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 4025 by jmarr, Thu Feb 6 23:20:32 2014 UTC vs.
Revision 4028 by gezelter, Tue Feb 18 21:26:04 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}  
44   \author{James M. Marr}
45   \author{J. Daniel Gezelter}
46   \email{gezelter@nd.edu}
# Line 58 | Line 58 | Nitrile Stark shift repsonses to electric fields have
58   \begin{doublespace}
59  
60   \begin{abstract}
61 < Nitrile Stark shift repsonses to electric fields have been used
62 < extensively in biology for the probing of local internal fields of
63 < structures like proteins and DNA. Intigration of these probes into
64 < different areas of interest are important for studing local structure
65 < and fields within confined, nanoscopic
66 < systems. 4-Cyano-4'-pentylbiphenyl (5CB) is a liquid crystal with a known
67 < macroscopic structure reordering from the isotropic to nematic
68 < phase with the application of an external
69 < field and as the name suggests has an inherent nitrile group. Through
70 < simulations of this molecule where application of
71 < large, nanoscale external fields were applied, the nitrile was invenstigated
72 < as a local field sensor. It was
73 < found that while most computational methods for nitrile spectral
74 < calculations rely on a correlation between local electric field and
75 < the nitrile bond, 5CB did not have an easily obtained
76 < correlation. Instead classical calculation through correlation of the
77 < cyanide bond displacement in time use enabled to show a spectral
78 < change in the formation of a red
79 < shifted peak from the main peak as an external field was applied. This indicates
80 < that local structure has a larger impact on the nitrile frequency then
81 < does the local electric field. By better understanding how nitrile
82 < groups respond to local and external fields it will help
83 < nitrile groups branch out beyond their biological
84 < applications to uses in electronics and surface sciences.
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
124 < 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
131 < of the molecules have important applications in display technologies
132 < in addition to their relevance in biological systems.\cite{LCapp}
133 <
134 < Many of the technological applications of the lyotropic mesogens
135 < manipulate the orientation and structuring of the liquid crystal
136 < through application of external electric fields.\cite{?}
137 < Macroscopically, this restructuring is visible in the interactions the
138 < bulk phase has with scattered or transmitted light.\cite{?}  
139 <
140 < 4-Cyano-4'-pentylbiphenyl (5CB), has been a model for field-induced
141 < phase changes due to the known electric field response of the liquid
142 < crystal\cite{Hatta:1991ee}.  It was discovered (along with three
143 < similar compounds) in 1973 in an effort to find a LC that had a
144 < melting point near room temperature.\cite{Gray:1973ca} It's known to
145 < have a crystalline to nematic phase transition at 18 C and a nematic
146 < to isotropic transition at 35 C.\cite{Gray:1973ca} Recently it has
147 < seen new life with the application of droplets of the molecule in
148 < water being used to study defect sites and nanoparticle
149 < strcuturing.\cite{PhysRevLett.111.227801}
150 <
151 < Nitrile groups can serve as very precise electric field reporters via
152 < their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
153 < triple bond between the nitrogen and the carbon atom is very sensitive
154 < to local field changes and is observed to have a direct impact on the
155 < peak position within the spectrum.  The Stark shift in the spectrum
156 < can be quantified and mapped into a field value that is impinging upon
157 < the nitrile bond. This has been used extensively in biological systems
158 < like proteins and enzymes.\cite{Tucker:2004qq,Webb:2008kn}
159 <
160 < To date, the nitrile electric field response of
161 < 4-Cyano-4'-n-alkylbiphenyl liquid crystals has not been investigated.
162 < While macroscopic electric fields applied across macro electrodes show
163 < the phase change of the molecule as a function of electric
164 < field,\cite{Lim:2006xq} the effect of a nanoscopic field application
165 < has not been probed. These previous studies have shown the nitrile
166 < group serves as an excellent indicator of the molecular orientation
167 < within the field applied. Lee {\it et al.}~showed the 180 degree change in field
168 < direction could be probed with the nitrile peak intensity as it
169 < decreased and increased with molecule alignment in the
170 < field.\cite{Lee:2006qd,Leyte:97}
171 <
172 < 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
186 < bond. This should be readily visible experimentally through Raman or
187 < IR spectroscopy.
188 <
189 < Herein, we show the computational investigation of these electric
190 < field effects through atomistic simulations of 5CB with external
191 < fields applied. These simulations are then coupled with ab intio and
192 < classical spectrum calculations to predict changes. These changes are
193 < easily varifiable with experiments and should be able to replicated
194 < experimentally.
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 + 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
217 < $\frac{V}{\text{\AA}}$. The three field environments were, 1) no field
218 <  applied, 2) 0.01 $\frac{V}{\text{\AA}}$ (0.5 V) and 3) 0.024
219 <    $\frac{V}{\text{\AA}}$ (1.2 V). Each field was applied in the
220 <    Z-axis of the simulation box. For the simplicity of this paper,
221 <    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 + 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}$, corresponding to a
224 + vibrational frequency of $2375 \mathrm{~cm}^{-1}$, a
225 + bit higher than the experimental frequency.  The flexible nitrile
226 + moiety required simulation time steps of 1~fs, so the additional
227 + flexibility was introducuced only after the rigid systems had come to
228 + equilibrium under the applied fields.  Whenever time correlation
229 + 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 + In Figure \ref{fig:orderParameter}, the field-induced phase change can
258 + be clearly seen over the course of a 60 ns equilibration run. 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 landed in the nematic-ordered state, it became stable for the
264 + remaining 50 ns of simulation time.  It is possible that the
265 + partial-field simulation is meta-stable and given enough time, it
266 + would eventually find a nematic-ordered phase, but the partial-field
267 + simulation was stable as an isotropic phase for the full duration of a
268 + 60 ns simulation.
269 + \begin{figure}
270 +  \includegraphics[trim = 5mm 10mm 3mm 10mm, clip, width=3.25in]{P2}
271 +  \caption{Ordering of each external field application over the course
272 +    of 60 ns with a sampling every 100 ps. Each trajectory was started
273 +    after equilibration with zero field applied.}
274 +  \label{fig:orderParameter}
275 + \end{figure}
276 +
277 + In figure \ref{fig:Cigars}, the field-induced isotropic-nematic
278 + transition is represented using ellipsoids aligned along the long-axis
279 + of each molecule.  The vector between the nitrogen of the nitrile
280 + group and the terminal tail atom is used to orient each
281 + ellipsoid.  Both the zero field and partial field simulations appear
282 + isotropic, while the full field simulations has clearly been
283 + orientationally ordered
284 + \begin{figure}
285 +  \includegraphics[width=7in]{Elip_3}
286 +  \caption{Ellipsoid reprsentation of 5CB at three different
287 +    field strengths, Zero Field (Left), Partial Field (Middle), and Full
288 +    Field (Right) Each image was created by taking the final
289 +    snapshot of each 60 ns run}
290 +  \label{fig:Cigars}
291 + \end{figure}
292 +
293 + \section{Analysis}
294 +
295   For quantum calculation of the nitrile bond frequency, Gaussian 09 was
296   used. A single 5CB molecule was selected for the center of the
297   cluster. For effects from molecules located near the chosen nitrile
# Line 267 | Line 339 | In order to characterize the orientational ordering of
339  
340   \section{Results}
341  
270 In order to characterize the orientational ordering of the system, the
271 primary quantity of interest is the nematic (orientational) order
272 parameter. This is determined using the tensor
273 \begin{equation}
274 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{e}_{i
275    \alpha} \hat{e}_{i \beta} - \delta_{\alpha \beta} \right)
276 \end{equation}
277 where $\alpha, \beta = x, y, z$, and $\hat{e}_i$ is the molecular
278 end-to-end unit vector for molecule $i$. The nematic order parameter
279 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
280 corresponding eigenvector defines the director axis for the phase.
281 $S$ takes on values close to 1 in highly ordered phases, but falls to
282 zero for isotropic fluids. In the context of 5CB, this value would be
283 close to zero for its isotropic phase and raise closer to one as it
284 moved to the nematic and crystalline phases.
285
286 This value indicates phases changes at temperature boundaries but also
287 when a phase change occurs due to external field applications. In
288 Figure 1, this phase change can be clearly seen over the course of 60
289 ns. Each system starts with an ordering paramter near 0.1 to 0.2,
290 which is an isotropic phase. Over the course 10 ns, the full external field
291 causes a shift in the ordering of the system to 0.5, the nematic phase
292 of the liquid crystal. This change is consistent over the full 50 ns
293 with no drop back into the isotropic phase. This change is clearly
294 field induced and stable over a long period of simulation time.
295 \begin{figure}
296  \includegraphics[trim = 5mm 10mm 3mm 10mm, clip, width=3.25in]{P2}
297  \caption{Ordering of each external field application over the course
298    of 60 ns with a sampling every 100 ps. Each trajectory was started
299    after equilibration with zero field applied.}
300  \label{fig:orderParameter}
301 \end{figure}
302
303 In the figure below, this phase change is represented nicely as
304 ellipsoids that are created by the vector formed between the nitrogen
305 of the nitrile group and the tail terminal carbon atom. These
306 illistrate the change from isotropic phase to nematic change. Both the
307 zero field and partial field images look mostly disordered. The
308 partial field does look somewhat orded but without any overall order
309 of the entire system. This is most likely a high point in the ordering
310 for the trajectory. The full field image on the other hand looks much
311 more ordered when compared to the two lower field simulations.
312 \begin{figure}
313  \includegraphics[width=7in]{Elip_3}
314  \caption{Ellipsoid reprsentation of 5CB at three different
315    field strengths, Zero Field (Left), Partial Field (Middle), and Full
316    Field (Right) Each image was created by taking the final
317    snapshot of each 60 ns run}
318  \label{fig:Cigars}
319 \end{figure}
320
342   This change in phase was followed by two courses of further
343   analysis. First was the replacement of the static nitrile bond with a
344   morse oscillator bond. This was then simulated for a period of time

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines