ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4023
Committed: Wed Feb 5 22:17:10 2014 UTC (10 years, 7 months ago) by jmarr
Content type: application/x-tex
File size: 22670 byte(s)
Log Message:
Stuff

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3
4 \usepackage{caption}
5 \usepackage{float}
6 \usepackage{geometry}
7 \usepackage{natbib}
8 \usepackage{setspace}
9 \usepackage{xkeyval}
10 \usepackage{amsmath}
11 \usepackage{amssymb}
12 \usepackage{times}
13 \usepackage{mathptm}
14 \usepackage{setspace}
15 %\usepackage{endfloat}
16 \usepackage{tabularx}
17 \usepackage{longtable}
18 \usepackage{graphicx}
19 \usepackage{multirow}
20 \usepackage{multicol}
21 \usepackage{achemso}
22 \usepackage{subcaption}
23 \usepackage[colorinlistoftodos]{todonotes}
24 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25 % \usepackage[square, comma, sort&compress]{natbib}
26 \usepackage{url}
27 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29 9.0in \textwidth 6.5in \brokenpenalty=10000
30
31 % double space list of tables and figures
32 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33 \setlength{\abovecaptionskip}{20 pt}
34 \setlength{\belowcaptionskip}{30 pt}
35
36 % \bibpunct{}{}{,}{s}{}{;}
37
38 %\citestyle{nature}
39 % \bibliographystyle{achemso}
40
41
42 \title{Nitrile vibrations as reporters of field-induced phase
43 transitions in liquid crystals}
44 \author{James M. Marr}
45 \author{J. Daniel Gezelter}
46 \email{gezelter@nd.edu}
47 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
48 Department of Chemistry and Biochemistry\\
49 University of Notre Dame\\
50 Notre Dame, Indiana 46556}
51
52 \date{\today}
53
54 \begin{document}
55
56 \maketitle
57
58 \begin{doublespace}
59
60 \begin{abstract}
61 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.
85 \end{abstract}
86
87 \newpage
88
89 \section{Introduction}
90
91 The fundamental characteristic of liquid crystal mesophases is that
92 they maintain some degree of orientational order while translational
93 order is limited or absent. This orientational order produces a
94 complex direction-dependent response to external perturbations like
95 electric fields and mechanical distortions. The anisotropy of the
96 macroscopic phases originates in the anisotropy of the constituent
97 molecules, which typically have highly non-spherical structures with a
98 significant degree of internal rigidity. In nematic phases, rod-like
99 molecules are orientationally ordered with isotropic distributions of
100 molecular centers of mass, while in smectic phases, the molecules
101 arrange themselves into layers with their long (symmetry) axis normal
102 ($S_{A}$) or tilted ($S_{C}$) with respect to the layer planes.
103
104 The behavior of the $S_{A}$ phase can be partially explained with
105 models mainly based on geometric factors and van der Waals
106 interactions. However, these simple models are insufficient to
107 describe liquid crystal phases which exhibit more complex polymorphic
108 nature. X-ray diffraction studies have shown that the ratio between
109 lamellar spacing ($s$) and molecular length ($l$) can take on a wide
110 range of values.\cite{Gray:1984hc,Leadbetter:1976vf,Hardouin:1980yq}
111 Typical $S_{A}$ phases have $s/l$ ratios on the order of $0.8$, while
112 for some compounds, e.g. the 4-alkyl-4'-cyanobiphenyls, the $s/l$
113 ratio is on the order of $1.4$. Molecules which form $S_{A}$ phases
114 can exhibit a wide variety of subphases like monolayers ($S_{A1}$),
115 uniform bilayers ($S_{A2}$), partial bilayers ($S_{\tilde A}$) as well
116 as interdigitated bilayers ($S_{A_{d}}$), and often have a terminal
117 cyano or nitro group. In particular lyotropic liquid crystals (those
118 exhibiting liquid crystal phase transition as a function of water
119 concentration) often have polar head groups or zwitterionic charge
120 separated groups that result in strong dipolar
121 interactions.\cite{Collings97} Because of their versatile polymorphic
122 nature, polar liquid crystalline materials have important
123 technological applications in addition to their immense relevance to
124 biological systems.\cite{Collings97}
125
126 Experimental studies by Levelut {\it et al.}~\cite{Levelut:1981eu}
127 revealed that terminal cyano or nitro groups usually induce permanent
128 longitudinal dipole moments on the molecules.
129
130 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 local 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}
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 et. al. showed the 180 degree change in field
165 direction could be probed with the nitrile peak intensity as it
166 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
170 response, the atomic scale response has not been studied. With the
171 advent of nano-electrodes and coupling them with atomic force
172 microscopy, control of electric fields applied across nanometer
173 distances is now possible\cite{citation1}. This application of
174 nanometer length is interesting in the case of a nitrile group on the
175 molecule. While macroscopic fields are insufficient to cause a Stark
176 effect, small fields across nanometer-sized gaps are of sufficient
177 strength. If one were to assume a gap of 5 nm between a lower
178 electrode having a nanoelectrode placed near it via an atomic force
179 microscope, a field of 1 V applied across the electrodes would
180 translate into a field of 2x10\textsuperscript{8} $\frac{V}{M}$. This
181 field is theoretically strong enough to cause a phase change from
182 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.
185
186 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.
187
188 \section{Computational Details}
189 The force field was mainly taken from Guo et al.\cite{Zhang:2011hh} A
190 deviation from this force field was made to create a rigid body from
191 the phenyl rings. Bond distances within the rigid body were taken from
192 equilibrium bond distances. While the phenyl rings were held rigid,
193 bonds, bends, torsions and inversion centers still included the rings.
194
195 Simulations were with boxes of 270 molecules locked at experimental
196 densities with periodic boundaries. The molecules were thermalized
197 from 0 kelvin to 300 kelvin. To equilibrate, each was first run in NVT
198 for 1 ns. This was followed by NVE for simulations used in the data
199 collection.
200
201 External electric fields were applied in a simplistic charge-field
202 interaction. Forces were calculated by multiplying the electric field
203 being applied by the charge at each atom. For the potential, the
204 origin of the box was used as a point of reference. This allows for a
205 potential value to be added to each atom as the molecule move in space
206 within the box. Fields values were applied in a manner representing
207 those that would be applable in an experimental set-up. The assumed
208 electrode seperation was 5 nm and the field was input as
209 $\frac{V}{\text{\AA}}$. The three field environments were, 1) no field
210 applied, 2) 0.01 $\frac{V}{\text{\AA}}$ (0.5 V) and 3) 0.024
211 $\frac{V}{\text{\AA}}$ (1.2 V). Each field was applied in the
212 Z-axis of the simulation box. For the simplicity of this paper,
213 each field will be called zero, partial and full, respectively.
214
215 For quantum calculation of the nitrile bond frequency, Gaussian 09 was
216 used. A single 5CB molecule was selected for the center of the
217 cluster. For effects from molecules located near the chosen nitrile
218 group, parts of molecules nearest to the nitrile group were
219 included. For the body not including the tail, any atom within 6~\AA
220 of the midpoint of the nitrile group was included. For the tail
221 structure, the whole tail was included if a tail atom was within 4~\AA
222 of the midpoint. If the tail did not include any atoms from the ring
223 structure, it was considered a propane molecule and included as
224 such. Once the clusters were generated, input files were created that
225 stretched the nitrile bond along its axis from 0.87 to 1.52~\AA at
226 increments of 0.05~\AA. This generated 13 single point energies to be
227 calculated. The Gaussian files were run with B3LYP/6-311++G(d,p) with
228 no other keywords for the zero field simulation. Simulations with
229 fields applied included the keyword ''Field=Z+5'' to match the
230 external field applied in molecular dynamic runs. Once completed, the central nitrile bond frequency
231 was calculated with a Morse fit. Using this fit and the solved energy
232 levels for a Morse oscillator, the frequency was found. Each set of
233 frequencies were then convolved together with a lorezian lineshape
234 function over each value. The width value used was 1.5. For the zero
235 field spectrum, 67 frequencies were used and for the full field, 59
236 frequencies were used.
237
238 Classical nitrile bond frequencies were found by replacing the rigid
239 cyanide bond with a flexible Morse oscillator bond
240 ($r_0= 1.157437$ \AA , $D_0 = 212.95$ and
241 $\beta = 2.67566$) . Once replaced, the
242 systems were allowed to re-equilibrate in NVT for 100 ps. After
243 re-equilibration, the system was run in NVE for 20 ps with a snapshot
244 spacing of 1 fs. These snapshot were then used in bond correlation
245 calculation to find the decay structure of the bond in time using the
246 average bond displacement in time,
247 \begin{equation}
248 C(t) = \langle \left(r(t) - r_0 \right) \cdot \left(r(0) - r_0 \right) \rangle
249 \end{equation}
250 %
251 where $r_0$ is the equilibrium bond distance and $r(t)$ is the
252 instantaneous bond displacement at time $t$. Once calculated,
253 smoothing was applied by adding an exponential decay on top of the
254 decay with a $\tau$ of 6000. Further smoothing
255 was applied by padding 20,000 zeros on each side of the symmetric
256 data. This was done five times by allowing the systems to run 1 ns
257 with a rigid bond followed by an equilibrium run with the bond
258 switched back to a Morse oscillator and a short production run of 20 ps.
259
260 \section{Results}
261
262 In order to characterize the orientational ordering of the system, the
263 primary quantity of interest is the nematic (orientational) order
264 parameter. This is determined using the tensor
265 \begin{equation}
266 Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^{N} \left(3 \hat{e}_{i
267 \alpha} \hat{e}_{i \beta} - \delta_{\alpha \beta} \right)
268 \end{equation}
269 where $\alpha, \beta = x, y, z$, and $\hat{e}_i$ is the molecular
270 end-to-end unit vector for molecule $i$. The nematic order parameter
271 $S$ is the largest eigenvalue of $Q_{\alpha \beta}$, and the
272 corresponding eigenvector defines the director axis for the phase.
273 $S$ takes on values close to 1 in highly ordered phases, but falls to
274 zero for isotropic fluids. In the context of 5CB, this value would be
275 close to zero for its isotropic phase and raise closer to one as it
276 moved to the nematic and crystalline phases.
277
278 This value indicates phases changes at temperature boundaries but also
279 when a phase change occurs due to external field applications. In
280 Figure 1, this phase change can be clearly seen over the course of 60
281 ns. Each system starts with an ordering paramter near 0.1 to 0.2,
282 which is an isotropic phase. Over the course 10 ns, the full external field
283 causes a shift in the ordering of the system to 0.5, the nematic phase
284 of the liquid crystal. This change is consistent over the full 50 ns
285 with no drop back into the isotropic phase. This change is clearly
286 field induced and stable over a long period of simulation time.
287 \begin{figure}
288 \includegraphics[trim = 5mm 10mm 3mm 10mm, clip, width=3.25in]{P2}
289 \caption{Ordering of each external field application over the course
290 of 60 ns with a sampling every 100 ps. Each trajectory was started
291 after equilibration with zero field applied.}
292 \label{fig:orderParameter}
293 \end{figure}
294
295 Interestingly, the field that is needed to switch the phase of 5CB
296 macroscopically is larger than 1 V. However, in this case, only a
297 voltage of 1.2 V was need to induce a phase change. This is impart due
298 to the distance the field is being applied across. At such a small
299 distance, the field is much larger than the macroscopic and thus
300 easily induces a field dependent phase change.
301
302 In the figure below, this phase change is represented nicely as
303 ellipsoids that are created by the vector formed between the nitrogen
304 of the nitrile group and the tail terminal carbon atom. These
305 illistrate the change from isotropic phase to nematic change. Both the
306 zero field and partial field images look mostly disordered. The
307 partial field does look somewhat orded but without any overall order
308 of the entire system. This is most likely a high point in the ordering
309 for the trajectory. The full field image on the other hand looks much
310 more ordered when compared to the two lower field simulations.
311 \begin{figure}
312 \includegraphics[width=7in]{Elip_3}
313 \caption{Ellipsoid reprsentation of 5CB at three different
314 field strengths, Zero Field (Left), Partial Field (Middle), and Full
315 Field (Right) Each image was created by taking the final
316 snapshot of each 60 ns run}
317 \label{fig:Cigars}
318 \end{figure}
319
320 This change in phase was followed by two courses of further
321 analysis. First was the replacement of the static nitrile bond with a
322 morse oscillator bond. This was then simulated for a period of time
323 and a classical spetrum was calculated. Second, ab intio calcualtions
324 were performed to investigate if the phase change caused any change
325 spectrum through quantum effects.
326
327 The classical nitrile spectrum can be seen in Figure 2. Most noticably
328 is the position of the two peaks. Obviously the experimental peak
329 position is near 2226 cm\textsuperscript{-1}. However, in this case
330 the peak position is shifted to the blue at a position of 2375
331 cm\textsuperscript{-1}. This shift is due solely to the choice of
332 oscillator strength in the Morse oscillator parameters. While this
333 shift makes the two spectra differ, it does not affect the ability to
334 qualitatively compare peak changes to possible experimental changes.
335 With this important fact out of the way, differences between the two
336 states are subtle but are very much present. The first and
337 most notable is the apperance for a strong band near 2300
338 cm\textsuperscript{-1}.
339 \begin{figure}
340 \includegraphics[width=3.25in]{2Spectra}
341 \caption{The classically calculated nitrile bond spetrum for no
342 external field application (black) and full external field
343 application (red)}
344 \label{fig:twoSpectra}
345 \end{figure}
346
347 Before Gaussian silumations were carried out, it was attempt to apply
348 the method developed by Cho et. al. This method involves the fitting
349 of multiple parameters to Gaussian calculated freuencies to find a
350 correlation between the potential around the bond and the
351 frequency. This is very similar to work done by Skinner et. al. with
352 water models like SPC/E. The general method is to find the shift in
353 the peak position through,
354 \begin{equation}
355 \delta\tilde{\nu} =\sum^{n}_{a=1} l_{a}\phi^{water}_{a}
356 \end{equation}
357 where $l_{a}$ are the fitting parameters and $\phi^{water}_{a}$ is the
358 potential from the surrounding water cluster. This $\phi^{water}_{a}$
359 takes the form,
360 \begin{equation}
361 \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{m} \sum_{j}
362 \frac{C^{H_{2}O}_{j \left(m \right) }}{r_{aj \left(m\right)}}
363 \end{equation}
364 where $C^{H_{2}O}_{j \left(m \right) }$ indicates the partial charge
365 on the $j$th site of the $m$th water molecule and $r_{aj \left(m\right)}$
366 is the distance between the site $a$ of the nitrile molecule and the $j$th
367 site of the $m$th water molecule. However, since these simulations
368 are done under the presence of external electric fields and in the
369 absence of water the equations must have a correction factor for the
370 external field change as well as the use of electric field site data
371 instead of charged site points. So by modifing the original
372 $\phi^{water}_{a}$ to $\phi^{5CB}_{a}$ we get,
373 \begin{equation}
374 \phi^{5CB}_{a} = \frac{1}{4\pi \epsilon_{0}} \left( \vec{E}\bullet
375 \left(\vec{r}_{a}-\vec{r}_{CN}\right) \right) + \phi^{5CB}_{0}
376 \end{equation}
377 where $\vec{E}$ is the electric field at each atom, $\left( \vec{r}_{a} -
378 \vec{r}_{CN} \right)$ is the vector between the nitrile bond and the
379 cooridinates described by Cho around the bond and $\phi^{5CB}_{0}$ is
380 the correction factor for the system of parameters. After these
381 changes, the correction factor was found for multiple values of an
382 external field being applied. However, the factor was no linear and
383 was overly large due to the fitting parameters being so small.
384
385 Due to this, Gaussian calculations were performed in lieu of this
386 method. A set of snapshots for the zero and full field simualtions,
387 they were first investigated for any dependence on the local, with
388 external field included, electric field. This was to see if a linear
389 or non-linear relationship between the two could be utilized for
390 generating spectra. This was done in part because of previous studies
391 showing the frequency dependence of nitrile bonds to the electric
392 fields generated locally between solvating water. It was seen that
393 little to no dependence could be directly shown. This data is not
394 shown.
395
396 Since no explicit dependence was observed between the calculated
397 frequency and the electric field, it was not a viable route for the
398 calculation of a nitrile spectrum. Instead, the frequencies were taken
399 and convolved together. These two spectra are seen below in Figure
400 4. While the spectrum without a field is lower in intensity and is
401 almost bimodel in distrobuiton, the external field spectrum is much
402 more unimodel. This tighter clustering has the affect of increasing the
403 intensity around 2226 cm\textsuperscript{-1} where the peak is
404 centered. The external field also has fewer frequencies of higher
405 energy in the spectrum. Unlike the the zero field, where some frequencies
406 reach as high as 2280 cm\textsuperscript{-1}.
407 \begin{figure}
408 \includegraphics[width=3.25in]{Convolved}
409 \caption{Lorentzian convolved Gaussian frequencies of the zero field
410 system (black) and the full field system (red)}
411 \label{fig:Con}
412 \end{figure}
413 \section{Discussion}
414 The absence of any electric field dependency of the freuquency with
415 the Gaussian simulations is strange. A large base of research has been
416 built upon the known tuning of the nitrile bond as local field
417 changes. This differences may be due to the absence of water. Many of
418 the nitrile bond fitting maps are done in the presence of liquid
419 water. Liquid water is known to have a very high internal field which
420 is much larger than the internal fields of neat 5CB. Even though the
421 application of Gaussian simulations followed by mappying to
422 some classical parameter is easy and straight forward, this system
423 illistrates how that 'go to' method can break down.
424
425 While this makes the application of nitrile Stark effects in
426 simulations of water absent simulations harder, these data show
427 that it is not a deal breaker. The classically calculated nitrile
428 spectrum shows changes in the spectra that will be easily seen through
429 experimental routes. It indicates a shifted peak lower in energy
430 should arise. This peak is a few wavenumbers from the larger peak and
431 almost 75 wavenumbers from the center. This seperation between the two
432 peaks means experimental results will have an easily resolved peak.
433
434 The Gaussian derived spectra do indicate that with an applied field
435 and subsiquent phase change does cause a narrowing of freuency
436 distrobution.
437 \section{Conclusions}
438 Field dependent changes in the phase of a system are
439 Jonathan K. Whitmer
440 cho stuff
441 \newpage
442
443 \bibliography{5CB}
444
445 \end{doublespace}
446 \end{document}