ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
(Generate patch)

Comparing trunk/5cb/5CB.tex (file contents):
Revision 4018 by jmarr, Mon Feb 3 21:46:13 2014 UTC vs.
Revision 4033 by gezelter, Wed Feb 19 19:48:47 2014 UTC

# Line 40 | Line 40
40  
41  
42   \title{Nitrile vibrations as reporters of field-induced phase
43 <  transitions in liquid crystals}  
43 >  transitions in 4-cyano-4'-pentylbiphenyl (5CB)}  
44   \author{James M. Marr}
45   \author{J. Daniel Gezelter}
46   \email{gezelter@nd.edu}
# Line 58 | Line 58 | Notre Dame, Indiana 46556}
58   \begin{doublespace}
59  
60   \begin{abstract}
61 <  The behavior of the spectral lineshape of the nitrile group in
62 <  4-Cyano-4'-pentylbiphenyl (5CB) in response to an applied electric
63 <  field has been simulated using both classical molecular dynamics
64 <  simulations and {\it ab initio} quantum mechanical calculations of
65 <  liquid-like clusters.  This nitrile group is a well-known reporter
66 <  of local field effects in other condensed phase settings, and our
67 <  simulations suggest that there is a significant response when 5CB
68 <  liquids are exposed to a relatively large external field.  However,
69 <  this response is due largely to the field-induced phase transition.
70 <  We observe a peak shift to the red of nearly 40
71 <  cm\textsuperscript{-1}. These results indicate that applied fields
72 <  can play a role in the observed peak shape and position even if
73 <  those fields are significantly weaker than the local electric fields
74 <  that are normally felt within polar liquids.
61 >  4-cyano-4'-pentylbiphenyl (5CB) is a liquid-crystal-forming compound
62 >  with a terminal nitrile group aligned with the long axis of the
63 >  molecule.  Simulations of condensed-phase 5CB were carried out both
64 >  with and without applied electric fields to provide an understanding
65 >  of the 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   \end{abstract}
77  
78   \newpage
79  
80   \section{Introduction}
81  
82 + 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 + spectrum can be quantified and mapped into a field value that is
88 + impinging upon the nitrile bond. This has been used extensively in
89 + biological systems like proteins and
90 + enzymes.\cite{Tucker:2004qq,Webb:2008kn}
91 +
92 + The response of nitrile groups to electric fields has now been
93 + investigated for a number of small molecules,\cite{Andrews:2000qv} as
94 + well as in biochemical settings, where nitrile groups can act as
95 + minimally invasive probes of structure and
96 + dynamics.\cite{Lindquist:2009fk,Fafarman:2010dq} The vibrational Stark
97 + effect has also been used to study the effects of electric fields on
98 + nitrile-containing self-assembled monolayers at metallic
99 + interfaces.\cite{Oklejas:2002uq,Schkolnik:2012ty}
100 +
101 + Recently 4-cyano-4'-pentylbiphenyl (5CB), a liquid crystalline
102 + molecule with a terminal nitrile group, has seen renewed interest as
103 + one way to impart order on the surfactant interfaces of
104 + nanodroplets,\cite{Moreno-Razo:2012rz} or to drive surface-ordering
105 + that can be used to promote particular kinds of
106 + self-assembly.\cite{PhysRevLett.111.227801} The nitrile group in 5CB
107 + is a particularly interesting case for studying electric field
108 + effects, as 5CB exhibits an isotropic to nematic phase transition that
109 + can be triggered by the application of an external field near room
110 + temperature.\cite{Gray:1973ca,Hatta:1991ee} This presents the
111 + possiblity that the field-induced changes in the local environment
112 + could have dramatic effects on the vibrations of this particular CN
113 + bond.  Although the infrared spectroscopy of 5CB has been
114 + well-investigated, particularly as a measure of the kinetics of the
115 + phase transition,\cite{Leyte:1997zl} the 5CB nitrile group has not yet
116 + seen the detailed theoretical treatment that biologically-relevant
117 + small molecules have
118 + received.\cite{Lindquist:2008bh,Lindquist:2008qf,Oh:2008fk,Choi:2008cr,Waegele:2010ve}
119 +
120   The fundamental characteristic of liquid crystal mesophases is that
121   they maintain some degree of orientational order while translational
122   order is limited or absent. This orientational order produces a
123   complex direction-dependent response to external perturbations like
124 < electric fields and mechanical distortions.  The anisotropy of the
124 > electric fields and mechanical distortions. The anisotropy of the
125   macroscopic phases originates in the anisotropy of the constituent
126   molecules, which typically have highly non-spherical structures with a
127 < significant degree of internal rigidity.  In nematic phases, rod-like
127 > significant degree of internal rigidity. In nematic phases, rod-like
128   molecules are orientationally ordered with isotropic distributions of
129 < molecular centers of mass, while in smectic phases, the molecules
130 < arrange themselves into layers with their long (symmetry) axis normal
131 < ($S_{A}$) or tilted ($S_{C}$) with respect to the layer planes.
129 > molecular centers of mass. For example, 5CB has a solid to nematic
130 > phase transition at 18C and a nematic to isotropic transition at
131 > 35C.\cite{Gray:1973ca}
132  
133 < The behavior of the $S_{A}$ phase can be partially explained with
134 < models mainly based on geometric factors and van der Waals
135 < interactions.  However, these simple models are insufficient to
136 < describe liquid crystal phases which exhibit more complex polymorphic
137 < nature.  X-ray diffraction studies have shown that the ratio between
138 < lamellar spacing ($s$) and molecular length ($l$) can take on a wide
139 < range of values.\cite{Gray:1984hc,Leadbetter:1976vf,Hardouin:1980yq}
140 < Typical $S_{A}$ phases have $s/l$ ratios on the order of $0.8$, while
141 < for some compounds, e.g. the 4-alkyl-4'-cyanobiphenyls, the $s/l$
142 < ratio is on the order of $1.4$.  Molecules which form $S_{A}$ phases
143 < can exhibit a wide variety of subphases like monolayers ($S_{A1}$),
144 < uniform bilayers ($S_{A2}$), partial bilayers ($S_{\tilde A}$) as well
145 < as interdigitated bilayers ($S_{A_{d}}$), and often have a terminal
146 < cyano or nitro group.  In particular lyotropic liquid crystals (those
147 < exhibiting liquid crystal phase transition as a function of water
148 < concentration) often have polar head groups or zwitterionic charge
149 < separated groups that result in strong dipolar
150 < interactions.\cite{Collings97} Because of their versatile polymorphic
151 < nature, polar liquid crystalline materials have important
152 < technological applications in addition to their immense relevance to
114 < biological systems.\cite{Collings97}
133 > In smectic phases, the molecules arrange themselves into layers with
134 > their long (symmetry) axis normal ($S_{A}$) or tilted ($S_{C}$) with
135 > respect to the layer planes. The behavior of the $S_{A}$ phase can be
136 > partially explained with models mainly based on geometric factors and
137 > van der Waals interactions. The Gay-Berne potential, in particular,
138 > has been widely used in the liquid crystal community to describe this
139 > anisotropic phase
140 > behavior.~\cite{Gay:1981yu,Berne72,Kushick:1976xy,Luckhurst90,Cleaver:1996rt}
141 > However, these simple models are insufficient to describe liquid
142 > crystal phases which exhibit more complex polymorphic nature.
143 > Molecules which form $S_{A}$ phases can exhibit a wide variety of
144 > subphases like monolayers ($S_{A1}$), uniform bilayers ($S_{A2}$),
145 > partial bilayers ($S_{\tilde A}$) as well as interdigitated bilayers
146 > ($S_{A_{d}}$), and often have a terminal cyano or nitro group.  In
147 > particular, lyotropic liquid crystals (those exhibiting liquid crystal
148 > phase transition as a function of water concentration), often have
149 > polar head groups or zwitterionic charge separated groups that result
150 > in strong dipolar interactions,\cite{Collings:1997rz} and terminal cyano
151 > groups (like the one in 5CB) can induce permanent longitudinal
152 > dipoles.\cite{Levelut:1981eu}
153  
154 < Experimental studies by Levelut {\it et al.}~\cite{Levelut:1981eu}
155 < revealed that terminal cyano or nitro groups usually induce permanent
156 < longitudinal dipole moments on the molecules.
154 > Macroscopic electric fields applied using electrodes on opposing sides
155 > of a sample of 5CB have demonstrated the phase change of the molecule
156 > as a function of electric field.\cite{Lim:2006xq} These previous
157 > studies have shown the nitrile group serves as an excellent indicator
158 > of the molecular orientation within the applied field. Lee {\it et
159 >  al.}~showed a 180 degree change in field direction could be probed
160 > with the nitrile peak intensity as it changed along with molecular
161 > alignment in the field.\cite{Lee:2006qd,Leyte:1997zl}
162  
163 < Liquid crystalline materials with dipole moments located at the ends
121 < of the molecules have important applications in display technologies
122 < in addition to their relevance in biological systems.\cite{LCapp}
123 <
124 < Many of the technological applications of the lyotropic mesogens
125 < manipulate the orientation and structuring of the liquid crystal
126 < through application of local electric fields.\cite{?}
127 < Macroscopically, this restructuring is visible in the interactions the
128 < bulk phase has with scattered or transmitted light.\cite{?}  
129 <
130 < 4-Cyano-4'-pentylbiphenyl (5CB), has been a model for field-induced
131 < phase changes due to the known electric field response of the liquid
132 < crystal\cite{Hatta:1991ee}.  It was discovered (along with three
133 < similar compounds) in 1973 in an effort to find a LC that had a
134 < melting point near room temperature.\cite{Gray:1973ca} It's known to
135 < have a crystalline to nematic phase transition at 18 C and a nematic
136 < to isotropic transition at 35 C.\cite{Gray:1973ca}
137 <
138 < Nitrile groups can serve as very precise electric field reporters via
139 < their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
140 < triple bond between the nitrogen and the carbon atom is very sensitive
141 < to local field changes and is observed to have a direct impact on the
142 < peak position within the spectrum.  The Stark shift in the spectrum
143 < can be quantified and mapped into a field value that is impinging upon
144 < the nitrile bond. This has been used extensively in biological systems
145 < like proteins and enzymes.\cite{Tucker:2004qq,Webb:2008kn}
146 <
147 < To date, the nitrile electric field response of
148 < 4-Cyano-4'-n-alkylbiphenyl liquid crystals has not been investigated.
149 < While macroscopic electric fields applied across macro electrodes show
150 < the phase change of the molecule as a function of electric
151 < field,\cite{Lim:2006xq} the effect of a microscopic field application
152 < has not been probed. These previous studies have shown the nitrile
153 < group serves as an excellent indicator of the molecular orientation
154 < within the field applied. Blank showed the 180 degree change in field
155 < direction could be probed with the nitrile peak intensity as it
156 < decreased and increased with molecule alignment in the
157 < field.\cite{Lee:2006qd,Leyte:97}
158 <
159 < While these macroscopic fields worked well at showing the bulk
163 > While these macroscopic fields work well at indicating the bulk
164   response, the atomic scale response has not been studied. With the
165   advent of nano-electrodes and coupling them with atomic force
166   microscopy, control of electric fields applied across nanometer
167 < distances is now possible\cite{citation1}. This application of
168 < nanometer length is interesting in the case of a nitrile group on the
169 < molecule. While macroscopic fields are insufficient to cause a Stark
170 < effect, small fields across nanometer-sized gaps are of sufficient
171 < strength. If one were to assume a gap of 5 nm between a lower
172 < electrode having a nanoelectrode placed near it via an atomic force
173 < microscope, a field of 1 V applied across the electrodes would
174 < translate into a field of 2x10\textsuperscript{8} $\frac{V}{M}$. This
175 < field is theoretically strong enough to cause a phase change from
176 < isotropic to nematic, as well as Stark tuning of the nitrile
173 < bond. This should be readily visible experimentally through Raman or
174 < IR spectroscopy.
167 > distances is now possible.\cite{citation1} While macroscopic fields
168 > are insufficient to cause a Stark effect without dielectric breakdown
169 > of the material, small fields across nanometer-sized gaps may be of
170 > sufficient strength.  For a gap of 5 nm between a lower electrode
171 > having a nanoelectrode placed near it via an atomic force microscope,
172 > a potential of 1 V applied across the electrodes is equivalent to a
173 > field of 2x10\textsuperscript{8} $\frac{V}{M}$.  This field is
174 > certainly strong enough to cause the isotropic-nematic phase change
175 > and as well as Stark tuning of the nitrile bond.  This should be
176 > readily visible experimentally through Raman or IR spectroscopy.
177  
178 < Herein, we show the computational investigation of these electric field effects through atomistic simulations. These are then coupled with ab intio and classical spectrum calculations to predict changes experiments should be able to replicate.
178 > In the sections that follow, we outline a series of coarse-grained
179 > classical molecular dynamics simulations of 5CB that were done in the
180 > presence of static electric fields. These simulations were then
181 > coupled with both {\it ab intio} calculations of CN-deformations and
182 > classical bond-length correlation functions to predict spectral
183 > shifts. These predictions made should be easily varifiable with
184 > scanning electrochemical microscopy experiments.
185  
186   \section{Computational Details}
187 < The force field was mainly taken from Guo et al.\cite{Zhang:2011hh} A
188 < deviation from this force field was made to create a rigid body from
189 < the phenyl rings. Bond distances within the rigid body were taken from
190 < equilibrium bond distances. While the phenyl rings were held rigid,
191 < bonds, bends, torsions and inversion centers still included the rings.
187 > The force field used for 5CB was taken from Guo {\it et
188 >  al.}\cite{Zhang:2011hh} However, for most of the simulations, each
189 > of the phenyl rings was treated as a rigid body to allow for larger
190 > time steps and very long simulation times.  The geometries of the
191 > rigid bodies were taken from equilibrium bond distances and angles.
192 > Although the phenyl rings were held rigid, bonds, bends, torsions and
193 > inversion centers that involved atoms in these substructures (but with
194 > connectivity to the rest of the molecule) were still included in the
195 > potential and force calculations.
196  
197 < Simulations were with boxes of 270 molecules locked at experimental
198 < densities with periodic boundaries. The molecules were thermalized
199 < from 0 kelvin to 300 kelvin. To equilibrate, each was first run in NVT
200 < for 1 ns. This was followed by NVE for simulations used in the data
201 < collection.
197 > Periodic simulations cells containing 270 molecules in random
198 > orientations were constructed and were locked at experimental
199 > densities.  Electrostatic interactions were computed using damped
200 > shifted force (DSF) electrostatics.\cite{Fennell:2006zl} The molecules
201 > were equilibrated for 1~ns at a temperature of 300K.  Simulations with
202 > applied fields were carried out in the microcanonical (NVE) ensemble
203 > with an energy corresponding to the average energy from the canonical
204 > (NVT) equilibration runs.  Typical applied-field runs were more than
205 > 60ns in length.
206  
207 < External electric fields were applied in a simplistic charge-field
208 < interaction. Forces were calculated by multiplying the electric field
209 < being applied by the charge at each atom. For the potential, the
210 < origin of the box was used as a point of reference. This allows for a
211 < potential value to be added to each atom as the molecule move in space
212 < within the box. Fields values were applied in a manner representing
213 < those that would be applable in an experimental set-up. The assumed
214 < electrode seperation was 5 nm and the field was input as
199 < $\frac{V}{\text{\AA}}$. The three field environments were, 1) no field
200 <  applied, 2) 0.01 $\frac{V}{\text{\AA}}$ (0.5 V) and 3) 0.024
201 <    $\frac{V}{\text{\AA}}$ (1.2 V). Each field was applied in the
202 <    Z-axis of the simulation box. For the simplicity of this paper,
203 <    each field will be called zero, partial and full, respectively.
207 > Static electric fields with magnitudes similar to what would be
208 > available in an experimental setup were applied to the different
209 > simulations.  With an assumed electrode seperation of 5 nm and an
210 > electrostatic potential that is limited by the voltage required to
211 > split water (1.23V), the maximum realistic field that could be applied
212 > is $\sim 0.024$ V/\AA.  Three field environments were investigated:
213 > (1) no field applied, (2) partial field = 0.01 V/\AA\ , and (3) full
214 > field = 0.024 V/\AA\ .
215  
216 < For quantum calculation of the nitrile bond frequency, Gaussian 09 was
217 < used. A single 5CB molecule was selected for the center of the
218 < cluster. For effects from molecules located near the chosen nitrile
219 < group, parts of molecules nearest to the nitrile group were
220 < included. For the body not including the tail, any atom within 6~\AA
221 < of the midpoint of the nitrile group was included. For the tail
222 < structure, the whole tail was included if a tail atom was within 4~\AA
223 < of the midpoint. If the tail did not include any atoms from the ring
224 < structure, it was considered a propane molecule and included as
225 < such. Once the clusters were generated, input files were created that
226 < stretched the nitrile bond along its axis from 0.87 to 1.52~\AA at
227 < increments of 0.05~\AA. This generated 13 single point energies to be
228 < calculated. The Gaussian files were run with B3LYP/6-311++G(d,p) with
229 < no other keywords for the zero field simulation. Simulations with
230 < fields applied included the keyword ''Field=Z+5'' to match the
231 < external field applied in molecular dynamic runs. Once completed, the central nitrile bond frequency
232 < was calculated with a Morse fit. Using this fit and the solved energy
233 < levels for a Morse oscillator, the frequency was found. Each set of
234 < frequencies were then convolved together with a guassian spread
235 < function over each value. The width value used was 1.5. For the zero
236 < field spectrum, 67 frequencies were used and for the full field, 59
237 < frequencies were used.
216 > After the systems had come to equilibrium under the applied fields,
217 > additional simulations were carried out with a flexible (Morse)
218 > nitrile bond,
219 > \begin{displaymath}
220 > V(r_\ce{CN}) = D_e \left(1 - e^{-\beta (r_\ce{CN}-r_e)}\right)^2
221 > \end{displaymath}
222 > where $r_e= 1.157437$ \AA , $D_e = 212.95 \mathrm{~kca~l} /
223 > \mathrm{mol}^{-1}$ and $\beta = 2.67566 $\AA~$^{-1}$.  These
224 > parameters correspond to a vibrational frequency of $2375
225 > \mathrm{~cm}^{-1}$, a bit higher than the experimental frequency. The
226 > flexible nitrile moiety required simulation time steps of 1~fs, so the
227 > additional flexibility was introducuced only after the rigid systems
228 > had come to equilibrium under the applied fields.  Whenever time
229 > correlation functions were computed from the flexible simulations,
230 > statistically-independent configurations were sampled from the last ns
231 > of the induced-field runs.  These configurations were then
232 > equilibrated with the flexible nitrile moiety for 100 ps, and time
233 > correlation functions were computed using data sampled from an
234 > additional 200 ps of run time carried out in the microcanonical
235 > ensemble.
236 >
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 > 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  
257 + 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 + 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 + system started exhibiting nematic ordering, the orientational order
264 + parameter became stable for the remaining 50 ns of simulation time.
265 + 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 + 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 +    no-field, partial field, and full field simulations over the
278 +    course of 60 ns. Each simulation was started from a
279 +    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 +  \label{fig:orderParameter}
284 + \end{figure}
285 +
286 +
287 + \section{Sampling the CN bond frequency}
288 +
289 + The vibrational frequency of the nitrile bond in 5CB is assumed to
290 + depend on features of the local solvent environment of the individual
291 + molecules as well as the bond's orientation relative to the applied
292 + field.  Therefore, the primary quantity of interest is the
293 + distribution of vibrational frequencies exhibited by the 5CB nitrile
294 + bond under the different electric fields.  Three distinct methods for
295 + mapping classical simulations onto vibrational spectra were brought to
296 + bear on these simulations:
297 + \begin{enumerate}
298 + \item Isolated 5CB molecules and their immediate surroundings were
299 +  extracted from the simulations, their nitrile bonds were stretched
300 +  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 +  Skinner {\it et al.}\cite{XXXX}
309 + \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 + 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 + alkyl chain).  Any molecule which had a body atom within 6~\AA of the
326 + midpoint of the target nitrile bond had its own molecular body (the
327 + 4-cyano-4'-pentylbiphenyl moiety) included in the configuration.  For
328 + the alkyl tail, the entire tail was included if any tail atom was
329 + within 4~\AA of the target nitrile bond.  If tail atoms (but no body
330 + atoms) were included within these distances, only the tail was
331 + included as a capped propane molecule.  
332 +
333 + \begin{figure}[H]
334 +  \includegraphics[width=\linewidth]{Figure2}
335 +  \caption{Cluster calculations were performed on randomly sampled 5CB
336 +    molecules from each of the simualtions. Surrounding molecular
337 +    bodies were included if any body atoms were within 6 \AA\ of the
338 +    target nitrile bond, and tails were included if they were within 4
339 +    \AA.  The CN bond on the target molecule was stretched and
340 +    compressed (left), and the resulting single point energies were
341 +    fit to Morse oscillators to obtain frequency distributions.}
342 +  \label{fig:cluster}
343 + \end{figure}
344 +
345 + Inferred hydrogen atom locations were generated, and cluster
346 + geometries were created that stretched the nitrile bond along from
347 + 0.87 to 1.52~\AA at increments of 0.05~\AA. This generated 13 single
348 + point energies to be calculated per gas phase cluster. Energies were
349 + computed with the B3LYP functional and 6-311++G(d,p) basis set.  For
350 + the cluster configurations that had been generated with applied
351 + fields, a field strength of 5 atomic units in the $z$ direction was
352 + applied to match the molecular dynamics runs.
353 +
354 + The relative energies for the stretched and compressed nitrile bond
355 + were used to fit a Morse oscillator, and the frequencies were obtained
356 + from the $0 \rightarrow 1$ transition for the exact energies. To
357 + obtain a spectrum, each of the frequencies was convoluted with a
358 + Lorentzian lineshape with a width of 1.5 $\mathrm{cm}^{-1}$.  Our
359 + available computing resources limited us to 67 clusters for the
360 + zero-field spectrum, and 59 for the full field.
361 +
362 + \subsection{CN frequencies from potential-frequency maps}
363 + Before Gaussian silumations were carried out, it was attempt to apply
364 + the method developed by Cho  {\it et al.}~\cite{Oh:2008fk} This method involves the fitting
365 + of multiple parameters to Gaussian calculated freuencies to find a
366 + correlation between the potential around the bond and the
367 + frequency. This is very similar to work done by Skinner  {\it et al.}~with
368 + water models like SPC/E. The general method is to find the shift in
369 + the peak position through,
370 + \begin{equation}
371 + \delta\tilde{\nu} =\sum^{n}_{a=1} l_{a}\phi^{water}_{a}
372 + \end{equation}
373 + where $l_{a}$ are the fitting parameters and $\phi^{water}_{a}$ is the
374 + potential from the surrounding water cluster. This $\phi^{water}_{a}$
375 + takes the form,
376 + \begin{equation}
377 + \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{m} \sum_{j}
378 + \frac{C^{H_{2}O}_{j \left(m \right) }}{r_{aj \left(m\right)}}
379 + \end{equation}
380 + where $C^{H_{2}O}_{j \left(m \right) }$ indicates the partial charge
381 + on the $j$th site of the $m$th water molecule and $r_{aj \left(m\right)}$
382 + is the distance between the site $a$ of the nitrile molecule and the $j$th
383 + site of the $m$th water molecule. However, since these simulations
384 + are done under the presence of external fields and in the
385 + absence of water, the equations need a correction factor for the shift
386 + caused by the external field. The equation is also reworked to use
387 + electric field site data instead of partial charges from surrounding
388 + atoms. So by modifing the original
389 + $\phi^{water}_{a}$ to $\phi^{5CB}_{a}$ we get,
390 + \begin{equation}
391 + \phi^{5CB}_{a} = \frac{1}{4\pi \epsilon_{0}} \left( \vec{E}\bullet
392 +  \left(\vec{r}_{a}-\vec{r}_{CN}\right) \right) + \phi^{5CB}_{0}
393 + \end{equation}
394 + where $\vec{E}$ is the electric field at each atom, $\left( \vec{r}_{a} -
395 +  \vec{r}_{CN} \right)$ is the vector between the nitrile bond and the
396 + cooridinates described by Cho around the bond and $\phi^{5CB}_{0}$ is
397 + the correction factor for the system of parameters. After these
398 + changes, the correction factor was found for multiple values of an
399 + external field being applied. However, the factor was no linear and
400 + was overly large due to the fitting parameters being so small.
401 +
402 +
403 + \subsection{CN frequencies from bond length autocorrelation functions}
404 +
405   Classical nitrile bond frequencies were found by replacing the rigid
406   cyanide bond with a flexible Morse oscillator bond
407   ($r_0= 1.157437$ \AA , $D_0 = 212.95$ and
# Line 241 | Line 418 | decay with a $\tau$ of 3000 (have to check this). Furt
418   where $r_0$ is the equilibrium bond distance and $r(t)$ is the
419   instantaneous bond displacement at time $t$. Once calculated,
420   smoothing was applied by adding an exponential decay on top of the
421 < decay with a $\tau$ of 3000 (have to check this). Further smoothing
421 > decay with a $\tau$ of 6000. Further smoothing
422   was applied by padding 20,000 zeros on each side of the symmetric
423   data. This was done five times by allowing the systems to run 1 ns
424   with a rigid bond followed by an equilibrium run with the bond
425 < switched back on and the short production run.
425 > switched back to a Morse oscillator and a short production run of 20 ps.
426  
250 \section{Results}
427  
252 In order to characterize the orientational ordering of the system, the
253 primary quantity of interest is the nematic (orientational) order
254 parameter. This is determined using the tensor
255 \begin{equation}
256 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{e}_{i
257    \alpha} \hat{e}_{i \beta} - \delta_{\alpha \beta} \right)
258 \end{equation}
259 where $\alpha, \beta = x, y, z$, and $\hat{e}_i$ is the molecular
260 end-to-end unit vector for molecule $i$. The nematic order parameter
261 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
262 corresponding eigenvector defines the director axis for the phase.
263 $S$ takes on values close to 1 in highly ordered phases, but falls to
264 zero for isotropic fluids. In the context of 5CB, this value would be
265 close to zero for its isotropic phase and raise closer to one as it
266 moved to the nematic and crystalline phases.
267
268 This value indicates phases changes at temperature boundaries but also
269 when a phase changes occurs due to external field applications. In
270 Figure 1, this phase change can be clearly seen over the course of 60
271 ns. Each system starts with an ordering paramter near 0.1 to 0.2,
272 which is an isotropic phase. Over the course 10 ns, the full external field
273 causes a shift in the ordering of the system to 0.5, the nematic phase
274 of the liquid crystal. This change is consistent over the full 50 ns
275 with no drop back into the isotropic phase. This change is clearly
276 field induced and stable over a long period of simulation time.
277
278 Interestingly, the field that is needed to switch the phase of 5CB
279 macroscopically is larger than 1 V. However, in this case, only a
280 voltage of 1.2 V was need to induce a phase change. This is impart due
281 to the distance the field is being applied across. At such a small
282 distance, the field is much larger than the macroscopic and thus
283 easily induces a field dependent phase change.
284
428   This change in phase was followed by two courses of further
429 < simulation. First, was replacement of the static nitrile bond with a
429 > analysis. First was the replacement of the static nitrile bond with a
430   morse oscillator bond. This was then simulated for a period of time
431 < and a classical spetrum was calculated. Second, ab intio calcualtions were performe to investigate
432 < if the phase change caused any change spectrum from quantum
433 < effects.
431 > and a classical spetrum was calculated. Second, ab intio calcualtions
432 > were performed to investigate if the phase change caused any change
433 > spectrum through quantum effects.
434  
435 < In respect to the classical calculations, it was first seen if previous
436 < studies of nitriles within water and neat simulation done by Cho
437 < et. al. could be used for the spectrum.
438 <
439 < After Gaussian calculations were performed on a set of snapshots, any
435 > The classical nitrile spectrum can be seen in Figure 2. Most noticably
436 > is the position of the two peaks. Obviously the experimental peak
437 > position is near 2226 cm\textsuperscript{-1}. However, in this case
438 > the peak position is shifted to the blue at a position of 2375
439 > cm\textsuperscript{-1}. This shift is due solely to the choice of
440 > oscillator strength in the Morse oscillator parameters. While this
441 > shift makes the two spectra differ, it does not affect the ability to
442 > qualitatively compare peak changes to possible experimental changes.
443 > With this important fact out of the way, differences between the two
444 > states are subtle but are very much present. The first and
445 > most notable is the apperance for a strong band near 2300
446 > cm\textsuperscript{-1}.
447   \begin{figure}
298  \includegraphics[trim = 5mm 10mm 3mm 10mm, clip, width=3.25in]{P2}
299  \caption{Ordering of each external field application over the course
300    of 60 ns with a sampling every 100 ps. Each trajectory was started
301  after equilibration with zero field applied.}
302  \label{fig:orderParameter}
303 \end{figure}
304 \begin{figure}
448    \includegraphics[width=3.25in]{2Spectra}
449    \caption{The classically calculated nitrile bond spetrum for no
450      external field application (black) and full external field
451      application (red)}
452    \label{fig:twoSpectra}
453   \end{figure}
454 +
455 +
456 + Due to this, Gaussian calculations were performed in lieu of this
457 + method. A set of snapshots for the zero and full field simualtions,
458 + they were first investigated for any dependence on the local, with
459 + external field included, electric field. This was to see if a linear
460 + or non-linear relationship between the two could be utilized for
461 + generating spectra. This was done in part because of previous studies
462 + showing the frequency dependence of nitrile bonds to the electric
463 + fields generated locally between solvating water. It was seen that
464 + little to no dependence could be directly shown. This data is not
465 + shown.
466 +
467 + Since no explicit dependence was observed between the calculated
468 + frequency and the electric field, it was not a viable route for the
469 + calculation of a nitrile spectrum. Instead, the frequencies were taken
470 + and convolved together with a lorentzian line shape applied around the
471 + frequency value. These spectra are seen below in Figure
472 + 4. While the spectrum without a field is lower in intensity and is
473 + almost bimodel in distrobution, the external field spectrum is much
474 + more unimodel. This tighter clustering has the affect of increasing the
475 + intensity around 2226 cm\textsuperscript{-1} where the peak is
476 + centered. The external field also has fewer frequencies of higher
477 + energy in the spectrum. Unlike the the zero field, where some frequencies
478 + reach as high as 2280 cm\textsuperscript{-1}.
479   \begin{figure}
480    \includegraphics[width=3.25in]{Convolved}
481 <  \caption{Gaussian frequencies added together with gaussian }
481 >  \caption{Lorentzian convolved Gaussian frequencies of the zero field
482 >  system (black) and the full field system (red)}
483    \label{fig:Con}
484   \end{figure}
316 \begin{figure}
317  \includegraphics[width=7in]{Elip_3}
318  \caption{Ellipsoid reprsentation of 5CB at three different
319          field strengths, Zero Field (Left), Partial Field (Middle), and Full
320        Field (Right)}
321  \label{fig:Cigars}
322 \end{figure}
323
485   \section{Discussion}
486 + Interestingly, the field that is needed to switch the phase of 5CB
487 + macroscopically is larger than 1 V. However, in this case, only a
488 + voltage of 1.2 V was need to induce a phase change. This is impart due
489 + to the short distance of 5 nm the field is being applied across. At such a small
490 + distance, the field is much larger than the macroscopic and thus
491 + easily induces a field dependent phase change. However, this field
492 + will not cause a breakdown of the 5CB since electrochemistry studies
493 + have shown that it can be used in the presence of fields as high as
494 + 500 V macroscopically. This large of a field near the surface of the
495 + elctrode would cause breakdown of 5CB if it could happen.
496  
497 + The absence of any electric field dependency of the freuquency with
498 + the Gaussian simulations is interesting. A large base of research has been
499 + built upon the known tuning of the nitrile bond as the local field
500 + changes. This difference may be due to the absence of water or a
501 + molecule that induces a large internal field. Liquid water is known to have a very high internal field which
502 + is much larger than the internal fields of neat 5CB. Even though the
503 + application of Gaussian simulations followed by mapping it to
504 + some classical parameter is easy and straight forward, this system
505 + illistrates how that 'go to' method can break down.
506 +
507 + While this makes the application of nitrile Stark effects in
508 + simulations without water harder, these data show
509 + that it is not a deal breaker. The classically calculated nitrile
510 + spectrum shows changes in the spectra that will be easily seen through
511 + experimental routes. It indicates a shifted peak lower in energy
512 + should arise. This peak is a few wavenumbers from the leading edge of
513 + the larger peak and almost 75 wavenumbers from the center. This
514 + seperation between the two peaks means experimental results will show
515 + an easily resolved peak.
516 +
517 + The Gaussian derived spectra do indicate an applied field
518 + and subsiquent phase change does cause a narrowing of freuency
519 + distrobution. With narrowing, it would indicate an increased
520 + homogeneous distrobution of the local field near the nitrile.
521   \section{Conclusions}
522 + Field dependent changes
523   \newpage
524  
525   \bibliography{5CB}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines