ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Water.tex
Revision: 3023
Committed: Tue Sep 26 03:07:59 2006 UTC (18 years, 7 months ago) by chrisfen
Content type: application/x-tex
File size: 61005 byte(s)
Log Message:
Refinements.  Just a little proof checking left...

File Contents

# Content
1 \chapter{\label{chap:water}SIMPLE MODELS FOR WATER}
2
3 One of the most important tasks in the simulation of biochemical
4 systems is the proper depiction of the aqueous environment around the
5 molecules of interest. In some cases (such as in the simulation of
6 phospholipid bilayers), the majority of the calculations that are
7 performed involve interactions with or between solvent molecules.
8 Thus, the motion and behavior of molecules in biochemical simulations
9 are highly dependent on the properties of the water model that is
10 chosen as the solvent.
11 \begin{figure}
12 \includegraphics[width=\linewidth]{./figures/waterModels.pdf}
13 \caption{Partial-charge geometries for the TIP3P, TIP4P, TIP5P, and
14 SPC/E rigid body water models.\cite{Jorgensen83,Mahoney00,Berendsen87}
15 In the case of the TIP models, the depiction of water improves with
16 increasing number of point charges; however, computational performance
17 simultaneously degrades due to the increasing number of distances and
18 interactions that need to be calculated.}
19 \label{fig:waterModels}
20 \end{figure}
21
22 As discussed in the previous chapter, water is typically modeled with
23 fixed geometries of point charges shielded by the repulsive part of a
24 Lennard-Jones interaction. Some of the common water models are shown
25 in figure \ref{fig:waterModels}. The various models all have their
26 benefits and drawbacks, and these primarily focus on the balance
27 between computational efficiency and the ability to accurately predict
28 the properties of bulk water. For example, the TIP5P model improves on
29 the structural and transport properties of water relative to the TIP3P
30 and TIP4P models, yet this comes at a greater than 50\% increase in
31 computational cost.\cite{Mahoney00,Mahoney01} This cost is entirely
32 due to the additional distance and electrostatic calculations that
33 come from the extra point charges in the model description. Thus, the
34 main criteria for choosing a water model are:
35
36 \begin{enumerate}
37 \item capturing the liquid state properties and
38 \item having the fewest number of points to insure efficient performance.
39 \end{enumerate}
40 As researchers have begun to study larger systems, such as entire
41 viruses, the choice has shifted towards efficiency over accuracy in
42 order to make the calculations feasible.\cite{Freddolino06}
43
44 \section{Soft Sticky Dipole Model for Water}
45
46 One recently-developed model that largely succeeds in retaining the
47 accuracy of bulk properties while greatly reducing the computational
48 cost is the Soft Sticky Dipole (SSD) water
49 model.\cite{Liu96,Liu96b,Chandra99,Tan03} The SSD model was developed
50 as a modified form of the hard-sphere water model proposed by Bratko,
51 Blum, and Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point}
52 model which has an interaction site that is both a point dipole and a
53 Lennard-Jones core. However, since the normal aligned and
54 anti-aligned geometries favored by point dipoles are poor mimics of
55 local structure in liquid water, a short ranged ``sticky'' potential
56 is also added. The sticky potential directs the molecules to assume
57 the proper hydrogen bond orientation in the first solvation shell.
58
59 The interaction between two SSD water molecules \emph{i} and \emph{j}
60 is given by the potential
61 \begin{equation}
62 u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
63 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ +
64 u_{ij}^{sp}
65 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j),
66 \end{equation}
67 where the ${\bf r}_{ij}$ is the position vector between molecules
68 \emph{i} and \emph{j} with magnitude $r_{ij}$, and
69 ${\bf \Omega}_i$ and ${\bf \Omega}_j$ represent the orientations of
70 the two molecules. The Lennard-Jones and dipole interactions are given
71 by the following familiar forms:
72 \begin{equation}
73 u_{ij}^{LJ}(r_{ij}) = 4\epsilon
74 \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right]
75 \ ,
76 \end{equation}
77 and
78 \begin{equation}
79 u_{ij}^{dp} = \frac{|\mu_i||\mu_j|}{4 \pi \epsilon_0 r_{ij}^3} \left(
80 \hat{\bf u}_i \cdot \hat{\bf u}_j - 3(\hat{\bf u}_i\cdot\hat{\bf
81 r}_{ij})(\hat{\bf u}_j\cdot\hat{\bf r}_{ij}) \right)\ ,
82 \end{equation}
83 where $\hat{\bf u}_i$ and $\hat{\bf u}_j$ are the unit vectors along
84 the dipoles of molecules $i$ and $j$ respectively. $|\mu_i|$ and
85 $|\mu_j|$ are the strengths of the dipole moments, and $\hat{\bf
86 r}_{ij}$ is the unit vector pointing from molecule $j$ to molecule
87 $i$.
88
89 The sticky potential is somewhat less familiar:
90 \begin{equation}
91 u_{ij}^{sp}
92 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
93 \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)
94 + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf
95 \Omega}_j)]\ .
96 \label{eq:stickyfunction}
97 \end{equation}
98 Here, $\nu_0$ is a strength parameter for the sticky potential, and
99 $s$ and $s^\prime$ are cubic switching functions which turn off the
100 sticky interaction beyond the first solvation shell. The $w$ function
101 can be thought of as an attractive potential with tetrahedral
102 geometry:
103 \begin{equation}
104 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
105 \end{equation}
106 while the $w^\prime$ function counters the normal aligned and
107 anti-aligned structures favored by point dipoles:
108 \begin{equation}
109 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,
110 \end{equation}
111 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
112 and $Y_3^{-2}$ spherical harmonics (a linear combination which
113 enhances the tetrahedral geometry for hydrogen bonded structures),
114 while $w^\prime$ is a purely empirical function. A more detailed
115 description of the functional parts and variables in this potential
116 can be found in the original SSD
117 articles.\cite{Liu96,Liu96b,Chandra99,Tan03}
118
119 Since SSD is a single-point {\it dipolar} model, the force
120 calculations are simplified significantly relative to the standard
121 {\it charged} multi-point models. In the original Monte Carlo
122 simulations using this model, Liu and Ichiye reported that using SSD
123 decreased computer time by a factor of 6-7 compared to other
124 models.\cite{Liu96b} What is most impressive is that this savings did
125 not come at the expense of accurate depiction of the liquid state
126 properties. Indeed, SSD maintains reasonable agreement with the Soper
127 data for the structural features of liquid water.\cite{Soper86,Liu96b}
128 Additionally, the dynamical properties exhibited by SSD agree with
129 experiment better than those of more computationally expensive models
130 (like TIP3P and SPC/E).\cite{Chandra99} The combination of speed and
131 accurate depiction of solvent properties makes SSD a very attractive
132 model for the simulation of large scale biochemical simulations.
133
134 It is important to note that the SSD model was originally developed
135 for use with the Ewald summation for handling long-range
136 electrostatics.\cite{Ewald21} In applying this water model in a
137 variety of molecular systems, it would be useful to know its
138 properties and behavior under the more computationally efficient
139 reaction field (RF) technique, the correction techniques discussed in
140 the previous chapter, or even a simple
141 cutoff.\cite{Onsager36,Fennell06} This chapter addresses these issues
142 by looking at the structural and transport behavior of SSD over a
143 variety of temperatures with the purpose of utilizing the RF
144 correction technique. We then suggest modifications to the parameters
145 that result in more realistic bulk phase behavior. It should be noted
146 that in a recent publication, some of the original investigators of
147 the SSD water model have suggested adjustments to the SSD water model
148 to address abnormal density behavior (also observed here), calling the
149 corrected model SSD1.\cite{Tan03} In the later sections of this
150 chapter, we compare our modified variants of SSD with both the
151 original SSD and SSD1 models and discuss how our changes improve the
152 depiction of water.
153
154 \section{Simulation Methods}{\label{sec:waterSimMethods}
155
156 Most of the calculations in this particular study were performed using
157 an internally developed simulation code that was one of the precursors
158 of the {\sc oopse} molecular dynamics (MD) package.\cite{Meineke05}
159 All of the capabilities of this code have been incorporated into {\sc
160 oopse}, and calculation results are consistent between the two
161 simulation packages. The later calculations involving the damped
162 shifted force ({\sc sf}) techniques were performed using {\sc oopse}.
163
164 In the primary simulations of this study, long-range dipole-dipole
165 corrections were accounted for by using either the reaction field
166 technique or a simple cubic switching function at the cutoff
167 radius. Interestingly, one of the early applications of a reaction
168 field was in Monte Carlo simulations of liquid water.\cite{Barker73}
169 In this method, the magnitude of the reaction field acting on dipole
170 $i$ is
171 \begin{equation}
172 \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
173 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
174 \label{eq:rfequation}
175 \end{equation}
176 where $\mathcal{R}$ is the cavity defined by the cutoff radius
177 ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
178 system, ${\bf\mu}_{j}$ is the dipole moment vector of particle $j$,
179 and $s(r_{ij})$ is a cubic switching function.\cite{Allen87} The
180 reaction field contribution to the total energy from particle $i$ is
181 given by $-\frac{1}{2}{\bf\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
182 on dipole $i$ by ${\bf\mu}_{i}\times\mathcal{E}_{i}$.\cite{Allen87} An
183 applied reaction field will alter the bulk orientational properties of
184 simulated water, and there is particular sensitivity of these
185 properties on changes in the length of the cutoff
186 radius.\cite{vanderSpoel98} This behavior makes the reaction field a
187 less attractive method than the Ewald sum; however, for very large
188 systems, the computational benefit of reaction field is significant.
189
190 In contrast to the simulations with a reaction field, we have also
191 performed a companion set of simulations {\it without} a surrounding
192 dielectric (i.e. using a simple cubic switching function at the cutoff
193 radius). As a result, we have developed two reparametrizations of SSD
194 which can be used either with or without an active reaction field.
195
196 To determine the preferred densities of the models, we performed
197 simulations in the isobaric-isothermal ({\it NPT}) ensemble. All
198 dynamical properties for these models were then obtained from
199 microcanonical ({\it NVE}) simulations done at densities matching the
200 {\it NPT} density for a particular target temperature. The constant
201 pressure simulations were implemented using an integral thermostat and
202 barostat as outlined by Hoover.\cite{Hoover85,Hoover86} All molecules
203 were treated as non-linear rigid bodies. Vibrational constraints are
204 not necessary in simulations of SSD, because there are no explicit
205 hydrogen atoms, and thus no molecular vibrational modes need to be
206 considered.
207
208 The symplectic splitting method proposed by Dullweber, Leimkuhler, and
209 McLachlan ({\sc dlm}, see section \ref{sec:IntroIntegrate}) was used
210 to carry out the integration of the equations of motion in place of
211 the more prevalent quaternion
212 method.\cite{Dullweber97,Evans77,Evans77b,Allen87} The reason behind
213 this decision was that, in {\it NVE} simulations, the energy drift
214 when using quaternions was substantially greater than when using the
215 {\sc dlm} method (Fig. \ref{fig:timeStepIntegration}). This steady
216 drift in the total energy has also been observed in other
217 studies.\cite{Kol97}
218
219 \begin{figure}
220 \centering
221 \includegraphics[width=\linewidth]{./figures/timeStepIntegration.pdf}
222 \caption{Energy conservation using both quaternion-based integration
223 and the {\sc dlm} method with increasing time step. The larger time
224 step plots are shifted from the true energy baseline (that of $\Delta
225 t$ = 0.1~fs) for clarity.}
226 \label{fig:timeStepIntegration}
227 \end{figure}
228
229 The {\sc dlm} method allows for Verlet style integration orientational
230 motion of rigid bodies via a sequence of rotation matrix
231 operations. Because these matrix operations are more costly than the
232 simpler arithmetic operations for quaternion propagation, typical SSD
233 particle simulations using {\sc dlm} are 5-10\% slower than
234 simulations using the quaternion method and an identical time
235 step. This additional expense is justified because of the ability to
236 use time steps that are more that twice as long and still achieve the
237 same energy conservation.
238
239 Figure \ref{fig:timeStepIntegration} shows the resulting energy drift
240 at various time steps for both {\sc dlm} and quaternion
241 integration. All of the 1000 SSD particle simulations started with the
242 same configuration, and the only difference was the method used to
243 handle orientational motion. At time steps of 0.1 and 0.5~fs, both
244 methods for propagating the orientational degrees of freedom conserve
245 energy fairly well, with the quaternion method showing a slight energy
246 drift over time in the 0.5~fs time step simulation. Time steps of 1 and
247 2~fs clearly demonstrate the benefits in energy conservation that come
248 with the {\sc dlm} method. Thus, while maintaining the same degree of
249 energy conservation, one can take considerably longer time steps,
250 leading to an overall reduction in computation time.
251
252 Energy drifts in water simulations using {\sc dlm} integration were
253 unnoticeable for time steps up to 3~fs. We observed a slight energy
254 drift on the order of 0.012~kcal/mol per nanosecond with a time step
255 of 4~fs. As expected, this drift increases dramatically with increasing
256 time step. To insure accuracy in our {\it NVE} simulations, time steps
257 were set at 2~fs and were also kept at this value for {\it NPT}
258 simulations.
259
260 Proton-disordered ice crystals in both the I$_\textrm{h}$ and
261 I$_\textrm{c}$ lattices were generated as starting points for all
262 simulations. The I$_\textrm{h}$ crystals were formed by first
263 arranging the centers of mass of the SSD particles into a
264 ``hexagonal'' ice lattice of 1024 particles. Because of the crystal
265 structure of I$_\textrm{h}$ ice, the simulation boxes were
266 orthorhombic in shape with an edge length ratio of approximately
267 1.00$\times$1.06$\times$1.23. We then allowed the particles to orient
268 freely about their fixed lattice positions with angular momenta values
269 randomly sampled at 400~K. The rotational temperature was then scaled
270 down in stages to slowly cool the crystals to 25~K. The particles were
271 then allowed to translate with fixed orientations at a constant
272 pressure of 1~atm for 50~ps at 25~K. Finally, all constraints were
273 removed and the ice crystals were allowed to equilibrate for 50~ps at
274 25~K and a constant pressure of 1~atm. This procedure resulted in
275 structurally stable I$_\textrm{h}$ ice crystals that obey the
276 Bernal-Fowler rules.\cite{Bernal33,Rahman72} This method was also
277 utilized in the making of diamond lattice I$_\textrm{c}$ ice crystals,
278 with each cubic simulation box consisting of either 512 or 1000
279 particles. Only isotropic volume fluctuations were performed under
280 constant pressure, so the ratio of edge lengths remained constant
281 throughout the simulations.
282
283 \section{SSD Density Behavior}
284
285 Melting studies were performed on the randomized ice crystals using
286 the {\it NPT} ensemble. During melting simulations, the melting
287 transition and the density maximum can both be observed, provided that
288 the density maximum occurs in the liquid and not the supercooled
289 regime. It should be noted that the calculated melting temperature
290 ($T_\textrm{m}$) will not be the true $T_\textrm{m}$ because of
291 super-heating due to the relatively short time scales in molecular
292 simulations. This behavior results in inflated $T_\textrm{m}$ values;
293 however, these values provide a reasonable initial estimate of
294 $T_\textrm{m}$.
295
296 An ensemble average from five separate melting simulations was
297 acquired, each starting from different ice crystals generated as
298 described previously. All simulations were equilibrated for 100~ps
299 prior to a 200~ps data collection run at each temperature setting. The
300 temperature range of study spanned from 25 to 400~K, with a maximum
301 degree increment of 25~K. For regions of interest along this stepwise
302 progression, the temperature increment was decreased from 25~K to 10
303 and 5~K. The above equilibration and production times were sufficient
304 in that the fluctuations in the volume autocorrelation function damped
305 out in all of the simulations in under 20~ps.
306
307 Our initial simulations focused on the original SSD water model, and
308 an average density versus temperature plot is shown in figure
309 \ref{fig:ssdDense}. Note that the density maximum when using a
310 reaction field appears between 255 and 265~K. There were smaller
311 fluctuations in the density at 260~K than at either 255 or 265~K, so we
312 report this value as the location of the density maximum. Figure
313 \ref{fig:ssdDense} was constructed using ice I$_\textrm{h}$ crystals
314 for the initial configuration; though not pictured, the simulations
315 starting from ice I$_\textrm{c}$ crystal configurations showed similar
316 results, with a liquid-phase density maximum at the same temperature.
317
318 \begin{figure}
319 \centering
320 \includegraphics[width=\linewidth]{./figures/ssdDense.pdf}
321 \caption{ Density versus temperature for TIP3P, SPC/E, TIP4P, SSD,
322 SSD with a reaction field, and
323 experiment.\cite{Jorgensen98b,Baez94,CRC80}. Note that using a
324 reaction field lowers the density more than the already lowered SSD
325 densities. The lower than expected densities for the SSD model
326 prompted the original reparametrization of SSD to SSD1.\cite{Tan03}}
327 \label{fig:ssdDense}
328 \end{figure}
329
330 The density maximum for SSD compares quite favorably to other simple
331 water models. Figure \ref{fig:ssdDense} also shows calculated
332 densities of several other models and experiment obtained from other
333 sources.\cite{Jorgensen98b,Baez94,CRC80} Of the water models shown,
334 SSD has a temperature closest to the experimentally observed density
335 maximum. Of the {\it charge-based} models in figure
336 \ref{fig:ssdDense}, TIP4P has a density maximum behavior most like
337 that seen in SSD. Though not included in this plot, it is useful to
338 note that TIP5P has a density maximum nearly identical to the
339 experimentally measured temperature (see section
340 \ref{sec:t5peDensity}.
341
342 Liquid state densities in water have been observed to be dependent on
343 the cutoff radius ($R_\textrm{c}$), both with and without the use of a
344 reaction field.\cite{vanderSpoel98} In order to address the possible
345 effect of $R_\textrm{c}$, simulations were performed with a cutoff
346 radius of 12~\AA\, complementing the 9~\AA\ $R_\textrm{c}$ used in the
347 previous SSD simulations. All of the resulting densities overlapped
348 within error and showed no significant trend toward lower or higher
349 densities in simulations both with and without reaction field.
350
351 The key feature to recognize in figure \ref{fig:ssdDense} is the
352 density scaling of SSD relative to other common models at any given
353 temperature. SSD assumes a lower density than any of the other listed
354 models at the same pressure. This behavior is especially apparent at
355 temperatures greater than 300~K. Lower than expected densities have
356 been observed for other systems using a reaction field for long-range
357 electrostatic interactions, so the most likely reason for the reduced
358 densities is the presence of the reaction
359 field.\cite{vanderSpoel98,Nezbeda02} In order to test the effect of
360 the reaction field on the density of the systems, the simulations were
361 repeated without a reaction field present. The results of these
362 simulations are also displayed in figure \ref{fig:ssdDense}. Without
363 the reaction field, the densities increase to more experimentally
364 reasonable values, especially around the freezing point of liquid
365 water. The shape of the curve is similar to the curve produced from
366 SSD simulations using reaction field, specifically the rapidly
367 decreasing densities at higher temperatures; however, a shift in the
368 density maximum location, down to 245~K, is observed. This is a more
369 accurate comparison to the other listed water models, in that no long
370 range corrections were applied in those
371 simulations.\cite{Baez94,Jorgensen98b} However, even without the
372 reaction field, the density around 300~K is still significantly lower
373 than experiment and comparable water models. This anomalous behavior
374 was what lead Tan {\it et al.} to recently reparametrize
375 SSD.\cite{Tan03} Throughout the remainder of the paper our
376 reparametrizations of SSD will be compared with their newer SSD1
377 model.
378
379 \section{SSD Transport Behavior}
380
381 Accurate dynamical properties of a water model are particularly
382 important when using the model to study permeation or transport across
383 biological membranes. In order to probe transport in bulk water, {\it
384 NVE} simulations were performed at the average densities obtained from
385 the {\it NPT} simulations at an identical target
386 temperature. Simulations started with randomized velocities and
387 underwent 50~ps of temperature scaling and 50~ps of constant energy
388 equilibration before a 200~ps data collection run. Diffusion constants
389 were calculated via linear fits to the long-time behavior of the
390 mean-square displacement as a function of time.\cite{Allen87} The
391 averaged results from five sets of {\it NVE} simulations are displayed
392 in figure \ref{fig:ssdDiffuse}, alongside experimental, SPC/E, and TIP5P
393 results.\cite{Gillen72,Holz00,Baez94,Mahoney01}
394
395 \begin{figure}
396 \centering
397 \includegraphics[width=\linewidth]{./figures/ssdDiffuse.pdf}
398 \caption{ Average self-diffusion constant as a function of temperature for
399 SSD, SPC/E, and TIP5P compared with experimental
400 data.\cite{Baez94,Mahoney01,Gillen72,Holz00} Of the three water
401 models shown, SSD has the least deviation from the experimental
402 values. The rapidly increasing diffusion constants for TIP5P and SSD
403 correspond to significant decreases in density at the higher
404 temperatures.}
405 \label{fig:ssdDiffuse}
406 \end{figure}
407
408 The observed values for the diffusion constant point out one of the
409 strengths of the SSD model. Of the three models shown, the SSD model
410 has the most accurate depiction of self-diffusion in both the
411 supercooled and liquid regimes. SPC/E does a respectable job by
412 reproducing values similar to experiment around 290~K; however, it
413 deviates at both higher and lower temperatures, failing to predict the
414 correct thermal trend. TIP5P and SSD both start off low at colder
415 temperatures and tend to diffuse too rapidly at higher temperatures.
416 This behavior at higher temperatures is not particularly surprising
417 since the densities of both TIP5P and SSD are lower than experimental
418 water densities at higher temperatures. When calculating the
419 diffusion coefficients for SSD at experimental densities (instead of
420 the densities from the {\it NPT} simulations), the resulting values
421 fall more in line with experiment at these temperatures.
422
423 \section{Structural Changes and Characterization}
424
425 By starting the simulations from the crystalline state, we can get an
426 estimation of the $T_\textrm{m}$ of the ice structure, and beyond the
427 melting point, we study the phase behavior of the liquid. The constant
428 pressure heat capacity ($C_\textrm{p}$) was monitored to locate
429 $T_\textrm{m}$ in each of the simulations. In the melting simulations
430 of the 1024 particle ice I$_\textrm{h}$ simulations, a large spike in
431 $C_\textrm{p}$ occurs at 245~K, indicating a first order phase
432 transition for the melting of these ice crystals (see figure
433 \ref{fig:ssdCp}. When the reaction field is turned off, the melting
434 transition occurs at 235~K. These melting transitions are considerably
435 lower than the experimental value of 273~K, indicating that the solid
436 ice I$_\textrm{h}$ is not thermodynamically preferred relative to the
437 liquid state at these lower temperatures.
438 \begin{figure}
439 \centering
440 \includegraphics[width=\linewidth]{./figures/ssdCp.pdf}
441 \caption{Heat capacity versus temperature for the SSD model with an
442 active reaction field. Note the large spike in $C_p$ around 245~K,
443 indicating a phase transition from the ordered crystal to disordered
444 liquid.}
445 \label{fig:ssdCp}
446 \end{figure}
447
448 \begin{figure}
449 \centering
450 \includegraphics[width=\linewidth]{./figures/fullContour.pdf}
451 \caption{ Contour plots of 2D angular pair correlation functions for
452 512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
453 signify regions of enhanced density while light areas signify
454 depletion relative to the bulk density. White areas have pair
455 correlation values below 0.5 and black areas have values above 1.5.}
456 \label{fig:contour}
457 \end{figure}
458
459 \begin{figure}
460 \centering
461 \includegraphics[width=2.5in]{./figures/corrDiag.pdf}
462 \caption{ An illustration of angles involved in the correlations
463 displayed in figure \ref{fig:contour}.}
464 \label{fig:corrAngle}
465 \end{figure}
466
467 Additional analysis of the melting process was performed using
468 two-dimensional structure and dipole angle correlations. Expressions
469 for these correlations are as follows:
470
471 \begin{equation}
472 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\ ,
473 \end{equation}
474 \begin{equation}
475 g_{\text{AB}}(r,\cos\omega) =
476 \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\ ,
477 \end{equation}
478 where $\theta$ and $\omega$ refer to the angles shown in figure
479 \ref{fig:corrAngle}. By binning over both distance and the cosine of the
480 desired angle between the two dipoles, the $g(r)$ can be analyzed to
481 determine the common dipole arrangements that constitute the peaks and
482 troughs in the standard one-dimensional $g(r)$ plots. Frames A and B
483 of figure \ref{fig:contour} show results from an ice I$_\textrm{c}$
484 simulation. The first peak in the $g(r)$ consists primarily of the
485 preferred hydrogen bonding arrangements as dictated by the tetrahedral
486 sticky potential - one peak for the hydrogen bond donor and the other
487 for the hydrogen bond acceptor. Due to the high degree of
488 crystallinity of the sample, the second and third solvation shells
489 show a repeated peak arrangement which decays at distances around the
490 fourth solvation shell, near the imposed cutoff for the Lennard-Jones
491 and dipole-dipole interactions. In the higher temperature simulation
492 shown in frames C and D, these long-range features deteriorate
493 rapidly. The first solvation shell still shows the strong effect of
494 the sticky-potential, although it covers a larger area, extending to
495 include a fraction of aligned dipole peaks within the first solvation
496 shell. The latter peaks lose due to thermal motion and as the
497 competing dipole force overcomes the sticky potential's tight
498 tetrahedral structuring of the crystal.
499
500 This complex interplay between dipole and sticky interactions was
501 remarked upon as a possible reason for the split second peak in the
502 oxygen-oxygen pair correlation function,
503 $g_\textrm{OO}(r)$.\cite{Liu96b} At low temperatures, the second
504 solvation shell peak appears to have two distinct components that
505 blend together to form one observable peak. At higher temperatures,
506 this split character alters to show the leading 4~\AA\ peak dominated
507 by equatorial anti-parallel dipole orientations. There is also a
508 tightly bunched group of axially arranged dipoles that most likely
509 consist of the smaller fraction of aligned dipole pairs. The trailing
510 component of the split peak at 5~\AA\ is dominated by aligned dipoles
511 that assume hydrogen bond arrangements similar to those seen in the
512 first solvation shell. This evidence indicates that the dipole pair
513 interaction begins to dominate outside of the range of the dipolar
514 repulsion term. The energetically favorable dipole arrangements
515 populate the region immediately outside this repulsion region (around
516 4~\AA), while arrangements that seek to satisfy both the sticky and
517 dipole forces locate themselves just beyond this initial buildup
518 (around 5~\AA).
519
520 This analysis indicates that the split second peak is primarily the
521 product of the dipolar repulsion term of the sticky potential. In
522 fact, the inner peak can be pushed out and merged with the outer split
523 peak just by extending the switching function ($s^\prime(r_{ij})$)
524 from its normal 4~\AA\ cutoff to values of 4.5 or even 5~\AA. This
525 type of correction is not recommended for improving the liquid
526 structure, since the second solvation shell would still be shifted too
527 far out. In addition, this would have an even more detrimental effect
528 on the system densities, leading to a liquid with a more open
529 structure and a density considerably lower than the already low SSD
530 density. A better correction would be to include the
531 quadrupole-quadrupole interactions for the water particles outside of
532 the first solvation shell, but this would remove the simplicity and
533 speed advantage of SSD.
534
535 \section{Adjusted Potentials: SSD/RF and SSD/E}
536
537 The propensity of SSD to adopt lower than expected densities under
538 varying conditions is troubling, especially at higher temperatures. In
539 order to correct this model for use with a reaction field, it is
540 necessary to adjust the force field parameters for the primary
541 intermolecular interactions. In undergoing a reparametrization, it is
542 important not to focus on just one property and neglect the others. In
543 this case, it would be ideal to correct the densities while
544 maintaining the accurate transport behavior.
545
546 The parameters available for tuning include the $\sigma$ and
547 $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
548 strength of the sticky potential ($\nu_0$), and the cutoff distances
549 for the sticky attractive and dipole repulsive cubic switching
550 function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
551 respectively). The results of the reparametrizations are shown in
552 table \ref{tab:ssdParams}. We are calling these reparametrizations the
553 Soft Sticky Dipole Reaction Field (SSD/RF - for use with a reaction
554 field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve
555 the liquid structure in simulations without a long-range correction).
556
557 \begin{table}
558 \caption{PARAMETERS FOR THE ORIGINAL AND ADJUSTED SSD MODELS}
559
560 \centering
561 \begin{tabular}{ llcccc }
562 \toprule
563 \toprule
564 Parameters & & SSD & SSD1 & SSD/E & SSD/RF \\
565 \midrule
566 $\sigma$ & (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
567 $\epsilon$ & (kcal mol$^{-1}$) & 0.152 & 0.152 & 0.152 & 0.152\\
568 $\mu$ & (D) & 2.35 & 2.35 & 2.42 & 2.48\\
569 $\nu_0$ & (kcal mol$^{-1}$) & 3.7284 & 3.6613 & 3.90 & 3.90\\
570 $\omega^\circ$ & & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
571 $r_l$ & (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
572 $r_u$ & (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
573 $r_l^\prime$ & (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
574 $r_u^\prime$ & (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
575 \bottomrule
576 \end{tabular}
577 \label{tab:ssdParams}
578 \end{table}
579
580 \begin{figure}
581 \centering
582 \includegraphics[width=4.5in]{./figures/newGofRCompare.pdf}
583 \caption{ Plots showing the experimental $g(r)$ (Ref. \cite{Hura00})
584 with SSD/E and SSD1 without reaction field (top), as well as SSD/RF
585 and SSD1 with reaction field turned on (bottom). The changes in
586 parameters have lowered and broadened the first peak of SSD/E and
587 SSD/RF, resulting in a better fit to the first solvation shell.}
588 \label{fig:gofrCompare}
589 \end{figure}
590
591 \begin{figure}
592 \centering
593 \includegraphics[width=\linewidth]{./figures/dualPotentials.pdf}
594 \caption{ Positive and negative isosurfaces of the sticky potential for
595 SSD and SSD1 (A) and SSD/E \& SSD/RF (B). Gold areas correspond to the
596 tetrahedral attractive component, and blue areas correspond to the
597 dipolar repulsive component.}
598 \label{fig:isosurface}
599 \end{figure}
600
601 In the original paper detailing the development of SSD, Liu and Ichiye
602 placed particular emphasis on an accurate description of the first
603 solvation shell. This resulted in a somewhat tall and narrow first
604 peak in $g(r)$ that integrated to give similar coordination numbers to
605 the experimental data obtained by Soper and
606 Phillips.\cite{Liu96b,Soper86} New experimental x-ray scattering data
607 from Hura {\it et al.} indicates a slightly lower and shifted first
608 peak in the $g_\textrm{OO}(r)$, so our adjustments to SSD were made
609 after taking into consideration the new experimental
610 findings.\cite{Hura00} Figure \ref{fig:gofrCompare} shows the
611 relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
612 the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
613 experimental results. Both modified water models have shorter peaks
614 that match more closely to the experimental peak (as seen in the
615 insets of figure \ref{fig:gofrCompare}). This structural alteration
616 was accomplished by the combined reduction in the Lennard-Jones
617 $\sigma$ variable and adjustment of the sticky potential strength and
618 cutoffs. As can be seen in table \ref{tab:ssdParams}, the cutoffs for
619 the tetrahedral attractive and dipolar repulsive terms were nearly
620 swapped with each other. Isosurfaces of the original and modified
621 sticky potentials are shown in figure \ref{fig:isosurface}. In these
622 isosurfaces, it is easy to see how altering the cutoffs changes the
623 repulsive and attractive character of the particles. With a reduced
624 repulsive surface, the particles can move closer to one another,
625 increasing the density for the overall system. This change in
626 interaction cutoff also results in a more gradual orientational motion
627 by allowing the particles to maintain preferred dipolar arrangements
628 before they begin to feel the pull of the tetrahedral
629 restructuring. As the particles move closer together, the dipolar
630 repulsion term becomes active and excludes unphysical nearest-neighbor
631 arrangements. This compares with how SSD and SSD1 exclude preferred
632 dipole alignments before the particles feel the pull of the ``hydrogen
633 bonds''. Aside from improving the shape of the first peak in the
634 $g(r)$, this modification improves the densities considerably by
635 allowing the persistence of full dipolar character below the previous
636 4~\AA\ cutoff.
637
638 While adjusting the location and shape of the first peak of $g(r)$
639 improves the densities, these changes alone are insufficient to bring
640 the system densities up to the values observed experimentally. To
641 further increase the densities, the dipole moments were increased in
642 both of our adjusted models. Since SSD is a dipole based model, the
643 structure and transport are very sensitive to changes in the dipole
644 moment. The original SSD simply used the dipole moment calculated from
645 the TIP3P water model, which at 2.35~D is significantly greater than
646 the experimental gas phase value of 1.84~D. The larger dipole moment
647 is a more realistic value and improves the dielectric properties of
648 the fluid. Both theoretical and experimental measurements indicate a
649 liquid phase dipole moment ranging from 2.4~D to values as high as
650 3.11~D, providing a substantial range of reasonable values for a
651 dipole moment.\cite{Sprik91,Gubskaya02,Badyal00,Barriol64} Moderately
652 increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF,
653 respectively, leads to significant changes in the density and
654 transport of the water models.
655
656 \subsection{Density Behavior}
657
658 In order to demonstrate the benefits of these reparametrizations, we
659 performed a series of {\it NPT} and {\it NVE} simulations to probe the
660 density and transport properties of the adapted models and compare the
661 results to the original SSD model. This comparison involved full {\it
662 NPT} melting sequences for both SSD/E and SSD/RF, as well as {\it NVE}
663 transport calculations at the calculated self-consistent
664 densities. Again, the results were obtained from five separate
665 simulations of 1024 particle systems, and the melting sequences were
666 started from different ice I$_\textrm{h}$ crystals constructed as
667 described previously. Each {\it NPT} simulation was equilibrated for
668 100~ps before a 200~ps data collection run at each temperature step,
669 and the final configuration from the previous temperature simulation
670 was used as a starting point. All {\it NVE} simulations had the same
671 thermalization, equilibration, and data collection times as stated
672 previously.
673
674 \begin{figure}
675 \centering
676 \includegraphics[width=\linewidth]{./figures/ssdeDense.pdf}
677 \caption{ Comparison of densities calculated with SSD/E to
678 SSD1 without a reaction field, TIP3P, SPC/E, TIP5P, and
679 experiment.\cite{Jorgensen98b,Baez94,Mahoney00,CRC80} Both SSD1 and
680 SSD/E show good agreement with experiment when the long-range
681 correction is neglected.}
682 \label{fig:ssdeDense}
683 \end{figure}
684
685 Figure \ref{fig:ssdeDense} shows the density profiles for SSD/E, SSD1,
686 TIP3P, TIP4P, and SPC/E alongside the experimental results. The
687 calculated densities for both SSD/E and SSD1 have increased
688 significantly over the original SSD model (see figure
689 \ref{fig:ssdDense}) and are in better agreement with the experimental
690 values. At 298~K, the densities of SSD/E and SSD1 without a long-range
691 correction are 0.996~g/cm$^3$ and 0.999~g/cm$^3$ respectively. These
692 both compare well with the experimental value of 0.997~g/cm$^3$, and
693 they are considerably better than the SSD value of 0.967~g/cm$^3$. The
694 changes to the dipole moment and sticky switching functions have
695 improved the structuring of the liquid (as seen in figure
696 \ref{fig:gofrCompare}), but they have shifted the density maximum to
697 much lower temperatures. This comes about via an increase in the
698 liquid disorder through the weakening of the sticky potential and
699 strengthening of the dipolar character. However, this increasing
700 disorder in the SSD/E model has little effect on the melting
701 transition. By monitoring $C_p$ throughout these simulations, we
702 observed a melting transition for SSD/E at 235~K, the same as SSD and
703 SSD1.
704
705 \begin{figure}
706 \centering
707 \includegraphics[width=\linewidth]{./figures/ssdrfDense.pdf}
708 \caption{ Comparison of densities calculated with SSD/RF to
709 SSD1 with a reaction field, TIP3P, SPC/E, TIP5P, and
710 experiment.\cite{Jorgensen98b,Baez94,Mahoney00,CRC80} This plot
711 shows the benefit afforded by the reparametrization for use with a
712 reaction field correction - SSD/RF provides significantly more
713 accurate densities than SSD1 when performing room temperature
714 simulations.}
715 \label{fig:ssdrfDense}
716 \end{figure}
717
718 Including the reaction field long-range correction results is a more
719 interesting comparison. A density profile including SSD/RF and SSD1
720 with an active reaction field is shown in figure \ref{fig:ssdrfDense}.
721 As observed in the simulations without a reaction field, the densities
722 of SSD/RF and SSD1 show a dramatic increase over normal SSD (see
723 figure \ref{fig:ssdDense}). At 298~K, SSD/RF has a density of 0.997
724 g/cm$^3$, directly in line with experiment and considerably better
725 than the original SSD value of 0.941~g/cm$^3$ and the SSD1 value of
726 0.972~g/cm$^3$. These results further emphasize the importance of
727 reparametrization in order to model the density properly under
728 different simulation conditions. Again, these changes have only a
729 minor effect on the melting point, which observed at 245~K for SSD/RF,
730 is identical to SSD and only 5~K lower than SSD1 with a reaction
731 field. Additionally, the difference in density maxima is not as
732 extreme, with SSD/RF showing a density maximum at 255~K, fairly close
733 to the density maxima of 260~K and 265~K, shown by SSD and SSD1
734 respectively.
735
736 \subsection{Transport Behavior}
737
738 \begin{figure}
739 \centering
740 \includegraphics[width=\linewidth]{./figures/ssdeDiffuse.pdf}
741 \caption{ The diffusion constants calculated from SSD/E and
742 SSD1 (both without a reaction field) along with experimental
743 results.\cite{Gillen72,Holz00} The {\it NVE} calculations were
744 performed at the average densities from the {\it NPT} simulations for
745 the respective models. SSD/E is slightly more mobile than experiment
746 at all of the temperatures, but it is closer to experiment at
747 biologically relevant temperatures than SSD1 without a long-range
748 correction.}
749 \label{fig:ssdeDiffuse}
750 \end{figure}
751
752 The reparametrization of the SSD water model, both for use with and
753 without an applied long-range correction, brought the densities up to
754 what is expected for proper simulation of liquid water. In addition to
755 improving the densities, it is important that the diffusive behavior
756 of SSD be maintained or improved. Figure \ref{fig:ssdeDiffuse}
757 compares the temperature dependence of the diffusion constant of SSD/E
758 to SSD1 without an active reaction field at the densities calculated
759 from their respective {\it NPT} simulations at 1 atm. The diffusion
760 constant for SSD/E is consistently higher than experiment, while SSD1
761 remains lower than experiment until relatively high temperatures
762 (around 360~K). Both models follow the shape of the experimental curve
763 below 300~K but tend to diffuse too rapidly at higher temperatures, as
764 seen in SSD1 crossing above 360~K. This increasing diffusion relative
765 to the experimental values is caused by the rapidly decreasing system
766 density with increasing temperature. Both SSD1 and SSD/E show this
767 deviation in particle mobility, but this trend has different
768 implications on the diffusive behavior of the models. While SSD1
769 shows more experimentally accurate diffusive behavior in the high
770 temperature regimes, SSD/E shows more accurate behavior in the
771 supercooled and biologically relevant temperature ranges. Thus, the
772 changes made to improve the liquid structure may have had an adverse
773 affect on the density maximum, but they improve the transport behavior
774 of SSD/E relative to SSD1 under the most commonly simulated
775 conditions.
776
777 \begin{figure}
778 \centering
779 \includegraphics[width=\linewidth]{./figures/ssdrfDiffuse.pdf}
780 \caption{ The diffusion constants calculated from SSD/RF and
781 SSD1 (both with an active reaction field) along with experimental
782 results.\cite{Gillen72,Holz00} The {\it NVE} calculations were
783 performed at the average densities from the {\it NPT} simulations for
784 both of the models. SSD/RF captures the self-diffusion of water
785 throughout most of this temperature range. The increasing diffusion
786 constants at high temperatures for both models can be attributed to
787 lower calculated densities than those observed in experiment.}
788 \label{fig:ssdrfDiffuse}
789 \end{figure}
790
791 In figure \ref{fig:ssdrfDiffuse}, the diffusion constants for SSD/RF are
792 compared to SSD1 with an active reaction field. Note that SSD/RF
793 tracks the experimental results quantitatively, identical within error
794 throughout most of the temperature range shown and exhibiting only a
795 slight increasing trend at higher temperatures. SSD1 tends to diffuse
796 more slowly at low temperatures and deviates to diffuse too rapidly at
797 temperatures greater than 330~K. As stated above, this deviation away
798 from the ideal trend is due to a rapid decrease in density at higher
799 temperatures. SSD/RF does not suffer from this problem as much as SSD1
800 because the calculated densities are closer to the experimental
801 values. These results again emphasize the importance of careful
802 reparametrization when using an altered long-range correction.
803
804 \subsection{Summary of Liquid State Properties}
805
806 \begin{table}
807 \caption{PROPERTIES OF THE SINGLE-POINT WATER MODELS COMPARED WITH
808 EXPERIMENTAL DATA AT AMBIENT CONDITIONS}
809 \footnotesize
810 \centering
811 \begin{tabular}{ llccccc }
812 \toprule
813 \toprule
814 & & SSD1 & SSD/E & SSD1 (RF) & SSD/RF & Experiment [Ref.] \\
815 \midrule
816 $\rho$ & (g cm$^{-3}$) & 0.999(1) & 0.996(1) & 0.972(2) & 0.997(1) & 0.997 \cite{CRC80}\\
817 $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 28.80(11) & 25.45(9) & 28.28(6) & 23.83(16) & 18.005 \cite{Wagner02}\\
818 $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 1.78(7) & 2.51(18) & 2.00(17) & 2.32(6) & 2.299 \cite{Mills73}\\
819 $n_C$ & & 3.9 & 4.3 & 3.8 & 4.4 & 4.7 \cite{Hura00}\\
820 $n_H$ & & 3.7 & 3.6 & 3.7 & 3.7 & 3.5 \cite{Soper86}\\
821 $\tau_1$ & (ps) & 10.9(6) & 7.3(4) & 7.5(7) & 7.2(4) & 5.7 \cite{Eisenberg69}\\
822 $\tau_2$ & (ps) & 4.7(4) & 3.1(2) & 3.5(3) & 3.2(2) & 2.3 \cite{Krynicki66}\\
823 \bottomrule
824 \end{tabular}
825 \label{tab:liquidProperties}
826 \end{table}
827
828 Table \ref{tab:liquidProperties} gives a synopsis of the liquid state
829 properties of the water models compared in this study along with the
830 experimental values for liquid water at ambient conditions. The
831 coordination number ($n_C$) and number of hydrogen bonds per particle
832 ($n_H$) were calculated by integrating the following relations:
833 \begin{equation}
834 n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2g_{\textrm{OO}}(r)dr,
835 \end{equation}
836 \begin{equation}
837 n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2g_{\textrm{OH}}(r)dr,
838 \end{equation}
839 where $\rho$ is the number density of the specified pair interactions,
840 $a$ and $b$ are the radial locations of the minima following the first
841 peak in $g_\textrm{OO}(r)$ or $g_\textrm{OH}(r)$ respectively. The
842 number of hydrogen bonds stays relatively constant across all of the
843 models, but the coordination numbers of SSD/E and SSD/RF show an
844 improvement over SSD1. This improvement is primarily due to extension
845 of the first solvation shell in the new parameter sets. Because $n_H$
846 and $n_C$ are nearly identical in SSD1, it appears that all molecules
847 in the first solvation shell are involved in hydrogen bonds. Since
848 $n_H$ and $n_C$ differ in the newly parameterized models, the
849 orientations in the first solvation shell are a bit more ``fluid''.
850 Therefore SSD1 over-structures the first solvation shell and our
851 reparametrizations have returned this shell to more realistic
852 liquid-like behavior.
853
854 The time constants for the orientational autocorrelation functions
855 are also displayed in Table \ref{tab:liquidProperties}. The dipolar
856 orientational time correlation functions ($C_{l}$) are described
857 by:
858 \begin{equation}
859 C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
860 \label{eq:reorientCorr}
861 \end{equation}
862 where $P_l$ are Legendre polynomials of order $l$ and
863 $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular
864 dipole.\cite{Rahman71} Note that this is identical to equation
865 (\ref{eq:OrientCorr}) were $\alpha$ is equal to $z$. From these
866 correlation functions, the orientational relaxation time of the dipole
867 vector can be calculated from an exponential fit in the long-time
868 regime ($t > \tau_l$).\cite{Rothschild84} Calculation of these time
869 constants were averaged over five detailed {\it NVE} simulations
870 performed at the ambient conditions for each of the respective
871 models. It should be noted that the commonly cited value of 1.9~ps for
872 $\tau_2$ was determined from the NMR data in Ref. \cite{Krynicki66} at
873 a temperature near 34$^\circ$C.\cite{Rahman71} Because of the strong
874 temperature dependence of $\tau_2$, it is necessary to recalculate it
875 at 298~K to make proper comparisons. The value shown in Table
876 \ref{tab:liquidProperties} was calculated from the same NMR data in the
877 fashion described in Ref. \cite{Krynicki66}. Similarly, $\tau_1$ was
878 recomputed for 298~K from the data in Ref. \cite{Eisenberg69}.
879 Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
880 and without an active reaction field. Turning on the reaction field
881 leads to much improved time constants for SSD1; however, these results
882 also include a corresponding decrease in system density.
883 Orientational relaxation times published in the original SSD dynamics
884 paper are smaller than the values observed here, and this difference
885 can be attributed to the use of the Ewald sum.\cite{Chandra99}
886
887 \subsection{SSD/RF and Damped Electrostatics}\label{sec:ssdrfDamped}
888
889 In section \ref{sec:dampingMultipoles}, a method was described for
890 applying the damped {\sc sf} or {\sc sp} techniques to for systems
891 containing point multipoles. The SSD family of water models is the
892 perfect test case because of the dipole-dipole (and
893 charge-dipole/quadrupole) interactions that are present. The {\sc sf}
894 and {\sc sp} techniques were presented as a pairwise replacement for
895 the Ewald summation. It has been suggested that models parametrized
896 for the Ewald summation (like TIP5P-E) would be appropriate for use
897 with a reaction field and vice versa.\cite{Rick04} Therefore, we
898 decided to test the SSD/RF water model with this damped electrostatic
899 technique in place of the reaction field to see how the calculated
900 properties change.
901
902 \begin{table}
903 \caption{PROPERTIES OF SSD/RF WHEN USING DIFFERENT ELECTROSTATIC
904 CORRECTION METHODS}
905 \footnotesize
906 \centering
907 \begin{tabular}{ llccc }
908 \toprule
909 \toprule
910 & & Reaction Field & Damped Electrostatics & Experiment [Ref.] \\
911 & & $\epsilon = 80$ & $\alpha = 0.2125$~\AA$^{-1}$ & \\
912 \midrule
913 $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 \cite{CRC80}\\
914 $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 \cite{Wagner02} \\
915 $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 \cite{Mills73}\\
916 $n_C$ & & 4.4 & 4.2 & 4.7 \cite{Hura00}\\
917 $n_H$ & & 3.7 & 3.7 & 3.5 \cite{Soper86}\\
918 $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 \cite{Eisenberg69}\\
919 $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 \cite{Krynicki66}\\
920 \bottomrule
921 \end{tabular}
922 \label{tab:dampedSSDRF}
923 \end{table}
924
925 The properties shown in table \ref{tab:dampedSSDRF} compare
926 surprisingly well. The average density shows a modest increase when
927 using damped electrostatics in place of the reaction field. This comes
928 about because we neglect the pressure effect due to the surroundings
929 outside of the cuttoff, instead relying on screening effects to
930 neutralize electrostatic interactions at long distances. The $C_p$
931 also shows a slight increase, indicating greater fluctuation in the
932 enthalpy at constant pressure. The only other differences between the
933 damped and reaction field results are the dipole reorientational time
934 constants, $\tau_1$ and $\tau_2$. When using damped electrostatics,
935 the water molecules relax more quickly and exhibit behavior very
936 similar to the experimental values. These results indicate that not
937 only is it reasonable to use damped electrostatics with SSD/RF, it is
938 recommended if capturing realistic dynamics is of primary
939 importance. This is an encouraging result because the damping methods
940 are more generally applicable than reaction field. Using damping,
941 SSD/RF can be used effectively with mixed systems, such as dissolved
942 ions, dissolved organic molecules, or even proteins.
943
944 In addition to the properties tabulated in table
945 \ref{tab:dampedSSDRF}, we calculated the static dielectric constant
946 from a 5~ns simulation of SSD/RF using the damped electrostatics. The
947 resulting value of 82.6(6) compares very favorably with the
948 experimental value of 78.3.\cite{Malmberg56} This value is closer to
949 the experimental value than what was expected according to figure
950 \ref{fig:dielectricMap}, raising some questions as to the accuracy of
951 the visual contours in the figure. This highlights the qualitative
952 nature of contour plotting.
953
954 \section{Tetrahedrally Restructured Elongated Dipole (TRED) Water Model}\label{sec:tredWater}
955
956 The SSD/RF model works well with damped electrostatics, but because of
957 its point multipole character, there is no charge neutralization
958 correction at $R_\textrm{c}$. This has the effect of increasing the
959 density, since there is no consideration of the ``surroundings''. In
960 an attempt to optimize a water model for these conditions, we decided
961 to both simplify the parameters of the SSD type models and break the
962 point dipole into a charge pair so that it will gain some effect from
963 the shifting action in the {\sc sf} technique. This process resulted
964 in a two-point model that we are calling the Tetrahedrally
965 Restructured Elongated Dipole (TRED) water model.
966
967 \begin{figure}
968 \centering
969 \includegraphics[width=2.75in]{./figures/tredLayout.pdf}
970 \caption{Geometry of TRED water. In this two-point model, the origin
971 is the center-of-mass of the water molecule with the same moments of
972 inertia as SSD/RF. The negatively charged site at the origin is also
973 both a Lennard-Jones and sticky interaction site.}
974 \label{fig:tredLayout}
975 \end{figure}
976 \begin{table}
977 \centering
978 \caption{PARAMETERS FOR THE TRED WATER MODEL}
979 \begin{tabular}{ llr }
980 \toprule
981 \toprule
982 $\sigma$ & (\AA) & 2.980 \\
983 $\epsilon$ & (kcal mol$^{-1}$) & 0.2045 \\
984 $q$ & ($e$) & 1.041 \\
985 $v_0$ & (kcal mol$^{-1}$) & 4.22 \\
986 $\omega^\circ$ & & 0.07715 \\
987 $r_l$ \& $r_l^\prime$ & (\AA) & 2.4 \\
988 $r_u$ \& $r_u^\prime$ & (\AA) & 4.0 \\
989 Q$_xx$ & (D \AA) & -1.682 \\
990 Q$_yy$ & (D \AA) & 1.762 \\
991 Q$_zz$ & (D \AA) & -0.080 \\
992 \bottomrule
993 \end{tabular}
994 \label{tab:tredParams}
995 \end{table}
996 \begin{figure}
997 \centering
998 \includegraphics[width=\linewidth]{./figures/tredGofR.pdf}
999 \caption{The $g_\textrm{OO}(r)$ for TRED water. The first peak has a
1000 closer to the experimental plot; however, the second and third peaks
1001 exhibit a more expanded structure similar to SSD/RF. The minimum
1002 following the first peak is located at 3.55~\AA , which is further out
1003 than both experiment and SSD/RF (3.42 and 3.3~\AA\ respectively),
1004 leading to a larger coordination number. If all the curves were
1005 integrated to the experimental minimum, the $n_C$ results would all be
1006 similar, with TRED having an $n_C$ only 0.2 larger than experiment.}
1007 \label{fig:tredGofR}
1008 \end{figure}
1009 The geometric layout of TRED water is shown in figure
1010 \ref{fig:tredLayout} and the electrostatic, Lennard-Jones, and sticky
1011 parameters are displayed in table \ref{tab:tredParams}. The negatively
1012 charged site is located at the center of mass of the molecule and is
1013 also home to the Lennard-Jones and sticky interactions. The charge
1014 separation distance of 0.5~\AA\ along the dipolar ($z$) axis combined
1015 with the charge magnitude ($q$) results in a dipole moment of
1016 2.5~D. This value is simply the value used for SSD/RF (2.48~D) rounded
1017 to the tenths place. We also unified the sticky parameters for the
1018 switching functions on the repulsive and attractive interactions in
1019 the interest of simplicity, and we left the quadrupole moment elements
1020 and $\omega^\circ$ unaltered. It should be noted that additional logic
1021 needs to be included into the electrostatic code when using TRED to
1022 insure that the charges of each water do not interact with the other
1023 water's quadrupole moment. Finally, the strength of the sticky
1024 interaction ($v_0$) was modified to improve the shape of the first
1025 peaks in $g_\textrm{OO}(r)$ and $g_\textrm{OH}(r)$, while the $\sigma$
1026 and $\epsilon$ values were varied to adjust the location of the first
1027 peak of $g_\textrm{OO}(r)$ (see figure \ref{fig:tredGofR}) and the
1028 density. The $\sigma$ and $\epsilon$ optimization was carried out by
1029 separating the Lennard-Jones potential into near entirely repulsive
1030 ($A$) and attractive ($C$) parts, such that
1031 \begin{equation}
1032 \sigma = \left(\frac{A}{C}\right)^\frac{1}{6},
1033 \end{equation}
1034 and
1035 \begin{equation}
1036 \epsilon = \frac{C^2}{4A}.
1037 \end{equation}
1038 The location of the peak is adjusted through $A$, while the
1039 interaction strength is modified through $C$. All of the above changes
1040 were made with the goal of capturing the experimental density and
1041 translational diffusion constant at 298~K and 1~atm.
1042
1043 Since TRED is a two-point water model, we expect its computational
1044 efficiency to fall some place in between the one-point and three-point
1045 water models. In comparative simulations, TRED was approximately 85\%
1046 slower than SSD/RF, while SPC/E was 225\% slower than SSD/RF. While
1047 TRED loses some of the performance advantage of SSD, it is still
1048 nearly twice as computationally efficient as SPC/E and TIP3P.
1049
1050 \begin{table}
1051 \caption{PROPERTIES OF TRED COMPARED WITH SSD/RF AND EXPERIMENT}
1052 \footnotesize
1053 \centering
1054 \begin{tabular}{ llccc }
1055 \toprule
1056 \toprule
1057 & & SSD/RF & TRED & Experiment [Ref.]\\
1058 & & $\alpha = 0.2125$~\AA$^{-1}$ & $\alpha = 0.2125$~\AA$^{-1}$ & \\
1059 \midrule
1060 $\rho$ & (g cm$^{-3}$) & 1.004(4) & 0.995(5) & 0.997 \cite{CRC80}\\
1061 $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 27(1) & 23(1) & 18.005 \cite{Wagner02} \\
1062 $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.33(2) & 2.30(5) & 2.299 \cite{Mills73}\\
1063 $n_C$ & & 4.2 & 5.3 & 4.7 \cite{Hura00}\\
1064 $n_H$ & & 3.7 & 4.1 & 3.5 \cite{Soper86}\\
1065 $\tau_1$ & (ps) & 5.86(8) & 6.0(1) & 5.7 \cite{Eisenberg69}\\
1066 $\tau_2$ & (ps) & 2.45(7) & 2.49(5) & 2.3 \cite{Krynicki66}\\
1067 $\epsilon_0$ & & 82.6(6) & 83(1) & 78.3 \cite{Malmberg56}\\
1068 $\tau_D$ & (ps) & 9.1(2) & 10.6(3) & 8.2(4) \cite{Kindt96}\\
1069 \bottomrule
1070 \end{tabular}
1071 \label{tab:tredProperties}
1072 \end{table}
1073 We calculated thermodynamic and dynamic properties in the same manner
1074 described in section \ref{sec:ssdrfDamped} for SSD/RF, and the results
1075 are presented in table \ref{tab:tredProperties}. These results show that
1076 TRED improves upon the $\rho$ and $C_p$ properties of SSD/RF with
1077 damped electrostatics while maintaining the excellent dynamics
1078 behavior of both the translational self-diffusivity and the
1079 reorientational time constants, $\tau_1$ and $\tau_2$. The structural
1080 results show some differences between the two models. Despite the
1081 improved peak shape for the first solvation shell (see figure
1082 \ref{fig:tredGofR}), $n_C$ and $n_H$ counts increase because of the
1083 further first minimum distance locations. This results in the
1084 integration extending over a larger water volume. If we integrate to
1085 the first minimum value of the experimental $g_\textrm{OO}(r)$
1086 (3.42~\AA ) in both the SSD/RF and TRED plots, the $n_C$ values for
1087 both are much closer to experiment (4.7 for SSD/RF and 4.9 for TRED).
1088
1089 \begin{figure}
1090 \centering
1091 \includegraphics[width=3in]{./figures/tredDielectric.pdf}
1092 \caption{Contour map of the dielectric constant for TRED as a function
1093 of damping parameter and cutoff radius. The dielectric behavior for TRED
1094 is very similar to SSD/RF (see figure \ref{fig:dielectricMap}D), which
1095 is to be expected due to the similar dipole moment and sticky interaction
1096 strength.}
1097 \label{fig:tredDielectric}
1098 \end{figure}
1099 A comparison of the dielectric behavior was also included at the
1100 bottom of table \ref{tab:tredProperties}. The static dielectric
1101 constant results for SSD/RF and TRED are identical within error. This
1102 is not surprising given the similar dipole moment, similar sticky
1103 interaction strength, and identical applied damping
1104 constant. Comparing the static dielectric constant contour map (figure
1105 \ref{fig:tredDielectric}) with the dielectric map for SSD/RF (figure
1106 \ref{fig:dielectricMap}D) highlights the similarities in how these
1107 models respond to dielectric damping and how the dipolar and monopolar
1108 electrostatic damping act in an equivalent fashion. Both these
1109 dielectric maps span a larger range than the 3, 4, and 5 point-charge
1110 water models; however, the SSD/RF range is greater than TRED,
1111 indicating that multipoles are a little more sensitive to damping than
1112 monopoles.
1113
1114 The final dielectric comparison comes through the Debye relaxation
1115 time ($\tau_D$) or the collective dipolar relaxation time when
1116 assuming a Debye treatment for the dielectric
1117 relaxation.\cite{Chandra99,Kindt96} This value is calculated through
1118 equation (\ref{eq:reorientCorr}) applied to the total system dipole
1119 moment. The values for both of the models are a slower than the
1120 experimental relaxation; however, they compare compare very well to
1121 experiment considering the Debye relaxation times calculated for the
1122 original SSD (11.95~ps) and the SPC/E (6.95~ps) and TIP3P (6.1~ps)
1123 values. The $\tau_D$ for TRED is about 1.5~ps slower than the $\tau_D$
1124 for SSD/RF, most likely due to the slower decay of the charge-charge
1125 interaction, even when screened by the same damping constant.
1126
1127 \section{Conclusions}
1128
1129 In the above sections, the density maximum and temperature dependence
1130 of the self-diffusion constant were studied for the SSD water model,
1131 both with and without the use of reaction field, via a series of {\it
1132 NPT} and {\it NVE} simulations. The constant pressure simulations
1133 showed a density maximum near 260~K. In most cases, the calculated
1134 densities were significantly lower than the densities obtained from
1135 other water models (and experiment). Analysis of self-diffusion showed
1136 SSD to capture the transport properties of water well in both the
1137 liquid and supercooled liquid regimes.
1138
1139 In order to correct the density behavior, we reparametrized the
1140 original SSD model for use both with and without a reaction field
1141 (SSD/RF and SSD/E), and made comparisons with SSD1, an alternate
1142 density corrected version of SSD. Both models improve the liquid
1143 structure, densities, and diffusive properties under their respective
1144 simulation conditions, indicating the necessity of reparametrization
1145 when changing the method of calculating long-range electrostatic
1146 interactions.
1147
1148 We also showed that SSD/RF performs well under the alternative damped
1149 electrostatic conditions, validating the multipolar damping work in
1150 the previous chapter. To improve the modeling of water when {\sc sf},
1151 the TRED water model was developed. This model maintains improves upon
1152 the thermodynamic properties of SSD/RF using damped electrostatics
1153 while maintaining the exceptional depiction of water dynamics.
1154
1155 The simple water models investigated here are excellent choices for
1156 representing water in large scale simulations of biochemical
1157 systems. They are more computationally efficient than the common
1158 charge based water models, and often exhibit more realistic bulk phase
1159 fluid properties. These models are one of the few cases in which
1160 maximizing efficiency does not result in a loss in realistic
1161 representation.