ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/5cb/5CB.tex
Revision: 4021
Committed: Tue Feb 4 22:54:16 2014 UTC (10 years, 4 months ago) by jmarr
Content type: application/x-tex
File size: 20489 byte(s)
Log Message:
eth

File Contents

# User Rev Content
1 gezelter 4007 \documentclass[journal = jpccck, manuscript = article]{achemso}
2     \setkeys{acs}{usetitle = true}
3    
4     \usepackage{caption}
5     \usepackage{float}
6     \usepackage{geometry}
7     \usepackage{natbib}
8     \usepackage{setspace}
9     \usepackage{xkeyval}
10     \usepackage{amsmath}
11     \usepackage{amssymb}
12     \usepackage{times}
13     \usepackage{mathptm}
14     \usepackage{setspace}
15 jmarr 4013 %\usepackage{endfloat}
16 gezelter 4007 \usepackage{tabularx}
17     \usepackage{longtable}
18     \usepackage{graphicx}
19     \usepackage{multirow}
20     \usepackage{multicol}
21     \usepackage{achemso}
22 jmarr 4013 \usepackage{subcaption}
23 gezelter 4007 \usepackage[colorinlistoftodos]{todonotes}
24     \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25     % \usepackage[square, comma, sort&compress]{natbib}
26     \usepackage{url}
27     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29     9.0in \textwidth 6.5in \brokenpenalty=10000
30    
31     % double space list of tables and figures
32     %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33     \setlength{\abovecaptionskip}{20 pt}
34     \setlength{\belowcaptionskip}{30 pt}
35    
36     % \bibpunct{}{}{,}{s}{}{;}
37    
38     %\citestyle{nature}
39     % \bibliographystyle{achemso}
40    
41    
42     \title{Nitrile vibrations as reporters of field-induced phase
43     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     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.
75     \end{abstract}
76    
77     \newpage
78    
79     \section{Introduction}
80    
81     The fundamental characteristic of liquid crystal mesophases is that
82     they maintain some degree of orientational order while translational
83     order is limited or absent. This orientational order produces a
84     complex direction-dependent response to external perturbations like
85     electric fields and mechanical distortions. The anisotropy of the
86     macroscopic phases originates in the anisotropy of the constituent
87     molecules, which typically have highly non-spherical structures with a
88     significant degree of internal rigidity. In nematic phases, rod-like
89     molecules are orientationally ordered with isotropic distributions of
90     molecular centers of mass, while in smectic phases, the molecules
91     arrange themselves into layers with their long (symmetry) axis normal
92     ($S_{A}$) or tilted ($S_{C}$) with respect to the layer planes.
93    
94     The behavior of the $S_{A}$ phase can be partially explained with
95     models mainly based on geometric factors and van der Waals
96     interactions. However, these simple models are insufficient to
97     describe liquid crystal phases which exhibit more complex polymorphic
98     nature. X-ray diffraction studies have shown that the ratio between
99     lamellar spacing ($s$) and molecular length ($l$) can take on a wide
100     range of values.\cite{Gray:1984hc,Leadbetter:1976vf,Hardouin:1980yq}
101     Typical $S_{A}$ phases have $s/l$ ratios on the order of $0.8$, while
102     for some compounds, e.g. the 4-alkyl-4'-cyanobiphenyls, the $s/l$
103     ratio is on the order of $1.4$. Molecules which form $S_{A}$ phases
104     can exhibit a wide variety of subphases like monolayers ($S_{A1}$),
105     uniform bilayers ($S_{A2}$), partial bilayers ($S_{\tilde A}$) as well
106     as interdigitated bilayers ($S_{A_{d}}$), and often have a terminal
107     cyano or nitro group. In particular lyotropic liquid crystals (those
108     exhibiting liquid crystal phase transition as a function of water
109     concentration) often have polar head groups or zwitterionic charge
110     separated groups that result in strong dipolar
111     interactions.\cite{Collings97} Because of their versatile polymorphic
112     nature, polar liquid crystalline materials have important
113     technological applications in addition to their immense relevance to
114     biological systems.\cite{Collings97}
115    
116     Experimental studies by Levelut {\it et al.}~\cite{Levelut:1981eu}
117     revealed that terminal cyano or nitro groups usually induce permanent
118     longitudinal dipole moments on the molecules.
119    
120     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
160     response, the atomic scale response has not been studied. With the
161     advent of nano-electrodes and coupling them with atomic force
162     microscopy, control of electric fields applied across nanometer
163     distances is now possible\cite{citation1}. This application of
164     nanometer length is interesting in the case of a nitrile group on the
165     molecule. While macroscopic fields are insufficient to cause a Stark
166     effect, small fields across nanometer-sized gaps are of sufficient
167     strength. If one were to assume a gap of 5 nm between a lower
168     electrode having a nanoelectrode placed near it via an atomic force
169     microscope, a field of 1 V applied across the electrodes would
170     translate into a field of 2x10\textsuperscript{8} $\frac{V}{M}$. This
171     field is theoretically strong enough to cause a phase change from
172     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.
175    
176     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.
177    
178     \section{Computational Details}
179     The force field was mainly taken from Guo et al.\cite{Zhang:2011hh} A
180     deviation from this force field was made to create a rigid body from
181     the phenyl rings. Bond distances within the rigid body were taken from
182     equilibrium bond distances. While the phenyl rings were held rigid,
183     bonds, bends, torsions and inversion centers still included the rings.
184    
185     Simulations were with boxes of 270 molecules locked at experimental
186     densities with periodic boundaries. The molecules were thermalized
187     from 0 kelvin to 300 kelvin. To equilibrate, each was first run in NVT
188     for 1 ns. This was followed by NVE for simulations used in the data
189     collection.
190    
191     External electric fields were applied in a simplistic charge-field
192     interaction. Forces were calculated by multiplying the electric field
193     being applied by the charge at each atom. For the potential, the
194     origin of the box was used as a point of reference. This allows for a
195     potential value to be added to each atom as the molecule move in space
196 jmarr 4008 within the box. Fields values were applied in a manner representing
197     those that would be applable in an experimental set-up. The assumed
198     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 jmarr 4017 Z-axis of the simulation box. For the simplicity of this paper,
203     each field will be called zero, partial and full, respectively.
204 gezelter 4007
205     For quantum calculation of the nitrile bond frequency, Gaussian 09 was
206     used. A single 5CB molecule was selected for the center of the
207     cluster. For effects from molecules located near the chosen nitrile
208     group, parts of molecules nearest to the nitrile group were
209 jmarr 4008 included. For the body not including the tail, any atom within 6~\AA
210 gezelter 4007 of the midpoint of the nitrile group was included. For the tail
211 jmarr 4008 structure, the whole tail was included if a tail atom was within 4~\AA
212 gezelter 4007 of the midpoint. If the tail did not include any atoms from the ring
213     structure, it was considered a propane molecule and included as
214     such. Once the clusters were generated, input files were created that
215 jmarr 4008 stretched the nitrile bond along its axis from 0.87 to 1.52~\AA at
216 gezelter 4007 increments of 0.05~\AA. This generated 13 single point energies to be
217     calculated. The Gaussian files were run with B3LYP/6-311++G(d,p) with
218 jmarr 4008 no other keywords for the zero field simulation. Simulations with
219     fields applied included the keyword ''Field=Z+5'' to match the
220     external field applied in molecular dynamic runs. Once completed, the central nitrile bond frequency
221 gezelter 4007 was calculated with a Morse fit. Using this fit and the solved energy
222 jmarr 4018 levels for a Morse oscillator, the frequency was found. Each set of
223 jmarr 4020 frequencies were then convolved together with a lorezian lineshape
224 jmarr 4018 function over each value. The width value used was 1.5. For the zero
225     field spectrum, 67 frequencies were used and for the full field, 59
226     frequencies were used.
227 gezelter 4007
228     Classical nitrile bond frequencies were found by replacing the rigid
229 jmarr 4008 cyanide bond with a flexible Morse oscillator bond
230     ($r_0= 1.157437$ \AA , $D_0 = 212.95$ and
231     $\beta = 2.67566$) . Once replaced, the
232 gezelter 4007 systems were allowed to re-equilibrate in NVT for 100 ps. After
233     re-equilibration, the system was run in NVE for 20 ps with a snapshot
234     spacing of 1 fs. These snapshot were then used in bond correlation
235     calculation to find the decay structure of the bond in time using the
236     average bond displacement in time,
237     \begin{equation}
238     C(t) = \langle \left(r(t) - r_0 \right) \cdot \left(r(0) - r_0 \right) \rangle
239     \end{equation}
240     %
241     where $r_0$ is the equilibrium bond distance and $r(t)$ is the
242     instantaneous bond displacement at time $t$. Once calculated,
243     smoothing was applied by adding an exponential decay on top of the
244     decay with a $\tau$ of 3000 (have to check this). Further smoothing
245     was applied by padding 20,000 zeros on each side of the symmetric
246     data. This was done five times by allowing the systems to run 1 ns
247     with a rigid bond followed by an equilibrium run with the bond
248     switched back on and the short production run.
249    
250     \section{Results}
251    
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 jmarr 4017 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 gezelter 4007
268 jmarr 4017 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 jmarr 4020 \begin{figure}
278     \includegraphics[trim = 5mm 10mm 3mm 10mm, clip, width=3.25in]{P2}
279     \caption{Ordering of each external field application over the course
280     of 60 ns with a sampling every 100 ps. Each trajectory was started
281     after equilibration with zero field applied.}
282     \label{fig:orderParameter}
283     \end{figure}
284 jmarr 4017
285     Interestingly, the field that is needed to switch the phase of 5CB
286     macroscopically is larger than 1 V. However, in this case, only a
287     voltage of 1.2 V was need to induce a phase change. This is impart due
288     to the distance the field is being applied across. At such a small
289     distance, the field is much larger than the macroscopic and thus
290     easily induces a field dependent phase change.
291    
292 jmarr 4020 In the figure below, this phase change is represented nicely as
293     ellipsoids that are created by the vector formed between the nitrogen
294     of the nitrile group and the tail terminal carbon atom. These
295     illistrate the change from isotropic phase to nematic change. Both the
296     zero field and partial field images look mostly disordered. The
297     partial field does look somewhat orded but without any overall order
298     of the entire system. This is most likely a high point in the ordering
299     for the trajectory. The full field image on the other hand looks much
300     more ordered when compared to the two lower field simulations.
301     \begin{figure}
302     \includegraphics[width=7in]{Elip_3}
303     \caption{Ellipsoid reprsentation of 5CB at three different
304     field strengths, Zero Field (Left), Partial Field (Middle), and Full
305     Field (Right) Each image was created by taking the final
306     snapshot of each 60 ns run}
307     \label{fig:Cigars}
308     \end{figure}
309    
310 jmarr 4017 This change in phase was followed by two courses of further
311 jmarr 4019 analysis. First was the replacement of the static nitrile bond with a
312 jmarr 4017 morse oscillator bond. This was then simulated for a period of time
313     and a classical spetrum was calculated. Second, ab intio calcualtions were performe to investigate
314 jmarr 4019 if the phase change caused any change spectrum through quantum
315 jmarr 4017 effects.
316    
317 jmarr 4019 The classical nitrile spectrum can be seen in Figure 2. Most noticably
318     is the position of the two peaks. Obviously the experimental peak
319     position is near 2226 cm\textsuperscript{-1}. However, in this case
320     the peak position is shifted to the blue at a position of 2375
321     cm\textsuperscript{-1}. This shift is due solely to the choice of
322     oscillator strength in the Mores oscillator parameters. While this
323     shift makes the two spectra differ, it does not affect the ability to
324     compare peak changes to experimental peak changes.
325     With this important fact out of the way, differences between the two
326     states are subtle but are very much present. The first and
327     most notable is the apperance for a strong band near 2300
328 jmarr 4020 cm\textsuperscript{-1}.
329 jmarr 4013 \begin{figure}
330     \includegraphics[width=3.25in]{2Spectra}
331 jmarr 4017 \caption{The classically calculated nitrile bond spetrum for no
332     external field application (black) and full external field
333     application (red)}
334 jmarr 4013 \label{fig:twoSpectra}
335     \end{figure}
336 jmarr 4020
337 jmarr 4021 Before Gaussian silumations were carried out, it was attempt to apply
338     the method developed by Cho et. al. This method involves the fitting
339     of multiple parameters to However, since these simulations
340     are done under the presence of external electric fields and in the
341     absence of water the equations had to be reworked. Originally, the
342     nitrile bond frequency was related to the potential of water around
343     the bond via
344     \begin{equation}
345     \phi_{a} = \frac{1}{4\pi \epsilon_{0}} \sum_{m} \sum_{j}
346     \frac{C^{H_{2}O}_{j \left(m \right) }}{r_{aj \left(m\right)}}
347     \end{equation}
348 jmarr 4020 After Gaussian calculations were performed on a set of snapshots for
349     the zero and full field simualtions, they were first investigated for
350     any dependence on the local, with external field included, electric
351     field. This was to see if a linear or non-linear relationship between
352     the two could be utilized for generating spectra. This was done in
353     part because of previous studies showing the frequency dependence of
354     nitrile bonds to the electric fields generated locally between
355     solvating water. It was seen that little to no dependence could be
356     directly shown. This data is not shown.
357    
358     Since no explicit dependence was observed between the calculated
359     frequency and the electric field, it was not a viable route for the
360     calculation of a nitrile spectrum. Instead, the frequencies were taken
361     and convolved together. These two spectra are seen below in Figure
362     4. While the spectrum without a field is lower in intensity and is
363     almost bimodel in distrobuiton, the external field spectrum is much
364     more unimodel. This narrowing has the affect of increasing the
365     intensity around 2226 cm\textsuperscript{-1} where the peak is
366     centered. The external field also has fewer frequencies higher to the
367     blue of the spectra. Unlike the the zero field, where some frequencies reach as high
368     as 2280 cm\textsuperscript{-1}.
369 jmarr 4013 \begin{figure}
370 jmarr 4018 \includegraphics[width=3.25in]{Convolved}
371 jmarr 4020 \caption{Lorentzian convolved Gaussian frequencies of the zero field
372     system (black) and the full field system (red)}
373 jmarr 4018 \label{fig:Con}
374     \end{figure}
375 gezelter 4007 \section{Discussion}
376 jmarr 4020 The absence of any electric field dependency of the freuquency with
377     the Gaussian simulations is strange. A large base of research has been
378     built upon the known tuning the nitrile bond as local field
379     changes. This differences may be due to the absence of water. Many of
380     the nitrile bond fitting maps are done in the presence of
381     water. Liquid water is known to have a very high internal field which
382     is much larger than the internal fields of neat 5CB. Even though the
383     application of running Gaussian simulations followed by mappying to
384     some classical parameter is easy and straight forward, this system
385     illistrates how that 'go to' method can break down.
386 gezelter 4007
387 jmarr 4020 While this makes the application of nitrile Stark effects in
388     simulations of liquid water absent simulations harder, these data show
389 jmarr 4021 that it is not a deal breaker. The classically calculated nitrile
390     spectrum shows changes in the spectra that will be easily seen through
391     experimental routes. It indicates a shifted peak lower in energy
392     should arise. This peak is a few wavenumbers from the larger peak and
393     almost 75 wavenmubers from the center. This seperation between the two
394     peaks means experimental results will show a well resolved peak.
395    
396     The Gaussian derived frequencies and subsiquent spectra also indicate
397     changes that can be observed experimentally.
398 gezelter 4007 \section{Conclusions}
399 jmarr 4021 Jonathan K. Whitmer
400     cho stuff
401 gezelter 4007 \newpage
402    
403     \bibliography{5CB}
404    
405     \end{doublespace}
406     \end{document}