ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4095
Committed: Tue Apr 1 14:44:23 2014 UTC (10 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 32506 byte(s)
Log Message:
Many changes in advance of publication

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