to impart order on the surfactant interfaces of
101 nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
102 that can be used to promote particular kinds of
103 self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
104 is a particularly interesting case for studying electric field
105 effects, as 5CB exhibits an isotropic to nematic phase transition that
106 can be triggered by the application of an external field near room
107 temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
108 possiblity that the field-induced changes in the local environment
109 could have dramatic effects on the vibrations of this particular CN
110 bond. Although the infrared spectroscopy of 5CB has been
111 well-investigated, particularly as a measure of the kinetics of the
112 phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
113 seen the detailed theoretical treatment that biologically-relevant
114 small molecules have
115 received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Morales:2009fp,Waegele:2010ve}
116
117 The fundamental characteristic of liquid crystal mesophases is that
118 they maintain some degree of orientational order while translational
119 order is limited or absent. This orientational order produces a
120 complex direction-dependent response to external perturbations like
121 electric fields and mechanical distortions. The anisotropy of the
122 macroscopic phases originates in the anisotropy of the constituent
123 molecules, which typically have highly non-spherical structures with a
124 significant degree of internal rigidity. In nematic phases, rod-like
125 molecules are orientationally ordered with isotropic distributions of
126 molecular centers of mass. For example, 5CB has a solid to nematic
127 phase transition at 18C and a nematic to isotropic transition at
128 35C.\cite{Gray:1973ca}
129
130 In smectic phases, the molecules arrange themselves into layers with
131 their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
132 respect to the layer planes. The behavior of the $S_{A}$ phase can be
133 partially explained with models mainly based on geometric factors and
134 van der Waals interactions. The Gay-Berne potential, in particular,
135 has been widely used in the liquid crystal community to describe this
136 anisotropic phase
137 behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
138 However, these simple models are insufficient to describe liquid
139 crystal phases which exhibit more complex polymorphic nature.
140 Molecules which form $S_{A}$ phases can exhibit a wide variety of
141 subphases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
142 partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
143 ($S_{A_{d}}$), and often have a terminal cyano or nitro group. In
144 particular, lyotropic liquid crystals (those exhibiting liquid crystal
145 phase transitions as a function of water concentration), often have
146 polar head groups or zwitterionic charge separated groups that result
147 in strong dipolar interactions,\cite{Collings:1997rz} and terminal
148 cyano groups (like the one in 5CB) can induce permanent longitudinal
149 dipoles.\cite{Levelut:1981eu} Modeling of the phase behavior of these
150 molecules either requires additional dipolar
151 interactions,\cite{Bose:2012eu} or a unified-atom approach utilizing
152 point charges on the sites that contribute to the dipole
153 moment.\cite{Zhang:2011hh}
154
155 Macroscopic electric fields applied using electrodes on opposing sides
156 of a sample of 5CB have demonstrated the phase change of the molecule
157 as a function of electric field.\cite{Lim:2006xq} These previous
158 studies have shown the nitrile group serves as an excellent indicator
159 of the molecular orientation within the applied field. Lee {\it et
160 al.}~showed a 180 degree change in field direction could be probed
161 with the nitrile peak intensity as it changed along with molecular
162 alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
163
164 While these macroscopic fields work well at indicating the bulk
165 response, the response at a molecular scale has not been studied. With
166 the advent of nano-electrodes and the ability to couple these
167 electrodes to atomic force microscopy, control of electric fields
168 applied across nanometer distances is now possible.\cite{C3AN01651J} In
169 special cases where the macroscopic fields are insufficient to cause
170 an observable Stark effect without dielectric breakdown of the
171 material, small potentials across nanometer-sized gaps may be of
172 sufficient strength. For a gap of 5 nm between a lower electrode
173 having a nanoelectrode placed near it via an atomic force microscope,
174 a potential of 1 V applied across the electrodes is equivalent to a
175 field of 2x10\textsuperscript{8} $\frac{V}{M}$. This field is
176 certainly strong enough to cause the isotropic-nematic phase change
177 and as well as a visible Stark tuning of the nitrile bond. We expect
178 that this would be readily visible experimentally through Raman or IR
179 spectroscopy.
180
181 In the sections that follow, we outline a series of coarse-grained
182 classical molecular dynamics simulations of 5CB that were done in the
183 presence of static electric fields. These simulations were then
184 coupled with both {\it ab intio} calculations of CN-deformations and
185 classical bond-length correlation functions to predict spectral
186 shifts. These predictions made should be easily varifiable with
187 scanning electrochemical microscopy experiments.
188
189 \section{Computational Details}
190 The force field used for 5CB was a united-atom model that was
191 parameterized by Guo {\it et al.}\cite{Zhang:2011hh} However, for most
192 of the simulations, each of the phenyl rings was treated as a rigid
193 body to allow for larger time steps and very long simulation times.
194 The geometries of the rigid bodies were taken from equilibrium bond
195 distances and angles. Although the individual phenyl rings were held
196 rigid, bonds, bends, torsions and inversion centers that involved
197 atoms in these substructures (but with connectivity to the rest of the
198 molecule) were still included in the potential and force calculations.
199
200 Periodic simulations cells containing 270 molecules in random
201 orientations were constructed and were locked at experimental
202 densities. Electrostatic interactions were computed using damped
203 shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
204 were equilibrated for 1~ns at a temperature of 300K. Simulations with
205 applied fields were carried out in the microcanonical (NVE) ensemble
206 with an energy corresponding to the average energy from the canonical
207 (NVT) equilibration runs. Typical applied-field equilibration runs
208 were more than 60ns in length.
209
210 Static electric fields with magnitudes similar to what would be
211 available in an experimental setup were applied to the different
212 simulations. With an assumed electrode seperation of 5 nm and an
213 electrostatic potential that is limited by the voltage required to
214 split water (1.23V), the maximum realistic field that could be applied
215 is $\sim 0.024$ V/\AA. Three field environments were investigated:
216 (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
217 field = 0.024 V/\AA\ .
218
219 After the systems had come to equilibrium under the applied fields,
220 additional simulations were carried out with a flexible (Morse)
221 nitrile bond,
222 \begin{displaymath}
223 V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
224 \label{eq:morse}
225 \end{displaymath}
226 where $r_e= 1.157437$ \AA , $D_e = 212.95 \mathrm{~kcal~} /
227 \mathrm{mol}^{-1}$ and $\beta = 2.67566 $\AA~$^{-1}$. These
228 parameters correspond to a vibrational frequency of $2358
229 \mathrm{~cm}^{-1}$, somewhat higher than the experimental
230 frequency. The flexible nitrile moiety required simulation time steps
231 of 1~fs, so the additional flexibility was introducuced only after the
232 rigid systems had come to equilibrium under the applied fields.
233 Whenever time correlation functions were computed from the flexible
234 simulations, statistically-independent configurations were sampled
235 from the last ns of the induced-field runs. These configurations were
236 then equilibrated with the flexible nitrile moiety for 100 ps, and
237 time correlation functions were computed using data sampled from an
238 additional 200 ps of run time carried out in the microcanonical
239 ensemble.
240
241 \section{Field-induced Nematic Ordering}
242
243 In order to characterize the orientational ordering of the system, the
244 primary quantity of interest is the nematic (orientational) order
245 parameter. This was determined using the tensor
246 \begin{equation}
247 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{u}_{i
248 \alpha} \hat{u}_{i \beta} - \delta_{\alpha \beta} \right)
249 \end{equation}
250 where $\alpha, \beta = x, y, z$, and $\hat{u}_i$ is the molecular
251 end-to-end unit vector for molecule $i$. The nematic order parameter
252 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
253 corresponding eigenvector defines the director axis for the phase.
254 $S$ takes on values close to 1 in highly ordered (smectic A) phases,
255 but falls to much smaller values ($0 \rightarrow 0.3$) for isotropic
256 fluids. Note that the nitrogen and the terminal chain atom were used
257 to define the vectors for each molecule, so the typical order
258 parameters are lower than if one defined a vector using only the rigid
259 core of the molecule. In nematic phases, typical values for $S$ are
260 close to 0.5.
261
262 The field-induced phase transition can be clearly seen over the course
263 of a 60 ns equilibration runs in figure \ref{fig:orderParameter}. All
264 three of the systems started in a random (isotropic) packing, with
265 order parameters near 0.2. Over the course 10 ns, the full field
266 causes an alignment of the molecules (due primarily to the interaction
267 of the nitrile group dipole with the electric field). Once this
268 system began exhibiting nematic ordering, the orientational order
269 parameter became stable for the remaining 150 ns of simulation time.
270 It is possible that the partial-field simulation is meta-stable and
271 given enough time, it would eventually find a nematic-ordered phase,
272 but the partial-field simulation was stable as an isotropic phase for
273 the full duration of a 60 ns simulation. Ellipsoidal renderings of the
274 final configurations of the runs shows that the full-field (0.024
275 V/\AA\ ) experienced a isotropic-nematic phase transition and has
276 ordered with a director axis that is parallel to the direction of the
277 applied field.
278
279 \begin{figure}[H]
280 \includegraphics[width=\linewidth]{Figure1}
281 \caption{Evolution of the orientational order parameters for the
282 no-field, partial field, and full field simulations over the
283 course of 60 ns. Each simulation was started from a
284 statistically-independent isotropic configuration. On the right
285 are ellipsoids representing the final configurations at three
286 different field strengths: zero field (bottom), partial field
287 (middle), and full field (top)}
288 \label{fig:orderParameter}
289 \end{figure}
290
291
292 \section{Sampling the CN bond frequency}
293
294 The vibrational frequency of the nitrile bond in 5CB depends on
295 features of the local solvent environment of the individual molecules
296 as well as the bond's orientation relative to the applied field. The
297 primary quantity of interest for interpreting the condensed phase
298 spectrum of this vibration is the distribution of frequencies
299 exhibited by the 5CB nitrile bond under the different electric fields.
300 There have been a number of elegant techniques for obtaining
301 vibrational lineshapes from classical simulations, including a
302 perturbation theory approach,\cite{Morales:2009fp} the use of an
303 optimized QM/MM approach coupled with the fluctuating frequency
304 approximation,\cite{Lindquist:2008qf} and empirical frequency
305 correlation maps.\cite{Oh:2008fk} Three distinct (and comparatively
306 primitive) methods for mapping classical simulations onto vibrational
307 spectra were brought to bear on the simulations in this work:
308 \begin{enumerate}
309 \item Isolated 5CB molecules and their immediate surroundings were
310 extracted from the simulations. These nitrile bonds were stretched
311 and single-point {\em ab initio} calculations were used to obtain
312 Morse-oscillator fits for the local vibrational motion along that
313 bond.
314 \item A static-field extension of the empirical frequency correlation
315 maps developed by Cho {\it et al.}~\cite{Oh:2008fk} for nitrile
316 moieties in water was attempted.
317 \item Classical bond-length autocorrelation functions were Fourier
318 transformed to directly obtain the vibrational spectrum from
319 molecular dynamics simulations.
320 \end{enumerate}
321
322 \subsection{CN frequencies from isolated clusters}
323 The size of the periodic condensed phase system prevented direct
324 computation of the complete library of nitrile bond frequencies using
325 {\it ab initio} methods. In order to sample the nitrile frequencies
326 present in the condensed-phase, individual molecules were selected
327 randomly to serve as the center of a local (gas phase) cluster. To
328 include steric, electrostatic, and other effects from molecules
329 located near the targeted nitrile group, portions of other molecules
330 nearest to the nitrile group were included in the quantum mechanical
331 calculations. The surrounding solvent molecules were divided into
332 ``body'' (the two phenyl rings and the nitrile bond) and ``tail'' (the
333 alkyl chain). Any molecule which had a body atom within 6~\AA\ of the
334 midpoint of the target nitrile bond had its own molecular body (the
335 4-cyano-biphenyl moiety) included in the configuration. Likewise, the
336 entire alkyl tail was included if any tail atom was within 4~\AA\ of
337 the target nitrile bond. If tail atoms (but no body atoms) were
338 included within these distances, only the tail was included as a
339 capped propane molecule.
340
341 \begin{figure}[H]
342 \includegraphics[width=\linewidth]{Figure2}
343 \caption{Cluster calculations were performed on randomly sampled 5CB
344 molecules (shown in red) from each of the simulations. Surrounding
345 molecular bodies were included if any body atoms were within 6
346 \AA\ of the target nitrile bond, and tails were included if they
347 were within 4 \AA. Included portions of these molecules are shown
348 in green. The CN bond on the target molecule was stretched and
349 compressed, and the resulting single point energies were fit to
350 Morse oscillators to obtain a distribution of frequencies.}
351 \label{fig:cluster}
352 \end{figure}
353
354 Inferred hydrogen atom locations were added to the cluster geometries,
355 and the nitrile bond was stretched from 0.87 to 1.52~\AA\ at
356 increments of 0.05~\AA. This generated 13 configurations per gas phase
357 cluster. Single-point energies were computed using the B3LYP
358 functional~\cite{Becke:1993kq,Lee:1988qf} and the 6-311++G(d,p) basis
359 set. For the cluster configurations that had been generated from
360 molecular dynamics running under applied fields, the density
361 functional calculations had a field of $5 \times 10^{-4}$ atomic units
362 ($E_h / (e a_0)$) applied in the $+z$ direction in order to match the
363 molecular dynamics simulations.
364
365 The energies for the stretched / compressed nitrile bond in each of
366 the clusters were used to fit Morse potentials, and the frequencies
367 were obtained from the $0 \rightarrow 1$ transition for the energy
368 levels for this potential.\cite{Morse:1929xy} To obtain a spectrum,
369 each of the frequencies was convoluted with a Lorentzian lineshape
370 with a width of 1.5 $\mathrm{cm}^{-1}$. Available computing resources
371 limited the sampling to 67 clusters for the zero-field spectrum, and
372 59 for the full field. Comparisons of the quantum mechanical spectrum
373 to the classical are shown in figure \ref{fig:spectrum}.
374
375 \subsection{CN frequencies from potential-frequency maps}
376
377 One approach which has been used to successfully analyze the spectrum
378 of nitrile and thiocyanate probes in aqueous environments was
379 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} This
380 method involves finding a multi-parameter fit that maps between the
381 local electrostatic potential at selected sites surrounding the
382 nitrile bond and the vibrational frequency of that bond obtained from
383 more expensive {\it ab initio} methods. This approach is similar in
384 character to the field-frequency maps developed by the Skinner group
385 for OH stretches in liquid water.\cite{Corcelli:2004ai,Auer:2007dp}
386
387 To use the potential-frequency maps, the local electrostatic
388 potential, $\phi_a$, is computed at 20 sites ($a = 1 \rightarrow 20$)
389 that surround the nitrile bond,
390 \begin{equation}
391 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{j}
392 \frac{q_j}{\left|r_{aj}\right|}.
393 \end{equation}
394 Here $q_j$ is the partial site on atom $j$ (residing on a different
395 molecule) and $r_{aj}$ is the distance between site $a$ and atom $j$.
396 The original map was parameterized in liquid water and comprises a set
397 of parameters, $l_a$, that predict the shift in nitrile peak
398 frequency,
399 \begin{equation}
400 \delta\tilde{\nu} =\sum^{20}_{a=1} l_{a}\phi_{a}.
401 \end{equation}
402
403 The simulations of 5CB were carried out in the presence of
404 externally-applied uniform electric fields. Although uniform fields
405 exert forces on charge sites, they only contribute to the potential if
406 one defines a reference point that can serve as an origin. One simple
407 modification to the potential at each of the probe sites is to use the
408 centroid of the \ce{CN} bond as the origin for that site,
409 \begin{equation}
410 \phi_a^\prime = \phi_a + \frac{1}{4\pi\epsilon_{0}} \vec{E} \cdot
411 \left(\vec{r}_a - \vec{r}_\ce{CN} \right)
412 \end{equation}
413 where $\vec{E}$ is the uniform electric field, $\left( \vec{r}_{a} -
414 \vec{r}_\ce{CN} \right)$ is the displacement between the
415 cooridinates described by Choi {\it et
416 al.}~\cite{Choi:2008cr,Oh:2008fk} and the \ce{CN} bond centroid.
417 $\phi_a^\prime$ then contains an effective potential contributed by
418 the uniform field in addition to the local potential contributions
419 from other molecules.
420
421 The sites $\{\vec{r}_a\}$ and weights $\left\{l_a \right\}$
422 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} are quite
423 symmetric around the \ce{CN} centroid, and even at large uniform field
424 values we observed nearly-complete cancellation of the potenial
425 contributions from the uniform field. In order to utilize the
426 potential-frequency maps for this problem, one would therefore need
427 extensive reparameterization of the maps to include explicit
428 contributions from the external field. This reparameterization is
429 outside the scope of the current work, but would make a useful
430 addition to the potential-frequency map approach.
431
432 \subsection{CN frequencies from bond length autocorrelation functions}
433
434 The distribution of nitrile vibrational frequencies can also be found
435 using classical time correlation functions. This was done by
436 replacing the rigid \ce{CN} bond with a flexible Morse oscillator
437 described in Eq. \ref{eq:morse}. Since the systems were perturbed by
438 the addition of a flexible high-frequency bond, they were allowed to
439 re-equilibrate in the canonical (NVT) ensemble for 100 ps with 1 fs
440 timesteps. After equilibration, each configuration was run in the
441 microcanonical (NVE) ensemble for 20 ps. Configurations sampled every
442 fs were then used to compute bond-length autocorrelation functions,
443 \begin{equation}
444 C(t) = \langle \delta r(t) \cdot \delta r(0) ) \rangle
445 \end{equation}
446 %
447 where $\delta r(t) = r(t) - r_0$ is the deviation from the equilibrium
448 bond distance at time $t$. Because the other atomic sites have very
449 small partial charges, this correlation function is an approximation
450 to the dipole autocorrelation function for the molecule, which would
451 be particularly relevant to computing the IR spectrum. Ten
452 statistically-independent correlation functions were obtained by
453 allowing the systems to run 10 ns with rigid \ce{CN} bonds followed by
454 120 ps equilibration and data collection using the flexible \ce{CN}
455 bonds. This process was repeated 10 times, and the total sampling
456 time, from sample preparation to final configurations, exceeded 150 ns
457 for each of the field strengths investigated.
458
459 The correlation functions were filtered using exponential apodization
460 functions,\cite{FILLER:1964yg} $f(t) = e^{-|t|/c}$, with a time
461 constant, $c =$ 3.5 ps, and were Fourier transformed to yield a
462 spectrum,
463 \begin{equation}
464 I(\omega) = \int_{-\infty}^{\infty} C(t) f(t) e^{-i \omega t} dt.
465 \end{equation}
466 The sample-averaged classical nitrile spectrum can be seen in Figure
467 \ref{fig:spectra}. Note that the Morse oscillator parameters listed
468 above yield a natural frequency of 2358 $\mathrm{cm}^{-1}$, somewhat
469 higher than the experimental peak near 2226 $\mathrm{cm}^{-1}$. This
470 shift does not effect the ability to qualitatively compare peaks from
471 the classical and quantum mechanical approaches, so the classical
472 spectra are shown as a shift relative to the natural oscillation of
473 the Morse bond.
474
475 \begin{figure}
476 \includegraphics[width=3.25in]{Convolved}
477 \includegraphics[width=3.25in]{2Spectra}
478 \caption{Quantum mechanical nitrile spectrum for the no-field simulation
479 (black) and the full field simulation (red). The lower panel
480 shows the corresponding classical bond-length autocorrelation
481 spectrum for the flexible nitrile measured relative to the natural
482 frequency for the flexible bond.}
483 \label{fig:spectra}
484 \end{figure}
485
486 Note that due to electrostatic interactions, the classical approach
487 implicitly couples \ce{CN} vibrations to the same vibrational mode on
488 other nearby molecules. This coupling is not handled in the {\it ab
489 initio} cluster approach.
490
491 \section{Discussion}
492
493 Our simulations show that the united-atom model can reproduce the
494 field-induced nematic ordering of the 4-cyano-4'-pentylbiphenyl.
495 Because we are simulating a very small electrode separation (5~nm), a
496 voltage drop as low as 1.2~V was sufficient to induce the phase
497 change. This potential is significantly smaller than 100~V that has
498 used within a 5~um gap for electrochemiluminescence of rubrene,\cite{Kojima19881789} and suggests
499 that by using close electrode separation, it would be relatively
500 straightforward to observe the nitrile Stark shift in 5CB.
501
502 Both the classical correlation function and the isolated cluster
503 approaches to estimating the IR spectrum show that a small population
504 of nitrile stretches shift by $\sim 40 \mathrm{cm}^{-1}$ to the red of
505 the unperturbed vibrational line. To understand the origin of this
506 shift, a more complete picture of the spatial ordering around the
507 nitrile bonds is required. We have computed the angle-dependent pair
508 distribution functions,
509 \begin{align}
510 g(r, \cos \omega) = & \frac{1}{\rho N} \left< \sum_{i}
511 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \omega_{ij} -
512 \cos \omega\right) \right> \\ \nonumber \\
513 g(r, \cos \theta) = & \frac{1}{\rho N} \left< \sum_{i}
514 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \theta_{i} -
515 \cos \theta \right) \right>
516 \end{align}
517 which provide information about the joint spatial and angular
518 correlations present in the system. The angles $\omega$ and $\theta$
519 are defined by vectors along the CN axis of each nitrile bond (see
520 figure \ref{fig:definition}).
521 \begin{figure}
522 \includegraphics[width=4in]{definition}
523 \caption{Definitions of the angles between two nitrile bonds.}
524 \label{fig:definition}
525 \end{figure}
526
527 The primary structural effect of the field-induced phase transition is
528 apparent in figure \ref{fig:gofromega}. The nematic ordering transfers
529 population from the perpendicular ($\cos\omega\approx 0$) and
530 anti-aligned ($\cos\omega\approx -1$) to the nitrile-alinged peak
531 near $\cos\omega\approx 1$, leaving most other features undisturbed. This
532 change is visible in the simulations as an increased population of
533 aligned nitrile bonds in the first solvation shell.
534 \begin{figure}
535 \includegraphics[width=\linewidth]{Figure4}
536 \caption{Contours of the angle-dependent pair distribution functions
537 for nitrile bonds on 5CB in the no field (upper panel) and full
538 field (lower panel) simulations. Dark areas signify regions of
539 enhanced density, while light areas signify depletion relative to
540 the bulk density.}
541 \label{fig:gofromega}
542 \end{figure}
543 Although it is certainly possible that the coupling between
544 closely-spaced nitrile pairs is responsible for some of the red-shift,
545 that is not the only structural change that is taking place. The
546 second two-dimensional pair distribution function, $g(r,\cos\theta)$,
547 shows that nematic ordering also transfers population that is directly
548 in line with the nitrile bond (see figure \ref{fig:gofrtheta}) to the
549 sides of the molecule, thereby freeing steric blockage can directly
550 influence the nitrile vibration. We are suggesting here that the
551 nematic ordering provides an anti-caging of the nitrile vibration, and
552 given that the oscillator is fairly anharmonic, this provides a
553 fraction of the nitrile bonds with a significant red-shift.
554 \begin{figure}
555 \includegraphics[width=\linewidth]{Figure6}
556 \caption{Contours of the angle-dependent pair distribution function,
557 $g(r,\cos \theta)$, for finding any other atom at a distance and
558 angular deviation from the center of a nitrile bond. The top edge
559 of each contour plot corresponds to local density along the
560 direction of the nitrogen in the CN bond, while the bottom is in
561 the direction of the carbon atom. Bottom panel: $g(z)$ data taken
562 by following the \ce{C -> N} vector for each nitrile bond shows
563 that the field-induced phase transition reduces the population of
564 atoms that are directly in line with the nitrogen motion.}
565 \label{fig:gofrtheta}
566 \end{figure}
567
568 The cause of this shift does not appear to be related to the alignment
569 of those nitrile bonds with the field, but rather to the change in
570 local environment that is brought about by the isotropic-nematic
571 transition. We have compared configurations for many of the cluster
572 calculations that exhibited the frequencies between (2190 and 2215
573 $\mathrm{cm}^{-1}$) , and have observed some similar features. The
574 lowest frequencies appear to come from configurations which have
575 nearly-empty pockets directly opposite the nitrogen atom from the
576 nitrile carbon. Because we have so few clusters, this is certainly not
577 quantitative confirmation of this effect.
578
579
580 While this makes the application of nitrile Stark effects in
581 simulations without water harder, these data show
582 that it is not a deal breaker. The classically calculated nitrile
583 spectrum shows changes in the spectra that will be easily seen through
584 experimental routes. It indicates a shifted peak lower in energy
585 should arise. This peak is a few wavenumbers from the leading edge of
586 the larger peak and almost 75 wavenumbers from the center. This
587 seperation between the two peaks means experimental results will show
588 an easily resolved peak.
589
590 The Gaussian derived spectra do indicate an applied field
591 and subsiquent phase change does cause a narrowing of freuency
592 distrobution. With narrowing, it would indicate an increased
593 homogeneous distrobution of the local field near the nitrile.
594
595
596
597 \section{Conclusions}
598 Field dependent changes
599
600 \section{Acknowledgements}
601 The authors thank Steven Corcelli for helpful comments and
602 suggestions. Support for this project was provided by the National
603 Science Foundation under grant CHE-0848243. Computational time was
604 provided by the Center for Research Computing (CRC) at the University
605 of Notre Dame.
606
607 \newpage
608
609 \bibliography{5CB}
610
611 \end{doublespace}
612 \end{document}