ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 1033
Committed: Fri Feb 6 19:15:44 2004 UTC (20 years, 4 months ago) by chrisfen
Content type: application/x-tex
File size: 66536 byte(s)
Log Message:
I hate revtex

File Contents

# Content
1 %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 \documentclass[preprint,aps]{revtex4}
3 %\documentclass[11pt]{article}
4 %\usepackage{endfloat}
5 \usepackage{amsmath}
6 \usepackage{epsf}
7 \usepackage{berkeley}
8 \usepackage{setspace}
9 \usepackage{tabularx}
10 \usepackage{graphicx}
11 %\usepackage[ref]{overcite}
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 \newcounter{captions}
22 \newcounter{figs}
23
24 \begin{document}
25
26 \title{On the structural and transport properties of the soft sticky
27 dipole (SSD) and related single point water models}
28
29 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu}}
30
31 \affiliation{Department of Chemistry and Biochemistry\\ University of Notre Dame\\
32 Notre Dame, Indiana 46556}
33
34 \date{\today}
35
36
37 \begin{abstract}
38 The density maximum and temperature dependence of the self-diffusion
39 constant were investigated for the soft sticky dipole (SSD) water
40 model and two related reparameterizations of this single-point model.
41 A combination of microcanonical and isobaric-isothermal molecular
42 dynamics simulations were used to calculate these properties, both
43 with and without the use of reaction field to handle long-range
44 electrostatics. The isobaric-isothermal (NPT) simulations of the
45 melting of both ice-$I_h$ and ice-$I_c$ showed a density maximum near
46 260~K. In most cases, the use of the reaction field resulted in
47 calculated densities which were were significantly lower than
48 experimental densities. Analysis of self-diffusion constants shows
49 that the original SSD model captures the transport properties of
50 experimental water very well in both the normal and super-cooled
51 liquid regimes. We also present our reparameterized versions of SSD
52 for use both with the reaction field or without any long-range
53 electrostatic corrections. These are called the SSD/RF and SSD/E
54 models respectively. These modified models were shown to maintain or
55 improve upon the experimental agreement with the structural and
56 transport properties that can be obtained with either the original SSD
57 or the density corrected version of the original model (SSD1).
58 Additionally, a novel low-density ice structure is presented
59 which appears to be the most stable ice structure for the entire SSD
60 family.
61 \end{abstract}
62
63 \maketitle
64
65 \newpage
66
67 %\narrowtext
68
69
70 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71 % BODY OF TEXT
72 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73
74 \section{Introduction}
75
76 One of the most important tasks in the simulation of biochemical
77 systems is the proper depiction of the aqueous environment of the
78 molecules of interest. In some cases (such as in the simulation of
79 phospholipid bilayers), the majority of the calculations that are
80 performed involve interactions with or between solvent molecules.
81 Thus, the properties one may observe in biochemical simulations are
82 going to be highly dependent on the physical properties of the water
83 model that is chosen.
84
85 There is an especially delicate balance between computational
86 efficiency and the ability of the water model to accurately predict
87 the properties of bulk
88 water.\cite{Jorgensen83,Berendsen87,Jorgensen00} For example, the
89 TIP5P model improves on the structural and transport properties of
90 water relative to the previous TIP models, yet this comes at a greater
91 than 50\% increase in computational
92 cost.\cite{Jorgensen01,Jorgensen00}
93
94 One recently developed model that largely succeeds in retaining the
95 accuracy of bulk properties while greatly reducing the computational
96 cost is the Soft Sticky Dipole (SSD) water
97 model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The SSD model
98 was developed by Ichiye \emph{et al.} as a modified form of the
99 hard-sphere water model proposed by Bratko, Blum, and
100 Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point} model
101 which has an interaction site that is both a point dipole and a
102 Lennard-Jones core. However, since the normal aligned and
103 anti-aligned geometries favored by point dipoles are poor mimics of
104 local structure in liquid water, a short ranged ``sticky'' potential
105 is also added. The sticky potential directs the molecules to assume
106 the proper hydrogen bond orientation in the first solvation shell.
107
108 The interaction between two SSD water molecules \emph{i} and \emph{j}
109 is given by the potential
110 \begin{equation}
111 u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
112 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ +
113 u_{ij}^{sp}
114 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j),
115 \end{equation}
116 where the ${\bf r}_{ij}$ is the position vector between molecules
117 \emph{i} and \emph{j} with magnitude $r_{ij}$, and
118 ${\bf \Omega}_i$ and ${\bf \Omega}_j$ represent the orientations of
119 the two molecules. The Lennard-Jones and dipole interactions are given
120 by the following familiar forms:
121 \begin{equation}
122 u_{ij}^{LJ}(r_{ij}) = 4\epsilon
123 \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right]
124 \ ,
125 \end{equation}
126 and
127 \begin{equation}
128 u_{ij}^{dp} = \frac{|\mu_i||\mu_j|}{4 \pi \epsilon_0 r_{ij}^3} \left(
129 \hat{\bf u}_i \cdot \hat{\bf u}_j - 3(\hat{\bf u}_i\cdot\hat{\bf
130 r}_{ij})(\hat{\bf u}_j\cdot\hat{\bf r}_{ij}) \right)\ ,
131 \end{equation}
132 where $\hat{\bf u}_i$ and $\hat{\bf u}_j$ are the unit vectors along
133 the dipoles of molecules $i$ and $j$ respectively. $|\mu_i|$ and
134 $|\mu_j|$ are the strengths of the dipole moments, and $\hat{\bf
135 r}_{ij}$ is the unit vector pointing from molecule $j$ to molecule
136 $i$.
137
138 The sticky potential is somewhat less familiar:
139 \begin{equation}
140 u_{ij}^{sp}
141 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
142 \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)
143 + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf
144 \Omega}_j)]\ .
145 \label{stickyfunction}
146 \end{equation}
147 Here, $\nu_0$ is a strength parameter for the sticky potential, and
148 $s$ and $s^\prime$ are cubic switching functions which turn off the
149 sticky interaction beyond the first solvation shell. The $w$ function
150 can be thought of as an attractive potential with tetrahedral
151 geometry:
152 \begin{equation}
153 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
154 \end{equation}
155 while the $w^\prime$ function counters the normal aligned and
156 anti-aligned structures favored by point dipoles:
157 \begin{equation}
158 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,
159 \end{equation}
160 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
161 and $Y_3^{-2}$ spherical harmonics (a linear combination which
162 enhances the tetrahedral geometry for hydrogen bonded structures),
163 while $w^\prime$ is a purely empirical function. A more detailed
164 description of the functional parts and variables in this potential
165 can be found in the original SSD
166 articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
167
168 Since SSD is a single-point {\it dipolar} model, the force
169 calculations are simplified significantly relative to the standard
170 {\it charged} multi-point models. In the original Monte Carlo
171 simulations using this model, Liu and Ichiye reported that using SSD
172 decreased computer time by a factor of 6-7 compared to other
173 models.\cite{Ichiye96} What is most impressive is that this savings
174 did not come at the expense of accurate depiction of the liquid state
175 properties. Indeed, SSD maintains reasonable agreement with the Soper
176 data for the structural features of liquid
177 water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
178 exhibited by SSD agree with experiment better than those of more
179 computationally expensive models (like TIP3P and
180 SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
181 of solvent properties makes SSD a very attractive model for the
182 simulation of large scale biochemical simulations.
183
184 One feature of the SSD model is that it was parameterized for
185 use with the Ewald sum to handle long-range interactions. This would
186 normally be the best way of handling long-range interactions in
187 systems that contain other point charges. However, our group has
188 recently become interested in systems with point dipoles as mimics for
189 neutral, but polarized regions on molecules (e.g. the zwitterionic
190 head group regions of phospholipids). If the system of interest does
191 not contain point charges, the Ewald sum and even particle-mesh Ewald
192 become computational bottlenecks. Their respective ideal
193 $N^\frac{3}{2}$ and $N\log N$ calculation scaling orders for $N$
194 particles can become prohibitive when $N$ becomes
195 large.\cite{Darden99} In applying this water model in these types of
196 systems, it would be useful to know its properties and behavior under
197 the more computationally efficient reaction field (RF) technique, or
198 even with a simple cutoff. This study addresses these issues by
199 looking at the structural and transport behavior of SSD over a
200 variety of temperatures with the purpose of utilizing the RF
201 correction technique. We then suggest modifications to the parameters
202 that result in more realistic bulk phase behavior. It should be noted
203 that in a recent publication, some of the original investigators of
204 the SSD water model have suggested adjustments to the SSD
205 water model to address abnormal density behavior (also observed here),
206 calling the corrected model SSD1.\cite{Ichiye03} In what
207 follows, we compare our reparamaterization of SSD with both the
208 original SSD and SSD1 models with the goal of improving
209 the bulk phase behavior of an SSD-derived model in simulations
210 utilizing the reaction field.
211
212 \section{Methods}
213
214 Long-range dipole-dipole interactions were accounted for in this study
215 by using either the reaction field technique or by resorting to a
216 simple cubic switching function at a cutoff radius. One of the early
217 applications of a reaction field was actually in Monte Carlo
218 simulations of liquid water.\cite{Barker73} Under this method, the
219 magnitude of the reaction field acting on dipole $i$ is
220 \begin{equation}
221 \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
222 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
223 \label{rfequation}
224 \end{equation}
225 where $\mathcal{R}$ is the cavity defined by the cutoff radius
226 ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
227 system (80 in the case of liquid water), ${\bf \mu}_{j}$ is the dipole
228 moment vector of particle $j$, and $s(r_{ij})$ is a cubic switching
229 function.\cite{AllenTildesley} The reaction field contribution to the
230 total energy by particle $i$ is given by $-\frac{1}{2}{\bf
231 \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf
232 \mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use of the reaction
233 field is known to alter the bulk orientational properties of simulated
234 water, and there is particular sensitivity of these properties on
235 changes in the length of the cutoff radius.\cite{Berendsen98} This
236 variable behavior makes reaction field a less attractive method than
237 the Ewald sum. However, for very large systems, the computational
238 benefit of reaction field is dramatic.
239
240 We have also performed a companion set of simulations {\it without} a
241 surrounding dielectric (i.e. using a simple cubic switching function
242 at the cutoff radius), and as a result we have two reparamaterizations
243 of SSD which could be used either with or without the reaction
244 field turned on.
245
246 Simulations to obtain the preferred densities of the models were
247 performed in the isobaric-isothermal (NPT) ensemble, while all
248 dynamical properties were obtained from microcanonical (NVE)
249 simulations done at densities matching the NPT density for a
250 particular target temperature. The constant pressure simulations were
251 implemented using an integral thermostat and barostat as outlined by
252 Hoover.\cite{Hoover85,Hoover86} All molecules were treated as
253 non-linear rigid bodies. Vibrational constraints are not necessary in
254 simulations of SSD, because there are no explicit hydrogen
255 atoms, and thus no molecular vibrational modes need to be considered.
256
257 Integration of the equations of motion was carried out using the
258 symplectic splitting method proposed by Dullweber, Leimkuhler, and
259 McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting
260 this integrator centers on poor energy conservation of rigid body
261 dynamics using traditional quaternion
262 integration.\cite{Evans77,Evans77b} In typical microcanonical ensemble
263 simulations, the energy drift when using quaternions was substantially
264 greater than when using the {\sc dlm} method (fig. \ref{timestep}).
265 This steady drift in the total energy has also been observed by Kol
266 {\it et al.}\cite{Laird97}
267
268 The key difference in the integration method proposed by Dullweber
269 \emph{et al.} is that the entire rotation matrix is propagated from
270 one time step to the next. The additional memory required by the
271 algorithm is inconsequential on modern computers, and translating the
272 rotation matrix into quaternions for storage purposes makes trajectory
273 data quite compact.
274
275 The {\sc dlm} method allows for Verlet style integration of both
276 translational and orientational motion of rigid bodies. In this
277 integration method, the orientational propagation involves a sequence
278 of matrix evaluations to update the rotation
279 matrix.\cite{Dullweber1997} These matrix rotations are more costly
280 than the simpler arithmetic quaternion propagation. With the same time
281 step, a 1000 SSD particle simulation shows an average 7\%
282 increase in computation time using the {\sc dlm} method in place of
283 quaternions. The additional expense per step is justified when one
284 considers the ability to use time steps that are nearly twice as large
285 under {\sc dlm} than would be usable under quaternion dynamics. The
286 energy conservation of the two methods using a number of different
287 time steps is illustrated in figure
288 \ref{timestep}.
289
290 %\begin{figure}
291 %\begin{center}
292 %\epsfxsize=6in
293 %\epsfbox{timeStep.epsi}
294 %\caption{Energy conservation using both quaternion-based integration and
295 %the {\sc dlm} method with increasing time step. The larger time step
296 %plots are shifted from the true energy baseline (that of $\Delta t$ =
297 %0.1~fs) for clarity.}
298 %\label{timestep}
299 %\end{center}
300 %\end{figure}
301
302 In figure \ref{timestep}, the resulting energy drift at various time
303 steps for both the {\sc dlm} and quaternion integration schemes is
304 compared. All of the 1000 SSD particle simulations started with
305 the same configuration, and the only difference was the method used to
306 handle orientational motion. At time steps of 0.1 and 0.5~fs, both
307 methods for propagating the orientational degrees of freedom conserve
308 energy fairly well, with the quaternion method showing a slight energy
309 drift over time in the 0.5~fs time step simulation. At time steps of 1
310 and 2~fs, the energy conservation benefits of the {\sc dlm} method are
311 clearly demonstrated. Thus, while maintaining the same degree of
312 energy conservation, one can take considerably longer time steps,
313 leading to an overall reduction in computation time.
314
315 Energy drift in the simulations using {\sc dlm} integration was
316 unnoticeable for time steps up to 3~fs. A slight energy drift on the
317 order of 0.012~kcal/mol per nanosecond was observed at a time step of
318 4~fs, and as expected, this drift increases dramatically with
319 increasing time step. To insure accuracy in our microcanonical
320 simulations, time steps were set at 2~fs and kept at this value for
321 constant pressure simulations as well.
322
323 Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices
324 were generated as starting points for all simulations. The $I_h$
325 crystals were formed by first arranging the centers of mass of the SSD
326 particles into a ``hexagonal'' ice lattice of 1024 particles. Because
327 of the crystal structure of $I_h$ ice, the simulation box assumed an
328 orthorhombic shape with an edge length ratio of approximately
329 1.00$\times$1.06$\times$1.23. The particles were then allowed to
330 orient freely about fixed positions with angular momenta randomized at
331 400~K for varying times. The rotational temperature was then scaled
332 down in stages to slowly cool the crystals to 25~K. The particles were
333 then allowed to translate with fixed orientations at a constant
334 pressure of 1 atm for 50~ps at 25~K. Finally, all constraints were
335 removed and the ice crystals were allowed to equilibrate for 50~ps at
336 25~K and a constant pressure of 1~atm. This procedure resulted in
337 structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
338 rules.\cite{Bernal33,Rahman72} This method was also utilized in the
339 making of diamond lattice $I_c$ ice crystals, with each cubic
340 simulation box consisting of either 512 or 1000 particles. Only
341 isotropic volume fluctuations were performed under constant pressure,
342 so the ratio of edge lengths remained constant throughout the
343 simulations.
344
345 \section{Results and discussion}
346
347 Melting studies were performed on the randomized ice crystals using
348 isobaric-isothermal (NPT) dynamics. During melting simulations, the
349 melting transition and the density maximum can both be observed,
350 provided that the density maximum occurs in the liquid and not the
351 supercooled regime. An ensemble average from five separate melting
352 simulations was acquired, each starting from different ice crystals
353 generated as described previously. All simulations were equilibrated
354 for 100~ps prior to a 200~ps data collection run at each temperature
355 setting. The temperature range of study spanned from 25 to 400~K, with
356 a maximum degree increment of 25~K. For regions of interest along this
357 stepwise progression, the temperature increment was decreased from
358 25~K to 10 and 5~K. The above equilibration and production times were
359 sufficient in that fluctuations in the volume autocorrelation function
360 were damped out in all simulations in under 20~ps.
361
362 \subsection{Density Behavior}
363
364 Our initial simulations focused on the original SSD water model,
365 and an average density versus temperature plot is shown in figure
366 \ref{dense1}. Note that the density maximum when using a reaction
367 field appears between 255 and 265~K. There were smaller fluctuations
368 in the density at 260~K than at either 255 or 265~K, so we report this
369 value as the location of the density maximum. Figure \ref{dense1} was
370 constructed using ice $I_h$ crystals for the initial configuration;
371 though not pictured, the simulations starting from ice $I_c$ crystal
372 configurations showed similar results, with a liquid-phase density
373 maximum in this same region (between 255 and 260~K).
374
375 %\begin{figure}
376 %\begin{center}
377 %\epsfxsize=6in
378 %\epsfbox{denseSSDnew.eps}
379 %\caption{Density versus temperature for TIP4P [Ref. \onlinecite{Jorgensen98b}],
380 % TIP3P [Ref. \onlinecite{Jorgensen98b}], SPC/E [Ref. \onlinecite{Clancy94}], SSD
381 % without Reaction Field, SSD, and experiment [Ref. \onlinecite{CRC80}]. The
382 % arrows indicate the change in densities observed when turning off the
383 % reaction field. The the lower than expected densities for the SSD
384 % model were what prompted the original reparameterization of SSD1
385 % [Ref. \onlinecite{Ichiye03}].}
386 %\label{dense1}
387 %\end{center}
388 %\end{figure}
389
390 The density maximum for SSD compares quite favorably to other
391 simple water models. Figure \ref{dense1} also shows calculated
392 densities of several other models and experiment obtained from other
393 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
394 models, SSD has a temperature closest to the experimentally
395 observed density maximum. Of the {\it charge-based} models in
396 Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
397 seen in SSD. Though not included in this plot, it is useful to
398 note that TIP5P has a density maximum nearly identical to the
399 experimentally measured temperature.
400
401 It has been observed that liquid state densities in water are
402 dependent on the cutoff radius used both with and without the use of
403 reaction field.\cite{Berendsen98} In order to address the possible
404 effect of cutoff radius, simulations were performed with a dipolar
405 cutoff radius of 12.0~\AA\ to complement the previous SSD
406 simulations, all performed with a cutoff of 9.0~\AA. All of the
407 resulting densities overlapped within error and showed no significant
408 trend toward lower or higher densities as a function of cutoff radius,
409 for simulations both with and without reaction field. These results
410 indicate that there is no major benefit in choosing a longer cutoff
411 radius in simulations using SSD. This is advantageous in that
412 the use of a longer cutoff radius results in a significant increase in
413 the time required to obtain a single trajectory.
414
415 The key feature to recognize in figure \ref{dense1} is the density
416 scaling of SSD relative to other common models at any given
417 temperature. SSD assumes a lower density than any of the other
418 listed models at the same pressure, behavior which is especially
419 apparent at temperatures greater than 300~K. Lower than expected
420 densities have been observed for other systems using a reaction field
421 for long-range electrostatic interactions, so the most likely reason
422 for the significantly lower densities seen in these simulations is the
423 presence of the reaction field.\cite{Berendsen98,Nezbeda02} In order
424 to test the effect of the reaction field on the density of the
425 systems, the simulations were repeated without a reaction field
426 present. The results of these simulations are also displayed in figure
427 \ref{dense1}. Without the reaction field, the densities increase
428 to more experimentally reasonable values, especially around the
429 freezing point of liquid water. The shape of the curve is similar to
430 the curve produced from SSD simulations using reaction field,
431 specifically the rapidly decreasing densities at higher temperatures;
432 however, a shift in the density maximum location, down to 245~K, is
433 observed. This is a more accurate comparison to the other listed water
434 models, in that no long range corrections were applied in those
435 simulations.\cite{Clancy94,Jorgensen98b} However, even without the
436 reaction field, the density around 300~K is still significantly lower
437 than experiment and comparable water models. This anomalous behavior
438 was what lead Tan {\it et al.} to recently reparameterize
439 SSD.\cite{Ichiye03} Throughout the remainder of the paper our
440 reparamaterizations of SSD will be compared with their newer SSD1
441 model.
442
443 \subsection{Transport Behavior}
444
445 Accurate dynamical properties of a water model are particularly
446 important when using the model to study permeation or transport across
447 biological membranes. In order to probe transport in bulk water,
448 constant energy (NVE) simulations were performed at the average
449 density obtained by the NPT simulations at an identical target
450 temperature. Simulations started with randomized velocities and
451 underwent 50~ps of temperature scaling and 50~ps of constant energy
452 equilibration before a 200~ps data collection run. Diffusion constants
453 were calculated via linear fits to the long-time behavior of the
454 mean-square displacement as a function of time. The averaged results
455 from five sets of NVE simulations are displayed in figure
456 \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
457 results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
458
459 %\begin{figure}
460 %\begin{center}
461 %\epsfxsize=6in
462 %\epsfbox{betterDiffuse.epsi}
463 %\caption{Average self-diffusion constant as a function of temperature for
464 %SSD, SPC/E [Ref. \onlinecite{Clancy94}], and TIP5P
465 %[Ref. \onlinecite{Jorgensen01}] compared with experimental data
466 %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. Of the three water models
467 %shown, SSD has the least deviation from the experimental values. The
468 %rapidly increasing diffusion constants for TIP5P and SSD correspond to
469 %significant decreases in density at the higher temperatures.}
470 %\label{diffuse}
471 %\end{center}
472 %\end{figure}
473
474 The observed values for the diffusion constant point out one of the
475 strengths of the SSD model. Of the three models shown, the SSD model
476 has the most accurate depiction of self-diffusion in both the
477 supercooled and liquid regimes. SPC/E does a respectable job by
478 reproducing values similar to experiment around 290~K; however, it
479 deviates at both higher and lower temperatures, failing to predict the
480 correct thermal trend. TIP5P and SSD both start off low at colder
481 temperatures and tend to diffuse too rapidly at higher temperatures.
482 This behavior at higher temperatures is not particularly surprising
483 since the densities of both TIP5P and SSD are lower than experimental
484 water densities at higher temperatures. When calculating the
485 diffusion coefficients for SSD at experimental densities
486 (instead of the densities from the NPT simulations), the resulting
487 values fall more in line with experiment at these temperatures.
488
489 \subsection{Structural Changes and Characterization}
490
491 By starting the simulations from the crystalline state, the melting
492 transition and the ice structure can be obtained along with the liquid
493 phase behavior beyond the melting point. The constant pressure heat
494 capacity (C$_\text{p}$) was monitored to locate the melting transition
495 in each of the simulations. In the melting simulations of the 1024
496 particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs
497 at 245~K, indicating a first order phase transition for the melting of
498 these ice crystals. When the reaction field is turned off, the melting
499 transition occurs at 235~K. These melting transitions are
500 considerably lower than the experimental value.
501
502 %\begin{figure}
503 %\begin{center}
504 %\epsfxsize=6in
505 %\epsfbox{corrDiag.eps}
506 %\caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
507 %\label{corrAngle}
508 %\end{center}
509 %\end{figure}
510
511 %\begin{figure}
512 %\begin{center}
513 %\epsfxsize=6in
514 %\epsfbox{fullContours.eps}
515 %\caption{Contour plots of 2D angular pair correlation functions for
516 %512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
517 %signify regions of enhanced density while light areas signify
518 %depletion relative to the bulk density. White areas have pair
519 %correlation values below 0.5 and black areas have values above 1.5.}
520 %\label{contour}
521 %\end{center}
522 %\end{figure}
523
524 Additional analysis of the melting process was performed using
525 two-dimensional structure and dipole angle correlations. Expressions
526 for these correlations are as follows:
527
528 \begin{equation}
529 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\ ,
530 \end{equation}
531 \begin{equation}
532 g_{\text{AB}}(r,\cos\omega) =
533 \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\ ,
534 \end{equation}
535 where $\theta$ and $\omega$ refer to the angles shown in figure
536 \ref{corrAngle}. By binning over both distance and the cosine of the
537 desired angle between the two dipoles, the $g(r)$ can be analyzed to
538 determine the common dipole arrangements that constitute the peaks and
539 troughs in the standard one-dimensional $g(r)$ plots. Frames A and B
540 of figure \ref{contour} show results from an ice $I_c$ simulation. The
541 first peak in the $g(r)$ consists primarily of the preferred hydrogen
542 bonding arrangements as dictated by the tetrahedral sticky potential -
543 one peak for the hydrogen bond donor and the other for the hydrogen
544 bond acceptor. Due to the high degree of crystallinity of the sample,
545 the second and third solvation shells show a repeated peak arrangement
546 which decays at distances around the fourth solvation shell, near the
547 imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
548 In the higher temperature simulation shown in frames C and D, these
549 long-range features deteriorate rapidly. The first solvation shell
550 still shows the strong effect of the sticky-potential, although it
551 covers a larger area, extending to include a fraction of aligned
552 dipole peaks within the first solvation shell. The latter peaks lose
553 due to thermal motion and as the competing dipole force overcomes the
554 sticky potential's tight tetrahedral structuring of the crystal.
555
556 This complex interplay between dipole and sticky interactions was
557 remarked upon as a possible reason for the split second peak in the
558 oxygen-oxygen pair correlation function,
559 $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures, the second
560 solvation shell peak appears to have two distinct components that
561 blend together to form one observable peak. At higher temperatures,
562 this split character alters to show the leading 4~\AA\ peak dominated
563 by equatorial anti-parallel dipole orientations. There is also a
564 tightly bunched group of axially arranged dipoles that most likely
565 consist of the smaller fraction of aligned dipole pairs. The trailing
566 component of the split peak at 5~\AA\ is dominated by aligned dipoles
567 that assume hydrogen bond arrangements similar to those seen in the
568 first solvation shell. This evidence indicates that the dipole pair
569 interaction begins to dominate outside of the range of the dipolar
570 repulsion term. The energetically favorable dipole arrangements
571 populate the region immediately outside this repulsion region (around
572 4~\AA), while arrangements that seek to satisfy both the sticky and
573 dipole forces locate themselves just beyond this initial buildup
574 (around 5~\AA).
575
576 From these findings, the split second peak is primarily the product of
577 the dipolar repulsion term of the sticky potential. In fact, the inner
578 peak can be pushed out and merged with the outer split peak just by
579 extending the switching function ($s^\prime(r_{ij})$) from its normal
580 4.0~\AA\ cutoff to values of 4.5 or even 5~\AA. This type of
581 correction is not recommended for improving the liquid structure,
582 since the second solvation shell would still be shifted too far
583 out. In addition, this would have an even more detrimental effect on
584 the system densities, leading to a liquid with a more open structure
585 and a density considerably lower than the already low SSD
586 density. A better correction would be to include the
587 quadrupole-quadrupole interactions for the water particles outside of
588 the first solvation shell, but this would remove the simplicity and
589 speed advantage of SSD.
590
591 \subsection{Adjusted Potentials: SSD/RF and SSD/E}
592
593 The propensity of SSD to adopt lower than expected densities under
594 varying conditions is troubling, especially at higher temperatures. In
595 order to correct this model for use with a reaction field, it is
596 necessary to adjust the force field parameters for the primary
597 intermolecular interactions. In undergoing a reparameterization, it is
598 important not to focus on just one property and neglect the other
599 important properties. In this case, it would be ideal to correct the
600 densities while maintaining the accurate transport behavior.
601
602 The parameters available for tuning include the $\sigma$ and
603 $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
604 strength of the sticky potential ($\nu_0$), and the cutoff distances
605 for the sticky attractive and dipole repulsive cubic switching
606 function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
607 respectively). The results of the reparameterizations are shown in
608 table \ref{params}. We are calling these reparameterizations the Soft
609 Sticky Dipole / Reaction Field (SSD/RF - for use with a reaction
610 field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve
611 the liquid structure in simulations without a long-range correction).
612
613 \begin{table}
614 \begin{center}
615 \caption{Parameters for the original and adjusted models}
616 \begin{tabular}{ l c c c c }
617 \hline \\[-3mm]
618 \ \ \ Parameters\ \ \ & \ \ \ SSD [Ref. \onlinecite{Ichiye96}] \ \ \
619 & \ SSD1 [Ref. \onlinecite{Ichiye03}]\ \ & \ SSD/E\ \ & \ \ SSD/RF \\
620 \hline \\[-3mm]
621 \ \ \ $\sigma$ (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
622 \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
623 \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
624 \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
625 \ \ \ $\omega^\circ$ & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
626 \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
627 \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
628 \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
629 \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
630 \end{tabular}
631 \label{params}
632 \end{center}
633 \end{table}
634
635 %\begin{figure}
636 %\begin{center}
637 %\epsfxsize=5in
638 %\epsfbox{GofRCompare.epsi}
639 %\caption{Plots comparing experiment [Ref. \onlinecite{Head-Gordon00_1}] with
640 %SSD/E and SSD1 without reaction field (top), as well as
641 %SSD/RF and SSD1 with reaction field turned on
642 %(bottom). The insets show the respective first peaks in detail. Note
643 %how the changes in parameters have lowered and broadened the first
644 %peak of SSD/E and SSD/RF.}
645 %\label{grcompare}
646 %\end{center}
647 %\end{figure}
648
649 %\begin{figure}
650 %\begin{center}
651 %\epsfxsize=6in
652 %\epsfbox{dualsticky_bw.eps}
653 %\caption{Positive and negative isosurfaces of the sticky potential for
654 %SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
655 %correspond to the tetrahedral attractive component, and darker areas
656 %correspond to the dipolar repulsive component.}
657 %\label{isosurface}
658 %\end{center}
659 %\end{figure}
660
661 In the original paper detailing the development of SSD, Liu and Ichiye
662 placed particular emphasis on an accurate description of the first
663 solvation shell. This resulted in a somewhat tall and narrow first
664 peak in $g(r)$ that integrated to give similar coordination numbers to
665 the experimental data obtained by Soper and
666 Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
667 data from the Head-Gordon lab indicates a slightly lower and shifted
668 first peak in the g$_\mathrm{OO}(r)$, so our adjustments to SSD were
669 made after taking into consideration the new experimental
670 findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
671 relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
672 the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
673 experimental results. Both modified water models have shorter peaks
674 that match more closely to the experimental peak (as seen in the
675 insets of figure \ref{grcompare}). This structural alteration was
676 accomplished by the combined reduction in the Lennard-Jones $\sigma$
677 variable and adjustment of the sticky potential strength and cutoffs.
678 As can be seen in table \ref{params}, the cutoffs for the tetrahedral
679 attractive and dipolar repulsive terms were nearly swapped with each
680 other. Isosurfaces of the original and modified sticky potentials are
681 shown in figure \ref{isosurface}. In these isosurfaces, it is easy to
682 see how altering the cutoffs changes the repulsive and attractive
683 character of the particles. With a reduced repulsive surface (darker
684 region), the particles can move closer to one another, increasing the
685 density for the overall system. This change in interaction cutoff
686 also results in a more gradual orientational motion by allowing the
687 particles to maintain preferred dipolar arrangements before they begin
688 to feel the pull of the tetrahedral restructuring. As the particles
689 move closer together, the dipolar repulsion term becomes active and
690 excludes unphysical nearest-neighbor arrangements. This compares with
691 how SSD and SSD1 exclude preferred dipole alignments before the
692 particles feel the pull of the ``hydrogen bonds''. Aside from
693 improving the shape of the first peak in the g(\emph{r}), this
694 modification improves the densities considerably by allowing the
695 persistence of full dipolar character below the previous 4.0~\AA\
696 cutoff.
697
698 While adjusting the location and shape of the first peak of $g(r)$
699 improves the densities, these changes alone are insufficient to bring
700 the system densities up to the values observed experimentally. To
701 further increase the densities, the dipole moments were increased in
702 both of our adjusted models. Since SSD is a dipole based model, the
703 structure and transport are very sensitive to changes in the dipole
704 moment. The original SSD simply used the dipole moment calculated from
705 the TIP3P water model, which at 2.35~D is significantly greater than
706 the experimental gas phase value of 1.84~D. The larger dipole moment
707 is a more realistic value and improves the dielectric properties of
708 the fluid. Both theoretical and experimental measurements indicate a
709 liquid phase dipole moment ranging from 2.4~D to values as high as
710 3.11~D, providing a substantial range of reasonable values for a
711 dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
712 increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF,
713 respectively, leads to significant changes in the density and
714 transport of the water models.
715
716 In order to demonstrate the benefits of these reparameterizations, a
717 series of NPT and NVE simulations were performed to probe the density
718 and transport properties of the adapted models and compare the results
719 to the original SSD model. This comparison involved full NPT melting
720 sequences for both SSD/E and SSD/RF, as well as NVE transport
721 calculations at the calculated self-consistent densities. Again, the
722 results are obtained from five separate simulations of 1024 particle
723 systems, and the melting sequences were started from different ice
724 $I_h$ crystals constructed as described previously. Each NPT
725 simulation was equilibrated for 100~ps before a 200~ps data collection
726 run at each temperature step, and the final configuration from the
727 previous temperature simulation was used as a starting point. All NVE
728 simulations had the same thermalization, equilibration, and data
729 collection times as stated previously.
730
731 %\begin{figure}
732 %\begin{center}
733 %\epsfxsize=6in
734 %\epsfbox{ssdeDense.epsi}
735 %\caption{Comparison of densities calculated with SSD/E to
736 %SSD1 without a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
737 %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}] and
738 %experiment [Ref. \onlinecite{CRC80}]. The window shows a expansion around
739 %300 K with error bars included to clarify this region of
740 %interest. Note that both SSD1 and SSD/E show good agreement with
741 %experiment when the long-range correction is neglected.}
742 %\label{ssdedense}
743 %\end{center}
744 %\end{figure}
745
746 Fig. \ref{ssdedense} shows the density profile for the SSD/E
747 model in comparison to SSD1 without a reaction field, other
748 common water models, and experimental results. The calculated
749 densities for both SSD/E and SSD1 have increased
750 significantly over the original SSD model (see
751 fig. \ref{dense1}) and are in better agreement with the experimental
752 values. At 298 K, the densities of SSD/E and SSD1 without
753 a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
754 0.999$\pm$0.001 g/cm$^3$ respectively. These both compare well with
755 the experimental value of 0.997 g/cm$^3$, and they are considerably
756 better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The
757 changes to the dipole moment and sticky switching functions have
758 improved the structuring of the liquid (as seen in figure
759 \ref{grcompare}, but they have shifted the density maximum to much
760 lower temperatures. This comes about via an increase in the liquid
761 disorder through the weakening of the sticky potential and
762 strengthening of the dipolar character. However, this increasing
763 disorder in the SSD/E model has little effect on the melting
764 transition. By monitoring $C_p$ throughout these simulations, the
765 melting transition for SSD/E was shown to occur at 235~K. The
766 same transition temperature observed with SSD and SSD1.
767
768 %\begin{figure}
769 %\begin{center}
770 %\epsfxsize=6in
771 %\epsfbox{ssdrfDense.epsi}
772 %\caption{Comparison of densities calculated with SSD/RF to
773 %SSD1 with a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
774 %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}], and
775 %experiment [Ref. \onlinecite{CRC80}]. The inset shows the necessity of
776 %reparameterization when utilizing a reaction field long-ranged
777 %correction - SSD/RF provides significantly more accurate
778 %densities than SSD1 when performing room temperature
779 %simulations.}
780 %\label{ssdrfdense}
781 %\end{center}
782 %\end{figure}
783
784 Including the reaction field long-range correction in the simulations
785 results in a more interesting comparison. A density profile including
786 SSD/RF and SSD1 with an active reaction field is shown in figure
787 \ref{ssdrfdense}. As observed in the simulations without a reaction
788 field, the densities of SSD/RF and SSD1 show a dramatic increase over
789 normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density
790 of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
791 considerably better than the original SSD value of 0.941$\pm$0.001
792 g/cm$^3$ and the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results
793 further emphasize the importance of reparameterization in order to
794 model the density properly under different simulation conditions.
795 Again, these changes have only a minor effect on the melting point,
796 which observed at 245~K for SSD/RF, is identical to SSD and only 5~K
797 lower than SSD1 with a reaction field. Additionally, the difference in
798 density maxima is not as extreme, with SSD/RF showing a density
799 maximum at 255~K, fairly close to the density maxima of 260~K and
800 265~K, shown by SSD and SSD1 respectively.
801
802 %\begin{figure}
803 %\begin{center}
804 %\epsfxsize=6in
805 %\epsfbox{ssdeDiffuse.epsi}
806 %\caption{The diffusion constants calculated from SSD/E and
807 %SSD1 (both without a reaction field) along with experimental results
808 %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The NVE calculations were
809 %performed at the average densities observed in the 1 atm NPT
810 %simulations for the respective models. SSD/E is slightly more mobile
811 %than experiment at all of the temperatures, but it is closer to
812 %experiment at biologically relevant temperatures than SSD1 without a
813 %long-range correction.}
814 %\label{ssdediffuse}
815 %\end{center}
816 %\end{figure}
817
818 The reparameterization of the SSD water model, both for use with and
819 without an applied long-range correction, brought the densities up to
820 what is expected for simulating liquid water. In addition to improving
821 the densities, it is important that the diffusive behavior of SSD be
822 maintained or improved. Figure \ref{ssdediffuse} compares the
823 temperature dependence of the diffusion constant of SSD/E to SSD1
824 without an active reaction field at the densities calculated from
825 their respective NPT simulations at 1 atm. The diffusion constant for
826 SSD/E is consistently higher than experiment, while SSD1 remains lower
827 than experiment until relatively high temperatures (around 360
828 K). Both models follow the shape of the experimental curve well below
829 300~K but tend to diffuse too rapidly at higher temperatures, as seen
830 in SSD1's crossing above 360~K. This increasing diffusion relative to
831 the experimental values is caused by the rapidly decreasing system
832 density with increasing temperature. Both SSD1 and SSD/E show this
833 deviation in particle mobility, but this trend has different
834 implications on the diffusive behavior of the models. While SSD1
835 shows more experimentally accurate diffusive behavior in the high
836 temperature regimes, SSD/E shows more accurate behavior in the
837 supercooled and biologically relevant temperature ranges. Thus, the
838 changes made to improve the liquid structure may have had an adverse
839 affect on the density maximum, but they improve the transport behavior
840 of SSD/E relative to SSD1 under the most commonly simulated
841 conditions.
842
843 %\begin{figure}
844 %\begin{center}
845 %\epsfxsize=6in
846 %\epsfbox{ssdrfDiffuse.epsi}
847 %\caption{The diffusion constants calculated from SSD/RF and
848 %SSD1 (both with an active reaction field) along with
849 %experimental results [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The
850 %NVE calculations were performed at the average densities observed in
851 %the 1 atm NPT simulations for both of the models. SSD/RF
852 %simulates the diffusion of water throughout this temperature range
853 %very well. The rapidly increasing diffusion constants at high
854 %temperatures for both models can be attributed to lower calculated
855 %densities than those observed in experiment.}
856 %\label{ssdrfdiffuse}
857 %\end{center}
858 %\end{figure}
859
860 In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
861 compared to SSD1 with an active reaction field. Note that SSD/RF
862 tracks the experimental results quantitatively, identical within error
863 throughout most of the temperature range shown and exhibiting only a
864 slight increasing trend at higher temperatures. SSD1 tends to diffuse
865 more slowly at low temperatures and deviates to diffuse too rapidly at
866 temperatures greater than 330~K. As stated above, this deviation away
867 from the ideal trend is due to a rapid decrease in density at higher
868 temperatures. SSD/RF does not suffer from this problem as much as SSD1
869 because the calculated densities are closer to the experimental
870 values. These results again emphasize the importance of careful
871 reparameterization when using an altered long-range correction.
872
873 \begin{table}
874 \begin{minipage}{\linewidth}
875 \renewcommand{\thefootnote}{\thempfootnote}
876 \begin{center}
877 \caption{Properties of the single-point water models compared with
878 experimental data at ambient conditions. Deviations of the of the
879 averages are given in parentheses.}
880 \begin{tabular}{ l c c c c c }
881 \hline \\[-3mm]
882 \ \ \ \ \ \ & \ \ \ SSD1 \ \ \ & \ \ SSD/E \ \ \ & \ \ SSD1 (RF) \ \
883 \ & \ \ SSD/RF \ \ \ & \ \ Expt. \\
884 \hline \\[-3mm]
885 \ \ $\rho$ (g/cm$^3$) & 0.999(0.001) & 0.996(0.001) & 0.972(0.002) & 0.997(0.001) & 0.997 \\
886 \ \ $C_p$ (cal/mol K) & 28.80(0.11) & 25.45(0.09) & 28.28(0.06) & 23.83(0.16) & 17.98 \\
887 \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78(0.7) & 2.51(0.18) & 2.00(0.17) & 2.32(0.06) & 2.299\cite{Mills73} \\
888 \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
889 4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in
890 Ref. \onlinecite{Head-Gordon00_1}} \\
891 \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
892 3.5\footnote{Calculated by integrating $g_{\text{OH}}(r)$ in
893 Ref. \onlinecite{Soper86}} \\
894 \ \ $\tau_1$ (ps) & 10.9(0.6) & 7.3(0.4) & 7.5(0.7) & 7.2(0.4) & 5.7\footnote{Calculated for 298 K from data in Ref. \onlinecite{Eisenberg69}} \\
895 \ \ $\tau_2$ (ps) & 4.7(0.4) & 3.1(0.2) & 3.5(0.3) & 3.2(0.2) & 2.3\footnote{Calculated for 298 K from data in
896 Ref. \onlinecite{Krynicki66}}
897 \end{tabular}
898 \label{liquidproperties}
899 \end{center}
900 \end{minipage}
901 \end{table}
902
903 Table \ref{liquidproperties} gives a synopsis of the liquid state
904 properties of the water models compared in this study along with the
905 experimental values for liquid water at ambient conditions. The
906 coordination number ($n_C$) and number of hydrogen bonds per particle
907 ($n_H$) were calculated by integrating the following relations:
908 \begin{equation}
909 n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr,
910 \end{equation}
911 \begin{equation}
912 n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr,
913 \end{equation}
914 where $\rho$ is the number density of the specified pair interactions,
915 $a$ and $b$ are the radial locations of the minima following the first
916 peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$ respectively. The number
917 of hydrogen bonds stays relatively constant across all of the models,
918 but the coordination numbers of SSD/E and SSD/RF show an
919 improvement over SSD1. This improvement is primarily due to
920 extension of the first solvation shell in the new parameter sets.
921 Because $n_H$ and $n_C$ are nearly identical in SSD1, it appears
922 that all molecules in the first solvation shell are involved in
923 hydrogen bonds. Since $n_H$ and $n_C$ differ in the newly
924 parameterized models, the orientations in the first solvation shell
925 are a bit more ``fluid''. Therefore SSD1 overstructures the
926 first solvation shell and our reparameterizations have returned this
927 shell to more realistic liquid-like behavior.
928
929 The time constants for the orientational autocorrelation functions
930 are also displayed in Table \ref{liquidproperties}. The dipolar
931 orientational time correlation functions ($C_{l}$) are described
932 by:
933 \begin{equation}
934 C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
935 \end{equation}
936 where $P_l$ are Legendre polynomials of order $l$ and
937 $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular
938 dipole.\cite{Rahman71} From these correlation functions, the
939 orientational relaxation time of the dipole vector can be calculated
940 from an exponential fit in the long-time regime ($t >
941 \tau_l$).\cite{Rothschild84} Calculation of these time constants were
942 averaged over five detailed NVE simulations performed at the ambient
943 conditions for each of the respective models. It should be noted that
944 the commonly cited value of 1.9 ps for $\tau_2$ was determined from
945 the NMR data in Ref. \onlinecite{Krynicki66} at a temperature near
946 34$^\circ$C.\cite{Rahman71} Because of the strong temperature
947 dependence of $\tau_2$, it is necessary to recalculate it at 298~K to
948 make proper comparisons. The value shown in Table
949 \ref{liquidproperties} was calculated from the same NMR data in the
950 fashion described in Ref. \onlinecite{Krynicki66}. Similarly, $\tau_1$ was
951 recomputed for 298~K from the data in Ref. \onlinecite{Eisenberg69}.
952 Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
953 and without an active reaction field. Turning on the reaction field
954 leads to much improved time constants for SSD1; however, these results
955 also include a corresponding decrease in system density.
956 Orientational relaxation times published in the original SSD dynamics
957 papers are smaller than the values observed here, and this difference
958 can be attributed to the use of the Ewald sum.\cite{Ichiye99}
959
960 \subsection{Additional Observations}
961
962 %\begin{figure}
963 %\begin{center}
964 %\epsfxsize=6in
965 %\epsfbox{icei_bw.eps}
966 %\caption{The most stable crystal structure assumed by the SSD family
967 %of water models. We refer to this structure as Ice-{\it i} to
968 %indicate its origins in computer simulation. This image was taken of
969 %the (001) face of the crystal.}
970 %\label{weirdice}
971 %\end{center}
972 %\end{figure}
973
974 While performing a series of melting simulations on an early iteration
975 of SSD/E not discussed in this paper, we observed
976 recrystallization into a novel structure not previously known for
977 water. After melting at 235~K, two of five systems underwent
978 crystallization events near 245~K. The two systems remained
979 crystalline up to 320 and 330~K, respectively. The crystal exhibits
980 an expanded zeolite-like structure that does not correspond to any
981 known form of ice. This appears to be an artifact of the point
982 dipolar models, so to distinguish it from the experimentally observed
983 forms of ice, we have denoted the structure
984 Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}). A large enough
985 portion of the sample crystallized that we have been able to obtain a
986 near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice}
987 shows the repeating crystal structure of a typical crystal at 5
988 K. Each water molecule is hydrogen bonded to four others; however, the
989 hydrogen bonds are bent rather than perfectly straight. This results
990 in a skewed tetrahedral geometry about the central molecule. In
991 figure \ref{isosurface}, it is apparent that these flexed hydrogen
992 bonds are allowed due to the conical shape of the attractive regions,
993 with the greatest attraction along the direct hydrogen bond
994 configuration. Though not ideal, these flexed hydrogen bonds are
995 favorable enough to stabilize an entire crystal generated around them.
996
997 Initial simulations indicated that Ice-{\it i} is the preferred ice
998 structure for at least the SSD/E model. To verify this, a comparison
999 was made between near ideal crystals of ice~$I_h$, ice~$I_c$, and
1000 Ice-{\it i} at constant pressure with SSD/E, SSD/RF, and
1001 SSD1. Near-ideal versions of the three types of crystals were cooled
1002 to 1 K, and enthalpies of formation of each were compared using all
1003 three water models. Enthalpies were estimated from the
1004 isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
1005 $P_{\text ext}$ is the applied pressure. A constant value of -60.158
1006 kcal / mol has been added to place our zero for the enthalpies of
1007 formation for these systems at the traditional state (elemental forms
1008 at standard temperature and pressure). With every model in the SSD
1009 family, Ice-{\it i} had the lowest calculated enthalpy of formation.
1010
1011 \begin{table}
1012 \begin{center}
1013 \caption{Enthalpies of Formation (in kcal / mol) of the three crystal
1014 structures (at 1 K) exhibited by the SSD family of water models}
1015 \begin{tabular}{ l c c c }
1016 \hline \\[-3mm]
1017 \ \ \ Water Model \ \ \ & \ \ \ Ice-$I_h$ \ \ \ & \ \ \ Ice-$I_c$ \ \ \ &
1018 \ \ \ \ Ice-{\it i} \\
1019 \hline \\[-3mm]
1020 \ \ \ SSD/E & -72.444 & -72.450 & -73.748 \\
1021 \ \ \ SSD/RF & -73.093 & -73.075 & -74.180 \\
1022 \ \ \ SSD1 & -72.654 & -72.569 & -73.575 \\
1023 \ \ \ SSD1 (RF) & -72.662 & -72.569 & -73.292 \\
1024 \end{tabular}
1025 \label{iceenthalpy}
1026 \end{center}
1027 \end{table}
1028
1029 In addition to these energetic comparisons, melting simulations were
1030 performed with Ice-{\it i} as the initial configuration using SSD/E,
1031 SSD/RF, and SSD1 both with and without a reaction field. The melting
1032 transitions for both SSD/E and SSD1 without reaction field occurred at
1033 temperature in excess of 375~K. SSD/RF and SSD1 with a reaction field
1034 showed more reasonable melting transitions near 325~K. These melting
1035 point observations clearly show that all of the SSD-derived models
1036 prefer the ice-{\it i} structure.
1037
1038 \section{Conclusions}
1039
1040 The density maximum and temperature dependence of the self-diffusion
1041 constant were studied for the SSD water model, both with and
1042 without the use of reaction field, via a series of NPT and NVE
1043 simulations. The constant pressure simulations showed a density
1044 maximum near 260 K. In most cases, the calculated densities were
1045 significantly lower than the densities obtained from other water
1046 models (and experiment). Analysis of self-diffusion showed SSD
1047 to capture the transport properties of water well in both the liquid
1048 and supercooled liquid regimes.
1049
1050 In order to correct the density behavior, the original SSD model was
1051 reparameterized for use both with and without a reaction field (SSD/RF
1052 and SSD/E), and comparisons were made with SSD1, Ichiye's density
1053 corrected version of SSD. Both models improve the liquid structure,
1054 densities, and diffusive properties under their respective simulation
1055 conditions, indicating the necessity of reparameterization when
1056 changing the method of calculating long-range electrostatic
1057 interactions. In general, however, these simple water models are
1058 excellent choices for representing explicit water in large scale
1059 simulations of biochemical systems.
1060
1061 The existence of a novel low-density ice structure that is preferred
1062 by the SSD family of water models is somewhat troubling, since
1063 liquid simulations on this family of water models at room temperature
1064 are effectively simulations of supercooled or metastable liquids. One
1065 way to destabilize this unphysical ice structure would be to make the
1066 range of angles preferred by the attractive part of the sticky
1067 potential much narrower. This would require extensive
1068 reparameterization to maintain the same level of agreement with the
1069 experiments.
1070
1071 Additionally, our initial calculations show that the Ice-{\it i}
1072 structure may also be a preferred crystal structure for at least one
1073 other popular multi-point water model (TIP3P), and that much of the
1074 simulation work being done using this popular model could also be at
1075 risk for crystallization into this unphysical structure. A future
1076 publication will detail the relative stability of the known ice
1077 structures for a wide range of popular water models.
1078
1079 \section{Acknowledgments}
1080 Support for this project was provided by the National Science
1081 Foundation under grant CHE-0134881. Computation time was provided by
1082 the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
1083 DMR-0079647.
1084
1085 \newpage
1086
1087 \bibliographystyle{jcp}
1088 \bibliography{nptSSD}
1089
1090 \newpage
1091
1092 \begin{list}
1093 {Figure \arabic{captions}: }{\usecounter{captions}
1094 \setlength{\rightmargin}{\leftmargin}}
1095
1096 \item Energy conservation using both quaternion-based integration and
1097 the {\sc dlm} method with increasing time step. The larger time step
1098 plots are shifted from the true energy baseline (that of $\Delta t$ =
1099 0.1~fs) for clarity.
1100
1101 \item Density versus temperature for TIP4P [Ref. \onlinecite{Jorgensen98b}],
1102 TIP3P [Ref. \onlinecite{Jorgensen98b}], SPC/E
1103 [Ref. \onlinecite{Clancy94}], SSD without Reaction Field, SSD, and
1104 experiment [Ref. \onlinecite{CRC80}]. The arrows indicate the change
1105 in densities observed when turning off the reaction field. The the
1106 lower than expected densities for the SSD model were what prompted the
1107 original reparameterization of SSD1 [Ref. \onlinecite{Ichiye03}].
1108
1109 \item Average self-diffusion constant as a function of temperature for
1110 SSD, SPC/E [Ref. \onlinecite{Clancy94}], and TIP5P
1111 [Ref. \onlinecite{Jorgensen01}] compared with experimental data
1112 [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. Of the three
1113 water models shown, SSD has the least deviation from the experimental
1114 values. The rapidly increasing diffusion constants for TIP5P and SSD
1115 correspond to significant decreases in density at the higher
1116 temperatures.
1117
1118 \item An illustration of angles involved in the correlations observed in
1119 Fig. \ref{contour}.
1120
1121 \item Contour plots of 2D angular pair correlation functions for
1122 512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
1123 signify regions of enhanced density while light areas signify
1124 depletion relative to the bulk density. White areas have pair
1125 correlation values below 0.5 and black areas have values above 1.5.
1126
1127 \item Plots comparing experiment [Ref. \onlinecite{Head-Gordon00_1}] with
1128 SSD/E and SSD1 without reaction field (top), as well as SSD/RF and
1129 SSD1 with reaction field turned on (bottom). The insets show the
1130 respective first peaks in detail. Note how the changes in parameters
1131 have lowered and broadened the first peak of SSD/E and SSD/RF.
1132
1133 \item Positive and negative isosurfaces of the sticky potential for
1134 SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
1135 correspond to the tetrahedral attractive component, and darker areas
1136 correspond to the dipolar repulsive component.
1137
1138 \item Comparison of densities calculated with SSD/E to
1139 SSD1 without a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1140 TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}] and
1141 experiment [Ref. \onlinecite{CRC80}]. The window shows a expansion around
1142 300 K with error bars included to clarify this region of
1143 interest. Note that both SSD1 and SSD/E show good agreement with
1144 experiment when the long-range correction is neglected.
1145
1146 \item Comparison of densities calculated with SSD/RF to
1147 SSD1 with a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1148 TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}], and
1149 experiment [Ref. \onlinecite{CRC80}]. The inset shows the necessity of
1150 reparameterization when utilizing a reaction field long-ranged
1151 correction - SSD/RF provides significantly more accurate
1152 densities than SSD1 when performing room temperature
1153 simulations.
1154
1155 \item The diffusion constants calculated from SSD/E and
1156 SSD1 (both without a reaction field) along with experimental results
1157 [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The NVE calculations were
1158 performed at the average densities observed in the 1 atm NPT
1159 simulations for the respective models. SSD/E is slightly more mobile
1160 than experiment at all of the temperatures, but it is closer to
1161 experiment at biologically relevant temperatures than SSD1 without a
1162 long-range correction.
1163
1164 \item The diffusion constants calculated from SSD/RF and
1165 SSD1 (both with an active reaction field) along with
1166 experimental results [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The
1167 NVE calculations were performed at the average densities observed in
1168 the 1 atm NPT simulations for both of the models. SSD/RF
1169 simulates the diffusion of water throughout this temperature range
1170 very well. The rapidly increasing diffusion constants at high
1171 temperatures for both models can be attributed to lower calculated
1172 densities than those observed in experiment.
1173
1174 \item The most stable crystal structure assumed by the SSD family
1175 of water models. We refer to this structure as Ice-{\it i} to
1176 indicate its origins in computer simulation. This image was taken of
1177 the (001) face of the crystal.
1178 \end{list}
1179
1180 \newpage
1181
1182 \begin{figure}
1183 \begin{center}
1184 \epsfxsize=6in
1185 \epsfbox{timeStep.epsi}
1186 %\caption{Energy conservation using both quaternion-based integration and
1187 %the {\sc dlm} method with increasing time step. The larger time step
1188 %plots are shifted from the true energy baseline (that of $\Delta t$ =
1189 %0.1~fs) for clarity.}
1190 \label{timestep}
1191 \end{center}
1192 \end{figure}
1193
1194 \newpage
1195
1196 \begin{figure}
1197 \begin{center}
1198 \epsfxsize=6in
1199 \epsfbox{denseSSDnew.eps}
1200 %\caption{Density versus temperature for TIP4P [Ref. \onlinecite{Jorgensen98b}],
1201 % TIP3P [Ref. \onlinecite{Jorgensen98b}], SPC/E [Ref. \onlinecite{Clancy94}], SSD
1202 % without Reaction Field, SSD, and experiment [Ref. \onlinecite{CRC80}]. The
1203 % arrows indicate the change in densities observed when turning off the
1204 % reaction field. The the lower than expected densities for the SSD
1205 % model were what prompted the original reparameterization of SSD1
1206 % [Ref. \onlinecite{Ichiye03}].}
1207 \label{dense1}
1208 \end{center}
1209 \end{figure}
1210
1211 \newpage
1212
1213 \begin{figure}
1214 \begin{center}
1215 \epsfxsize=6in
1216 \epsfbox{betterDiffuse.epsi}
1217 %\caption{Average self-diffusion constant as a function of temperature for
1218 %SSD, SPC/E [Ref. \onlinecite{Clancy94}], and TIP5P
1219 %[Ref. \onlinecite{Jorgensen01}] compared with experimental data
1220 %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. Of the three water models
1221 %shown, SSD has the least deviation from the experimental values. The
1222 %rapidly increasing diffusion constants for TIP5P and SSD correspond to
1223 %significant decreases in density at the higher temperatures.}
1224 \label{diffuse}
1225 \end{center}
1226 \end{figure}
1227
1228 \newpage
1229
1230 \begin{figure}
1231 \begin{center}
1232 \epsfxsize=6in
1233 \epsfbox{corrDiag.eps}
1234 %\caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
1235 \label{corrAngle}
1236 \end{center}
1237 \end{figure}
1238
1239 \newpage
1240
1241 \begin{figure}
1242 \begin{center}
1243 \epsfxsize=6in
1244 \epsfbox{fullContours.eps}
1245 %\caption{Contour plots of 2D angular pair correlation functions for
1246 %512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
1247 %signify regions of enhanced density while light areas signify
1248 %depletion relative to the bulk density. White areas have pair
1249 %correlation values below 0.5 and black areas have values above 1.5.}
1250 \label{contour}
1251 \end{center}
1252 \end{figure}
1253
1254 \newpage
1255
1256 \begin{figure}
1257 \begin{center}
1258 \epsfxsize=6in
1259 \epsfbox{GofRCompare.epsi}
1260 %\caption{Plots comparing experiment [Ref. \onlinecite{Head-Gordon00_1}] with
1261 %SSD/E and SSD1 without reaction field (top), as well as
1262 %SSD/RF and SSD1 with reaction field turned on
1263 %(bottom). The insets show the respective first peaks in detail. Note
1264 %how the changes in parameters have lowered and broadened the first
1265 %peak of SSD/E and SSD/RF.}
1266 \label{grcompare}
1267 \end{center}
1268 \end{figure}
1269
1270 \newpage
1271
1272 \begin{figure}
1273 \begin{center}
1274 \epsfxsize=7in
1275 \epsfbox{dualsticky_bw.eps}
1276 %\caption{Positive and negative isosurfaces of the sticky potential for
1277 %SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
1278 %correspond to the tetrahedral attractive component, and darker areas
1279 %correspond to the dipolar repulsive component.}
1280 \label{isosurface}
1281 \end{center}
1282 \end{figure}
1283
1284 \newpage
1285
1286 \begin{figure}
1287 \begin{center}
1288 \epsfxsize=6in
1289 \epsfbox{ssdeDense.epsi}
1290 %\caption{Comparison of densities calculated with SSD/E to
1291 %SSD1 without a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1292 %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}] and
1293 %experiment [Ref. \onlinecite{CRC80}]. The window shows a expansion around
1294 %300 K with error bars included to clarify this region of
1295 %interest. Note that both SSD1 and SSD/E show good agreement with
1296 %experiment when the long-range correction is neglected.}
1297 \label{ssdedense}
1298 \end{center}
1299 \end{figure}
1300
1301 \newpage
1302
1303 \begin{figure}
1304 \begin{center}
1305 \epsfxsize=6in
1306 \epsfbox{ssdrfDense.epsi}
1307 %\caption{Comparison of densities calculated with SSD/RF to
1308 %SSD1 with a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1309 %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}], and
1310 %experiment [Ref. \onlinecite{CRC80}]. The inset shows the necessity of
1311 %reparameterization when utilizing a reaction field long-ranged
1312 %correction - SSD/RF provides significantly more accurate
1313 %densities than SSD1 when performing room temperature
1314 %simulations.}
1315 \label{ssdrfdense}
1316 \end{center}
1317 \end{figure}
1318
1319 \newpage
1320
1321 \begin{figure}
1322 \begin{center}
1323 \epsfxsize=6in
1324 \epsfbox{ssdeDiffuse.epsi}
1325 %\caption{The diffusion constants calculated from SSD/E and
1326 %SSD1 (both without a reaction field) along with experimental results
1327 %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The NVE calculations were
1328 %performed at the average densities observed in the 1 atm NPT
1329 %simulations for the respective models. SSD/E is slightly more mobile
1330 %than experiment at all of the temperatures, but it is closer to
1331 %experiment at biologically relevant temperatures than SSD1 without a
1332 %long-range correction.}
1333 \label{ssdediffuse}
1334 \end{center}
1335 \end{figure}
1336
1337 \newpage
1338
1339 \begin{figure}
1340 \begin{center}
1341 \epsfxsize=6in
1342 \epsfbox{ssdrfDiffuse.epsi}
1343 %\caption{The diffusion constants calculated from SSD/RF and
1344 %SSD1 (both with an active reaction field) along with
1345 %experimental results [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The
1346 %NVE calculations were performed at the average densities observed in
1347 %the 1 atm NPT simulations for both of the models. SSD/RF
1348 %simulates the diffusion of water throughout this temperature range
1349 %very well. The rapidly increasing diffusion constants at high
1350 %temperatures for both models can be attributed to lower calculated
1351 %densities than those observed in experiment.}
1352 \label{ssdrfdiffuse}
1353 \end{center}
1354 \end{figure}
1355
1356 \newpage
1357
1358 \begin{figure}
1359 \begin{center}
1360 \epsfxsize=6in
1361 \epsfbox{icei_bw.eps}
1362 %\caption{The most stable crystal structure assumed by the SSD family
1363 %of water models. We refer to this structure as Ice-{\it i} to
1364 %indicate its origins in computer simulation. This image was taken of
1365 %the (001) face of the crystal.}
1366 \label{weirdice}
1367 \end{center}
1368 \end{figure}
1369
1370 \end{document}