ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4112
Committed: Fri May 9 16:02:26 2014 UTC (10 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 33814 byte(s)
Log Message:
Last

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
39 \keywords{}
40
41 \begin{document}
42
43
44
45 \begin{tocentry}
46 %\includegraphics[width=9cm]{Elip_3}
47 \includegraphics[width=9cm]{cluster.pdf}
48 \end{tocentry}
49
50 \begin{abstract}
51 4-cyano-4'-pentylbiphenyl (5CB) is a liquid crystal forming compound
52 with a terminal nitrile group aligned with the long axis of the
53 molecule. Simulations of condensed-phase 5CB were carried out both
54 with and without applied electric fields to provide an understanding
55 of the Stark shift of the terminal nitrile group. A field-induced
56 isotropic-nematic phase transition was observed in the simulations,
57 and the effects of this transition on the distribution of nitrile
58 frequencies were computed. Classical bond displacement correlation
59 functions exhibit a $\sim~3~\mathrm{cm}^{-1}$ red shift of a portion
60 of the main nitrile peak, and this shift was observed only when the
61 fields were large enough to induce orientational ordering of the
62 bulk phase. Distributions of frequencies obtained via cluster-based
63 fits to quantum mechanical energies of nitrile bond deformations
64 exhibit a similar $\sim~2.7~\mathrm{cm}^{-1}$ red shift. Joint
65 spatial-angular distribution functions indicate that phase-induced
66 anti-caging of the nitrile bond is contributing to the change in the
67 nitrile spectrum.
68 \end{abstract}
69
70 \newpage
71
72 \section{Introduction}
73
74 Because the triple bond between nitrogen and carbon is sensitive to
75 local electric fields, nitrile groups can report on field strengths
76 via their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
77 response of nitrile groups to electric fields has now been
78 investigated for a number of small molecules,\cite{Andrews:2000qv} as
79 well as in biochemical settings, where nitrile groups can act as
80 minimally invasive probes of structure and
81 dynamics.\cite{Tucker:2004qq,Webb:2008kn,Lindquist:2009fk,Fafarman:2010dq}
82 The vibrational Stark effect has also been used to study the effects
83 of electric fields on nitrile-containing self-assembled monolayers at
84 metallic interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
85
86 Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
87 molecule with a terminal nitrile group, has seen renewed interest as
88 one way to impart order on the surfactant interfaces of
89 nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
90 that can be used to promote particular kinds of
91 self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
92 is a particularly interesting case for studying electric field
93 effects, as 5CB exhibits an isotropic to nematic phase transition that
94 can be triggered by the application of an external field near room
95 temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
96 possibility that the field-induced changes in the local environment
97 could have dramatic effects on the vibrations of this particular nitrile
98 bond. Although the infrared spectroscopy of 5CB has been
99 well-investigated, particularly as a measure of the kinetics of the
100 phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
101 seen the detailed theoretical treatment that biologically-relevant
102 small molecules have
103 received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Morales:2009fp,Waegele:2010ve}
104
105 The fundamental characteristic of liquid crystal mesophases is that
106 they maintain some degree of orientational order while translational
107 order is limited or absent. This orientational order produces a
108 complex direction-dependent response to external perturbations like
109 electric fields and mechanical distortions. The anisotropy of the
110 macroscopic phases originates in the anisotropy of the constituent
111 molecules, which typically have highly non-spherical structures with a
112 significant degree of internal rigidity. In nematic phases, rod-like
113 molecules are orientationally ordered with isotropic distributions of
114 molecular centers of mass. For example, 5CB has a solid to nematic
115 phase transition at 18C and a nematic to isotropic transition at
116 35C.\cite{Gray:1973ca}
117
118 In smectic phases, the molecules arrange themselves into layers with
119 their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
120 respect to the layer planes. The behavior of the $S_{A}$ phase can be
121 explained with models based solely on geometric factors and van der
122 Waals interactions. The Gay-Berne potential, in particular, has been
123 widely used in the liquid crystal community to describe this
124 anisotropic phase
125 behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
126 However, these simple models are insufficient to describe liquid
127 crystal phases which exhibit more complex polymorphic nature.
128 Molecules which form $S_{A}$ phases can exhibit a wide variety of
129 sub-phases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
130 partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
131 ($S_{A_{d}}$), and often have a terminal cyano or nitro group. In
132 particular, lyotropic liquid crystals (those exhibiting liquid crystal
133 phase transitions as a function of water concentration), often have
134 polar head groups or zwitterionic charge separated groups that result
135 in strong dipolar interactions,\cite{Collings:1997rz} and terminal
136 cyano groups (like the one in 5CB) can induce permanent longitudinal
137 dipoles.\cite{Levelut:1981eu} Modeling of the phase behavior of these
138 molecules either requires additional dipolar
139 interactions,\cite{Bose:2012eu} or a unified-atom approach utilizing
140 point charges on the sites that contribute to the dipole
141 moment.\cite{Zhang:2011hh}
142
143 Macroscopic electric fields applied using electrodes on opposing sides
144 of a sample of 5CB have demonstrated the phase change of the molecule
145 as a function of electric field.\cite{Lim:2006xq} These previous
146 studies have shown the nitrile group serves as an excellent indicator
147 of the molecular orientation within the applied field. Lee {\it et
148 al.}~showed a 180 degree change in field direction could be probed
149 with the nitrile peak intensity as it changed along with molecular
150 alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
151
152 While these macroscopic fields work well at indicating the bulk
153 response, the response at a molecular scale has not been studied. With
154 the advent of nano-electrodes and the ability to couple these
155 electrodes to atomic force microscopy, control of electric fields
156 applied across nanometer distances is now possible.\cite{C3AN01651J}
157 In special cases where the macroscopic fields are insufficient to
158 cause an observable Stark effect without dielectric breakdown of the
159 material, small potentials across nanometer-sized gaps may have
160 sufficient strength. For a gap of 5 nm between a lower electrode
161 having a nanoelectrode placed near it via an atomic force microscope,
162 a potential of 1 V applied across the electrodes is equivalent to a
163 field of $2 \times 10^8~\mathrm{V/m}$. This field is certainly strong
164 enough to cause the isotropic-nematic phase change and an observable
165 Stark tuning of the nitrile bond. We expect that this would be readily
166 visible experimentally through Raman or IR spectroscopy.
167
168 In the sections that follow, we outline a series of coarse-grained
169 (united atom) classical molecular dynamics simulations of 5CB that
170 were done in the presence of static electric fields. These simulations
171 were then coupled with both {\it ab intio} calculations of
172 CN-deformations and classical bond-length correlation functions to
173 predict spectral shifts. These predictions should be verifiable via
174 scanning electrochemical microscopy.
175
176 \section{Computational Details}
177 The force-field used to model 5CB was a united-atom model that was
178 parameterized by Guo {\it et al.}\cite{Zhang:2011hh} However, for most
179 of the simulations, both of the phenyl rings and the nitrile bond were
180 treated as rigid bodies to allow for larger time steps and longer
181 simulation times. The geometries of the rigid bodies were taken from
182 equilibrium bond distances and angles. Although the individual phenyl
183 rings were held rigid, bonds, bends, torsions and inversion centers
184 that involved atoms in these substructures (but with connectivity to
185 the rest of the molecule) were still included in the potential and
186 force calculations.
187
188 Periodic simulations cells containing 270 molecules in random
189 orientations were constructed and were locked at experimental
190 densities. Electrostatic interactions were computed using damped
191 shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
192 were equilibrated for 1~ns at a temperature of 300K. Simulations with
193 applied fields were carried out in the microcanonical (NVE) ensemble
194 with an energy corresponding to the average energy from the canonical
195 (NVT) equilibration runs. Typical applied-field equilibration runs
196 were more than 60~ns in length.
197
198 Static electric fields with magnitudes similar to what would be
199 available in an experimental setup were applied to the different
200 simulations. With an assumed electrode separation of 5 nm and an
201 electrostatic potential that is limited by the voltage required to
202 split water (1.23V), the maximum realistic field that could be applied
203 is $\sim 0.024$ V/\AA. Three field environments were investigated:
204 (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
205 field = 0.024 V/\AA\ .
206
207 After the systems had come to equilibrium under the applied fields,
208 additional simulations were carried out with a flexible (Morse)
209 nitrile bond,
210 \begin{equation}
211 V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
212 \label{eq:morse}
213 \end{equation}
214 where $r_e= 1.157$ \AA (the fixed CN bond length from the force field
215 of Guo {\it et al.}\cite{Zhang:2011hh}), $D_e = 212.95 \mathrm{~kcal~}
216 / \mathrm{mol}^{-1}$ (the average bond energy for CN triple bonds) and
217 $\beta = 2.526 $\AA~$^{-1}$. These parameters correspond to a
218 vibrational frequency of $\approx 2226 \mathrm{~cm}^{-1}$, which is
219 very close to the frequency of the nitrile peak in the vibrational
220 spectrum of neat 5CB. The flexible nitrile moiety required simulation
221 time steps of 1~fs, so the additional flexibility was introduced only
222 after the rigid systems had come to equilibrium under the applied
223 fields. Whenever time correlation functions were computed from the
224 flexible simulations, statistically-independent configurations
225 (separated in time by 10 ns) were sampled from the last 110 ns of the
226 induced-field runs. These configurations were then equilibrated with
227 the flexible nitrile moiety for 100 ps, and time correlation functions
228 were computed using data sampled from an additional 20 ps of run time
229 carried out in the microcanonical ensemble.
230
231 \section{Field-induced Nematic Ordering}
232
233 In order to characterize the orientational ordering of the system, the
234 primary quantity of interest is the nematic (orientational) order
235 parameter. This was determined using the tensor
236 \begin{equation}
237 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{u}_{i
238 \alpha} \hat{u}_{i \beta} - \delta_{\alpha \beta} \right)
239 \end{equation}
240 where $\alpha, \beta = x, y, z$, and $\hat{u}_i$ is the molecular
241 end-to-end unit vector for molecule $i$. The nematic order parameter
242 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
243 corresponding eigenvector defines the director axis for the phase.
244 $S$ takes on values close to 1 in highly ordered (smectic A) phases,
245 but falls to much smaller values ($0 \rightarrow 0.3$) for isotropic
246 fluids. Note that the nitrogen and the terminal chain atom were used
247 to define the vector for each molecule, so the typical order
248 parameters are lower than if one defined a vector using only the rigid
249 core of the molecule. In nematic phases, typical values for $S$ are
250 close to 0.5.
251
252 The field-induced phase transition can be clearly seen over the course
253 of a 60 ns equilibration runs in figure \ref{fig:orderParameter}. All
254 three of the systems started in a random (isotropic) packing, with
255 order parameters near 0.2. Over the course 10 ns, the full field
256 causes an alignment of the molecules (due primarily to the interaction
257 of the nitrile group dipole with the electric field). Once this
258 system began exhibiting nematic ordering, the orientational order
259 parameter became stable for the remaining 150 ns of simulation time.
260 It is possible that the partial-field simulation is meta-stable and
261 given enough time, it would eventually find a nematic-ordered phase,
262 but the partial-field simulation was stable as an isotropic phase for
263 the full duration of the 60 ns simulation. Ellipsoidal renderings of
264 the final configurations of the runs show that the full-field (0.024
265 V/\AA\ ) experienced a isotropic-nematic phase transition and has
266 ordered with a director axis that is parallel to the direction of the
267 applied field.
268
269 \begin{figure}[H]
270 \includegraphics[width=\linewidth]{orderParameter.pdf}
271 \caption{Evolution of the orientational order parameters for the
272 no-field, partial field, and full field simulations over the
273 course of 60 ns. Each simulation was started from a
274 statistically-independent isotropic configuration. On the right
275 are ellipsoids representing the final configurations at three
276 different field strengths: zero field (bottom), partial field
277 (middle), and full field (top)}
278 \label{fig:orderParameter}
279 \end{figure}
280
281
282 \section{Sampling the CN bond frequency}
283
284 The vibrational frequency of the nitrile bond in 5CB depends on
285 features of the local solvent environment of the individual molecules
286 as well as the bond's orientation relative to the applied field. The
287 primary quantity of interest for interpreting the condensed phase
288 spectrum of this vibration is the distribution of frequencies
289 exhibited by the 5CB nitrile bond under the different electric fields.
290 There have been a number of elegant techniques for obtaining
291 vibrational line shapes from classical simulations, including a
292 perturbation theory approach,\cite{Morales:2009fp} the use of an
293 optimized QM/MM approach coupled with the fluctuating frequency
294 approximation,\cite{Lindquist:2008qf} and empirical frequency
295 correlation maps.\cite{Choi:2008cr,Oh:2008fk} Three distinct (and
296 comparatively primitive) methods for mapping classical simulations
297 onto vibrational spectra were brought to bear on the simulations in
298 this work:
299 \begin{enumerate}
300 \item Isolated 5CB molecules and their immediate surroundings were
301 extracted from the simulations. These nitrile bonds were stretched
302 by displacing the nitrogen along the CN bond vector with the carbon
303 atom remaining stationary. Single-point {\em ab initio} calculations
304 were used to obtain Morse-oscillator fits for the local vibrational
305 motion along that bond.
306 \item The empirical frequency correlation maps developed by Choi {\it
307 et al.}~\cite{Choi:2008cr,Oh:2008fk} for nitrile moieties in water
308 were utilized by adding an electric field contribution to the local
309 electrostatic potential.
310 \item Classical bond-length autocorrelation functions were Fourier
311 transformed to directly obtain the vibrational spectrum from
312 molecular dynamics simulations.
313 \end{enumerate}
314
315 \subsection{CN frequencies from isolated clusters}
316 The size of the condensed phase liquid crystal system prevented direct
317 computation of the complete library of nitrile bond frequencies using
318 {\it ab initio} methods. In order to sample the nitrile frequencies
319 present in the condensed-phase, individual molecules were selected
320 randomly to serve as the center of a local (gas phase) cluster. To
321 include steric, electrostatic, and other effects from molecules
322 located near the targeted nitrile group, portions of other molecules
323 nearest to the nitrile group were included in the quantum mechanical
324 calculations. Steric interactions are generally shorter ranged than
325 electrostatic interactions, so portions of surrounding molecules that
326 cause electrostatic perturbations to the central nitrile (e.g. the
327 biphenyl core and nitrile moieties) must be included if they fall
328 anywhere near the CN bond. Portions of these molecules that interact
329 primarily via dispersion and steric repulsion (e.g. the alkyl tails)
330 can be truncated at a shorter distance.
331
332 The surrounding solvent molecules were therefore divided into ``body''
333 (the two phenyl rings and the nitrile bond) and ``tail'' (the alkyl
334 chain). Any molecule which had a body atom within 6~\AA\ of the
335 midpoint of the target nitrile bond had its own molecular body (the
336 4-cyano-biphenyl moiety) included in the configuration. Likewise, the
337 entire alkyl tail was included if any tail atom was within 4~\AA\ of
338 the target nitrile bond. If tail atoms (but no body atoms) were
339 included within these distances, only the tail was included as a
340 capped propane molecule.
341
342 \begin{figure}[H]
343 \includegraphics[width=\linewidth]{cluster.pdf}
344 \caption{Cluster calculations were performed on randomly sampled 5CB
345 molecules (shown in red) from the full-field and no-field
346 simulations. Surrounding molecular bodies were included if any
347 body atoms were within 6 \AA\ of the target nitrile bond, and
348 tails were included if they were within 4 \AA. Included portions
349 of these molecules are shown in green. The CN bond on the target
350 molecule was stretched and compressed, and the resulting single
351 point energies were fit to Morse oscillators to obtain a
352 distribution of frequencies.}
353 \label{fig:cluster}
354 \end{figure}
355
356 Inferred hydrogen atom locations were added to the cluster geometries,
357 and the nitrile bond was stretched from 0.87 to 1.52~\AA\ at
358 increments of 0.05~\AA. This generated 13 configurations per gas phase
359 cluster. Single-point energies were computed using the B3LYP
360 functional~\cite{Becke:1993kq,Lee:1988qf} and the 6-311++G(d,p) basis
361 set. For the cluster configurations that had been generated from
362 molecular dynamics running under applied fields, the density
363 functional calculations had a field of $5 \times 10^{-4}$ atomic units
364 ($E_h / (e a_0)$) applied in the $+z$ direction in order to match the
365 molecular dynamics simulations.
366
367 The energies for the stretched / compressed nitrile bond in each of
368 the clusters were used to fit Morse potentials, and the frequencies
369 were obtained from the $0 \rightarrow 1$ transition for the energy
370 levels for this potential.\cite{Morse:1929xy} To obtain a spectrum,
371 each of the frequencies was convoluted with a Lorentzian line shape
372 with a width of 1.5 $\mathrm{cm}^{-1}$. This linewidth corresponds to
373 a vibrational lifetime of $\sim 3.5$ ps, which is within the reported
374 ranges ($\sim 1 - 5$ ps) for CN stretching vibrational lifetimes in
375 other molecules.\cite{Ghosh:2009qf,Ha:2009xy,Waegele:2010ve}.
376 Available computing resources limited the sampling to 100 clusters for
377 both the no-field and full-field spectra. Comparisons of the quantum
378 mechanical spectrum to the classical are shown in figure
379 \ref{fig:spectra}. The mean frequencies obtained from the
380 distributions give a field-induced red shift of
381 $2.68~\mathrm{cm}^{-1}$.
382
383 \subsection{CN frequencies from potential-frequency maps}
384
385 One approach which has been used to successfully analyze the spectrum
386 of nitrile and thiocyanate probes in aqueous environments was
387 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} This
388 method involves finding a multi-parameter fit that maps between the
389 local electrostatic potential at selected sites surrounding the
390 nitrile bond and the vibrational frequency of that bond obtained from
391 more expensive {\it ab initio} methods. This approach is similar in
392 character to the field-frequency maps developed by the Skinner group
393 for OH stretches in liquid water.\cite{Corcelli:2004ai,Auer:2007dp}
394
395 To use the potential-frequency maps, the local electrostatic
396 potential, $\phi_a$, is computed at 20 sites ($a = 1 \rightarrow 20$)
397 that surround the nitrile bond,
398 \begin{equation}
399 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{j}
400 \frac{q_j}{\left|r_{aj}\right|}.
401 \end{equation}
402 Here $q_j$ is the partial charge on atom $j$ (residing on a different
403 molecule) and $r_{aj}$ is the distance between site $a$ and atom $j$.
404 The original map was parameterized in liquid water and comprises a set
405 of parameters, $l_a$, that predict the shift in nitrile peak
406 frequency,
407 \begin{equation}
408 \delta\tilde{\nu} =\sum^{20}_{a=1} l_{a}\phi_{a}.
409 \end{equation}
410
411 The simulations of 5CB were carried out in the presence of
412 externally-applied uniform electric fields. Although uniform fields
413 exert forces on charge sites, they only contribute to the potential if
414 one defines a reference point that can serve as an origin. One simple
415 modification to the potential at each of the probe sites is to use the
416 centroid of the \ce{CN} bond as the origin for that site,
417 \begin{equation}
418 \phi_a^\prime = \phi_a + \frac{1}{4\pi\epsilon_{0}} \vec{E} \cdot
419 \left(\vec{r}_a - \vec{r}_\ce{CN} \right)
420 \end{equation}
421 where $\vec{E}$ is the uniform electric field, $\left( \vec{r}_{a} -
422 \vec{r}_\ce{CN} \right)$ is the displacement between the
423 coordinates described by Choi {\it et
424 al.}~\cite{Choi:2008cr,Oh:2008fk} and the \ce{CN} bond centroid.
425 $\phi_a^\prime$ then contains an effective potential contributed by
426 the uniform field in addition to the local potential contributions
427 from other molecules.
428
429 The sites $\{\vec{r}_a\}$ and weights $\left\{l_a \right\}$
430 developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} are quite
431 symmetric around the \ce{CN} centroid, and even at large uniform field
432 values we observed nearly-complete cancellation of the potential
433 contributions from the uniform field.
434
435 The frequency shifts were computed for 4000 configurations sampled
436 every 1 ps after the systems had equilibrated. The potential
437 frequency map produces a small blue shift of 0.34 cm$^{-1}$, and the
438 frequency shifts are quite narrowly distributed. However, the
439 parameters for the potential frequency maps were derived for nitrile
440 bonds in aqueous solutions, where the magnitudes of the local fields
441 and electrostatic potentials are much larger than they would be in
442 neat 5CB.
443
444 We note that in 5CB there does not appear to be a particularly strong
445 correlation between the electric field strengths observed at the
446 nitrile centroid and the calculated vibrational frequencies. In
447 Fig. \ref{fig:fieldMap} we show the calculated frequencies plotted
448 against the field magnitude as well as the parallel and perpendicular
449 components of that field.
450
451 \begin{figure}
452 \includegraphics[width=\linewidth]{fieldMap.pdf}
453 \caption{The observed cluster frequencies have no apparent
454 correlation with the electric field felt at the centroid of the
455 nitrile bond. Upper panel: vibrational frequencies plotted
456 against the component of the field parallel to the CN bond.
457 Middle panel: plotted against the magnitude of the field
458 components perpendicular to the CN bond. Lower panel: plotted
459 against the total field magnitude.}
460 \label{fig:fieldMap}
461 \end{figure}
462
463
464 \subsection{CN frequencies from bond length autocorrelation functions}
465
466 The distribution of nitrile vibrational frequencies can also be found
467 using classical time correlation functions. This was done by
468 replacing the rigid \ce{CN} bond with a flexible Morse oscillator
469 described in Eq. \ref{eq:morse}. Since the systems were perturbed by
470 the addition of a flexible high-frequency bond, they were allowed to
471 re-equilibrate in the canonical (NVT) ensemble for 100 ps with 1 fs
472 time steps. After equilibration, each configuration was run in the
473 microcanonical (NVE) ensemble for 20 ps. Configurations sampled every
474 fs were then used to compute bond-length autocorrelation functions,
475 \begin{equation}
476 C(t) = \langle \delta r(t) \cdot \delta r(0) ) \rangle
477 \end{equation}
478 %
479 where $\delta r(t) = r(t) - r_0$ is the deviation from the equilibrium
480 bond distance at time $t$. Because the other atomic sites have very
481 small partial charges, this correlation function is an approximation
482 to the dipole autocorrelation function for the molecule, which would
483 be particularly relevant to computing the IR spectrum. Eleven
484 statistically-independent correlation functions were obtained by
485 allowing the systems to run 10 ns with rigid \ce{CN} bonds followed by
486 120 ps equilibration and data collection using the flexible \ce{CN}
487 bonds. This process was repeated 11 times, and the total sampling
488 time, from sample preparation to final configurations, exceeded 160 ns
489 for each of the field strengths investigated.
490
491 The correlation functions were filtered using exponential apodization
492 functions,\cite{FILLER:1964yg} $f(t) = e^{-|t|/c}$, with a time
493 constant, $c =$ 3.5 ps, and were Fourier transformed to yield a
494 spectrum,
495 \begin{equation}
496 I(\omega) = \int_{-\infty}^{\infty} C(t) f(t) e^{-i \omega t} dt.
497 \end{equation}
498 This time constant was chosen to match the Lorentzian linewidth that
499 was used for computing the quantum mechanical spectra, and falls
500 within the range of reported lifetimes for CN vibrations in other
501 nitrile-containing molecules. The sample-averaged classical nitrile
502 spectrum can be seen in Figure \ref{fig:spectra}. The Morse oscillator
503 parameters listed above yield a natural frequency of 2226
504 $\mathrm{cm}^{-1}$ (close to the experimental value). To compare peaks
505 from the classical and quantum mechanical approaches, both are
506 displayed on an axis centered on the experimental nitrile frequency.
507
508 \begin{figure}
509 \includegraphics[width=\linewidth]{spectra.pdf}
510 \caption{Spectrum of nitrile frequency shifts for the no-field
511 (black) and the full-field (red) simulations. Upper panel:
512 frequency shifts obtained from {\it ab initio} cluster
513 calculations. Lower panel: classical bond-length autocorrelation
514 spectrum for the flexible nitrile measured relative to the natural
515 frequency for the flexible bond. The dashed lines indicate the
516 mean frequencies for each of the distributions. The cluster
517 calculations exhibit a $2.68~\mathrm{cm}^{-1}$ field-induced red
518 shift, while the classical correlation functions predict a red
519 shift of $2.29~\mathrm{cm}^{-1}$.}
520 \label{fig:spectra}
521 \end{figure}
522
523 The classical approach includes both intramolecular and electrostatic
524 interactions, and so it implicitly couples \ce{CN} vibrations to other
525 vibrations within the molecule as well as to nitrile vibrations on
526 other nearby molecules. The classical frequency spectrum is
527 significantly broader because of this coupling. The {\it ab initio}
528 cluster approach exercises only the targeted nitrile bond, with no
529 additional coupling to other degrees of freedom. As a result the
530 quantum calculations are quite narrowly peaked around the experimental
531 nitrile frequency. Although the spectra are quite noisy, the main
532 effect seen in both distributions is a moderate shift to the red
533 ($2.29~\mathrm{cm}^{-1}$ classical and $2.68~\mathrm{cm}^{-1}$
534 quantum) after the electrostatic field had induced the nematic phase
535 transition.
536
537 \section{Discussion}
538 Our simulations show that the united-atom model can reproduce the
539 field-induced nematic ordering of the 4-cyano-4'-pentylbiphenyl.
540 Because we are simulating a very small electrode separation (5~nm), a
541 voltage drop as low as 1.2~V was sufficient to induce the phase
542 change. This potential is significantly smaller than 100~V that was
543 used with a 5~$\mu$m gap to study the electrochemiluminescence of
544 rubrene in neat 5CB,\cite{Kojima19881789} and suggests that by using
545 electrodes separated by a nanometer-scale gap, it will be relatively
546 straightforward to observe the nitrile Stark shift in 5CB.
547
548 Both the classical correlation function and the isolated cluster
549 approaches to estimating the IR spectrum show that a population of
550 nitrile stretches shift by $\sim~3~\mathrm{cm}^{-1}$ to the red of
551 the unperturbed vibrational line. To understand the origin of this
552 shift, a more complete picture of the spatial ordering around the
553 nitrile bonds is required. We have computed the angle-dependent pair
554 distribution functions,
555 \begin{align}
556 g(r, \cos \omega) = & \frac{1}{\rho N} \left< \sum_{i} \sum_{j}
557 \delta \left(r - r_{ij}\right) \delta\left(\cos \omega_{ij} -
558 \cos \omega\right) \right> \\ \nonumber \\
559 g(r, \cos \theta) = & \frac{1}{\rho N} \left< \sum_{i}
560 \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \theta_{i} -
561 \cos \theta \right) \right>
562 \end{align}
563 which provide information about the joint spatial and angular
564 correlations present in the system. The angles $\omega$ and $\theta$
565 are defined by vectors along the CN axis of each nitrile bond (see
566 figure \ref{fig:definition}).
567 \begin{figure}
568 \includegraphics[width=4in]{definition.pdf}
569 \caption{Definitions of the angles between two nitrile bonds.}
570 \label{fig:definition}
571 \end{figure}
572
573 The primary structural effect of the field-induced phase transition is
574 apparent in figure \ref{fig:gofromega}. The nematic ordering transfers
575 population from the perpendicular ($\cos\omega\approx 0$) and
576 anti-aligned ($\cos\omega\approx -1$) to the nitrile-aligned peak
577 near $\cos\omega\approx 1$, leaving most other features undisturbed. This
578 change is visible in the simulations as an increased population of
579 aligned nitrile bonds in the first solvation shell.
580
581 \begin{figure}
582 \includegraphics[width=\linewidth]{gofrOmega.pdf}
583 \caption{Contours of the angle-dependent pair distribution functions
584 for nitrile bonds on 5CB in the no field (upper panel) and full
585 field (lower panel) simulations. Dark areas signify regions of
586 enhanced density, while light areas signify depletion relative to
587 the bulk density.}
588 \label{fig:gofromega}
589 \end{figure}
590
591 Although it is certainly possible that the coupling between
592 closely-spaced nitrile pairs is responsible for some of the red-shift,
593 that is not the only structural change that is taking place. The
594 second two-dimensional pair distribution function, $g(r,\cos\theta)$,
595 shows that nematic ordering also transfers population that is directly
596 in line with the nitrile bond (see figure \ref{fig:gofrtheta}) to the
597 sides of the molecule, thereby freeing steric blockage which can
598 directly influence the nitrile vibration. This is confirmed by
599 observing the one-dimensional $g(z)$ obtained by following the \ce{C
600 -> N} vector for each nitrile bond and observing the local density
601 ($\rho(z)/\rho$) of other atoms at a distance $z$ along this
602 direction. The full-field simulation shows a significant drop in the
603 first peak of $g(z)$, indicating that the nematic ordering has moved
604 density away from the region that is directly in line with the
605 nitrogen side of the CN bond.
606
607 \begin{figure}
608 \includegraphics[width=\linewidth]{gofrTheta.pdf}
609 \caption{Contours of the angle-dependent pair distribution function,
610 $g(r,\cos \theta)$, for finding any other atom at a distance and
611 angular deviation from the center of a nitrile bond. The top edge
612 of each contour plot corresponds to local density along the
613 direction of the nitrogen in the CN bond, while the bottom is in
614 the direction of the carbon atom. Bottom panel: $g(z)$ data taken
615 by following the \ce{C -> N} vector for each nitrile bond shows
616 that the field-induced phase transition reduces the population of
617 atoms that are directly in line with the nitrogen motion.}
618 \label{fig:gofrtheta}
619 \end{figure}
620
621 We are suggesting an anti-caging mechanism here -- the nematic
622 ordering provides additional space directly inline with the nitrile
623 vibration, and since the oscillator is fairly anharmonic, this freedom
624 provides a fraction of the nitrile bonds with a significant red-shift.
625
626 The cause of this shift does not appear to be related to the alignment
627 of those nitrile bonds with the field, but rather to the change in
628 local steric environment that is brought about by the
629 isotropic-nematic transition. We have compared configurations for many
630 of the cluster that exhibited the lowest frequencies (between 2190 and
631 2215 $\mathrm{cm}^{-1}$) and have observed some similar structural
632 features. The lowest frequencies appear to come from configurations
633 which have nearly-empty pockets directly opposite the nitrogen atom
634 from the nitrile carbon. However, because we do not have a
635 particularly large cluster population to interrogate, this is
636 certainly not quantitative confirmation of this effect.
637
638 The prediction of a small red-shift of the nitrile peak in 5CB in
639 response to a field-induced nematic ordering is the primary result of
640 this work, and although the proposed anti-caging mechanism is somewhat
641 speculative, this work provides some impetus for further theory and
642 experiments.
643
644 \section{Acknowledgements}
645 The authors thank Steven Corcelli and Zac Schultz for helpful comments
646 and suggestions. Support for this project was provided by the National
647 Science Foundation under grant CHE-0848243. Computational time was
648 provided by the Center for Research Computing (CRC) at the University
649 of Notre Dame.
650
651 \newpage
652
653 \bibliography{5CB}
654
655 \end{document}