ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 757
Committed: Wed Sep 10 15:38:43 2003 UTC (20 years, 9 months ago) by chrisfen
Content type: application/x-tex
File size: 48730 byte(s)
Log Message:
First CVS commit.  Minor changes in the introduction.  More changes to follow...

Chris

File Contents

# Content
1 \documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 %\documentclass[prb,aps,times,tabularx,preprint]{revtex4}
3 \usepackage{amsmath}
4 \usepackage{graphicx}
5
6 %\usepackage{endfloat}
7 %\usepackage{berkeley}
8 %\usepackage{epsf}
9 %\usepackage[ref]{overcite}
10 %\usepackage{setspace}
11 %\usepackage{tabularx}
12 %\usepackage{graphicx}
13 %\usepackage{curves}
14 %\usepackage{amsmath}
15 %\pagestyle{plain}
16 %\pagenumbering{arabic}
17 %\oddsidemargin 0.0cm \evensidemargin 0.0cm
18 %\topmargin -21pt \headsep 10pt
19 %\textheight 9.0in \textwidth 6.5in
20 %\brokenpenalty=10000
21 %\renewcommand{\baselinestretch}{1.2}
22 %\renewcommand\citemid{\ } % no comma in optional reference note
23
24 \begin{document}
25
26 \title{On the temperature dependent structural and transport properties of the soft sticky dipole (SSD) and related single point water models}
27
28 \author{Christopher J. Fennell and J. Daniel Gezelter{\thefootnote}
29 \footnote[1]{Corresponding author. \ Electronic mail: gezelter@nd.edu}}
30
31 \address{Department of Chemistry and Biochemistry\\ University of Notre Dame\\
32 Notre Dame, Indiana 46556}
33
34 \date{\today}
35
36 \begin{abstract}
37 NVE and NPT molecular dynamics simulations were performed in order to
38 investigate the density maximum and temperature dependent transport
39 for the SSD water model, both with and without the use of reaction
40 field. The constant pressure simulations of the melting of both $I_h$
41 and $I_c$ ice showed a density maximum near 260 K. In most cases, the
42 calculated densities were significantly lower than the densities
43 calculated in simulations of other water models. Analysis of particle
44 diffusion showed SSD to capture the transport properties of
45 experimental very well in both the normal and super-cooled liquid
46 regimes. In order to correct the density behavior, SSD was
47 reparameterized for use both with and without a long-range interaction
48 correction, SSD/RF and SSD/E respectively. In addition to correcting
49 the abnormally low densities, these new versions were show to maintain
50 or improve upon the transport and structural features of the original
51 water model.
52 \end{abstract}
53
54 \maketitle
55
56 %\narrowtext
57
58
59 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60 % BODY OF TEXT
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62
63 \section{Introduction}
64
65 One of the most important tasks in simulations of biochemical systems
66 is the proper depiction of water and water solvation. In fact, the
67 bulk of the calculations performed in solvated simulations are of
68 interactions with or between solvent molecules. Thus, the outcomes of
69 these types of simulations are highly dependent on the physical
70 properties of water, both as individual molecules and in
71 groups/bulk. Due to the fact that explicit solvent accounts for a
72 massive portion of the calculations, it necessary to simplify the
73 solvent to some extent in order to complete simulations in a
74 reasonable amount of time. In the case of simulating water in
75 bio-molecular studies, the balance between accurate properties and
76 computational efficiency is especially delicate, and it has resulted
77 in a variety of different water
78 models.\cite{Jorgensen83,Berendsen87,Jorgensen00} Many of these models
79 get specific properties correct or better than their predecessors, but
80 this is often at a cost of some other properties or of computer
81 time. As an example, compare TIP3P or TIP4P to TIP5P. TIP5P succeeds
82 in improving the structural and transport properties over its
83 predecessors, yet this comes at a greater than 50\% increase in
84 computational cost.\cite{Jorgensen01,Jorgensen00} One recently
85 developed model that succeeds in both retaining accuracy of system
86 properties and simplifying calculations to increase computational
87 efficiency is the Soft Sticky Dipole water model.\cite{Ichiye96}
88
89 The Soft Sticky Dipole (SSD)\ water model was developed by Ichiye
90 \emph{et al.} as a modified form of the hard-sphere water model
91 proposed by Bratko, Blum, and Luzar.\cite{Bratko85,Bratko95} SSD
92 consists of a single point dipole with a Lennard-Jones core and a
93 sticky potential that directs the particles to assume the proper
94 hydrogen bond orientation in the first solvation shell. Thus, the
95 interaction between two SSD water molecules \emph{i} and \emph{j} is
96 given by the potential
97 \begin{equation}
98 u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
99 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
100 u_{ij}^{sp}
101 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
102 \end{equation}
103 where the $\mathbf{r}_{ij}$ is the position vector between molecules
104 \emph{i} and \emph{j} with magnitude equal to the distance $r_ij$, and
105 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
106 orientations of the respective molecules. The Lennard-Jones, dipole,
107 and sticky parts of the potential are giving by the following
108 equations,
109 \begin{equation}
110 u_{ij}^{LJ}(r_{ij}) = 4\epsilon \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right],
111 \end{equation}
112 \begin{equation}
113 u_{ij}^{dp} = \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r_{ij}^3}-\frac{3(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})(\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r_{ij}^5}\ ,
114 \end{equation}
115 \begin{equation}
116 \begin{split}
117 u_{ij}^{sp}
118 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)
119 &=
120 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\\
121 & \quad \ +
122 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
123 \end{split}
124 \end{equation}
125 where $\boldsymbol{\mu}_i$ and $\boldsymbol{\mu}_j$ are the dipole
126 unit vectors of particles \emph{i} and \emph{j} with magnitude 2.35 D,
127 $\nu_0$ scales the strength of the overall sticky potential, $s$ and
128 $s^\prime$ are cubic switching functions. The $w$ and $w^\prime$
129 functions take the following forms,
130 \begin{equation}
131 w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
132 \end{equation}
133 \begin{equation}
134 w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
135 \end{equation}
136 where $w^0 = 0.07715$. The $w$ function is the tetrahedral attractive
137 term that promotes hydrogen bonding orientations within the first
138 solvation shell, and $w^\prime$ is a dipolar repulsion term that
139 repels unrealistic dipolar arrangements within the first solvation
140 shell. A more detailed description of the functional parts and
141 variables in this potential can be found in other
142 articles.\cite{Ichiye96,Ichiye99}
143
144 Being that this is a one-site point dipole model, the actual force
145 calculations are simplified significantly. In the original Monte Carlo
146 simulations using this model, Ichiye \emph{et al.} reported a
147 calculation speed up of up to an order of magnitude over other
148 comparable models while maintaining the structural behavior of
149 water.\cite{Ichiye96} In the original molecular dynamics studies of
150 SSD, it was shown that it actually improves upon the prediction of
151 water's dynamical properties 3 and 4-point models.\cite{Ichiye99} This
152 attractive combination of speed and accurate depiction of solvent
153 properties makes SSD a model of interest for the simulation of large
154 scale biological systems, such as membrane phase behavior, a specific
155 interest within our group.
156
157 One of the key limitations of this water model, however, is that it
158 has been parameterized for use with the Ewald Sum technique for the
159 handling of long-ranged interactions. When studying very large
160 systems, the Ewald summation and even particle-mesh Ewald become
161 computational burdens with their respective ideal $N^\frac{3}{2}$ and
162 $N\log N$ calculation scaling orders for $N$ particles.\cite{Darden99}
163
164 Up to this point, a detailed look at the model's structure and ion
165 solvation abilities has been performed.\cite{Ichiye96} In addition, a
166 thorough investigation of the dynamic properties of SSD was performed
167 by Chandra and Ichiye focusing on translational and orientational
168 properties at 298 K.\cite{Ichiye99} This study focuses on determining
169 the density maximum for SSD utilizing both microcanonical and
170 isobaric-isothermal ensemble molecular dynamics, while using the
171 reaction field method for handling long-ranged dipolar interactions. A
172 reaction field method has been previously implemented in Monte Carlo
173 simulations by Liu and Ichiye in order to study the static dielectric
174 constant for the model.\cite{Ichiye96b} This paper will expand the
175 scope of these original simulations to look on how the reaction field
176 affects the physical and dynamic properties of SSD systems.
177
178 \section{Methods}
179
180 As stated previously, in this study the long-range dipole-dipole
181 interactions were accounted for using the reaction field method. The
182 magnitude of the reaction field acting on dipole \emph{i} is given by
183 \begin{equation}
184 \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
185 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} \boldsymbol{\mu}_{j} f(r_{ij})\ ,
186 \label{rfequation}
187 \end{equation}
188 where $\mathcal{R}$ is the cavity defined by the cutoff radius
189 ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
190 system (80 in this case), $\boldsymbol{\mu}_{j}$ is the dipole moment
191 vector of particle \emph{j}, and $f(r_{ij})$ is a cubic switching
192 function.\cite{AllenTildesley} The reaction field contribution to the
193 total energy by particle \emph{i} is given by
194 $-\frac{1}{2}\boldsymbol{\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
195 on dipole \emph{i} by
196 $\boldsymbol{\mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use
197 of reaction field is known to alter the orientational dynamic
198 properties, such as the dielectric relaxation time, based on changes
199 in the length of the cutoff radius.\cite{Berendsen98} This variable
200 behavior makes reaction field a less attractive method than other
201 methods, like the Ewald summation; however, for the simulation of
202 large-scale system, the computational cost benefit of reaction field
203 is dramatic. To address some of the dynamical property alterations due
204 to the use of reaction field, simulations were also performed without
205 a surrounding dielectric and suggestions are proposed on how to make
206 SSD more compatible with a reaction field.
207
208 Simulations were performed in both the isobaric-isothermal and
209 microcanonical ensembles. The constant pressure simulations were
210 implemented using an integral thermostat and barostat as outlined by
211 Hoover.\cite{Hoover85,Hoover86} For the constant pressure
212 simulations, the \emph{Q} parameter for the was set to 5.0 amu
213 \(\cdot\)\AA\(^{2}\), and the relaxation time (\(\tau\))\ was set at
214 100 ps.
215
216 Integration of the equations of motion was carried out using the
217 symplectic splitting method proposed by Dullweber \emph{et
218 al.}.\cite{Dullweber1997} The reason for this integrator selection
219 deals with poor energy conservation of rigid body systems using
220 quaternions. While quaternions work well for orientational motion in
221 alternate ensembles, the microcanonical ensemble has a constant energy
222 requirement that is actually quite sensitive to errors in the
223 equations of motion. The original implementation of this code utilized
224 quaternions for rotational motion propagation; however, a detailed
225 investigation showed that they resulted in a steady drift in the total
226 energy, something that has been observed by others.\cite{Laird97}
227
228 The key difference in the integration method proposed by Dullweber
229 \emph{et al.} is that the entire rotation matrix is propagated from
230 one time step to the next. In the past, this would not have been as
231 feasible a option, being that the rotation matrix for a single body is
232 nine elements long as opposed to 3 or 4 elements for Euler angles and
233 quaternions respectively. System memory has become much less of an
234 issue in recent times, and this has resulted in substantial benefits
235 in energy conservation. There is still the issue of an additional 5 or
236 6 additional elements for describing the orientation of each particle,
237 which will increase dump files substantially. Simply translating the
238 rotation matrix into its component Euler angles or quaternions for
239 storage purposes relieves this burden.
240
241 The symplectic splitting method allows for Verlet style integration of
242 both linear and angular motion of rigid bodies. In the integration
243 method, the orientational propagation involves a sequence of matrix
244 evaluations to update the rotation matrix.\cite{Dullweber1997} These
245 matrix rotations end up being more costly computationally than the
246 simpler arithmetic quaternion propagation. On average, a 1000 SSD
247 particle simulation shows a 7\% increase in simulation time using the
248 symplectic step method in place of quaternions. This cost is more than
249 justified when comparing the energy conservation of the two methods as
250 illustrated in figure \ref{timestep}.
251
252 \begin{figure}
253 \includegraphics[width=61mm, angle=-90]{timeStep.epsi}
254 \caption{Energy conservation using quaternion based integration versus
255 the symplectic step method proposed by Dullweber \emph{et al.} with
256 increasing time step. For each time step, the dotted line is total
257 energy using the symplectic step integrator, and the solid line comes
258 from the quaternion integrator. The larger time step plots are shifted
259 up from the true energy baseline for clarity.}
260 \label{timestep}
261 \end{figure}
262
263 In figure \ref{timestep}, the resulting energy drift at various time
264 steps for both the symplectic step and quaternion integration schemes
265 is compared. All of the 1000 SSD particle simulations started with the
266 same configuration, and the only difference was the method for
267 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
268 methods for propagating particle rotation conserve energy fairly well,
269 with the quaternion method showing a slight energy drift over time in
270 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
271 energy conservation benefits of the symplectic step method are clearly
272 demonstrated.
273
274 Energy drift in these SSD particle simulations was unnoticeable for
275 time steps up to three femtoseconds. A slight energy drift on the
276 order of 0.012 kcal/mol per nanosecond was observed at a time step of
277 four femtoseconds, and as expected, this drift increases dramatically
278 with increasing time step. To insure accuracy in the constant energy
279 simulations, time steps were set at 2 fs and kept at this value for
280 constant pressure simulations as well.
281
282 Ice crystals in both the $I_h$ and $I_c$ lattices were generated as
283 starting points for all the simulations. The $I_h$ crystals were
284 formed by first arranging the center of masses of the SSD particles
285 into a ``hexagonal'' ice lattice of 1024 particles. Because of the
286 crystal structure of $I_h$ ice, the simulation box assumed a
287 rectangular shape with a edge length ratio of approximately
288 1.00$\times$1.06$\times$1.23. The particles were then allowed to
289 orient freely about fixed positions with angular momenta randomized at
290 400 K for varying times. The rotational temperature was then scaled
291 down in stages to slowly cool the crystals down to 25 K. The particles
292 were then allowed translate with fixed orientations at a constant
293 pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
294 removed and the ice crystals were allowed to equilibrate for 50 ps at
295 25 K and a constant pressure of 1 atm. This procedure resulted in
296 structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
297 rules\cite{Bernal33,Rahman72}. This method was also utilized in the
298 making of diamond lattice $I_c$ ice crystals, with each cubic
299 simulation box consisting of either 512 or 1000 particles. Only
300 isotropic volume fluctuations were performed under constant pressure,
301 so the ratio of edge lengths remained constant throughout the
302 simulations.
303
304 \section{Results and discussion}
305
306 Melting studies were performed on the randomized ice crystals using
307 constant pressure and temperature dynamics. This involved an initial
308 randomization of velocities about the starting temperature of 25 K for
309 varying amounts of time. The systems were all equilibrated for 100 ps
310 prior to a 200 ps data collection run at each temperature setting,
311 ranging from 25 to 400 K, with a maximum degree increment of 25 K. For
312 regions of interest along this stepwise progression, the temperature
313 increment was decreased from 25 K to 10 and then 5 K. The above
314 equilibration and production times were sufficient in that the system
315 volume fluctuations dampened out in all but the very cold simulations
316 (below 225 K). In order to further improve statistics, five separate
317 simulation progressions were performed, and the averaged results from
318 the $I_h$ melting simulations are shown in figure \ref{dense1}.
319
320 \begin{figure}
321 \includegraphics[width=65mm, angle=-90]{1hdense.epsi}
322 \caption{Average density of SSD water at increasing temperatures
323 starting from ice $I_h$ lattice.}
324 \label{dense1}
325 \end{figure}
326
327 \subsection{Density Behavior}
328 In the initial average density versus temperature plot, the density
329 maximum clearly appears between 255 and 265 K. The calculated
330 densities within this range were nearly indistinguishable, as can be
331 seen in the zoom of this region of interest, shown in figure
332 \ref{dense1}. The greater certainty of the average value at 260 K makes
333 a good argument for the actual density maximum residing at this
334 midpoint value. Figure \ref{dense1} was constructed using ice $I_h$
335 crystals for the initial configuration; and though not pictured, the
336 simulations starting from ice $I_c$ crystal configurations showed
337 similar results, with a liquid-phase density maximum in this same
338 region (between 255 and 260 K). In addition, the $I_c$ crystals are
339 more fragile than the $I_h$ crystals, leading them to deform into a
340 dense glassy state at lower temperatures. This resulted in an overall
341 low temperature density maximum at 200 K, but they still retained a
342 common liquid state density maximum with the $I_h$ simulations.
343
344 \begin{figure}
345 \includegraphics[width=65mm,angle=-90]{dense2.eps}
346 \caption{Density versus temperature for TIP4P\cite{Jorgensen98b},
347 TIP3P\cite{Jorgensen98b}, SPC/E\cite{Clancy94}, SSD without Reaction
348 Field, SSD, and Experiment\cite{CRC80}. }
349 \label{dense2}
350 \end{figure}
351
352 The density maximum for SSD actually compares quite favorably to other
353 simple water models. Figure \ref{dense2} shows a plot of these
354 findings with the density progression of several other models and
355 experiment obtained from other
356 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
357 models, SSD has results closest to the experimentally observed water
358 density maximum. Of the listed water models, TIP4P has a density
359 maximum behavior most like that seen in SSD. Though not shown, it is
360 useful to note that TIP5P has a water density maximum nearly identical
361 to experiment.
362
363 Possibly of more importance is the density scaling of SSD relative to
364 other common models at any given temperature (Fig. \ref{dense2}). Note
365 that the SSD model assumes a lower density than any of the other
366 listed models at the same pressure, behavior which is especially
367 apparent at temperatures greater than 300 K. Lower than expected
368 densities have been observed for other systems with the use of a
369 reaction field for long-range electrostatic interactions, so the most
370 likely reason for these significantly lower densities in these
371 simulations is the presence of the reaction field.\cite{Berendsen98}
372 In order to test the effect of the reaction field on the density of
373 the systems, the simulations were repeated for the temperature region
374 of interest without a reaction field present. The results of these
375 simulations are also displayed in figure \ref{dense2}. Without
376 reaction field, these densities increase considerably to more
377 experimentally reasonable values, especially around the freezing point
378 of liquid water. The shape of the curve is similar to the curve
379 produced from SSD simulations using reaction field, specifically the
380 rapidly decreasing densities at higher temperatures; however, a slight
381 shift in the density maximum location, down to 245 K, is
382 observed. This is probably a more accurate comparison to the other
383 listed water models in that no long range corrections were applied in
384 those simulations.\cite{Clancy94,Jorgensen98b}
385
386 It has been observed that densities are dependent on the cutoff radius
387 used for a variety of water models in simulations both with and
388 without the use of reaction field.\cite{Berendsen98} In order to
389 address the possible affect of cutoff radius, simulations were
390 performed with a dipolar cutoff radius of 12.0 \AA\ to compliment the
391 previous SSD simulations, all performed with a cutoff of 9.0 \AA. All
392 the resulting densities overlapped within error and showed no
393 significant trend in lower or higher densities as a function of cutoff
394 radius, both for simulations with and without reaction field. These
395 results indicate that there is no major benefit in choosing a longer
396 cutoff radius in simulations using SSD. This is comforting in that the
397 use of a longer cutoff radius results in a near doubling of the time
398 required to compute a single trajectory.
399
400 \subsection{Transport Behavior}
401 Of importance in these types of studies are the transport properties
402 of the particles and how they change when altering the environmental
403 conditions. In order to probe transport, constant energy simulations
404 were performed about the average density uncovered by the constant
405 pressure simulations. Simulations started with randomized velocities
406 and underwent 50 ps of temperature scaling and 50 ps of constant
407 energy equilibration before obtaining a 200 ps trajectory. Diffusion
408 constants were calculated via root-mean square deviation analysis. The
409 averaged results from 5 sets of these NVE simulations is displayed in
410 figure \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
411 results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
412
413 \begin{figure}
414 \includegraphics[width=65mm, angle=-90]{betterDiffuse.epsi}
415 \caption{Average diffusion coefficient over increasing temperature for
416 SSD, SPC/E\cite{Clancy94}, TIP5P\cite{Jorgensen01}, and Experimental
417 data from Gillen \emph{et al.}\cite{Gillen72}, and from
418 Mills\cite{Mills73}.}
419 \label{diffuse}
420 \end{figure}
421
422 The observed values for the diffusion constant point out one of the
423 strengths of the SSD model. Of the three experimental models shown,
424 the SSD model has the most accurate depiction of the diffusion trend
425 seen in experiment in both the supercooled and normal regimes. SPC/E
426 does a respectable job by getting similar values as SSD and experiment
427 around 290 K; however, it deviates at both higher and lower
428 temperatures, failing to predict the experimental trend. TIP5P and SSD
429 both start off low at the colder temperatures and tend to diffuse too
430 rapidly at the higher temperatures. This type of trend at the higher
431 temperatures is not surprising in that the densities of both TIP5P and
432 SSD are lower than experimental water at temperatures higher than room
433 temperature. When calculating the diffusion coefficients for SSD at
434 experimental densities, the resulting values fall more in line with
435 experiment at these temperatures, albeit not at standard
436 pressure. Results under these conditions can be found later in this
437 paper.
438
439 \subsection{Structural Changes and Characterization}
440 By starting the simulations from the crystalline state, the melting
441 transition and the ice structure can be studied along with the liquid
442 phase behavior beyond the melting point. To locate the melting
443 transition, the constant pressure heat capacity (C$_\text{p}$) was
444 monitored in each of the simulations. In the melting simulations of
445 the 1024 particle ice $I_h$ simulations, a large spike in C$_\text{p}$
446 occurs at 245 K, indicating a first order phase transition for the
447 melting of these ice crystals. When the reaction field is turned off,
448 the melting transition occurs at 235 K. These melting transitions are
449 considerably lower than the experimental value, but this is not
450 surprising in that SSD is a simple rigid body model with a fixed
451 dipole.
452
453 \begin{figure}
454 \includegraphics[width=85mm]{fullContours.eps}
455 \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
456 100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
457 clarity: dark areas signify peaks while light areas signify
458 depressions. White areas have g(\emph{r}) values below 0.5 and black
459 areas have values above 1.5.}
460 \label{contour}
461 \end{figure}
462
463 Additional analyses for understanding the melting phase-transition
464 process were performed via two-dimensional structure and dipole angle
465 correlations. Expressions for the correlations are as follows:
466
467 \begin{figure}
468 \includegraphics[width=45mm]{corrDiag.eps}
469 \caption{Two dimensional illustration of the angles involved in the
470 correlations observed in figure \ref{contour}.}
471 \label{corrAngle}
472 \end{figure}
473
474 \begin{multline}
475 g_{\text{AB}}(r,\cos\theta) = \\
476 \frac{V}{N_\text{A}N_\text{B}}\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\theta-\cos\theta_{ij})\delta(r-\left|\mathbf{r}_{ij}\right|)\rangle\ ,
477 \end{multline}
478 \begin{multline}
479 g_{\text{AB}}(r,\cos\omega) = \\
480 \frac{V}{N_\text{A}N_\text{B}}\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\omega-\cos\omega_{ij})\delta(r-\left|\mathbf{r}_{ij}\right|)\rangle\ ,
481 \end{multline}
482 where $\theta$ and $\omega$ refer to the angles shown in the above
483 illustration. By binning over both distance and the cosine of the
484 desired angle between the two dipoles, the g(\emph{r}) can be
485 dissected to determine the common dipole arrangements that constitute
486 the peaks and troughs. Frames A and B of figure \ref{contour} show a
487 relatively crystalline state of an ice $I_c$ simulation. The first
488 peak of the g(\emph{r}) primarily consists of the preferred hydrogen
489 bonding arrangements as dictated by the tetrahedral sticky potential,
490 one peak for the donating and the other for the accepting hydrogen
491 bonds. Due to the high degree of crystallinity of the sample, the
492 second and third solvation shells show a repeated peak arrangement
493 which decays at distances around the fourth solvation shell, near the
494 imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
495 In the higher temperature simulation shown in frames C and D, the
496 repeated peak features are significantly blurred. The first solvation
497 shell still shows the strong effect of the sticky-potential, although
498 it covers a larger area, extending to include a fraction of aligned
499 dipole peaks within the first solvation shell. The latter peaks lose
500 definition as thermal motion and the competing dipole force overcomes
501 the sticky potential's tight tetrahedral structuring of the fluid.
502
503 This complex interplay between dipole and sticky interactions was
504 remarked upon as a possible reason for the split second peak in the
505 oxygen-oxygen g(\emph{r}).\cite{Ichiye96} At low temperatures, the
506 second solvation shell peak appears to have two distinct parts that
507 blend together to form one observable peak. At higher temperatures,
508 this split character alters to show the leading 4 \AA\ peak dominated
509 by equatorial anti-parallel dipole orientations, and there is tightly
510 bunched group of axially arranged dipoles that most likely consist of
511 the smaller fraction aligned dipole pairs. The trailing part of the
512 split peak at 5 \AA\ is dominated by aligned dipoles that range
513 primarily within the axial to the chief hydrogen bond arrangements
514 similar to those seen in the first solvation shell. This evidence
515 indicates that the dipole pair interaction begins to dominate outside
516 of the range of the dipolar repulsion term, with the primary
517 energetically favorable dipole arrangements populating the region
518 immediately outside of it's range (around 4 \AA), and arrangements
519 that seek to ideally satisfy both the sticky and dipole forces locate
520 themselves just beyond this region (around 5 \AA).
521
522 From these findings, the split second peak is primarily the product of
523 the dipolar repulsion term of the sticky potential. In fact, the
524 leading of the two peaks can be pushed out and merged with the outer
525 split peak just by extending the switching function cutoff
526 ($s^\prime(r_{ij})$) from its normal 4.0 \AA\ to values of 4.5 or even
527 5 \AA. This type of correction is not recommended for improving the
528 liquid structure, because the second solvation shell will still be
529 shifted too far out. In addition, this would have an even more
530 detrimental effect on the system densities, leading to a liquid with a
531 more open structure and a density considerably lower than the normal
532 SSD behavior shown previously. A better correction would be to include
533 the quadrupole-quadrupole interactions for the water particles outside
534 of the first solvation shell, but this reduces the simplicity and
535 speed advantage of SSD, so it is not the most desirable path to take.
536
537 \subsection{Adjusted Potentials: SSD/E and SSD/RF}
538 The propensity of SSD to adopt lower than expected densities under
539 varying conditions is troubling, especially at higher temperatures. In
540 order to correct this behavior, it's necessary to adjust the force
541 field parameters for the primary intermolecular interactions. In
542 undergoing a reparameterization, it is important not to focus on just
543 one property and neglect the other important properties. In this case,
544 it would be ideal to correct the densities while maintaining the
545 accurate transport properties.
546
547 The possible parameters for tuning include the $\sigma$ and $\epsilon$
548 Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
549 attractive and dipole repulsive terms with their respective
550 cutoffs. To alter the attractive and repulsive terms of the sticky
551 potential independently, it is necessary to separate the terms as
552 follows:
553 \begin{equation}
554 \begin{split}
555 u_{ij}^{sp}
556 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) &=
557 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\\
558 & \quad \ + \frac{\nu_0^\prime}{2}
559 [s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)],
560 \end{split}
561 \end{equation}
562
563 where $\nu_0$ scales the strength of the tetrahedral attraction and
564 $\nu_0^\prime$ acts in an identical fashion on the dipole repulsion
565 term. For purposes of the reparameterization, the separation was
566 performed, but the final parameters were adjusted so that it is
567 unnecessary to separate the terms when implementing the adjusted water
568 potentials. The results of the reparameterizations are shown in table
569 \ref{params}. Note that both the tetrahedral attractive and dipolar
570 repulsive don't share the same lower cutoff ($r_l$) in the newly
571 parameterized potentials - soft sticky dipole enhanced (SSD/E) and
572 soft sticky dipole reaction field (SSD/RF).
573
574 \begin{table}
575 \caption{Parameters for the original and adjusted models}
576 \begin{tabular}{ l c c c }
577 \hline \\[-3mm]
578 \ Parameters & \ \ \ SSD$^\dagger$\ \ \ \ & \ SSD/E\ \ & \ SSD/RF\ \ \\
579 \hline \\[-3mm]
580 \ \ \ $\sigma$ (\AA) & 3.051 & 3.035 & 3.019\\
581 \ \ \ $\epsilon$ (kcal/mol)\ \ & 0.152 & 0.152 & 0.152\\
582 \ \ \ $\mu$ (D) & 2.35 & 2.418 & 2.480\\
583 \ \ \ $\nu_0$ (kcal/mol)\ \ & 3.7284 & 3.90 & 3.90\\
584 \ \ \ $r_l$ (\AA) & 2.75 & 2.40 & 2.40\\
585 \ \ \ $r_u$ (\AA) & 3.35 & 3.80 & 3.80\\
586 \ \ \ $\nu_0^\prime$ (kcal/mol)\ \ & 3.7284 & 3.90 & 3.90\\
587 \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75\\
588 \ \ \ $r_u^\prime$ (\AA) & 4.00 & 3.35 & 3.35\\
589 \\[-2mm]$^\dagger$ ref. \onlinecite{Ichiye96}
590 \end{tabular}
591 \label{params}
592 \end{table}
593
594 \begin{figure}
595 \includegraphics[width=85mm]{gofrCompare.epsi}
596 \caption{Plots comparing experiment\cite{Head-Gordon00_1} with SSD/E
597 and SSD without reaction field (top), as well as SSD/RF and SSD with
598 reaction field turned on (bottom). The insets show the respective
599 first peaks in detail. Solid Line - experiment, dashed line - SSD/E
600 and SSD/RF, and dotted line - SSD (with and without reaction field).}
601 \label{grcompare}
602 \end{figure}
603
604 \begin{figure}
605 \includegraphics[width=85mm]{dualsticky.ps}
606 \caption{Isosurfaces of the sticky potential for SSD (left) and SSD/E \&
607 SSD/RF (right). Light areas correspond to the tetrahedral attractive
608 part, and the darker areas correspond to the dipolar repulsive part.}
609 \label{isosurface}
610 \end{figure}
611
612 In the paper detailing the development of SSD, Liu and Ichiye placed
613 particular emphasis on an accurate description of the first solvation
614 shell. This resulted in a somewhat tall and sharp first peak that
615 integrated to give similar coordination numbers to the experimental
616 data obtained by Soper and Phillips.\cite{Ichiye96,Soper86} New
617 experimental x-ray scattering data from the Head-Gordon lab indicates
618 a slightly lower and shifted first peak in the g$_\mathrm{OO}(r)$, so
619 adjustments to SSD were made while taking into consideration the new
620 experimental findings.\cite{Head-Gordon00_1} Figure \ref{grcompare}
621 shows the relocation of the first peak of the oxygen-oxygen
622 g(\emph{r}) by comparing the original SSD (with and without reaction
623 field), SSD-E, and SSD-RF to the new experimental results. Both the
624 modified water models have shorter peaks that are brought in more
625 closely to the experimental peak (as seen in the insets of figure
626 \ref{grcompare}). This structural alteration was accomplished by a
627 reduction in the Lennard-Jones $\sigma$ variable as well as adjustment
628 of the sticky potential strength and cutoffs. The cutoffs for the
629 tetrahedral attractive and dipolar repulsive terms were nearly swapped
630 with each other. Isosurfaces of the original and modified sticky
631 potentials are shown in figure \cite{isosurface}. In these
632 isosurfaces, it is easy to see how altering the cutoffs changes the
633 repulsive and attractive character of the particles. With a reduced
634 repulsive surface (the darker region), the particles can move closer
635 to one another, increasing the density for the overall system. This
636 change in interaction cutoff also results in a more gradual
637 orientational motion by allowing the particles to maintain preferred
638 dipolar arrangements before they begin to feel the pull of the
639 tetrahedral restructuring. Upon moving closer together, the dipolar
640 repulsion term becomes active and excludes the unphysical
641 arrangements. This compares with the original SSD's excluding dipolar
642 before the particles feel the pull of the ``hydrogen bonds''. Aside
643 from improving the shape of the first peak in the g(\emph{r}), this
644 improves the densities considerably by allowing the persistence of
645 full dipolar character below the previous 4.0 \AA\ cutoff.
646
647 While adjusting the location and shape of the first peak of
648 g(\emph{r}) improves the densities to some degree, these changes alone
649 are insufficient to bring the system densities up to the values
650 observed experimentally. To finish bringing up the densities, the
651 dipole moments were increased in both the adjusted models. Being a
652 dipole based model, the structure and transport are very sensitive to
653 changes in the dipole moment. The original SSD simply used the dipole
654 moment calculated from the TIP3P water model, which at 2.35 D is
655 significantly greater than the experimental gas phase value of 1.84
656 D. The larger dipole moment is a more realistic value and improve the
657 dielectric properties of the fluid. Both theoretical and experimental
658 measurements indicate a liquid phase dipole moment ranging from 2.4 D
659 to values as high as 3.11 D, so there is quite a range available for
660 adjusting the dipole
661 moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} The increasing of
662 the dipole moments to 2.418 and 2.48 D for SSD/E and SSD/RF
663 respectively is moderate in the range of the experimental values;
664 however, it leads to significant changes in the density and transport
665 of the water models.
666
667 In order to demonstrate the benefits of this reparameterization, a
668 series of NPT and NVE simulations were performed to probe the density
669 and transport properties of the adapted models and compare the results
670 to the original SSD model. This comparison involved full NPT melting
671 sequences for both SSD/E and SSD/RF, as well as NVE transport
672 calculations at both self-consistent and experimental
673 densities. Again, the results come from five separate simulations of
674 1024 particle systems, and the melting sequences were started from
675 different ice $I_h$ crystals constructed as stated previously. Like
676 before, all of the NPT simulations were equilibrated for 100 ps before
677 a 200 ps data collection run, and they used the previous temperature's
678 final configuration as a starting point. All of the NVE simulations
679 had the same thermalization, equilibration, and data collection times
680 stated earlier in this paper.
681
682 \begin{figure}
683 \includegraphics[width=85mm]{ssdecompare.epsi}
684 \caption{Comparison of densities calculated with SSD/E to SSD without a
685 reaction field, TIP4P\cite{Jorgensen98b}, TIP3P\cite{Jorgensen98b},
686 SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The upper plot
687 includes error bars, and the calculated results from the other
688 references were removed for clarity.}
689 \label{ssdedense}
690 \end{figure}
691
692 Figure \ref{ssdedense} shows the density profile for the SSD/E water
693 model in comparison to the original SSD without a reaction field,
694 experiment, and the other common water models considered
695 previously. The calculated densities have increased significantly over
696 the original SSD model and match the experimental value just below 298
697 K. At 298 K, the density of SSD/E is 0.995$\pm$0.001 g/cm$^3$, which
698 compares well with the experimental value of 0.997 g/cm$^3$ and is
699 considerably better than the SSD value of 0.967$\pm$0.003
700 g/cm$^3$. The increased dipole moment in SSD/E has helped to flatten
701 out the curve at higher temperatures, only the improvement is marginal
702 at best. This steep drop in densities is due to the dipolar rather
703 than charge based interactions which decay more rapidly at longer
704 distances.
705
706 By monitoring C$\text{p}$ throughout these simulations, the melting
707 transition for SSD/E was observed at 230 K, about 5 degrees lower than
708 SSD. The resulting density maximum is located at 240 K, again about 5
709 degrees lower than the SSD value of 245 K. Though there is a decrease
710 in both of these values, the corrected densities near room temperature
711 justify the modifications taken.
712
713 \begin{figure}
714 \includegraphics[width=85mm]{ssdrfcompare.epsi}
715 \caption{Comparison of densities calculated with SSD/RF to SSD with a
716 reaction field, TIP4P\cite{Jorgensen98b}, TIP3P\cite{Jorgensen98b},
717 SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The upper plot
718 includes error bars, and the calculated results from the other
719 references were removed for clarity.}
720 \label{ssdrfdense}
721 \end{figure}
722
723 Figure \ref{ssdrfdense} shows a density comparison between SSD/RF and
724 SSD with an active reaction field. Like in the simulations of SSD/E,
725 the densities show a dramatic increase over normal SSD. At 298 K,
726 SSD/RF has a density of 0.997$\pm$0.001 g/cm$^3$, right in line with
727 experiment and considerably better than the SSD value of
728 0.941$\pm$0.001 g/cm$^3$. The melting point is observed at 240 K,
729 which is 5 degrees lower than SSD with a reaction field, and the
730 density maximum at 255 K, again 5 degrees lower than SSD. The density
731 at higher temperature still drops off more rapidly than the charge
732 based models but is in better agreement than SSD/E.
733
734 The reparameterization of the SSD water model, both for use with and
735 without an applied long-range correction, brought the densities up to
736 what is expected for simulating liquid water. In addition to improving
737 the densities, it is important that particle transport be maintained
738 or improved. Figure \ref{ssdediffuse} compares the temperature
739 dependence of the diffusion constant of SSD/E to SSD without an active
740 reaction field, both at the densities calculated at 1 atm and at the
741 experimentally calculated densities for super-cooled and liquid
742 water. In the upper plot, the diffusion constant for SSD/E is
743 consistently a little faster than experiment, while SSD starts off
744 slower than experiment and crosses to merge with SSD/E at high
745 temperatures. Both models follow the experimental trend well, but
746 diffuse too rapidly at higher temperatures. This abnormally fast
747 diffusion is caused by the decreased system density. Since the
748 densities of SSD/E don't deviate as much from experiment as those of
749 SSD, it follows the experimental trend more closely. This observation
750 is backed up by looking at the lower plot. The diffusion constants for
751 SSD/E track with the experimental values while SSD deviates on the low
752 side of the trend with increasing temperature. This is again a product
753 of SSD/E having densities closer to experiment, and not deviating to
754 lower densities with increasing temperature as rapidly.
755
756 \begin{figure}
757 \includegraphics[width=85mm]{ssdediffuse.epsi}
758 \caption{Plots of the diffusion constants calculated from SSD/E and SSD,
759 both without a reaction field along with experimental results from
760 Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
761 upper plot is at densities calculated from the NPT simulations at a
762 pressure of 1 atm, while the lower plot is at the experimentally
763 calculated densities.}
764 \label{ssdediffuse}
765 \end{figure}
766
767 \begin{figure}
768 \includegraphics[width=85mm]{ssdrfdiffuse.epsi}
769 \caption{Plots of the diffusion constants calculated from SSD/RF and SSD,
770 both with an active reaction field along with experimental results
771 from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
772 upper plot is at densities calculated from the NPT simulations at a
773 pressure of 1 atm, while the lower plot is at the experimentally
774 calculated densities.}
775 \label{ssdrfdiffuse}
776 \end{figure}
777
778 In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
779 compared with SSD with an active reaction field. In the upper plot,
780 SSD/RF tracks with the experimental results incredibly well, identical
781 within error throughout the temperature range and only showing a
782 slight increasing trend at higher temperatures. SSD also tracks
783 experiment well, only it tends to diffuse a little more slowly at low
784 temperatures and deviates to diffuse too rapidly at high
785 temperatures. As was stated in the SSD/E comparisons, this deviation
786 away from the ideal trend is due to a rapid decrease in density at
787 higher temperatures. SSD/RF doesn't suffer from this problem as much
788 as SSD, because the calculated densities are more true to
789 experiment. This is again emphasized in the lower plot, where SSD/RF
790 tracks the experimental diffusion exactly while SSD's diffusion
791 constants are slightly too low due to its need for a lower density at
792 the specified temperature.
793
794 \subsection{Additional Observations}
795
796 While performing the melting sequences of SSD/E, some interesting
797 observations were made. After melting at 230 K, two of the systems
798 underwent crystallization events near 245 K. As the heating process
799 continued, the two systems remained crystalline until finally melting
800 between 320 and 330 K. These simulations were excluded from the data
801 set shown in figure \ref{ssdedense} and replaced with two additional
802 melting sequences that did not undergo this anomalous phase
803 transition, while this crystallization event was investigated
804 separately.
805
806 \begin{figure}
807 \includegraphics[width=85mm]{povIce.ps}
808 \caption{Crystal structure of an ice 0 lattice shown from the (001) face.}
809 \label{weirdice}
810 \end{figure}
811
812 The final configurations of these two melting sequences shows an
813 expanded zeolite-like crystal structure that does not correspond to
814 any known form of ice. For convenience and to help distinguish it from
815 the experimentally observed forms of ice, this crystal structure will
816 henceforth be referred to as ice-zero (ice 0). The crystallinity was
817 extensive enough than a near ideal crystal structure could be
818 obtained. Figure \ref{weirdice} shows the repeating crystal structure
819 of a typical crystal at 5 K. The unit cell contains eight molecules,
820 and figure \ref{unitcell} shows a unit cell built from the water
821 particle center of masses that can be used to construct a repeating
822 lattice, similar to figure \ref{weirdice}. Each molecule is hydrogen
823 bonded to four other water molecules; however, the hydrogen bonds are
824 flexed rather than perfectly straight. This results in a skewed
825 tetrahedral geometry about the central molecule. Looking back at
826 figure \ref{isosurface}, it is easy to see how these flexed hydrogen
827 bonds are allowed in that the attractive regions are conical in shape,
828 with the greatest attraction in the central region. Though not ideal,
829 these flexed hydrogen bonds are favorable enough to stabilize an
830 entire crystal generated around them. In fact, the imperfect ice 0
831 crystals were so stable that they melted at greater than room
832 temperature.
833
834 \begin{figure}
835 \includegraphics[width=65mm]{ice0cell.eps}
836 \caption{Simple unit cell for constructing ice 0. In this cell, $c$ is
837 equal to $0.4714\times a$, and a typical value for $a$ is 8.25 \AA.}
838 \label{unitcell}
839 \end{figure}
840
841 The initial simulations indicated that ice 0 is the preferred ice
842 structure for at least SSD/E. To verify this, a comparison was made
843 between near ideal crystals of ice $I_h$, ice $I_c$, and ice 0 at
844 constant pressure with SSD/E, SSD/RF, and SSD. Near ideal versions of
845 the three types of crystals were cooled to ~1 K, and the potential
846 energies of each were compared using all three water models. With
847 every water model, ice 0 turned out to have the lowest potential
848 energy: 5\% lower than $I_h$ with SSD, 6.5\% lower with SSD/E, and
849 7.5\% lower with SSD/RF. In all three of these water models, ice $I_c$
850 was observed to be ~2\% less stable than ice $I_h$. In addition to
851 having the lowest potential energy, ice 0 was the most expanded of the
852 three ice crystals, ~5\% less dense than ice $I_h$ with all of the
853 water models. In all three water models, ice $I_c$ was observed to be
854 ~2\% more dense than ice $I_h$.
855
856 In addition to the low temperature comparisons, melting sequences were
857 performed with ice 0 as the initial configuration using SSD/E, SSD/RF,
858 and SSD both with and without a reaction field. The melting
859 transitions for both SSD/E and SSD without a reaction field occurred
860 at temperature in excess of 375 K. SSD/RF and SSD with a reaction
861 field had more reasonable melting transitions, down near 325 K. These
862 melting point observations emphasize how preferred this crystal
863 structure is over the most common types of ice when using these single
864 point water models.
865
866 Recognizing that the above tests show ice 0 to be both the most stable
867 and lowest density crystal structure for these single point water
868 models, it is interesting to speculate on the favorability of this
869 crystal structure with the different charge based models. As a quick
870 test, these 3 crystal types were converted from SSD type particles to
871 TIP3P waters and read into CHARMM.\cite{Karplus83} Identical energy
872 minimizations were performed on all of these crystals to compare the
873 system energies. Again, ice 0 was observed to have the lowest total
874 system energy. The total energy of ice 0 was ~2\% lower than ice
875 $I_h$, which was in turn ~3\% lower than ice $I_c$. From these initial
876 results, we would not be surprised if results from the other common
877 water models show ice 0 to be the lowest energy crystal structure. A
878 continuation on work studing ice 0 with multipoint water models will
879 be published in a coming article.
880
881 \section{Conclusions}
882 The density maximum and temperature dependent transport for the SSD
883 water model, both with and without the use of reaction field, were
884 studied via a series of NPT and NVE simulations. The constant pressure
885 simulations of the melting of both $I_h$ and $I_c$ ice showed a
886 density maximum near 260 K. In most cases, the calculated densities
887 were significantly lower than the densities calculated in simulations
888 of other water models. Analysis of particle diffusion showed SSD to
889 capture the transport properties of experimental very well in both the
890 normal and super-cooled liquid regimes. In order to correct the
891 density behavior, SSD was reparameterized for use both with and
892 without a long-range interaction correction, SSD/RF and SSD/E
893 respectively. In addition to correcting the abnormally low densities,
894 these new versions were show to maintain or improve upon the transport
895 and structural features of the original water model, all while
896 maintaining the fast performance of the SSD water model. This work
897 shows these simple water models, and in particular SSD/E and SSD/RF,
898 to be excellent choices to represent explicit water in future
899 simulations of biochemical systems.
900
901 \section{Acknowledgments}
902 The authors would like to thank the National Science Foundation for
903 funding under grant CHE-0134881. Computation time was provided by the
904 Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant DMR
905 00 79647.
906
907 \bibliographystyle{jcp}
908
909 \bibliography{nptSSD}
910
911 %\pagebreak
912
913 \end{document}