ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4039
Committed: Thu Feb 20 21:23:54 2014 UTC (10 years, 6 months ago) by gezelter
Content type: application/x-tex
File size: 29409 byte(s)
Log Message:
New stuff

File Contents

# User Rev Content
1 gezelter 4007 \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 jmarr 4013 %\usepackage{endfloat}
16 gezelter 4007 \usepackage{tabularx}
17     \usepackage{longtable}
18     \usepackage{graphicx}
19     \usepackage{multirow}
20     \usepackage{multicol}
21     \usepackage{achemso}
22 jmarr 4013 \usepackage{subcaption}
23 gezelter 4007 \usepackage[colorinlistoftodos]{todonotes}
24     \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25     % \usepackage[square, comma, sort&compress]{natbib}
26     \usepackage{url}
27     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29     9.0in \textwidth 6.5in \brokenpenalty=10000
30    
31     % double space list of tables and figures
32     %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33     \setlength{\abovecaptionskip}{20 pt}
34     \setlength{\belowcaptionskip}{30 pt}
35    
36     % \bibpunct{}{}{,}{s}{}{;}
37    
38     %\citestyle{nature}
39     % \bibliographystyle{achemso}
40    
41    
42     \title{Nitrile vibrations as reporters of field-induced phase
43 gezelter 4033 transitions in 4-cyano-4'-pentylbiphenyl (5CB)}
44 gezelter 4007 \author{James M. Marr}
45     \author{J. Daniel Gezelter}
46     \email{gezelter@nd.edu}
47     \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
48     Department of Chemistry and Biochemistry\\
49     University of Notre Dame\\
50     Notre Dame, Indiana 46556}
51    
52     \date{\today}
53    
54     \begin{document}
55    
56     \maketitle
57    
58     \begin{doublespace}
59    
60     \begin{abstract}
61 gezelter 4028 4-cyano-4'-pentylbiphenyl (5CB) is a liquid-crystal-forming compound
62 gezelter 4026 with a terminal nitrile group aligned with the long axis of the
63     molecule. Simulations of condensed-phase 5CB were carried out both
64 gezelter 4027 with and without applied electric fields to provide an understanding
65 gezelter 4028 of the the Stark shift of the terminal nitrile group. A
66     field-induced isotropic-nematic phase transition was observed in the
67     simulations, and the effects of this transition on the distribution
68     of nitrile frequencies were computed. Classical bond displacement
69     correlation functions exhibit a $\sim 40 \mathrm{~cm}^{-1}$ red
70     shift of a portion of the main nitrile peak, and this shift was
71     observed only when the fields were large enough to induce
72     orientational ordering of the bulk phase. Our simulations appear to
73     indicate that phase-induced changes to the local surroundings are a
74     larger contribution to the change in the nitrile spectrum than
75     direct field contributions.
76 gezelter 4007 \end{abstract}
77    
78     \newpage
79    
80     \section{Introduction}
81    
82 gezelter 4028 Nitrile groups can serve as very precise electric field reporters via
83     their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
84     triple bond between the nitrogen and the carbon atom is very sensitive
85     to local field changes and has been observed to have a direct impact
86     on the peak position within the spectrum. The Stark shift in the
87 gezelter 4039 spectrum can be quantified and mapped onto a field that is impinging
88     upon the nitrile bond. This has been used extensively in biological
89     systems like proteins and enzymes.\cite{Tucker:2004qq,Webb:2008kn}
90 gezelter 4028
91     The response of nitrile groups to electric fields has now been
92     investigated for a number of small molecules,\cite{Andrews:2000qv} as
93     well as in biochemical settings, where nitrile groups can act as
94     minimally invasive probes of structure and
95     dynamics.\cite{Lindquist:2009fk,Fafarman:2010dq} The vibrational Stark
96     effect has also been used to study the effects of electric fields on
97     nitrile-containing self-assembled monolayers at metallic
98     interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
99    
100     Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
101     molecule with a terminal nitrile group, has seen renewed interest as
102     one way to impart order on the surfactant interfaces of
103     nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
104     that can be used to promote particular kinds of
105     self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
106     is a particularly interesting case for studying electric field
107     effects, as 5CB exhibits an isotropic to nematic phase transition that
108     can be triggered by the application of an external field near room
109     temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
110     possiblity that the field-induced changes in the local environment
111     could have dramatic effects on the vibrations of this particular CN
112     bond. Although the infrared spectroscopy of 5CB has been
113     well-investigated, particularly as a measure of the kinetics of the
114     phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
115     seen the detailed theoretical treatment that biologically-relevant
116     small molecules have
117     received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Waegele:2010ve}
118    
119 gezelter 4007 The fundamental characteristic of liquid crystal mesophases is that
120     they maintain some degree of orientational order while translational
121     order is limited or absent. This orientational order produces a
122     complex direction-dependent response to external perturbations like
123 gezelter 4028 electric fields and mechanical distortions. The anisotropy of the
124 gezelter 4007 macroscopic phases originates in the anisotropy of the constituent
125     molecules, which typically have highly non-spherical structures with a
126 gezelter 4028 significant degree of internal rigidity. In nematic phases, rod-like
127 gezelter 4007 molecules are orientationally ordered with isotropic distributions of
128 gezelter 4028 molecular centers of mass. For example, 5CB has a solid to nematic
129     phase transition at 18C and a nematic to isotropic transition at
130     35C.\cite{Gray:1973ca}
131 gezelter 4007
132 gezelter 4028 In smectic phases, the molecules arrange themselves into layers with
133     their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
134     respect to the layer planes. The behavior of the $S_{A}$ phase can be
135     partially explained with models mainly based on geometric factors and
136     van der Waals interactions. The Gay-Berne potential, in particular,
137     has been widely used in the liquid crystal community to describe this
138     anisotropic phase
139     behavior.~\cite{Gay:1981yu,Berne72,Kushick:1976xy,Luckhurst90,Cleaver:1996rt}
140     However, these simple models are insufficient to describe liquid
141     crystal phases which exhibit more complex polymorphic nature.
142     Molecules which form $S_{A}$ phases can exhibit a wide variety of
143     subphases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
144     partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
145     ($S_{A_{d}}$), and often have a terminal cyano or nitro group. In
146     particular, lyotropic liquid crystals (those exhibiting liquid crystal
147     phase transition as a function of water concentration), often have
148     polar head groups or zwitterionic charge separated groups that result
149     in strong dipolar interactions,\cite{Collings:1997rz} and terminal cyano
150     groups (like the one in 5CB) can induce permanent longitudinal
151     dipoles.\cite{Levelut:1981eu}
152 gezelter 4007
153 gezelter 4028 Macroscopic electric fields applied using electrodes on opposing sides
154     of a sample of 5CB have demonstrated the phase change of the molecule
155     as a function of electric field.\cite{Lim:2006xq} These previous
156     studies have shown the nitrile group serves as an excellent indicator
157     of the molecular orientation within the applied field. Lee {\it et
158     al.}~showed a 180 degree change in field direction could be probed
159     with the nitrile peak intensity as it changed along with molecular
160     alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
161 gezelter 4007
162 gezelter 4028 While these macroscopic fields work well at indicating the bulk
163 gezelter 4007 response, the atomic scale response has not been studied. With the
164     advent of nano-electrodes and coupling them with atomic force
165     microscopy, control of electric fields applied across nanometer
166 gezelter 4028 distances is now possible.\cite{citation1} While macroscopic fields
167     are insufficient to cause a Stark effect without dielectric breakdown
168     of the material, small fields across nanometer-sized gaps may be of
169     sufficient strength. For a gap of 5 nm between a lower electrode
170     having a nanoelectrode placed near it via an atomic force microscope,
171     a potential of 1 V applied across the electrodes is equivalent to a
172     field of 2x10\textsuperscript{8} $\frac{V}{M}$. This field is
173     certainly strong enough to cause the isotropic-nematic phase change
174     and as well as Stark tuning of the nitrile bond. This should be
175     readily visible experimentally through Raman or IR spectroscopy.
176 gezelter 4007
177 gezelter 4028 In the sections that follow, we outline a series of coarse-grained
178     classical molecular dynamics simulations of 5CB that were done in the
179     presence of static electric fields. These simulations were then
180     coupled with both {\it ab intio} calculations of CN-deformations and
181     classical bond-length correlation functions to predict spectral
182     shifts. These predictions made should be easily varifiable with
183     scanning electrochemical microscopy experiments.
184 gezelter 4007
185     \section{Computational Details}
186 gezelter 4027 The force field used for 5CB was taken from Guo {\it et
187     al.}\cite{Zhang:2011hh} However, for most of the simulations, each
188     of the phenyl rings was treated as a rigid body to allow for larger
189     time steps and very long simulation times. The geometries of the
190     rigid bodies were taken from equilibrium bond distances and angles.
191     Although the phenyl rings were held rigid, bonds, bends, torsions and
192 gezelter 4028 inversion centers that involved atoms in these substructures (but with
193     connectivity to the rest of the molecule) were still included in the
194     potential and force calculations.
195 gezelter 4007
196 gezelter 4028 Periodic simulations cells containing 270 molecules in random
197     orientations were constructed and were locked at experimental
198     densities. Electrostatic interactions were computed using damped
199     shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
200     were equilibrated for 1~ns at a temperature of 300K. Simulations with
201     applied fields were carried out in the microcanonical (NVE) ensemble
202     with an energy corresponding to the average energy from the canonical
203     (NVT) equilibration runs. Typical applied-field runs were more than
204     60ns in length.
205 gezelter 4007
206 gezelter 4027 Static electric fields with magnitudes similar to what would be
207     available in an experimental setup were applied to the different
208     simulations. With an assumed electrode seperation of 5 nm and an
209     electrostatic potential that is limited by the voltage required to
210     split water (1.23V), the maximum realistic field that could be applied
211 gezelter 4028 is $\sim 0.024$ V/\AA. Three field environments were investigated:
212     (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
213     field = 0.024 V/\AA\ .
214 gezelter 4007
215 gezelter 4027 After the systems had come to equilibrium under the applied fields,
216 gezelter 4028 additional simulations were carried out with a flexible (Morse)
217     nitrile bond,
218     \begin{displaymath}
219     V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
220 gezelter 4036 \label{eq:morse}
221 gezelter 4028 \end{displaymath}
222 gezelter 4036 where $r_e= 1.157437$ \AA , $D_e = 212.95 \mathrm{~kcal~} /
223 gezelter 4029 \mathrm{mol}^{-1}$ and $\beta = 2.67566 $\AA~$^{-1}$. These
224 gezelter 4036 parameters correspond to a vibrational frequency of $2358
225 gezelter 4039 \mathrm{~cm}^{-1}$, somewhat higher than the experimental
226     frequency. The flexible nitrile moiety required simulation time steps
227     of 1~fs, so the additional flexibility was introducuced only after the
228     rigid systems had come to equilibrium under the applied fields.
229     Whenever time correlation functions were computed from the flexible
230     simulations, statistically-independent configurations were sampled
231     from the last ns of the induced-field runs. These configurations were
232     then equilibrated with the flexible nitrile moiety for 100 ps, and
233     time correlation functions were computed using data sampled from an
234 gezelter 4028 additional 200 ps of run time carried out in the microcanonical
235     ensemble.
236 gezelter 4027
237     \section{Field-induced Nematic Ordering}
238    
239     In order to characterize the orientational ordering of the system, the
240     primary quantity of interest is the nematic (orientational) order
241     parameter. This was determined using the tensor
242     \begin{equation}
243     Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{e}_{i
244     \alpha} \hat{e}_{i \beta} - \delta_{\alpha \beta} \right)
245     \end{equation}
246     where $\alpha, \beta = x, y, z$, and $\hat{e}_i$ is the molecular
247     end-to-end unit vector for molecule $i$. The nematic order parameter
248     $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
249     corresponding eigenvector defines the director axis for the phase.
250     $S$ takes on values close to 1 in highly ordered (smectic A) phases,
251 gezelter 4028 but falls to zero for isotropic fluids. Note that the nitrogen and
252     the terminal chain atom were used to define the vectors for each
253     molecule, so the typical order parameters are lower than if one
254     defined a vector using only the rigid core of the molecule. In
255     nematic phases, typical values for $S$ are close to 0.5.
256 gezelter 4027
257 gezelter 4029 The field-induced phase transition can be clearly seen over the course
258     of a 60 ns equilibration runs in figure \ref{fig:orderParameter}. All
259 gezelter 4027 three of the systems started in a random (isotropic) packing, with
260     order parameters near 0.2. Over the course 10 ns, the full field
261     causes an alignment of the molecules (due primarily to the interaction
262     of the nitrile group dipole with the electric field). Once this
263 gezelter 4039 system began exhibiting nematic ordering, the orientational order
264     parameter became stable for the remaining 150 ns of simulation time.
265 gezelter 4029 It is possible that the partial-field simulation is meta-stable and
266     given enough time, it would eventually find a nematic-ordered phase,
267     but the partial-field simulation was stable as an isotropic phase for
268 gezelter 4032 the full duration of a 60 ns simulation. Ellipsoidal renderings of the
269     final configurations of the runs shows that the full-field (0.024
270     V/\AA\ ) experienced a isotropic-nematic phase transition and has
271     ordered with a director axis that is parallel to the direction of the
272     applied field.
273    
274     \begin{figure}[H]
275     \includegraphics[width=\linewidth]{Figure1}
276     \caption{Evolution of the orientational order parameters for the
277 gezelter 4029 no-field, partial field, and full field simulations over the
278     course of 60 ns. Each simulation was started from a
279 gezelter 4032 statistically-independent isotropic configuration. On the right
280     are ellipsoids representing the final configurations at three
281     different field strengths: zero field (bottom), partial field
282     (middle), and full field (top)}
283 gezelter 4027 \label{fig:orderParameter}
284     \end{figure}
285    
286    
287 gezelter 4029 \section{Sampling the CN bond frequency}
288 gezelter 4027
289 gezelter 4035 The vibrational frequency of the nitrile bond in 5CB depends on
290     features of the local solvent environment of the individual molecules
291     as well as the bond's orientation relative to the applied field. The
292     primary quantity of interest for interpreting the condensed phase
293     spectrum of this vibration is the distribution of frequencies
294     exhibited by the 5CB nitrile bond under the different electric fields.
295     Three distinct methods for mapping classical simulations onto
296     vibrational spectra were brought to bear on these simulations:
297 gezelter 4029 \begin{enumerate}
298     \item Isolated 5CB molecules and their immediate surroundings were
299 gezelter 4035 extracted from the simulations. These nitrile bonds were stretched
300 gezelter 4029 and single-point {\em ab initio} calculations were used to obtain
301     Morse-oscillator fits for the local vibrational motion along that
302     bond.
303     \item The potential - frequency maps developed by Cho {\it et
304     al.}~\cite{Oh:2008fk} for nitrile moieties in water were
305     investigated. This method involves mapping the electrostatic
306     potential around the bond to the vibrational frequency, and is
307     similar in approach to field-frequency maps that were pioneered by
308 gezelter 4032 Skinner {\it et al.}\cite{XXXX}
309 gezelter 4029 \item Classical bond-length autocorrelation functions were Fourier
310     transformed to directly obtain the vibrational spectrum from
311     molecular dynamics simulations.
312     \end{enumerate}
313    
314     \subsection{CN frequencies from isolated clusters}
315 gezelter 4033 The size of the periodic condensed phase system prevented direct
316     computation of the complete library of nitrile bond frequencies using
317     {\it ab initio} methods. In order to sample the nitrile frequencies
318     present in the condensed-phase, individual molecules were selected
319     randomly to serve as the center of a local (gas phase) cluster. To
320     include steric, electrostatic, and other effects from molecules
321     located near the targeted nitrile group, portions of other molecules
322     nearest to the nitrile group were included in the quantum mechanical
323     calculations. The surrounding solvent molecules were divided into
324     ``body'' (the two phenyl rings and the nitrile bond) and ``tail'' (the
325 gezelter 4039 alkyl chain). Any molecule which had a body atom within 6~\AA\ of the
326 gezelter 4033 midpoint of the target nitrile bond had its own molecular body (the
327 gezelter 4039 4-cyano-biphenyl moiety) included in the configuration. Likewise, the
328     entire alkyl tail was included if any tail atom was within 4~\AA\ of
329     the target nitrile bond. If tail atoms (but no body atoms) were
330 gezelter 4035 included within these distances, only the tail was included as a
331     capped propane molecule.
332 gezelter 4029
333 gezelter 4033 \begin{figure}[H]
334     \includegraphics[width=\linewidth]{Figure2}
335     \caption{Cluster calculations were performed on randomly sampled 5CB
336 gezelter 4035 molecules (shown in red) from each of the simulations. Surrounding
337     molecular bodies were included if any body atoms were within 6
338     \AA\ of the target nitrile bond, and tails were included if they
339     were within 4 \AA. Included portions of these molecules are shown
340     in green. The CN bond on the target molecule was stretched and
341     compressed, and the resulting single point energies were fit to
342 gezelter 4039 Morse oscillators to obtain a distribution of frequencies.}
343 gezelter 4033 \label{fig:cluster}
344     \end{figure}
345 gezelter 4032
346 gezelter 4035 Inferred hydrogen atom locations were added to the cluster geometries,
347     and the nitrile bond was stretched from 0.87 to 1.52~\AA\ at
348     increments of 0.05~\AA. This generated 13 configurations per gas phase
349     cluster. Single-point energies were computed using the B3LYP
350     functional~\cite{Becke:1993kq,Lee:1988qf} and the 6-311++G(d,p) basis
351     set. For the cluster configurations that had been generated from
352     molecular dynamics running under applied fields, the density
353     functional calculations had a field of $5 \times 10^{-4}$ atomic units
354     ($E_h / (e a_0)$) applied in the $+z$ direction in order to match the
355     molecular dynamics simulations.
356 gezelter 4007
357 gezelter 4035 The energies for the stretched / compressed nitrile bond in each of
358 gezelter 4039 the clusters were used to fit Morse potentials, and the frequencies
359 gezelter 4035 were obtained from the $0 \rightarrow 1$ transition for the energy
360     levels for this potential.\cite{Morse:1929xy} To obtain a spectrum,
361     each of the frequencies was convoluted with a Lorentzian lineshape
362     with a width of 1.5 $\mathrm{cm}^{-1}$. Available computing resources
363     limited the sampling to 67 clusters for the zero-field spectrum, and
364     59 for the full field. Comparisons of the quantum mechanical spectrum
365     to the classical are shown in figure \ref{fig:spectrum}.
366 gezelter 4033
367 gezelter 4029 \subsection{CN frequencies from potential-frequency maps}
368 gezelter 4039
369 gezelter 4035 One approach which has been used to successfully analyze the spectrum
370     of nitrile and thiocyanate probes in aqueous environments was
371     developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} This
372     method involves finding a multi-parameter fit that maps between the
373     local electrostatic potential at selected sites surrounding the
374     nitrile bond and the vibrational frequency of that bond obtained from
375     more expensive {\it ab initio} methods. This approach is similar in
376     character to the field-frequency maps developed by Skinner {\it et
377     al.} for OH stretches in liquid water.\cite{XXXX}
378    
379     To use the potential-frequency maps, the local electrostatic
380 gezelter 4039 potential, $\phi_a$, is computed at 20 sites ($a = 1 \rightarrow 20$)
381 gezelter 4035 that surround the nitrile bond,
382 gezelter 4029 \begin{equation}
383 gezelter 4035 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{j}
384     \frac{q_j}{\left|r_{aj}\right|}.
385 gezelter 4029 \end{equation}
386 gezelter 4036 Here $q_j$ is the partial site on atom $j$ (residing on a different
387     molecule) and $r_{aj}$ is the distance between site $a$ and atom $j$.
388     The original map was parameterized in liquid water and comprises a set
389     of parameters, $l_a$, that predict the shift in nitrile peak
390     frequency,
391 gezelter 4029 \begin{equation}
392 gezelter 4036 \delta\tilde{\nu} =\sum^{20}_{a=1} l_{a}\phi_{a}.
393 gezelter 4029 \end{equation}
394 gezelter 4035
395 gezelter 4039 The simulations of 5CB were carried out in the presence of
396 gezelter 4036 externally-applied uniform electric fields. Although uniform fields
397     exert forces on charge sites, they only contribute to the potential if
398     one defines a reference point that can serve as an origin. One simple
399 gezelter 4039 modification to the potential at each of the probe sites is to use the
400 gezelter 4036 centroid of the \ce{CN} bond as the origin for that site,
401 gezelter 4029 \begin{equation}
402 gezelter 4036 \phi_a^\prime = \phi_a + \frac{1}{4\pi\epsilon_{0}} \vec{E} \cdot
403     \left(\vec{r}_a - \vec{r}_\ce{CN} \right)
404 gezelter 4029 \end{equation}
405 gezelter 4036 where $\vec{E}$ is the uniform electric field, $\left( \vec{r}_{a} -
406     \vec{r}_\ce{CN} \right)$ is the displacement between the
407     cooridinates described by Choi {\it et
408     al.}~\cite{Choi:2008cr,Oh:2008fk} and the \ce{CN} bond centroid.
409     $\phi_a^\prime$ then contains an effective potential contributed by
410     the uniform field in addition to the local potential contributions
411     from other molecules.
412 gezelter 4029
413 gezelter 4039 The sites $\{\vec{r}_a\}$ and weights $\left\{l_a \right\}$
414     developed by Choi {\it et al.}~\cite{Choi:2008cr,Oh:2008fk} are quite
415     symmetric around the \ce{CN} centroid, and even at large uniform field
416     values we observed nearly-complete cancellation of the potenial
417     contributions from the uniform field. In order to utilize the
418     potential-frequency maps for this problem, one would therefore need
419     extensive reparameterization of the maps to include explicit
420     contributions from the external field. This reparameterization is
421     outside the scope of the current work, but would make a useful
422     addition to the potential-frequency map approach.
423 gezelter 4029
424     \subsection{CN frequencies from bond length autocorrelation functions}
425    
426 gezelter 4039 The distribution of nitrile vibrational frequencies can also be found
427 gezelter 4036 using classical time correlation functions. This was done by
428     replacing the rigid \ce{CN} bond with a flexible Morse oscillator
429     described in Eq. \ref{eq:morse}. Since the systems were perturbed by
430     the addition of a flexible high-frequency bond, they were allowed to
431     re-equilibrate in the canonical (NVT) ensemble for 100 ps with 1 fs
432     timesteps. After equilibration, each configuration was run in the
433     microcanonical (NVE) ensemble for 20 ps. Configurations sampled every
434     fs were then used to compute bond-length autocorrelation functions,
435 gezelter 4007 \begin{equation}
436 gezelter 4036 C(t) = \langle \delta r(t) \cdot \delta r(0) ) \rangle
437 gezelter 4007 \end{equation}
438     %
439 gezelter 4036 where $\delta r(t) = r(t) - r_0$ is the deviation from the equilibrium
440     bond distance at time $t$. Ten statistically-independent correlation
441     functions were obtained by allowing the systems to run 10 ns with
442     rigid \ce{CN} bonds followed by 120 ps equilibration and data
443 gezelter 4039 collection using the flexible \ce{CN} bonds, and repeating this
444     process. The total sampling time, from sample preparation to final
445     configurations, exceeded 150 ns for each of the field strengths
446     investigated.
447 gezelter 4007
448 gezelter 4036 The correlation functions were filtered using exponential apodization
449 gezelter 4039 functions,\cite{FILLER:1964yg} $f(t) = e^{-c |t|}$, with a time
450     constant, $c =$ 6 ps, and were Fourier transformed to yield a
451     spectrum,
452 gezelter 4036 \begin{equation}
453     I(\omega) = \int_{-\infty}^{\infty} C(t) f(t) e^{-i \omega t} dt.
454     \end{equation}
455     The sample-averaged classical nitrile spectrum can be seen in Figure
456     \ref{fig:spectra}. Note that the Morse oscillator parameters listed
457 gezelter 4039 above yield a natural frequency of 2358 $\mathrm{cm}^{-1}$, somewhat
458     higher than the experimental peak near 2226 $\mathrm{cm}^{-1}$. This
459     shift does not effect the ability to qualitatively compare peaks from
460     the classical and quantum mechanical approaches, so the classical
461     spectra are shown as a shift relative to the natural oscillation of
462     the Morse bond.
463 gezelter 4007
464 jmarr 4013 \begin{figure}
465 gezelter 4036 \includegraphics[width=3.25in]{Convolved}
466 jmarr 4013 \includegraphics[width=3.25in]{2Spectra}
467 gezelter 4039 \caption{Quantum mechanical nitrile spectrum for the no-field simulation
468     (black) and the full field simulation (red). The lower panel
469     shows the corresponding classical bond-length autocorrelation
470     spectrum for the flexible nitrile measured relative to the natural
471     frequency for the flexible bond.}
472 gezelter 4036 \label{fig:spectra}
473 jmarr 4013 \end{figure}
474 jmarr 4020
475 gezelter 4036 Note that due to electrostatic interactions, the classical approach
476     implicitly couples \ce{CN} vibrations to the same vibrational mode on
477     other nearby molecules. This coupling is not handled in the {\it ab
478     initio} cluster approach.
479 jmarr 4020
480 gezelter 4036 \section{Discussion}
481    
482 jmarr 4023
483 gezelter 4039 Observation of Field-induced nematic ordering
484     Ordering corresponds to shift of a portion of the nitrile spectrum to
485     the red.
486     At the same time, the system exhibits an increase in aligned and anti-a
487    
488    
489    
490 jmarr 4020 Since no explicit dependence was observed between the calculated
491     frequency and the electric field, it was not a viable route for the
492     calculation of a nitrile spectrum. Instead, the frequencies were taken
493 jmarr 4024 and convolved together with a lorentzian line shape applied around the
494 gezelter 4036 frequency value. These spectra are seen below in Figure 4. While the
495     spectrum without a field is lower in intensity and is almost bimodel
496     in distrobution, the external field spectrum is much more
497     unimodel. This tighter clustering has the affect of increasing the
498 jmarr 4020 intensity around 2226 cm\textsuperscript{-1} where the peak is
499 jmarr 4023 centered. The external field also has fewer frequencies of higher
500 gezelter 4036 energy in the spectrum. Unlike the the zero field, where some
501     frequencies reach as high as 2280 cm\textsuperscript{-1}.
502    
503 jmarr 4024 Interestingly, the field that is needed to switch the phase of 5CB
504     macroscopically is larger than 1 V. However, in this case, only a
505     voltage of 1.2 V was need to induce a phase change. This is impart due
506 gezelter 4036 to the short distance of 5 nm the field is being applied across. At
507     such a small distance, the field is much larger than the macroscopic
508     and thus easily induces a field dependent phase change. However, this
509     field will not cause a breakdown of the 5CB since electrochemistry
510     studies have shown that it can be used in the presence of fields as
511     high as 500 V macroscopically. This large of a field near the surface
512     of the elctrode would cause breakdown of 5CB if it could happen.
513 jmarr 4024
514 jmarr 4020 The absence of any electric field dependency of the freuquency with
515 jmarr 4025 the Gaussian simulations is interesting. A large base of research has been
516 jmarr 4024 built upon the known tuning of the nitrile bond as the local field
517     changes. This difference may be due to the absence of water or a
518     molecule that induces a large internal field. Liquid water is known to have a very high internal field which
519 jmarr 4020 is much larger than the internal fields of neat 5CB. Even though the
520 jmarr 4024 application of Gaussian simulations followed by mapping it to
521 jmarr 4020 some classical parameter is easy and straight forward, this system
522     illistrates how that 'go to' method can break down.
523 gezelter 4007
524 jmarr 4020 While this makes the application of nitrile Stark effects in
525 jmarr 4024 simulations without water harder, these data show
526 jmarr 4021 that it is not a deal breaker. The classically calculated nitrile
527     spectrum shows changes in the spectra that will be easily seen through
528     experimental routes. It indicates a shifted peak lower in energy
529 jmarr 4024 should arise. This peak is a few wavenumbers from the leading edge of
530     the larger peak and almost 75 wavenumbers from the center. This
531     seperation between the two peaks means experimental results will show
532     an easily resolved peak.
533 jmarr 4021
534 jmarr 4024 The Gaussian derived spectra do indicate an applied field
535 jmarr 4023 and subsiquent phase change does cause a narrowing of freuency
536 jmarr 4025 distrobution. With narrowing, it would indicate an increased
537     homogeneous distrobution of the local field near the nitrile.
538 gezelter 4039
539    
540     The angle-dependent pair distribution function,
541     \begin{equation}
542     g(r, \cos \omega) = \frac{1}{\rho N} \left< \sum_{i}
543     \sum_{j} \delta \left(r - r_{ij}\right) \delta\left(\cos \omega_{ij} - \cos \omega\right) \right>
544     \end{equation}
545     provides information about the spatial and angular correlations in the
546     system. The angle $\omega$ is defined by vectors along the CN axis of
547     each nitrile bond (see figure \ref{fig:definition}).
548    
549     \begin{figure}
550     \includegraphics[width=\linewidth]{definition}
551     \caption{Definitions of the angles between two nitrile bonds. All
552     pairs of CN bonds in the simulation have three angles ($\theta_i$,
553     $\theta_j$ and $\omega$). $\cos\omega$ values range from -1
554     (anti-aligned) to +1 for aligned nitrile bonds.}
555     \label{fig:definition}
556     \end{figure}
557    
558     In figure \ref{fig:gofromega} the effects of the field-induced phase
559     transition are clear. The nematic ordering transfers population from
560     the perpendicular or unaligned region in the center of the plot to the
561     nitrile-alinged peak near $\cos\omega = 1$. Most other features are
562     undisturbed. This increased population of aligned nitrile bonds in
563     the close solvation shells is also the population that contributes
564     most heavily to the low-frequency peaks in the vibrational spectrum.
565    
566     \begin{figure}
567     \includegraphics[width=\linewidth]{Figure4}
568     \caption{Contours of the angle-dependent pair distribution functions
569     for nitrile bonds on 5CB in the zero-field (upper panel) and full
570     field (lower panel) simulations. Dark areas signify regions of
571     enhanced density, while light areas signify depletion relative to
572     the bulk density.}
573     \label{fig:gofromega}
574     \end{figure}
575    
576    
577 gezelter 4007 \section{Conclusions}
578 jmarr 4024 Field dependent changes
579 gezelter 4036
580     \section{Acknowledgements}
581     The authors thank Steven Corcelli for helpful comments and
582     suggestions. Support for this project was provided by the National
583     Science Foundation under grant CHE-0848243. Computational time was
584     provided by the Center for Research Computing (CRC) at the University
585     of Notre Dame.
586    
587 gezelter 4007 \newpage
588    
589     \bibliography{5CB}
590    
591     \end{doublespace}
592     \end{document}