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 4026 by gezelter, Fri Feb 14 18:54:39 2014 UTC vs.
Revision 4032 by gezelter, Wed Feb 19 19:14:24 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 | Notre Dame, Indiana 46556}
58   \begin{doublespace}
59  
60   \begin{abstract}
61 <  4-Cyano-4'-pentylbiphenyl (5CB) is a liquid-crystal-forming compound
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 the presence of static electric fields to provide
65 <  an understanding of the various contributions to the Stark shift of
66 <  the terminal nitrile group.  A field-induced isotropic-nematic phase
67 <  transition was observed in the simulations, and the effects of this
68 <  transition on the distribution of nitrile frequencies were
69 <  computed. Classical bond displacement correlation functions
70 <  exhibited a ($\approx 40 \mathrm{cm}^{-1}$ red shift of a fraction
71 <  of the main nitrile peak, and this shift was observed only when the
72 <  fields were large enough to induce orientational ordering of the
73 <  bulk phase.  Our simulations appear to indicate that phase-induced
74 <  changes to the local surroundings are a larger contribution to the
75 <  change in the nitrile spectrum than the direct field contribution.
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 The Stark shift of nitrile groups in response to applied electric
82 fields have been used extensively in biology for probing the internal
83 fields of structures like proteins and DNA.  Integration of these
84 probes into different materials is also important for studying local
85 structure in confined fluids. This work centers on the vibrational
86 response of the terminal nitrile group in 4-Cyano-4'-pentylbiphenyl
87 (5CB), a liquid crystalline molecule with a isotropic to nematic phase
88 that can be triggered by the application of an external field.
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
123 < 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.  Liquid crystalline
157 < materials with dipole moments located at the ends of the molecules
158 < have important applications in display technologies in addition to
159 < their relevance in biological systems.\cite{LCapp}
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 < Many of the technological applications of the lyotropic mesogens
133 < manipulate the orientation and structuring of the liquid crystal
134 < through application of external electric fields.\cite{?}
135 < Macroscopically, this restructuring is visible in the interactions the
136 < bulk phase has with scattered or transmitted light.\cite{?}  
137 <
138 < 4-Cyano-4'-pentylbiphenyl (5CB), has been a model for field-induced
139 < phase changes due to the known electric field response of the liquid
140 < crystal\cite{Hatta:1991ee}.  It was discovered (along with three
141 < similar compounds) in 1973 in an effort to find a LC that had a
142 < melting point near room temperature.\cite{Gray:1973ca} It's known to
143 < have a crystalline to nematic phase transition at 18 C and a nematic
144 < to isotropic transition at 35 C.\cite{Gray:1973ca} Recently there has
145 < been renewed interest in 5CB in nanodroplets to understand defect
146 < sites and nanoparticle structuring.\cite{PhysRevLett.111.227801}
147 <
148 < Nitrile groups can serve as very precise electric field reporters via
149 < their distinctive Raman and IR signatures.\cite{Boxer:2009xw} The
150 < triple bond between the nitrogen and the carbon atom is very sensitive
151 < to local field changes and is observed to have a direct impact on the
152 < peak position within the spectrum.  The Stark shift in the spectrum
153 < can be quantified and mapped into a field value that is impinging upon
154 < the nitrile bond. This has been used extensively in biological systems
155 < like proteins and enzymes.\cite{Tucker:2004qq,Webb:2008kn}
156 <
157 < To date, the nitrile electric field response of
158 < 4-Cyano-4'-n-alkylbiphenyl liquid crystals has not been investigated.
159 < While macroscopic electric fields applied across macro electrodes show
160 < the phase change of the molecule as a function of electric
161 < field,\cite{Lim:2006xq} the effect of a nanoscopic field application
162 < has not been probed. These previous studies have shown the nitrile
163 < group serves as an excellent indicator of the molecular orientation
164 < within the field applied. Lee {\it et al.}~showed the 180 degree
165 < change in field direction could be probed with the nitrile peak
166 < intensity as it decreased and increased with molecule alignment in the
167 < field.\cite{Lee:2006qd,Leyte:97}
168 <
169 < 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
183 < bond. This should be readily visible experimentally through Raman or
184 < 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 investigate these electric field effects using atomistic
179 < simulations of 5CB with applied external fields. These simulations are
180 < then coupled with both {\it ab intio} calculations of CN-deformations
181 < and classical correlation functions to predict spectral shifts. These
182 < predictions should be easily varifiable with scanning electrochemical
183 < microscopy experiments.
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
214 < $\frac{V}{\text{\AA}}$. The three field environments were, 1) no field
215 <  applied, 2) 0.01 $\frac{V}{\text{\AA}}$ (0.5 V) and 3) 0.024
216 <    $\frac{V}{\text{\AA}}$ (1.2 V). Each field was applied in the
217 <    Z-axis of the simulation box. For the simplicity of this paper,
218 <    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 lorezian lineshape
235 < function over each value. The width value used was 1.5. For the zero
240 < field spectrum, 67 frequencies were used and for the full field, 59
241 < 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 < Classical nitrile bond frequencies were found by replacing the rigid
244 < cyanide bond with a flexible Morse oscillator bond
245 < ($r_0= 1.157437$ \AA , $D_0 = 212.95$ and
246 < $\beta = 2.67566$) . Once replaced, the
247 < systems were allowed to re-equilibrate in NVT for 100 ps. After
248 < re-equilibration, the system was run in NVE for 20 ps with a snapshot
249 < spacing of 1 fs. These snapshot were then used in bond correlation
250 < calculation to find the decay structure of the bond in time using the
251 < average bond displacement in time,
252 < \begin{equation}
253 < C(t) = \langle \left(r(t) - r_0 \right) \cdot \left(r(0) - r_0 \right) \rangle
254 < \end{equation}
255 < %
256 < where $r_0$ is the equilibrium bond distance and $r(t)$ is the
257 < instantaneous bond displacement at time $t$. Once calculated,
258 < smoothing was applied by adding an exponential decay on top of the
259 < decay with a $\tau$ of 6000. Further smoothing
260 < was applied by padding 20,000 zeros on each side of the symmetric
261 < data. This was done five times by allowing the systems to run 1 ns
262 < with a rigid bond followed by an equilibrium run with the bond
263 < switched back to a Morse oscillator and a short production run of 20 ps.
237 > \section{Field-induced Nematic Ordering}
238  
265 \section{Results}
266
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 is determined using the tensor
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)
# Line 275 | Line 247 | $S$ takes on values close to 1 in highly ordered phase
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 phases, but falls to
251 < zero for isotropic fluids. In the context of 5CB, this value would be
252 < close to zero for its isotropic phase and raise closer to one as it
253 < moved to the nematic and crystalline phases.
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 < This value indicates phases changes at temperature boundaries but also
258 < when a phase change occurs due to external field applications. In
259 < Figure 1, this phase change can be clearly seen over the course of 60
260 < ns. Each system starts with an ordering paramter near 0.1 to 0.2,
261 < which is an isotropic phase. Over the course 10 ns, the full external field
262 < causes a shift in the ordering of the system to 0.5, the nematic phase
263 < of the liquid crystal. This change is consistent over the full 50 ns
264 < with no drop back into the isotropic phase. This change is clearly
265 < field induced and stable over a long period of simulation time.
266 < \begin{figure}
267 <  \includegraphics[trim = 5mm 10mm 3mm 10mm, clip, width=3.25in]{P2}
268 <  \caption{Ordering of each external field application over the course
269 <    of 60 ns with a sampling every 100 ps. Each trajectory was started
270 <    after equilibration with zero field applied.}
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  
300 In the figure below, this phase change is represented nicely as
301 ellipsoids that are created by the vector formed between the nitrogen
302 of the nitrile group and the tail terminal carbon atom. These
303 illistrate the change from isotropic phase to nematic change. Both the
304 zero field and partial field images look mostly disordered. The
305 partial field does look somewhat orded but without any overall order
306 of the entire system. This is most likely a high point in the ordering
307 for the trajectory. The full field image on the other hand looks much
308 more ordered when compared to the two lower field simulations.
309 \begin{figure}
310  \includegraphics[width=7in]{Elip_3}
311  \caption{Ellipsoid reprsentation of 5CB at three different
312    field strengths, Zero Field (Left), Partial Field (Middle), and Full
313    Field (Right) Each image was created by taking the final
314    snapshot of each 60 ns run}
315  \label{fig:Cigars}
316 \end{figure}
286  
287 < This change in phase was followed by two courses of further
319 < analysis. First was the replacement of the static nitrile bond with a
320 < morse oscillator bond. This was then simulated for a period of time
321 < and a classical spetrum was calculated. Second, ab intio calcualtions
322 < were performed to investigate if the phase change caused any change
323 < spectrum through quantum effects.
287 > \section{Sampling the CN bond frequency}
288  
289 < The classical nitrile spectrum can be seen in Figure 2. Most noticably
290 < is the position of the two peaks. Obviously the experimental peak
291 < position is near 2226 cm\textsuperscript{-1}. However, in this case
292 < the peak position is shifted to the blue at a position of 2375
293 < cm\textsuperscript{-1}. This shift is due solely to the choice of
294 < oscillator strength in the Morse oscillator parameters. While this
295 < shift makes the two spectra differ, it does not affect the ability to
296 < qualitatively compare peak changes to possible experimental changes.
297 < With this important fact out of the way, differences between the two
298 < states are subtle but are very much present. The first and
299 < most notable is the apperance for a strong band near 2300
300 < cm\textsuperscript{-1}.
301 < \begin{figure}
302 <  \includegraphics[width=3.25in]{2Spectra}
303 <  \caption{The classically calculated nitrile bond spetrum for no
304 <    external field application (black) and full external field
305 <    application (red)}
306 <  \label{fig:twoSpectra}
307 < \end{figure}
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 condensed phase system prevents direct computation of
316 + the nitrile bond frequencies using {\it ab initio} methods.  In order
317 + to sample the nitrile frequencies present in the condensed-phase,
318 + individual molecules were selected randomly to serve as the center of
319 + a local (gas phase) cluster.  To include steric, electrostatic, and
320 + other effects from molecules located near the targeted nitrile group,
321 + portions of other molecules nearest to the nitrile group were included
322 + in the calculations.  The surrounding solvent molecules were divided
323 + into ``body'' (the two phenyl rings and the nitrile bond) and ``tail''
324 + (the alkyl chain).  Any molecule which had a body atom within 6~\AA of
325 + the midpoint of the target nitrile group
326 +
327 +
328 +
329 + or the body not including
330 + the tail, any atom within 6~\AA of the midpoint of the nitrile group
331 + was included. For the tail structure, the whole tail was included if a
332 + tail atom was within 4~\AA of the midpoint. If the tail did not
333 + include any atoms from the ring structure, it was considered a propane
334 + molecule and included as such. Once the clusters were generated, input
335 + files were created that stretched the nitrile bond along its axis from
336 + 0.87 to 1.52~\AA at increments of 0.05~\AA. This generated 13 single
337 + point energies to be calculated. The Gaussian files were run with
338 + B3LYP/6-311++G(d,p) with no other keywords for the zero field
339 + simulation. Simulations with fields applied included the keyword
340 + ''Field=Z+5'' to match the external field applied in molecular dynamic
341 + runs. Once completed, the central nitrile bond frequency was
342 + calculated with a Morse fit. Using this fit and the solved energy
343 + levels for a Morse oscillator, the frequency was found. Each set of
344 + frequencies were then convolved together with a lorezian lineshape
345 + function over each value. The width value used was 1.5. For the zero
346 + field spectrum, 67 frequencies were used and for the full field, 59
347 + frequencies were used.
348 +
349 + \subsection{CN frequencies from potential-frequency maps}
350   Before Gaussian silumations were carried out, it was attempt to apply
351   the method developed by Cho  {\it et al.}~\cite{Oh:2008fk} This method involves the fitting
352   of multiple parameters to Gaussian calculated freuencies to find a
# Line 381 | Line 386 | Due to this, Gaussian calculations were performed in l
386   external field being applied. However, the factor was no linear and
387   was overly large due to the fitting parameters being so small.
388  
389 +
390 + \subsection{CN frequencies from bond length autocorrelation functions}
391 +
392 + Classical nitrile bond frequencies were found by replacing the rigid
393 + cyanide bond with a flexible Morse oscillator bond
394 + ($r_0= 1.157437$ \AA , $D_0 = 212.95$ and
395 + $\beta = 2.67566$) . Once replaced, the
396 + systems were allowed to re-equilibrate in NVT for 100 ps. After
397 + re-equilibration, the system was run in NVE for 20 ps with a snapshot
398 + spacing of 1 fs. These snapshot were then used in bond correlation
399 + calculation to find the decay structure of the bond in time using the
400 + average bond displacement in time,
401 + \begin{equation}
402 + C(t) = \langle \left(r(t) - r_0 \right) \cdot \left(r(0) - r_0 \right) \rangle
403 + \end{equation}
404 + %
405 + where $r_0$ is the equilibrium bond distance and $r(t)$ is the
406 + instantaneous bond displacement at time $t$. Once calculated,
407 + smoothing was applied by adding an exponential decay on top of the
408 + decay with a $\tau$ of 6000. Further smoothing
409 + was applied by padding 20,000 zeros on each side of the symmetric
410 + data. This was done five times by allowing the systems to run 1 ns
411 + with a rigid bond followed by an equilibrium run with the bond
412 + switched back to a Morse oscillator and a short production run of 20 ps.
413 +
414 +
415 + This change in phase was followed by two courses of further
416 + analysis. First was the replacement of the static nitrile bond with a
417 + morse oscillator bond. This was then simulated for a period of time
418 + and a classical spetrum was calculated. Second, ab intio calcualtions
419 + were performed to investigate if the phase change caused any change
420 + spectrum through quantum effects.
421 +
422 + The classical nitrile spectrum can be seen in Figure 2. Most noticably
423 + is the position of the two peaks. Obviously the experimental peak
424 + position is near 2226 cm\textsuperscript{-1}. However, in this case
425 + the peak position is shifted to the blue at a position of 2375
426 + cm\textsuperscript{-1}. This shift is due solely to the choice of
427 + oscillator strength in the Morse oscillator parameters. While this
428 + shift makes the two spectra differ, it does not affect the ability to
429 + qualitatively compare peak changes to possible experimental changes.
430 + With this important fact out of the way, differences between the two
431 + states are subtle but are very much present. The first and
432 + most notable is the apperance for a strong band near 2300
433 + cm\textsuperscript{-1}.
434 + \begin{figure}
435 +  \includegraphics[width=3.25in]{2Spectra}
436 +  \caption{The classically calculated nitrile bond spetrum for no
437 +    external field application (black) and full external field
438 +    application (red)}
439 +  \label{fig:twoSpectra}
440 + \end{figure}
441 +
442 +
443   Due to this, Gaussian calculations were performed in lieu of this
444   method. A set of snapshots for the zero and full field simualtions,
445   they were first investigated for any dependence on the local, with

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines