ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 1027
Committed: Thu Feb 5 18:42:59 2004 UTC (20 years, 4 months ago) by chrisfen
Content type: application/x-tex
File size: 55424 byte(s)
Log Message:
More corrections and fixing of experimental numbers

File Contents

# Content
1 %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 \documentclass[11pt]{article}
3 \usepackage{endfloat}
4 \usepackage{amsmath}
5 \usepackage{epsf}
6 \usepackage{berkeley}
7 \usepackage{setspace}
8 \usepackage{tabularx}
9 \usepackage{graphicx}
10 \usepackage[ref]{overcite}
11 %\usepackage{berkeley}
12 %\usepackage{curves}
13 \pagestyle{plain}
14 \pagenumbering{arabic}
15 \oddsidemargin 0.0cm \evensidemargin 0.0cm
16 \topmargin -21pt \headsep 10pt
17 \textheight 9.0in \textwidth 6.5in
18 \brokenpenalty=10000
19 \renewcommand{\baselinestretch}{1.2}
20 \renewcommand\citemid{\ } % no comma in optional reference note
21
22 \begin{document}
23
24 \title{On the structural and transport properties of the soft sticky
25 dipole (SSD) and related single point water models}
26
27 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
28 Department of Chemistry and Biochemistry\\ University of Notre Dame\\
29 Notre Dame, Indiana 46556}
30
31 \date{\today}
32
33 \maketitle
34
35 \begin{abstract}
36 The density maximum and temperature dependence of the self-diffusion
37 constant were investigated for the soft sticky dipole (SSD) water
38 model and two related re-parameterizations of this single-point model.
39 A combination of microcanonical and isobaric-isothermal molecular
40 dynamics simulations were used to calculate these properties, both
41 with and without the use of reaction field to handle long-range
42 electrostatics. The isobaric-isothermal (NPT) simulations of the
43 melting of both ice-$I_h$ and ice-$I_c$ showed a density maximum near
44 260 K. In most cases, the use of the reaction field resulted in
45 calculated densities which were were significantly lower than
46 experimental densities. Analysis of self-diffusion constants shows
47 that the original SSD model captures the transport properties of
48 experimental water very well in both the normal and super-cooled
49 liquid regimes. We also present our re-parameterized versions of SSD
50 for use both with the reaction field or without any long-range
51 electrostatic corrections. These are called the SSD/RF and SSD/E
52 models respectively. These modified models were shown to maintain or
53 improve upon the experimental agreement with the structural and
54 transport properties that can be obtained with either the original SSD
55 or the density corrected version of the original model (SSD1).
56 Additionally, a novel low-density ice structure is presented
57 which appears to be the most stable ice structure for the entire SSD
58 family.
59 \end{abstract}
60
61 \newpage
62
63 %\narrowtext
64
65
66 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67 % BODY OF TEXT
68 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69
70 \section{Introduction}
71
72 One of the most important tasks in the simulation of biochemical
73 systems is the proper depiction of the aqueous environment of the
74 molecules of interest. In some cases (such as in the simulation of
75 phospholipid bilayers), the majority of the calculations that are
76 performed involve interactions with or between solvent molecules.
77 Thus, the properties one may observe in biochemical simulations are
78 going to be highly dependent on the physical properties of the water
79 model that is chosen.
80
81 There is an especially delicate balance between computational
82 efficiency and the ability of the water model to accurately predict
83 the properties of bulk
84 water.\cite{Jorgensen83,Berendsen87,Jorgensen00} For example, the
85 TIP5P model improves on the structural and transport properties of
86 water relative to the previous TIP models, yet this comes at a greater
87 than 50\% increase in computational
88 cost.\cite{Jorgensen01,Jorgensen00}
89
90 One recently developed model that largely succeeds in retaining the
91 accuracy of bulk properties while greatly reducing the computational
92 cost is the Soft Sticky Dipole (SSD) water
93 model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The SSD model was
94 developed by Ichiye \emph{et al.} as a modified form of the
95 hard-sphere water model proposed by Bratko, Blum, and
96 Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point} model which
97 has an interaction site that is both a point dipole along with a
98 Lennard-Jones core. However, since the normal aligned and
99 anti-aligned geometries favored by point dipoles are poor mimics of
100 local structure in liquid water, a short ranged ``sticky'' potential
101 is also added. The sticky potential directs the molecules to assume
102 the proper hydrogen bond orientation in the first solvation
103 shell.
104
105 The interaction between two SSD water molecules \emph{i} and \emph{j}
106 is given by the potential
107 \begin{equation}
108 u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
109 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ +
110 u_{ij}^{sp}
111 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j),
112 \end{equation}
113 where the ${\bf r}_{ij}$ is the position vector between molecules
114 \emph{i} and \emph{j} with magnitude $r_{ij}$, and
115 ${\bf \Omega}_i$ and ${\bf \Omega}_j$ represent the orientations of
116 the two molecules. The Lennard-Jones and dipole interactions are given
117 by the following familiar forms:
118 \begin{equation}
119 u_{ij}^{LJ}(r_{ij}) = 4\epsilon
120 \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right]
121 \ ,
122 \end{equation}
123 and
124 \begin{equation}
125 u_{ij}^{dp} = \frac{|\mu_i||\mu_j|}{4 \pi \epsilon_0 r_{ij}^3} \left(
126 \hat{\bf u}_i \cdot \hat{\bf u}_j - 3(\hat{\bf u}_i\cdot\hat{\bf
127 r}_{ij})(\hat{\bf u}_j\cdot\hat{\bf r}_{ij}) \right)\ ,
128 \end{equation}
129 where $\hat{\bf u}_i$ and $\hat{\bf u}_j$ are the unit vectors along
130 the dipoles of molecules $i$ and $j$ respectively. $|\mu_i|$ and
131 $|\mu_j|$ are the strengths of the dipole moments, and $\hat{\bf
132 r}_{ij}$ is the unit vector pointing from molecule $j$ to molecule
133 $i$.
134
135 The sticky potential is somewhat less familiar:
136 \begin{equation}
137 u_{ij}^{sp}
138 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
139 \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)
140 + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf
141 \Omega}_j)]\ .
142 \label{stickyfunction}
143 \end{equation}
144 Here, $\nu_0$ is a strength parameter for the sticky potential, and
145 $s$ and $s^\prime$ are cubic switching functions which turn off the
146 sticky interaction beyond the first solvation shell. The $w$ function
147 can be thought of as an attractive potential with tetrahedral
148 geometry:
149 \begin{equation}
150 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
151 \end{equation}
152 while the $w^\prime$ function counters the normal aligned and
153 anti-aligned structures favored by point dipoles:
154 \begin{equation}
155 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^\circ,
156 \end{equation}
157 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
158 and $Y_3^{-2}$ spherical harmonics (a linear combination which
159 enhances the tetrahedral geometry for hydrogen bonded structures),
160 while $w^\prime$ is a purely empirical function. A more detailed
161 description of the functional parts and variables in this potential
162 can be found in the original SSD
163 articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
164
165 Since SSD is a single-point {\it dipolar} model, the force
166 calculations are simplified significantly relative to the standard
167 {\it charged} multi-point models. In the original Monte Carlo
168 simulations using this model, Ichiye {\it et al.} reported that using
169 SSD decreased computer time by a factor of 6-7 compared to other
170 models.\cite{Ichiye96} What is most impressive is that this savings
171 did not come at the expense of accurate depiction of the liquid state
172 properties. Indeed, SSD maintains reasonable agreement with the Soper
173 data for the structural features of liquid
174 water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
175 exhibited by SSD agree with experiment better than those of more
176 computationally expensive models (like TIP3P and
177 SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
178 of solvent properties makes SSD a very attractive model for the
179 simulation of large scale biochemical simulations.
180
181 One feature of the SSD model is that it was parameterized for use with
182 the Ewald sum to handle long-range interactions. This would normally
183 be the best way of handling long-range interactions in systems that
184 contain other point charges. However, our group has recently become
185 interested in systems with point dipoles as mimics for neutral, but
186 polarized regions on molecules (e.g. the zwitterionic head group
187 regions of phospholipids). If the system of interest does not contain
188 point charges, the Ewald sum and even particle-mesh Ewald become
189 computational bottlenecks. Their respective ideal $N^\frac{3}{2}$ and
190 $N\log N$ calculation scaling orders for $N$ particles can become
191 prohibitive when $N$ becomes large.\cite{Darden99} In applying this
192 water model in these types of systems, it would be useful to know its
193 properties and behavior under the more computationally efficient
194 reaction field (RF) technique, or even with a simple cutoff. This
195 study addresses these issues by looking at the structural and
196 transport behavior of SSD over a variety of temperatures with the
197 purpose of utilizing the RF correction technique. We then suggest
198 modifications to the parameters that result in more realistic bulk
199 phase behavior. It should be noted that in a recent publication, some
200 of the original investigators of the SSD water model have suggested
201 adjustments to the SSD water model to address abnormal density
202 behavior (also observed here), calling the corrected model
203 SSD1.\cite{Ichiye03} In what follows, we compare our
204 reparamaterization of SSD with both the original SSD and SSD1 models
205 with the goal of improving the bulk phase behavior of an SSD-derived
206 model in simulations utilizing the Reaction Field.
207
208 \section{Methods}
209
210 Long-range dipole-dipole interactions were accounted for in this study
211 by using either the reaction field method or by resorting to a simple
212 cubic switching function at a cutoff radius. The reaction field
213 method was actually first used in Monte Carlo simulations of liquid
214 water.\cite{Barker73} Under this method, the magnitude of the reaction
215 field acting on dipole $i$ is
216 \begin{equation}
217 \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
218 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} f(r_{ij}),
219 \label{rfequation}
220 \end{equation}
221 where $\mathcal{R}$ is the cavity defined by the cutoff radius
222 ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
223 system (80 in the case of liquid water), ${\bf \mu}_{j}$ is the dipole
224 moment vector of particle $j$, and $f(r_{ij})$ is a cubic switching
225 function.\cite{AllenTildesley} The reaction field contribution to the
226 total energy by particle $i$ is given by $-\frac{1}{2}{\bf
227 \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf
228 \mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use of the reaction
229 field is known to alter the bulk orientational properties, such as the
230 dielectric relaxation time. There is particular sensitivity of this
231 property on changes in the length of the cutoff
232 radius.\cite{Berendsen98} This variable behavior makes reaction field
233 a less attractive method than the Ewald sum. However, for very large
234 systems, the computational benefit of reaction field is dramatic.
235
236 We have also performed a companion set of simulations {\it without} a
237 surrounding dielectric (i.e. using a simple cubic switching function
238 at the cutoff radius), and as a result we have two reparamaterizations
239 of SSD which could be used either with or without the reaction field
240 turned on.
241
242 Simulations to obtain the preferred density were performed in the
243 isobaric-isothermal (NPT) ensemble, while all dynamical properties
244 were obtained from microcanonical (NVE) simulations done at densities
245 matching the NPT density for a particular target temperature. The
246 constant pressure simulations were implemented using an integral
247 thermostat and barostat as outlined by Hoover.\cite{Hoover85,Hoover86}
248 All molecules were treated as non-linear rigid bodies. Vibrational
249 constraints are not necessary in simulations of SSD, because there are
250 no explicit hydrogen atoms, and thus no molecular vibrational modes
251 need to be considered.
252
253 Integration of the equations of motion was carried out using the
254 symplectic splitting method proposed by Dullweber, Leimkuhler, and
255 McLachlan (DLM).\cite{Dullweber1997} Our reason for selecting this
256 integrator centers on poor energy conservation of rigid body dynamics
257 using traditional quaternion integration.\cite{Evans77,Evans77b} In
258 typical microcanonical ensemble simulations, the energy drift when
259 using quaternions was substantially greater than when using the DLM
260 method (fig. \ref{timestep}). This steady drift in the total energy
261 has also been observed by Kol {\it et al.}\cite{Laird97}
262
263 The key difference in the integration method proposed by Dullweber
264 \emph{et al.} is that the entire rotation matrix is propagated from
265 one time step to the next. The additional memory required by the
266 algorithm is inconsequential on modern computers, and translating the
267 rotation matrix into quaternions for storage purposes makes trajectory
268 data quite compact.
269
270 The DML method allows for Verlet style integration of both
271 translational and orientational motion of rigid bodies. In this
272 integration method, the orientational propagation involves a sequence
273 of matrix evaluations to update the rotation
274 matrix.\cite{Dullweber1997} These matrix rotations are more costly
275 than the simpler arithmetic quaternion propagation. With the same time
276 step, a 1000 SSD particle simulation shows an average 7\% increase in
277 computation time using the DML method in place of quaternions. The
278 additional expense per step is justified when one considers the
279 ability to use time steps that are nearly twice as large under DML
280 than would be usable under quaternion dynamics. The energy
281 conservation of the two methods using a number of different time steps
282 is illustrated in figure
283 \ref{timestep}.
284
285 \begin{figure}
286 \begin{center}
287 \epsfxsize=6in
288 \epsfbox{timeStep.epsi}
289 \caption{Energy conservation using both quaternion based integration and
290 the symplectic splitting method proposed by Dullweber \emph{et al.}
291 with increasing time step. The larger time step plots are shifted from
292 the true energy baseline (that of $\Delta t$ = 0.1 fs) for clarity.}
293 \label{timestep}
294 \end{center}
295 \end{figure}
296
297 In figure \ref{timestep}, the resulting energy drift at various time
298 steps for both the DML and quaternion integration schemes is compared.
299 All of the 1000 SSD particle simulations started with the same
300 configuration, and the only difference was the method used to handle
301 orientational motion. At time steps of 0.1 and 0.5 fs, both methods
302 for propagating the orientational degrees of freedom conserve energy
303 fairly well, with the quaternion method showing a slight energy drift
304 over time in the 0.5 fs time step simulation. At time steps of 1 and 2
305 fs, the energy conservation benefits of the DML method are clearly
306 demonstrated. Thus, while maintaining the same degree of energy
307 conservation, one can take considerably longer time steps, leading to
308 an overall reduction in computation time.
309
310 Energy drift in the simulations using DML integration was unnoticeable
311 for time steps up to 3 fs. A slight energy drift on the order of 0.012
312 kcal/mol per nanosecond was observed at a time step of 4 fs, and as
313 expected, this drift increases dramatically with increasing time
314 step. To insure accuracy in our microcanonical simulations, time steps
315 were set at 2 fs and kept at this value for constant pressure
316 simulations as well.
317
318 Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices
319 were generated as starting points for all simulations. The $I_h$
320 crystals were formed by first arranging the centers of mass of the SSD
321 particles into a ``hexagonal'' ice lattice of 1024 particles. Because
322 of the crystal structure of $I_h$ ice, the simulation box assumed an
323 orthorhombic shape with an edge length ratio of approximately
324 1.00$\times$1.06$\times$1.23. The particles were then allowed to
325 orient freely about fixed positions with angular momenta randomized at
326 400 K for varying times. The rotational temperature was then scaled
327 down in stages to slowly cool the crystals to 25 K. The particles were
328 then allowed to translate with fixed orientations at a constant
329 pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
330 removed and the ice crystals were allowed to equilibrate for 50 ps at
331 25 K and a constant pressure of 1 atm. This procedure resulted in
332 structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
333 rules.\cite{Bernal33,Rahman72} This method was also utilized in the
334 making of diamond lattice $I_c$ ice crystals, with each cubic
335 simulation box consisting of either 512 or 1000 particles. Only
336 isotropic volume fluctuations were performed under constant pressure,
337 so the ratio of edge lengths remained constant throughout the
338 simulations.
339
340 \section{Results and discussion}
341
342 Melting studies were performed on the randomized ice crystals using
343 isobaric-isothermal (NPT) dynamics. During melting simulations, the
344 melting transition and the density maximum can both be observed,
345 provided that the density maximum occurs in the liquid and not the
346 supercooled regime. An ensemble average from five separate melting
347 simulations was acquired, each starting from different ice crystals
348 generated as described previously. All simulations were equilibrated
349 for 100 ps prior to a 200 ps data collection run at each temperature
350 setting. The temperature range of study spanned from 25 to 400 K, with
351 a maximum degree increment of 25 K. For regions of interest along this
352 stepwise progression, the temperature increment was decreased from 25
353 K to 10 and 5 K. The above equilibration and production times were
354 sufficient in that fluctuations in the volume autocorrelation function
355 were damped out in all simulations in under 20 ps.
356
357 \subsection{Density Behavior}
358
359 Our initial simulations focused on the original SSD water model, and
360 an average density versus temperature plot is shown in figure
361 \ref{dense1}. Note that the density maximum when using a reaction
362 field appears between 255 and 265 K. There were smaller fluctuations
363 in the density at 260 K than at either 255 or 265, so we report this
364 value as the location of the density maximum. Figure \ref{dense1} was
365 constructed using ice $I_h$ crystals for the initial configuration;
366 though not pictured, the simulations starting from ice $I_c$ crystal
367 configurations showed similar results, with a liquid-phase density
368 maximum in this same region (between 255 and 260 K).
369
370 \begin{figure}
371 \begin{center}
372 \epsfxsize=6in
373 \epsfbox{denseSSD.eps}
374 \caption{Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
375 TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], SSD
376 without Reaction Field, SSD, and experiment [Ref. \citen{CRC80}]. The
377 arrows indicate the change in densities observed when turning off the
378 reaction field. The the lower than expected densities for the SSD
379 model were what prompted the original reparameterization of SSD1
380 [Ref. \citen{Ichiye03}].}
381 \label{dense1}
382 \end{center}
383 \end{figure}
384
385 The density maximum for SSD compares quite favorably to other simple
386 water models. Figure \ref{dense1} also shows calculated densities of
387 several other models and experiment obtained from other
388 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
389 models, SSD has a temperature closest to the experimentally observed
390 density maximum. Of the {\it charge-based} models in
391 Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
392 seen in SSD. Though not included in this plot, it is useful
393 to note that TIP5P has a density maximum nearly identical to the
394 experimentally measured temperature.
395
396 It has been observed that liquid state densities in water are
397 dependent on the cutoff radius used both with and without the use of
398 reaction field.\cite{Berendsen98} In order to address the possible
399 effect of cutoff radius, simulations were performed with a dipolar
400 cutoff radius of 12.0 \AA\ to complement the previous SSD simulations,
401 all performed with a cutoff of 9.0 \AA. All of the resulting densities
402 overlapped within error and showed no significant trend toward lower
403 or higher densities as a function of cutoff radius, for simulations
404 both with and without reaction field. These results indicate that
405 there is no major benefit in choosing a longer cutoff radius in
406 simulations using SSD. This is advantageous in that the use of a
407 longer cutoff radius results in a significant increase in the time
408 required to obtain a single trajectory.
409
410 The key feature to recognize in figure \ref{dense1} is the density
411 scaling of SSD relative to other common models at any given
412 temperature. SSD assumes a lower density than any of the other listed
413 models at the same pressure, behavior which is especially apparent at
414 temperatures greater than 300 K. Lower than expected densities have
415 been observed for other systems using a reaction field for long-range
416 electrostatic interactions, so the most likely reason for the
417 significantly lower densities seen in these simulations is the
418 presence of the reaction field.\cite{Berendsen98,Nezbeda02} In order
419 to test the effect of the reaction field on the density of the
420 systems, the simulations were repeated without a reaction field
421 present. The results of these simulations are also displayed in figure
422 \ref{dense1}. Without the reaction field, the densities increase
423 to more experimentally reasonable values, especially around the
424 freezing point of liquid water. The shape of the curve is similar to
425 the curve produced from SSD simulations using reaction field,
426 specifically the rapidly decreasing densities at higher temperatures;
427 however, a shift in the density maximum location, down to 245 K, is
428 observed. This is a more accurate comparison to the other listed water
429 models, in that no long range corrections were applied in those
430 simulations.\cite{Clancy94,Jorgensen98b} However, even without the
431 reaction field, the density around 300 K is still significantly lower
432 than experiment and comparable water models. This anomalous behavior
433 was what lead Tan {\it et al.} to recently reparameterize
434 SSD.\cite{Ichiye03} Throughout the remainder of the paper our
435 reparamaterizations of SSD will be compared with the newer SSD1 model.
436
437 \subsection{Transport Behavior}
438
439 Accurate dynamical properties of a water model are particularly
440 important when using the model to study permeation or transport across
441 biological membranes. In order to probe transport in bulk water,
442 constant energy (NVE) simulations were performed at the average
443 density obtained by the NPT simulations at an identical target
444 temperature. Simulations started with randomized velocities and
445 underwent 50 ps of temperature scaling and 50 ps of constant energy
446 equilibration before a 200 ps data collection run. Diffusion constants
447 were calculated via linear fits to the long-time behavior of the
448 mean-square displacement as a function of time. The averaged results
449 from five sets of NVE simulations are displayed in figure
450 \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
451 results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
452
453 \begin{figure}
454 \begin{center}
455 \epsfxsize=6in
456 \epsfbox{betterDiffuse.epsi}
457 \caption{Average self-diffusion constant as a function of temperature for
458 SSD, SPC/E [Ref. \citen{Clancy94}], TIP5P [Ref. \citen{Jorgensen01}],
459 and Experimental data [Refs. \citen{Gillen72} and \citen{Holz00}]. Of
460 the three water models shown, SSD has the least deviation from the
461 experimental values. The rapidly increasing diffusion constants for
462 TIP5P and SSD correspond to significant decrease in density at the
463 higher temperatures.}
464 \label{diffuse}
465 \end{center}
466 \end{figure}
467
468 The observed values for the diffusion constant point out one of the
469 strengths of the SSD model. Of the three models shown, the SSD model
470 has the most accurate depiction of self-diffusion in both the
471 supercooled and liquid regimes. SPC/E does a respectable job by
472 reproducing values similar to experiment around 290 K; however, it
473 deviates at both higher and lower temperatures, failing to predict the
474 correct thermal trend. TIP5P and SSD both start off low at colder
475 temperatures and tend to diffuse too rapidly at higher temperatures.
476 This behavior at higher temperatures is not particularly surprising
477 since the densities of both TIP5P and SSD are lower than experimental
478 water densities at higher temperatures. When calculating the
479 diffusion coefficients for SSD at experimental densities (instead of
480 the densities from the NPT simulations), the resulting values fall
481 more in line with experiment at these temperatures.
482
483 \subsection{Structural Changes and Characterization}
484
485 By starting the simulations from the crystalline state, the melting
486 transition and the ice structure can be obtained along with the liquid
487 phase behavior beyond the melting point. The constant pressure heat
488 capacity (C$_\text{p}$) was monitored to locate the melting transition
489 in each of the simulations. In the melting simulations of the 1024
490 particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs
491 at 245 K, indicating a first order phase transition for the melting of
492 these ice crystals. When the reaction field is turned off, the melting
493 transition occurs at 235 K. These melting transitions are
494 considerably lower than the experimental value.
495
496 \begin{figure}
497 \begin{center}
498 \epsfxsize=6in
499 \epsfbox{corrDiag.eps}
500 \caption{Two dimensional illustration of angles involved in the
501 correlations observed in Fig. \ref{contour}.}
502 \label{corrAngle}
503 \end{center}
504 \end{figure}
505
506 \begin{figure}
507 \begin{center}
508 \epsfxsize=6in
509 \epsfbox{fullContours.eps}
510 \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
511 100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
512 clarity: dark areas signify peaks while light areas signify
513 depressions. White areas have $g(r)$ values below 0.5 and black
514 areas have values above 1.5.}
515 \label{contour}
516 \end{center}
517 \end{figure}
518
519 Additional analysis of the melting process was performed using
520 two-dimensional structure and dipole angle correlations. Expressions
521 for these correlations are as follows:
522
523 \begin{equation}
524 g_{\text{AB}}(r,\cos\theta) = \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|{\bf r}_{ij}\right|)\rangle\ ,
525 \end{equation}
526 \begin{equation}
527 g_{\text{AB}}(r,\cos\omega) =
528 \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|{\bf r}_{ij}\right|)\rangle\ ,
529 \end{equation}
530 where $\theta$ and $\omega$ refer to the angles shown in figure
531 \ref{corrAngle}. By binning over both distance and the cosine of the
532 desired angle between the two dipoles, the $g(r)$ can be analyzed to
533 determine the common dipole arrangements that constitute the peaks and
534 troughs in the standard one-dimensional $g(r)$ plots. Frames A and B
535 of figure \ref{contour} show results from an ice $I_c$ simulation. The
536 first peak in the $g(r)$ consists primarily of the preferred hydrogen
537 bonding arrangements as dictated by the tetrahedral sticky potential -
538 one peak for the hydrogen bond donor and the other for the hydrogen
539 bond acceptor. Due to the high degree of crystallinity of the sample,
540 the second and third solvation shells show a repeated peak arrangement
541 which decays at distances around the fourth solvation shell, near the
542 imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
543 In the higher temperature simulation shown in frames C and D, these
544 long-range features deteriorate rapidly. The first solvation shell
545 still shows the strong effect of the sticky-potential, although it
546 covers a larger area, extending to include a fraction of aligned
547 dipole peaks within the first solvation shell. The latter peaks lose
548 due to thermal motion and as the competing dipole force overcomes the
549 sticky potential's tight tetrahedral structuring of the crystal.
550
551 This complex interplay between dipole and sticky interactions was
552 remarked upon as a possible reason for the split second peak in the
553 oxygen-oxygen $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures,
554 the second solvation shell peak appears to have two distinct
555 components that blend together to form one observable peak. At higher
556 temperatures, this split character alters to show the leading 4 \AA\
557 peak dominated by equatorial anti-parallel dipole orientations. There
558 is also a tightly bunched group of axially arranged dipoles that most
559 likely consist of the smaller fraction of aligned dipole pairs. The
560 trailing component of the split peak at 5 \AA\ is dominated by aligned
561 dipoles that assume hydrogen bond arrangements similar to those seen
562 in the first solvation shell. This evidence indicates that the dipole
563 pair interaction begins to dominate outside of the range of the
564 dipolar repulsion term. The energetically favorable dipole
565 arrangements populate the region immediately outside this repulsion
566 region (around 4 \AA), while arrangements that seek to satisfy both
567 the sticky and dipole forces locate themselves just beyond this
568 initial buildup (around 5 \AA).
569
570 From these findings, the split second peak is primarily the product of
571 the dipolar repulsion term of the sticky potential. In fact, the inner
572 peak can be pushed out and merged with the outer split peak just by
573 extending the switching function ($s^\prime(r_{ij})$) from its normal
574 4.0 \AA\ cutoff to values of 4.5 or even 5 \AA. This type of
575 correction is not recommended for improving the liquid structure,
576 since the second solvation shell would still be shifted too far
577 out. In addition, this would have an even more detrimental effect on
578 the system densities, leading to a liquid with a more open structure
579 and a density considerably lower than the already low SSD density. A
580 better correction would be to include the quadrupole-quadrupole
581 interactions for the water particles outside of the first solvation
582 shell, but this would remove the simplicity and speed advantage of
583 SSD.
584
585 \subsection{Adjusted Potentials: SSD/RF and SSD/E}
586
587 The propensity of SSD to adopt lower than expected densities under
588 varying conditions is troubling, especially at higher temperatures. In
589 order to correct this model for use with a reaction field, it is
590 necessary to adjust the force field parameters for the primary
591 intermolecular interactions. In undergoing a reparameterization, it is
592 important not to focus on just one property and neglect the other
593 important properties. In this case, it would be ideal to correct the
594 densities while maintaining the accurate transport behavior.
595
596 The parameters available for tuning include the $\sigma$ and
597 $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
598 strength of the sticky potential ($\nu_0$), and the sticky attractive
599 and dipole repulsive cubic switching function cutoffs ($r_l$, $r_u$
600 and $r_l^\prime$, $r_u^\prime$ respectively). The results of the
601 reparameterizations are shown in table \ref{params}. We are calling
602 these reparameterizations the Soft Sticky Dipole / Reaction Field
603 (SSD/RF - for use with a reaction field) and Soft Sticky Dipole
604 Extended (SSD/E - an attempt to improve the liquid structure in
605 simulations without a long-range correction).
606
607 \begin{table}
608 \begin{center}
609 \caption{Parameters for the original and adjusted models}
610 \begin{tabular}{ l c c c c }
611 \hline \\[-3mm]
612 \ \ \ Parameters\ \ \ & \ \ \ SSD [Ref. \citen{Ichiye96}] \ \ \
613 & \ SSD1 [Ref. \citen{Ichiye03}]\ \ & \ SSD/E\ \ & \ SSD/RF \\
614 \hline \\[-3mm]
615 \ \ \ $\sigma$ (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
616 \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
617 \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
618 \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
619 \ \ \ $\omega^\circ$ & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
620 \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
621 \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
622 \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
623 \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
624 \end{tabular}
625 \label{params}
626 \end{center}
627 \end{table}
628
629 \begin{figure}
630 \begin{center}
631 \epsfxsize=5in
632 \epsfbox{GofRCompare.epsi}
633 \caption{Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with SSD/E
634 and SSD1 without reaction field (top), as well as SSD/RF and SSD1 with
635 reaction field turned on (bottom). The insets show the respective
636 first peaks in detail. Note how the changes in parameters have lowered
637 and broadened the first peak of SSD/E and SSD/RF.}
638 \label{grcompare}
639 \end{center}
640 \end{figure}
641
642 \begin{figure}
643 \begin{center}
644 \epsfxsize=6in
645 \epsfbox{dualsticky_bw.eps}
646 \caption{Isosurfaces of the sticky potential for SSD1 (left) and SSD/E \&
647 SSD/RF (right). Light areas correspond to the tetrahedral attractive
648 component, and darker areas correspond to the dipolar repulsive
649 component.}
650 \label{isosurface}
651 \end{center}
652 \end{figure}
653
654 In the original paper detailing the development of SSD, Liu and Ichiye
655 placed particular emphasis on an accurate description of the first
656 solvation shell. This resulted in a somewhat tall and narrow first
657 peak in $g(r)$ that integrated to give similar coordination numbers to
658 the experimental data obtained by Soper and
659 Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
660 data from the Head-Gordon lab indicates a slightly lower and shifted
661 first peak in the g$_\mathrm{OO}(r)$, so our adjustments to SSD were
662 made while taking into consideration the new experimental
663 findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
664 relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
665 the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
666 experimental results. Both modified water models have shorter peaks
667 that match more closely to the experimental peak (as seen in the
668 insets of figure \ref{grcompare}). This structural alteration was
669 accomplished by the combined reduction in the Lennard-Jones $\sigma$
670 variable and adjustment of the sticky potential strength and cutoffs.
671 As can be seen in table \ref{params}, the cutoffs for the tetrahedral
672 attractive and dipolar repulsive terms were nearly swapped with each
673 other. Isosurfaces of the original and modified sticky potentials are
674 shown in figure \ref{isosurface}. In these isosurfaces, it is easy to
675 see how altering the cutoffs changes the repulsive and attractive
676 character of the particles. With a reduced repulsive surface (darker
677 region), the particles can move closer to one another, increasing the
678 density for the overall system. This change in interaction cutoff also
679 results in a more gradual orientational motion by allowing the
680 particles to maintain preferred dipolar arrangements before they begin
681 to feel the pull of the tetrahedral restructuring. As the particles
682 move closer together, the dipolar repulsion term becomes active and
683 excludes unphysical nearest-neighbor arrangements. This compares with
684 how SSD and SSD1 exclude preferred dipole alignments before the
685 particles feel the pull of the ``hydrogen bonds''. Aside from
686 improving the shape of the first peak in the g(\emph{r}), this
687 modification improves the densities considerably by allowing the
688 persistence of full dipolar character below the previous 4.0 \AA\
689 cutoff.
690
691 While adjusting the location and shape of the first peak of $g(r)$
692 improves the densities, these changes alone are insufficient to bring
693 the system densities up to the values observed experimentally. To
694 further increase the densities, the dipole moments were increased in
695 both of our adjusted models. Since SSD is a dipole based model, the
696 structure and transport are very sensitive to changes in the dipole
697 moment. The original SSD simply used the dipole moment calculated from
698 the TIP3P water model, which at 2.35 D is significantly greater than
699 the experimental gas phase value of 1.84 D. The larger dipole moment
700 is a more realistic value and improves the dielectric properties of
701 the fluid. Both theoretical and experimental measurements indicate a
702 liquid phase dipole moment ranging from 2.4 D to values as high as
703 3.11 D, providing a substantial range of reasonable values for a
704 dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
705 increasing the dipole moments to 2.42 and 2.48 D for SSD/E and SSD/RF,
706 respectively, leads to significant changes in the density and
707 transport of the water models.
708
709 In order to demonstrate the benefits of these reparameterizations, a
710 series of NPT and NVE simulations were performed to probe the density
711 and transport properties of the adapted models and compare the results
712 to the original SSD model. This comparison involved full NPT melting
713 sequences for both SSD/E and SSD/RF, as well as NVE transport
714 calculations at the calculated self-consistent densities. Again, the
715 results are obtained from five separate simulations of 1024 particle
716 systems, and the melting sequences were started from different ice
717 $I_h$ crystals constructed as described previously. Each NPT
718 simulation was equilibrated for 100 ps before a 200 ps data collection
719 run at each temperature step, and the final configuration from the
720 previous temperature simulation was used as a starting point. All NVE
721 simulations had the same thermalization, equilibration, and data
722 collection times as stated previously.
723
724 \begin{figure}
725 \begin{center}
726 \epsfxsize=6in
727 \epsfbox{ssdeDense.epsi}
728 \caption{Comparison of densities calculated with SSD/E to SSD1 without a
729 reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
730 [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and
731 experiment [Ref. \citen{CRC80}]. The window shows a expansion around
732 300 K with error bars included to clarify this region of
733 interest. Note that both SSD1 and SSD/E show good agreement with
734 experiment when the long-range correction is neglected.}
735 \label{ssdedense}
736 \end{center}
737 \end{figure}
738
739 Fig. \ref{ssdedense} shows the density profile for the SSD/E model
740 in comparison to SSD1 without a reaction field, other common water
741 models, and experimental results. The calculated densities for both
742 SSD/E and SSD1 have increased significantly over the original SSD
743 model (see fig. \ref{dense1}) and are in better agreement with the
744 experimental values. At 298 K, the densities of SSD/E and SSD1 without
745 a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
746 0.999$\pm$0.001 g/cm$^3$ respectively. These both compare well with
747 the experimental value of 0.997 g/cm$^3$, and they are considerably
748 better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The changes to
749 the dipole moment and sticky switching functions have improved the
750 structuring of the liquid (as seen in figure \ref{grcompare}, but they
751 have shifted the density maximum to much lower temperatures. This
752 comes about via an increase in the liquid disorder through the
753 weakening of the sticky potential and strengthening of the dipolar
754 character. However, this increasing disorder in the SSD/E model has
755 little effect on the melting transition. By monitoring $C_p$
756 throughout these simulations, the melting transition for SSD/E was
757 shown to occur at 235 K. The same transition temperature observed
758 with SSD and SSD1.
759
760 \begin{figure}
761 \begin{center}
762 \epsfxsize=6in
763 \epsfbox{ssdrfDense.epsi}
764 \caption{Comparison of densities calculated with SSD/RF to SSD1 with a
765 reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
766 [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and
767 experiment [Ref. \citen{CRC80}]. The inset shows the necessity of
768 reparameterization when utilizing a reaction field long-ranged
769 correction - SSD/RF provides significantly more accurate densities
770 than SSD1 when performing room temperature simulations.}
771 \label{ssdrfdense}
772 \end{center}
773 \end{figure}
774
775 Including the reaction field long-range correction in the simulations
776 results in a more interesting comparison. A density profile including
777 SSD/RF and SSD1 with an active reaction field is shown in figure
778 \ref{ssdrfdense}. As observed in the simulations without a reaction
779 field, the densities of SSD/RF and SSD1 show a dramatic increase over
780 normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density
781 of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
782 considerably better than the original SSD value of 0.941$\pm$0.001
783 g/cm$^3$ and the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results
784 further emphasize the importance of reparameterization in order to
785 model the density properly under different simulation conditions.
786 Again, these changes have only a minor effect on the melting point,
787 which observed at 245 K for SSD/RF, is identical to SSD and only 5 K
788 lower than SSD1 with a reaction field. Additionally, the difference in
789 density maxima is not as extreme, with SSD/RF showing a density
790 maximum at 255 K, fairly close to the density maxima of 260 K and 265
791 K, shown by SSD and SSD1 respectively.
792
793 \begin{figure}
794 \begin{center}
795 \epsfxsize=6in
796 \epsfbox{ssdeDiffuse.epsi}
797 \caption{The diffusion constants calculated from SSD/E and SSD1,
798 both without a reaction field, along with experimental results
799 [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations
800 were performed at the average densities observed in the 1 atm NPT
801 simulations for the respective models. SSD/E is slightly more mobile
802 than experiment at all of the temperatures, but it is closer to
803 experiment at biologically relevant temperatures than SSD1 without a
804 long-range correction.}
805 \label{ssdediffuse}
806 \end{center}
807 \end{figure}
808
809 The reparameterization of the SSD water model, both for use with and
810 without an applied long-range correction, brought the densities up to
811 what is expected for simulating liquid water. In addition to improving
812 the densities, it is important that the excellent diffusive behavior
813 of SSD be maintained or improved. Figure \ref{ssdediffuse} compares
814 the temperature dependence of the diffusion constant of SSD/E to SSD1
815 without an active reaction field at the densities calculated from
816 their respective NPT simulations at 1 atm. The diffusion constant for
817 SSD/E is consistently higher than experiment, while SSD1 remains lower
818 than experiment until relatively high temperatures (around 360
819 K). Both models follow the shape of the experimental curve well below
820 300 K but tend to diffuse too rapidly at higher temperatures, as seen
821 in SSD1's crossing above 360 K. This increasing diffusion relative to
822 the experimental values is caused by the rapidly decreasing system
823 density with increasing temperature. Both SSD1 and SSD/E show this
824 deviation in particle mobility, but this trend has different
825 implications on the diffusive behavior of the models. While SSD1
826 shows more experimentally accurate diffusive behavior in the high
827 temperature regimes, SSD/E shows more accurate behavior in the
828 supercooled and biologically relevant temperature ranges. Thus, the
829 changes made to improve the liquid structure may have had an adverse
830 affect on the density maximum, but they improve the transport behavior
831 of SSD/E relative to SSD1 under the most commonly simulated
832 conditions.
833
834 \begin{figure}
835 \begin{center}
836 \epsfxsize=6in
837 \epsfbox{ssdrfDiffuse.epsi}
838 \caption{The diffusion constants calculated from SSD/RF and SSD1,
839 both with an active reaction field, along with experimental results
840 [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations
841 were performed at the average densities observed in the 1 atm NPT
842 simulations for both of the models. Note how accurately SSD/RF
843 simulates the diffusion of water throughout this temperature
844 range. The more rapidly increasing diffusion constants at high
845 temperatures for both models is attributed to lower calculated
846 densities than those observed in experiment.}
847 \label{ssdrfdiffuse}
848 \end{center}
849 \end{figure}
850
851 In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
852 compared to SSD1 with an active reaction field. Note that SSD/RF
853 tracks the experimental results quantitatively, identical within error
854 throughout most of the temperature range shown and exhibiting only a
855 slight increasing trend at higher temperatures. SSD1 tends to diffuse
856 more slowly at low temperatures and deviates to diffuse too rapidly at
857 temperatures greater than 330 K. As stated above, this deviation away
858 from the ideal trend is due to a rapid decrease in density at higher
859 temperatures. SSD/RF does not suffer from this problem as much as SSD1
860 because the calculated densities are closer to the experimental
861 values. These results again emphasize the importance of careful
862 reparameterization when using an altered long-range correction.
863
864 \begin{table}
865 \begin{center}
866 \caption{Calculated and experimental properties of the single point waters and liquid water at 298 K and 1 atm. (a) Ref. [\citen{Mills73}]. (b) Calculated by integrating the data in ref. \citen{Head-Gordon00_1}. (c) Calculated by integrating the data in ref. \citen{Soper86}. (d) Calculated for 298 K from data in ref. [\citen{Eisenberg69}]. (e) Calculated for 298 K from data in ref. \citen{Krynicki66}.}
867 \begin{tabular}{ l c c c c c }
868 \hline \\[-3mm]
869 \ \ \ \ \ \ & \ \ \ SSD1 \ \ \ & \ SSD/E \ \ \ & \ SSD1 (RF) \ \
870 \ & \ SSD/RF \ \ \ & \ Expt. \\
871 \hline \\[-3mm]
872 \ \ \ $\rho$ (g/cm$^3$) & 0.999 $\pm$0.001 & 0.996 $\pm$0.001 & 0.972 $\pm$0.002 & 0.997 $\pm$0.001 & 0.997 \\
873 \ \ \ $C_p$ (cal/mol K) & 28.80 $\pm$0.11 & 25.45 $\pm$0.09 & 28.28 $\pm$0.06 & 23.83 $\pm$0.16 & 17.98 \\
874 \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 & 2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299$^\text{a}$ \\
875 \ \ \ Coordination Number & 3.9 & 4.3 & 3.8 & 4.4 & 4.7$^\text{b}$ \\
876 \ \ \ H-bonds per particle & 3.7 & 3.6 & 3.7 & 3.7 & 3.5$^\text{c}$ \\
877 \ \ \ $\tau_1$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 & 7.2 $\pm$0.4 & 5.7$^\text{d}$ \\
878 \ \ \ $\tau_2$ (ps) & 4.7 $\pm$0.4 & 3.1 $\pm$0.2 & 3.5 $\pm$0.3 & 3.2 $\pm$0.2 & 2.3$^\text{e}$ \\
879 \end{tabular}
880 \label{liquidproperties}
881 \end{center}
882 \end{table}
883
884 Table \ref{liquidproperties} gives a synopsis of the liquid state
885 properties of the water models compared in this study along with the
886 experimental values for liquid water at ambient conditions. The
887 coordination number ($N_C$) and hydrogen bonds per particle ($N_H$)
888 were calculated by integrating the following relations:
889 \begin{equation}
890 N_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr,
891 \end{equation}
892 \begin{equation}
893 N_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr,
894 \end{equation}
895 where $\rho$ is the number density of the specified pair interactions,
896 $a$ and $b$ are the radial locations of the minima following the first
897 solvation shell peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$
898 respectively. The number of hydrogen bonds stays relatively constant
899 across all of the models, but the coordination numbers of SSD/E and
900 SSD/RF show an improvement over SSD1. This improvement is primarily
901 due to the widening of the first solvation shell peak, allowing the
902 first minima to push outward. Comparing the coordination number with
903 the number of hydrogen bonds can lead to more insight into the
904 structural character of the liquid. Because of the near identical
905 values for SSD1, it appears to be a little too exclusive, in that all
906 molecules in the first solvation shell are involved in forming ideal
907 hydrogen bonds. The differing numbers for the newly parameterized
908 models indicate the allowance of more fluid configurations in addition
909 to the formation of an acceptable number of ideal hydrogen bonds.
910
911 The time constants for the self orientational autocorrelation function
912 are also displayed in Table \ref{liquidproperties}. The dipolar
913 orientational time correlation function ($\Gamma_{l}$) is described
914 by:
915 \begin{equation}
916 \Gamma_{l}(t) = \langle P_l[\mathbf{u}_j(0)\cdot\mathbf{u}_j(t)]\rangle,
917 \end{equation}
918 where $P_l$ is a Legendre polynomial of order $l$ and $\mathbf{u}_j$
919 is the unit vector of the particle dipole.\cite{Rahman71} From these
920 correlation functions, the orientational relaxation time of the dipole
921 vector can be calculated from an exponential fit in the long-time
922 regime ($t > \tau_l$).\cite{Rothschild84} Calculation of these
923 time constants were averaged from five detailed NVE simulations
924 performed at the STP density for each of the respective models. It
925 should be noted that the commonly cited value for $\tau_2$ of 1.9 ps
926 was determined from the NMR data in reference \citen{Krynicki66} at a
927 temperature near 34$^\circ$C.\cite{Rahman71} Because of the strong
928 temperature dependence of $\tau_2$, it is necessary to recalculate it
929 at 298 K to make proper comparisons. The value shown in Table
930 \ref{liquidproperties} was calculated from the same NMR data in the
931 fashion described in reference \citen{Krynicki66}. Similarly, $\tau_1$
932 was recomputed for 298 K from the data in ref \citen{Eisenberg69}.
933 Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
934 and without an active reaction field. Turning on the reaction field
935 leads to much improved time constants for SSD1; however, these results
936 also include a corresponding decrease in system density. Numbers
937 published from the original SSD dynamics studies are shorter than the
938 values observed here, and this difference can be attributed to the use
939 of the Ewald sum technique versus a reaction field.\cite{Ichiye99}
940
941 \subsection{Additional Observations}
942
943 \begin{figure}
944 \begin{center}
945 \epsfxsize=6in
946 \epsfbox{icei_bw.eps}
947 \caption{A water lattice built from the crystal structure assumed by
948 SSD/E when undergoing an extremely restricted temperature NPT
949 simulation. This form of ice is referred to as ice-{\it i} to
950 emphasize its simulation origins. This image was taken of the (001)
951 face of the crystal.}
952 \label{weirdice}
953 \end{center}
954 \end{figure}
955
956 While performing a series of melting simulations on an early iteration
957 of SSD/E not discussed in this paper, we observed recrystallization
958 into a novel structure not previously known for water. After melting
959 at 235 K, two of five systems underwent crystallization events near
960 245 K. The two systems remained crystalline up to 320 and 330 K,
961 respectively. The crystal exhibits an expanded zeolite-like structure
962 that does not correspond to any known form of ice. This appears to be
963 an artifact of the point dipolar models, so to distinguish it from the
964 experimentally observed forms of ice, we have denoted the structure
965 Ice-$\sqrt{\smash[b]{-\text{I}}}$ (ice-{\it i}). A large enough
966 portion of the sample crystallized that we have been able to obtain a
967 near ideal crystal structure of ice-{\it i}. Figure \ref{weirdice}
968 shows the repeating crystal structure of a typical crystal at 5
969 K. Each water molecule is hydrogen bonded to four others; however, the
970 hydrogen bonds are bent rather than perfectly straight. This results
971 in a skewed tetrahedral geometry about the central molecule. In
972 figure \ref{isosurface}, it is apparent that these flexed hydrogen
973 bonds are allowed due to the conical shape of the attractive regions,
974 with the greatest attraction along the direct hydrogen bond
975 configuration. Though not ideal, these flexed hydrogen bonds are
976 favorable enough to stabilize an entire crystal generated around them.
977
978 Initial simulations indicated that ice-{\it i} is the preferred ice
979 structure for at least the SSD/E model. To verify this, a comparison
980 was made between near ideal crystals of ice~$I_h$, ice~$I_c$, and
981 ice-{\it i} at constant pressure with SSD/E, SSD/RF, and
982 SSD1. Near-ideal versions of the three types of crystals were cooled
983 to 1 K, and the enthalpies of each were compared using all three water
984 models. With every model in the SSD family, ice-{\it i} had the lowest
985 calculated enthalpy: 5\% lower than $I_h$ with SSD1, 6.5\% lower with
986 SSD/E, and 7.5\% lower with SSD/RF. The enthalpy data is summarized
987 in Table \ref{iceenthalpy}.
988
989 \begin{table}
990 \begin{center}
991 \caption{Enthalpies (in kcal / mol) of the three crystal structures (at 1
992 K) exhibited by the SSD family of water models}
993 \begin{tabular}{ l c c c }
994 \hline \\[-3mm]
995 \ \ \ Water Model \ \ \ & \ \ \ Ice-$I_h$ \ \ \ & \ Ice-$I_c$\ \ & \
996 Ice-{\it i} \\
997 \hline \\[-3mm]
998 \ \ \ SSD/E & -12.286 & -12.292 & -13.590 \\
999 \ \ \ SSD/RF & -12.935 & -12.917 & -14.022 \\
1000 \ \ \ SSD1 & -12.496 & -12.411 & -13.417 \\
1001 \ \ \ SSD1 (RF) & -12.504 & -12.411 & -13.134 \\
1002 \end{tabular}
1003 \label{iceenthalpy}
1004 \end{center}
1005 \end{table}
1006
1007 In addition to these energetic comparisons, melting simulations were
1008 performed with ice-{\it i} as the initial configuration using SSD/E,
1009 SSD/RF, and SSD1 both with and without a reaction field. The melting
1010 transitions for both SSD/E and SSD1 without reaction field occurred at
1011 temperature in excess of 375~K. SSD/RF and SSD1 with a reaction field
1012 showed more reasonable melting transitions near 325~K. These melting
1013 point observations clearly show that all of the SSD-derived models
1014 prefer the ice-{\it i} structure.
1015
1016 \section{Conclusions}
1017
1018 The density maximum and temperature dependence of the self-diffusion
1019 constant were studied for the SSD water model, both with and without
1020 the use of reaction field, via a series of NPT and NVE
1021 simulations. The constant pressure simulations showed a density
1022 maximum near 260 K. In most cases, the calculated densities were
1023 significantly lower than the densities obtained from other water
1024 models (and experiment). Analysis of self-diffusion showed SSD to
1025 capture the transport properties of water well in both the liquid and
1026 supercooled liquid regimes.
1027
1028 In order to correct the density behavior, the original SSD model was
1029 reparameterized for use both with and without a reaction field (SSD/RF
1030 and SSD/E), and comparisons were made with SSD1, Ichiye's density
1031 corrected version of SSD. Both models improve the liquid structure,
1032 densities, and diffusive properties under their respective simulation
1033 conditions, indicating the necessity of reparameterization when
1034 changing the method of calculating long-range electrostatic
1035 interactions. In general, however, these simple water models are
1036 excellent choices for representing explicit water in large scale
1037 simulations of biochemical systems.
1038
1039 The existence of a novel low-density ice structure that is preferred
1040 by the SSD family of water models is somewhat troubling, since liquid
1041 simulations on this family of water models at room temperature are
1042 effectively simulations of supercooled or metastable liquids. One
1043 way to destabilize this unphysical ice structure would be to make the
1044 range of angles preferred by the attractive part of the sticky
1045 potential much narrower. This would require extensive
1046 reparameterization to maintain the same level of agreement with the
1047 experiments.
1048
1049 Additionally, our initial calculations show that the ice-{\it i}
1050 structure may also be a preferred crystal structure for at least one
1051 other popular multi-point water model (TIP3P), and that much of the
1052 simulation work being done using this popular model could also be at
1053 risk for crystallization into this unphysical structure. A future
1054 publication will detail the relative stability of the known ice
1055 structures for a wide range of popular water models.
1056
1057 \section{Acknowledgments}
1058 Support for this project was provided by the National Science
1059 Foundation under grant CHE-0134881. Computation time was provided by
1060 the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
1061 DMR-0079647.
1062
1063 \newpage
1064
1065 \bibliographystyle{jcp}
1066 \bibliography{nptSSD}
1067
1068 %\pagebreak
1069
1070 \end{document}