ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4096
Committed: Tue Apr 1 16:05:03 2014 UTC (10 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 32501 byte(s)
Log Message:
Spelling and sentence structure

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]{cluster/cluster.pdf}
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 portion
56 of the main nitrile peak, and this shift was observed only when the
57 fields were large enough to induce orientational ordering of the
58 bulk phase. Distributions of frequencies obtained via cluster-based
59 fits to quantum mechanical energies of nitrile bond deformations
60 exhibit a similar $\sim~2.7~\mathrm{cm}^{-1}$ red shift. Joint
61 spatial-angular distribution functions indicate that phase-induced
62 anti-caging of the nitrile bond is contributing to the change in the
63 nitrile spectrum.
64 \end{abstract}
65
66 \newpage
67
68 \section{Introduction}
69
70 Because the triple bond between nitrogen and carbon is sensitive to
71 local fields, nitrile groups can report on field strengths via their
72 distinctive Raman and IR signatures.\cite{Boxer:2009xw} The response
73 of nitrile groups to electric fields has now been investigated for a
74 number of small molecules,\cite{Andrews:2000qv} as well as in
75 biochemical settings, where nitrile groups can act as minimally
76 invasive probes of structure and
77 dynamics.\cite{Tucker:2004qq,Webb:2008kn,Lindquist:2009fk,Fafarman:2010dq}
78 The vibrational Stark effect has also been used to study the effects
79 of electric fields on nitrile-containing self-assembled monolayers at
80 metallic interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
81
82 Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
83 molecule with a terminal nitrile group, has seen renewed interest as
84 one way to impart order on the surfactant interfaces of
85 nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
86 that can be used to promote particular kinds of
87 self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
88 is a particularly interesting case for studying electric field
89 effects, as 5CB exhibits an isotropic to nematic phase transition that
90 can be triggered by the application of an external field near room
91 temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
92 possibility that the field-induced changes in the local environment
93 could have dramatic effects on the vibrations of this particular CN
94 bond. Although the infrared spectroscopy of 5CB has been
95 well-investigated, particularly as a measure of the kinetics of the
96 phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
97 seen the detailed theoretical treatment that biologically-relevant
98 small molecules have
99 received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Morales:2009fp,Waegele:2010ve}
100
101 The fundamental characteristic of liquid crystal mesophases is that
102 they maintain some degree of orientational order while translational
103 order is limited or absent. This orientational order produces a
104 complex direction-dependent response to external perturbations like
105 electric fields and mechanical distortions. The anisotropy of the
106 macroscopic phases originates in the anisotropy of the constituent
107 molecules, which typically have highly non-spherical structures with a
108 significant degree of internal rigidity. In nematic phases, rod-like
109 molecules are orientationally ordered with isotropic distributions of
110 molecular centers of mass. For example, 5CB has a solid to nematic
111 phase transition at 18C and a nematic to isotropic transition at
112 35C.\cite{Gray:1973ca}
113
114 In smectic phases, the molecules arrange themselves into layers with
115 their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
116 respect to the layer planes. The behavior of the $S_{A}$ phase can be
117 explained with models based solely on geometric factors and van der
118 Waals interactions. The Gay-Berne potential, in particular, has been
119 widely used in the liquid crystal community to describe this
120 anisotropic phase
121 behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
122 However, these simple models are insufficient to describe liquid
123 crystal phases which exhibit more complex polymorphic nature.
124 Molecules which form $S_{A}$ phases can exhibit a wide variety of
125 sub-phases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
126 partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
127 ($S_{A_{d}}$), and often have a terminal cyano or nitro group. In
128 particular, lyotropic liquid crystals (those exhibiting liquid crystal
129 phase transitions as a function of water concentration), often have
130 polar head groups or zwitterionic charge separated groups that result
131 in strong dipolar interactions,\cite{Collings:1997rz} and terminal
132 cyano groups (like the one in 5CB) can induce permanent longitudinal
133 dipoles.\cite{Levelut:1981eu} Modeling of the phase behavior of these
134 molecules either requires additional dipolar
135 interactions,\cite{Bose:2012eu} or a unified-atom approach utilizing
136 point charges on the sites that contribute to the dipole
137 moment.\cite{Zhang:2011hh}
138
139 Macroscopic electric fields applied using electrodes on opposing sides
140 of a sample of 5CB have demonstrated the phase change of the molecule
141 as a function of electric field.\cite{Lim:2006xq} These previous
142 studies have shown the nitrile group serves as an excellent indicator
143 of the molecular orientation within the applied field. Lee {\it et
144 al.}~showed a 180 degree change in field direction could be probed
145 with the nitrile peak intensity as it changed along with molecular
146 alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
147
148 While these macroscopic fields work well at indicating the bulk
149 response, the response at a molecular scale has not been studied. With
150 the advent of nano-electrodes and the ability to couple these
151 electrodes to atomic force microscopy, control of electric fields
152 applied across nanometer distances is now possible.\cite{C3AN01651J}
153 In special cases where the macroscopic fields are insufficient to
154 cause an observable Stark effect without dielectric breakdown of the
155 material, small potentials across nanometer-sized gaps may have
156 sufficient strength. For a gap of 5 nm between a lower electrode
157 having a nanoelectrode placed near it via an atomic force microscope,
158 a potential of 1 V applied across the electrodes is equivalent to a
159 field of $2 \times 10^8~\mathrm{V/m}$. This field is certainly strong
160 enough to cause the isotropic-nematic phase change and as well as a
161 visible Stark tuning of the nitrile bond. We expect that this would be
162 readily visible experimentally through Raman or IR spectroscopy.
163
164 In the sections that follow, we outline a series of coarse-grained
165 (united atom) classical molecular dynamics simulations of 5CB that
166 were done in the presence of static electric fields. These simulations
167 were then coupled with both {\it ab intio} calculations of
168 CN-deformations and classical bond-length correlation functions to
169 predict spectral shifts. These predictions should be verifiable via
170 scanning electrochemical microscopy.
171
172 \section{Computational Details}
173 The force-field used to model 5CB was a united-atom model that was
174 parameterized by Guo {\it et al.}\cite{Zhang:2011hh} However, for most
175 of the simulations, each of the phenyl rings was treated as a rigid
176 body to allow for larger time steps and longer simulation times. The
177 geometries of the rigid bodies were taken from equilibrium bond
178 distances and angles. Although the individual phenyl rings were held
179 rigid, bonds, bends, torsions and inversion centers that involved
180 atoms in these substructures (but with connectivity to the rest of the
181 molecule) were still included in the potential and force calculations.
182
183 Periodic simulations cells containing 270 molecules in random
184 orientations were constructed and were locked at experimental
185 densities. Electrostatic interactions were computed using damped
186 shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
187 were equilibrated for 1~ns at a temperature of 300K. Simulations with
188 applied fields were carried out in the microcanonical (NVE) ensemble
189 with an energy corresponding to the average energy from the canonical
190 (NVT) equilibration runs. Typical applied-field equilibration runs
191 were more than 60~ns in length.
192
193 Static electric fields with magnitudes similar to what would be
194 available in an experimental setup were applied to the different
195 simulations. With an assumed electrode separation of 5 nm and an
196 electrostatic potential that is limited by the voltage required to
197 split water (1.23V), the maximum realistic field that could be applied
198 is $\sim 0.024$ V/\AA. Three field environments were investigated:
199 (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
200 field = 0.024 V/\AA\ .
201
202 After the systems had come to equilibrium under the applied fields,
203 additional simulations were carried out with a flexible (Morse)
204 nitrile bond,
205 \begin{displaymath}
206 V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
207 \label{eq:morse}
208 \end{displaymath}
209 where $r_e= 1.157437$ \AA , $D_e = 212.95 \mathrm{~kcal~} /
210 \mathrm{mol}^{-1}$ and $\beta = 2.67566 $\AA~$^{-1}$. These
211 parameters correspond to a vibrational frequency of $2358
212 \mathrm{~cm}^{-1}$, somewhat higher than the experimental
213 frequency. The flexible nitrile moiety required simulation time steps
214 of 1~fs, so the additional flexibility was introduced only after the
215 rigid systems had come to equilibrium under the applied fields.
216 Whenever time correlation functions were computed from the flexible
217 simulations, statistically-independent configurations were sampled
218 from the last ns of the induced-field runs. These configurations were
219 then equilibrated with the flexible nitrile moiety for 100 ps, and
220 time correlation functions were computed using data sampled from an
221 additional 200 ps of run time carried out in the microcanonical
222 ensemble.
223
224 \section{Field-induced Nematic Ordering}
225
226 In order to characterize the orientational ordering of the system, the
227 primary quantity of interest is the nematic (orientational) order
228 parameter. This was determined using the tensor
229 \begin{equation}
230 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{u}_{i
231 \alpha} \hat{u}_{i \beta} - \delta_{\alpha \beta} \right)
232 \end{equation}
233 where $\alpha, \beta = x, y, z$, and $\hat{u}_i$ is the molecular
234 end-to-end unit vector for molecule $i$. The nematic order parameter
235 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
236 corresponding eigenvector defines the director axis for the phase.
237 $S$ takes on values close to 1 in highly ordered (smectic A) phases,
238 but falls to much smaller values ($0 \rightarrow 0.3$) for isotropic
239 fluids. Note that the nitrogen and the terminal chain atom were used
240 to define the vectors for each molecule, so the typical order
241 parameters are lower than if one defined a vector using only the rigid
242 core of the molecule. In nematic phases, typical values for $S$ are
243 close to 0.5.
244
245 The field-induced phase transition can be clearly seen over the course
246 of a 60 ns equilibration runs in figure \ref{fig:orderParameter}. All
247 three of the systems started in a random (isotropic) packing, with
248 order parameters near 0.2. Over the course 10 ns, the full field
249 causes an alignment of the molecules (due primarily to the interaction
250 of the nitrile group dipole with the electric field). Once this
251 system began exhibiting nematic ordering, the orientational order
252 parameter became stable for the remaining 150 ns of simulation time.
253 It is possible that the partial-field simulation is meta-stable and
254 given enough time, it would eventually find a nematic-ordered phase,
255 but the partial-field simulation was stable as an isotropic phase for
256 the full duration of a 60 ns simulation. Ellipsoidal renderings of the
257 final configurations of the runs shows that the full-field (0.024
258 V/\AA\ ) experienced a isotropic-nematic phase transition and has
259 ordered with a director axis that is parallel to the direction of the
260 applied field.
261
262 \begin{figure}[H]
263 \includegraphics[width=\linewidth]{orderParameter/orderParameter.pdf}
264 \caption{Evolution of the orientational order parameters for the
265 no-field, partial field, and full field simulations over the
266 course of 60 ns. Each simulation was started from a
267 statistically-independent isotropic configuration. On the right
268 are ellipsoids representing the final configurations at three
269 different field strengths: zero field (bottom), partial field
270 (middle), and full field (top)}
271 \label{fig:orderParameter}
272 \end{figure}
273
274
275 \section{Sampling the CN bond frequency}
276
277 The vibrational frequency of the nitrile bond in 5CB depends on
278 features of the local solvent environment of the individual molecules
279 as well as the bond's orientation relative to the applied field. The
280 primary quantity of interest for interpreting the condensed phase
281 spectrum of this vibration is the distribution of frequencies
282 exhibited by the 5CB nitrile bond under the different electric fields.
283 There have been a number of elegant techniques for obtaining
284 vibrational line shapes from classical simulations, including a
285 perturbation theory approach,\cite{Morales:2009fp} the use of an
286 optimized QM/MM approach coupled with the fluctuating frequency
287 approximation,\cite{Lindquist:2008qf} and empirical frequency
288 correlation maps.\cite{Oh:2008fk} Three distinct (and comparatively
289 primitive) methods for mapping classical simulations onto vibrational
290 spectra were brought to bear on the simulations in this work:
291 \begin{enumerate}
292 \item Isolated 5CB molecules and their immediate surroundings were
293 extracted from the simulations. These nitrile bonds were stretched
294 and single-point {\em ab initio} calculations were used to obtain
295 Morse-oscillator fits for the local vibrational motion along that
296 bond.
297 \item A static-field extension of the empirical frequency correlation
298 maps developed by Cho {\it et al.}~\cite{Oh:2008fk} for nitrile
299 moieties in water was attempted.
300 \item Classical bond-length autocorrelation functions were Fourier
301 transformed to directly obtain the vibrational spectrum from
302 molecular dynamics simulations.
303 \end{enumerate}
304
305 \subsection{CN frequencies from isolated clusters}
306 The size of the condensed phase liquid crystal system prevented direct
307 computation of the complete library of nitrile bond frequencies using
308 {\it ab initio} methods. In order to sample the nitrile frequencies
309 present in the condensed-phase, individual molecules were selected
310 randomly to serve as the center of a local (gas phase) cluster. To
311 include steric, electrostatic, and other effects from molecules
312 located near the targeted nitrile group, portions of other molecules
313 nearest to the nitrile group were included in the quantum mechanical
314 calculations. The surrounding solvent molecules were divided into
315 ``body'' (the two phenyl rings and the nitrile bond) and ``tail'' (the
316 alkyl chain). Any molecule which had a body atom within 6~\AA\ of the
317 midpoint of the target nitrile bond had its own molecular body (the
318 4-cyano-biphenyl moiety) included in the configuration. Likewise, the
319 entire alkyl tail was included if any tail atom was within 4~\AA\ of
320 the target nitrile bond. If tail atoms (but no body atoms) were
321 included within these distances, only the tail was included as a
322 capped propane molecule.
323
324 \begin{figure}[H]
325 \includegraphics[width=\linewidth]{cluster/cluster.pdf}
326 \caption{Cluster calculations were performed on randomly sampled 5CB
327 molecules (shown in red) from the full-field and no-field
328 simulations. Surrounding molecular bodies were included if any
329 body atoms were within 6 \AA\ of the target nitrile bond, and
330 tails were included if they were within 4 \AA. Included portions
331 of these molecules are shown in green. The CN bond on the target
332 molecule was stretched and compressed, and the resulting single
333 point energies were fit to Morse oscillators to obtain a
334 distribution of frequencies.}
335 \label{fig:cluster}
336 \end{figure}
337
338 Inferred hydrogen atom locations were added to the cluster geometries,
339 and the nitrile bond was stretched from 0.87 to 1.52~\AA\ at
340 increments of 0.05~\AA. This generated 13 configurations per gas phase
341 cluster. Single-point energies were computed using the B3LYP
342 functional~\cite{Becke:1993kq,Lee:1988qf} and the 6-311++G(d,p) basis
343 set. For the cluster configurations that had been generated from
344 molecular dynamics running under applied fields, the density
345 functional calculations had a field of $5 \times 10^{-4}$ atomic units
346 ($E_h / (e a_0)$) applied in the $+z$ direction in order to match the
347 molecular dynamics simulations.
348
349 The energies for the stretched / compressed nitrile bond in each of
350 the clusters were used to fit Morse potentials, and the frequencies
351 were obtained from the $0 \rightarrow 1$ transition for the energy
352 levels for this potential.\cite{Morse:1929xy} To obtain a spectrum,
353 each of the frequencies was convoluted with a Lorentzian line shape
354 with a width of 1.5 $\mathrm{cm}^{-1}$. Available computing resources
355 limited the sampling to 100 clusters for both the zero-field and full
356 field spectra. Comparisons of the quantum mechanical spectrum to the
357 classical are shown in figure \ref{fig:spectra}. The mean frequencies
358 obtained from the distributions give a field-induced red shift of
359 $2.68~\mathrm{cm}^{-1}$.
360
361 \subsection{CN frequencies from potential-frequency maps}
362
363 One approach which has been used to successfully analyze the spectrum
364 of nitrile and thiocyanate probes in aqueous environments was
365 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} This
366 method involves finding a multi-parameter fit that maps between the
367 local electrostatic potential at selected sites surrounding the
368 nitrile bond and the vibrational frequency of that bond obtained from
369 more expensive {\it ab initio} methods. This approach is similar in
370 character to the field-frequency maps developed by the Skinner group
371 for OH stretches in liquid water.\cite{Corcelli:2004ai,Auer:2007dp}
372
373 To use the potential-frequency maps, the local electrostatic
374 potential, $\phi_a$, is computed at 20 sites ($a = 1 \rightarrow 20$)
375 that surround the nitrile bond,
376 \begin{equation}
377 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{j}
378 \frac{q_j}{\left|r_{aj}\right|}.
379 \end{equation}
380 Here $q_j$ is the partial site on atom $j$ (residing on a different
381 molecule) and $r_{aj}$ is the distance between site $a$ and atom $j$.
382 The original map was parameterized in liquid water and comprises a set
383 of parameters, $l_a$, that predict the shift in nitrile peak
384 frequency,
385 \begin{equation}
386 \delta\tilde{\nu} =\sum^{20}_{a=1} l_{a}\phi_{a}.
387 \end{equation}
388
389 The simulations of 5CB were carried out in the presence of
390 externally-applied uniform electric fields. Although uniform fields
391 exert forces on charge sites, they only contribute to the potential if
392 one defines a reference point that can serve as an origin. One simple
393 modification to the potential at each of the probe sites is to use the
394 centroid of the \ce{CN} bond as the origin for that site,
395 \begin{equation}
396 \phi_a^\prime = \phi_a + \frac{1}{4\pi\epsilon_{0}} \vec{E} \cdot
397 \left(\vec{r}_a - \vec{r}_\ce{CN} \right)
398 \end{equation}
399 where $\vec{E}$ is the uniform electric field, $\left( \vec{r}_{a} -
400 \vec{r}_\ce{CN} \right)$ is the displacement between the
401 coordinates described by Choi {\it et
402 al.}~\cite{Choi:2008cr,Oh:2008fk} and the \ce{CN} bond centroid.
403 $\phi_a^\prime$ then contains an effective potential contributed by
404 the uniform field in addition to the local potential contributions
405 from other molecules.
406
407 The sites $\{\vec{r}_a\}$ and weights $\left\{l_a \right\}$
408 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} are quite
409 symmetric around the \ce{CN} centroid, and even at large uniform field
410 values we observed nearly-complete cancellation of the potential
411 contributions from the uniform field. In order to utilize the
412 potential-frequency maps for this problem, one would therefore need
413 extensive reparameterization of the maps to include explicit
414 contributions from the external field. This reparameterization is
415 outside the scope of the current work, but would make a useful
416 addition to the potential-frequency map approach.
417
418 We note that in 5CB there does not appear to be a particularly strong
419 correlation between the electric field observed at the nitrile
420 centroid and the calculated vibrational frequency. In
421 Fig. \ref{fig:fieldMap} we show the calculated frequencies plotted
422 against the field magnitude and the parallel and perpendicular
423 components of the field.
424
425 \begin{figure}
426 \includegraphics[width=\linewidth]{fieldMap/fieldMap.pdf}
427 \caption{The observed cluster frequencies have no apparent
428 correlation with the electric field felt at the centroid of the
429 nitrile bond. Upper panel: vibrational frequencies plotted
430 against the component of the field parallel to the CN bond.
431 Middle panel: mapped to the magnitude of the field components
432 perpendicular to the CN bond. Lower panel: mapped to the total
433 field magnitude.}
434 \label{fig:fieldMap}
435 \end{figure}
436
437
438 \subsection{CN frequencies from bond length autocorrelation functions}
439
440 The distribution of nitrile vibrational frequencies can also be found
441 using classical time correlation functions. This was done by
442 replacing the rigid \ce{CN} bond with a flexible Morse oscillator
443 described in Eq. \ref{eq:morse}. Since the systems were perturbed by
444 the addition of a flexible high-frequency bond, they were allowed to
445 re-equilibrate in the canonical (NVT) ensemble for 100 ps with 1 fs
446 time steps. After equilibration, each configuration was run in the
447 microcanonical (NVE) ensemble for 20 ps. Configurations sampled every
448 fs were then used to compute bond-length autocorrelation functions,
449 \begin{equation}
450 C(t) = \langle \delta r(t) \cdot \delta r(0) ) \rangle
451 \end{equation}
452 %
453 where $\delta r(t) = r(t) - r_0$ is the deviation from the equilibrium
454 bond distance at time $t$. Because the other atomic sites have very
455 small partial charges, this correlation function is an approximation
456 to the dipole autocorrelation function for the molecule, which would
457 be particularly relevant to computing the IR spectrum. Ten
458 statistically-independent correlation functions were obtained by
459 allowing the systems to run 10 ns with rigid \ce{CN} bonds followed by
460 120 ps equilibration and data collection using the flexible \ce{CN}
461 bonds. This process was repeated 10 times, and the total sampling
462 time, from sample preparation to final configurations, exceeded 150 ns
463 for each of the field strengths investigated.
464
465 The correlation functions were filtered using exponential apodization
466 functions,\cite{FILLER:1964yg} $f(t) = e^{-|t|/c}$, with a time
467 constant, $c =$ 3.5 ps, and were Fourier transformed to yield a
468 spectrum,
469 \begin{equation}
470 I(\omega) = \int_{-\infty}^{\infty} C(t) f(t) e^{-i \omega t} dt.
471 \end{equation}
472 The sample-averaged classical nitrile spectrum can be seen in Figure
473 \ref{fig:spectra}. Note that the Morse oscillator parameters listed
474 above yield a natural frequency of 2358 $\mathrm{cm}^{-1}$, somewhat
475 higher than the experimental peak near 2226 $\mathrm{cm}^{-1}$. This
476 shift does not effect the ability to qualitatively compare peaks from
477 the classical and quantum mechanical approaches, so the classical
478 spectra are shown as a shift relative to the natural oscillation of
479 the Morse bond.
480
481 \begin{figure}
482 \includegraphics[width=\linewidth]{spectra/spectra.pdf}
483 \caption{Spectrum of nitrile frequency shifts for the no-field
484 (black) and the full-field (red) simulations. Upper panel:
485 frequency shifts obtained from {\it ab initio} cluster
486 calculations. Lower panel: classical bond-length autocorrelation
487 spectrum for the flexible nitrile measured relative to the natural
488 frequency for the flexible bond. The dashed lines indicate the
489 mean frequencies for each of the distributions. The cluster
490 calculations exhibit a $2.68~\mathrm{cm}^{-1}$ field-induced red
491 shift, while the classical correlation functions predict a red
492 shift of $3.05~\mathrm{cm}^{-1}$.}
493 \label{fig:spectra}
494 \end{figure}
495
496 The classical approach includes both intramolecular and electrostatic
497 interactions, and so it implicitly couples \ce{CN} vibrations to other
498 vibrations within the molecule as well as to nitrile vibrations on
499 other nearby molecules. The classical frequency spectrum is
500 significantly broader because of this coupling. The {\it ab initio}
501 cluster approach exercises only the targeted nitrile bond, with no
502 additional coupling to other degrees of freedom. As a result the
503 quantum calculations are quite narrowly peaked around the experimental
504 nitrile frequency. Although the spectra are quite noisy, the main
505 effect seen in both distributions is a moderate shift to the red
506 ($3.05~\mathrm{cm}^{-1}$ classical and $2.68~\mathrm{cm}^{-1}$
507 quantum) when the full electrostatic field had induced the nematic
508 phase transition.
509
510 \section{Discussion}
511 Our simulations show that the united-atom model can reproduce the
512 field-induced nematic ordering of the 4-cyano-4'-pentylbiphenyl.
513 Because we are simulating a very small electrode separation (5~nm), a
514 voltage drop as low as 1.2~V was sufficient to induce the phase
515 change. This potential is significantly smaller than 100~V that was
516 used with a 5~$\mu$m gap to study the electrochemiluminescence of
517 rubrene in neat 5CB,\cite{Kojima19881789} and suggests that by using
518 electrodes separated by a nanometer-scale gap, it will be relatively
519 straightforward to observe the nitrile Stark shift in 5CB.
520
521 Both the classical correlation function and the isolated cluster
522 approaches to estimating the IR spectrum show that a population of
523 nitrile stretches shift by $\sim~3~\mathrm{cm}^{-1}$ to the red of
524 the unperturbed vibrational line. To understand the origin of this
525 shift, a more complete picture of the spatial ordering around the
526 nitrile bonds is required. We have computed the angle-dependent pair
527 distribution functions,
528 \begin{align}
529 g(r, \cos \omega) = & \frac{1}{\rho N} \left< \sum_{i} \sum_{j}
530 \delta \left(r - r_{ij}\right) \delta\left(\cos \omega_{ij} -
531 \cos \omega\right) \right> \\ \nonumber \\
532 g(r, \cos \theta) = & \frac{1}{\rho N} \left< \sum_{i}
533 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \theta_{i} -
534 \cos \theta \right) \right>
535 \end{align}
536 which provide information about the joint spatial and angular
537 correlations present in the system. The angles $\omega$ and $\theta$
538 are defined by vectors along the CN axis of each nitrile bond (see
539 figure \ref{fig:definition}).
540 \begin{figure}
541 \includegraphics[width=4in]{definition/definition.pdf}
542 \caption{Definitions of the angles between two nitrile bonds.}
543 \label{fig:definition}
544 \end{figure}
545
546 The primary structural effect of the field-induced phase transition is
547 apparent in figure \ref{fig:gofromega}. The nematic ordering transfers
548 population from the perpendicular ($\cos\omega\approx 0$) and
549 anti-aligned ($\cos\omega\approx -1$) to the nitrile-aligned peak
550 near $\cos\omega\approx 1$, leaving most other features undisturbed. This
551 change is visible in the simulations as an increased population of
552 aligned nitrile bonds in the first solvation shell.
553
554 \begin{figure}
555 \includegraphics[width=\linewidth]{gofrOmega/gofrOmega.pdf}
556 \caption{Contours of the angle-dependent pair distribution functions
557 for nitrile bonds on 5CB in the no field (upper panel) and full
558 field (lower panel) simulations. Dark areas signify regions of
559 enhanced density, while light areas signify depletion relative to
560 the bulk density.}
561 \label{fig:gofromega}
562 \end{figure}
563
564 Although it is certainly possible that the coupling between
565 closely-spaced nitrile pairs is responsible for some of the red-shift,
566 that is not the only structural change that is taking place. The
567 second two-dimensional pair distribution function, $g(r,\cos\theta)$,
568 shows that nematic ordering also transfers population that is directly
569 in line with the nitrile bond (see figure \ref{fig:gofrtheta}) to the
570 sides of the molecule, thereby freeing steric blockage can directly
571 influence the nitrile vibration. This is confirmed by observing the
572 one-dimensional $g(z)$ obtained by following the \ce{C -> N} vector
573 for each nitrile bond and observing the local density ($\rho(z)/\rho$)
574 of other atoms at a distance $z$ along this direction. The full-field
575 simulation shows a significant drop in the first peak of $g(z)$,
576 indicating that the nematic ordering has moved density away from the
577 region that is directly in line with the nitrogen side of the CN bond.
578
579 \begin{figure}
580 \includegraphics[width=\linewidth]{gofrTheta/gofrTheta.pdf}
581 \caption{Contours of the angle-dependent pair distribution function,
582 $g(r,\cos \theta)$, for finding any other atom at a distance and
583 angular deviation from the center of a nitrile bond. The top edge
584 of each contour plot corresponds to local density along the
585 direction of the nitrogen in the CN bond, while the bottom is in
586 the direction of the carbon atom. Bottom panel: $g(z)$ data taken
587 by following the \ce{C -> N} vector for each nitrile bond shows
588 that the field-induced phase transition reduces the population of
589 atoms that are directly in line with the nitrogen motion.}
590 \label{fig:gofrtheta}
591 \end{figure}
592
593 We are suggesting an anti-caging mechanism here -- the nematic
594 ordering provides additional space directly inline with the nitrile
595 vibration, and since the oscillator is fairly anharmonic, this freedom
596 provides a fraction of the nitrile bonds with a significant red-shift.
597
598 The cause of this shift does not appear to be related to the alignment
599 of those nitrile bonds with the field, but rather to the change in
600 local steric environment that is brought about by the
601 isotropic-nematic transition. We have compared configurations for many
602 of the cluster that exhibited the lowest frequencies (between 2190 and
603 2215 $\mathrm{cm}^{-1}$) and have observed some similar structural
604 features. The lowest frequencies appear to come from configurations
605 which have nearly-empty pockets directly opposite the nitrogen atom
606 from the nitrile carbon. However, because we do not have a
607 particularly large cluster population to interrogate, this is
608 certainly not quantitative confirmation of this effect.
609
610 The prediction of a small red-shift of the nitrile peak in 5CB in
611 response to a field-induced nematic ordering is the primary result of
612 this work, and although the proposed anti-caging mechanism is somewhat
613 speculative, this work provides some impetus for further theory and
614 experiments.
615
616 \section{Acknowledgements}
617 The authors thank Steven Corcelli and Zac Schultz for helpful comments
618 and suggestions. Support for this project was provided by the National
619 Science Foundation under grant CHE-0848243. Computational time was
620 provided by the Center for Research Computing (CRC) at the University
621 of Notre Dame.
622
623 \newpage
624
625 \bibliography{5CB}
626
627 \end{document}