ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/waterChapter.tex
Revision: 2986
Committed: Wed Aug 30 22:14:37 2006 UTC (17 years, 10 months ago) by chrisfen
Content type: application/x-tex
File size: 61022 byte(s)
Log Message:
I may have messed some stuff up...

File Contents

# User Rev Content
1 chrisfen 2977 \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 it 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     criteria for choosing a water model are capturing the liquid state
35     properties and having the fewest number of points to insure efficient
36     performance. As researchers have begun to study larger systems, such
37     as entire viruses, the choice readily shifts towards efficiency over
38     accuracy in order to make the calculations
39     feasible.\cite{Freddolino06}
40    
41     \section{Soft Sticky Dipole Model for Water}
42    
43     One recently developed model that largely succeeds in retaining the
44     accuracy of bulk properties while greatly reducing the computational
45     cost is the Soft Sticky Dipole (SSD) water
46     model.\cite{Liu96,Liu96b,Chandra99,Tan03} The SSD model was developed
47     as a modified form of the hard-sphere water model proposed by Bratko,
48     Blum, and Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point}
49     model which has an interaction site that is both a point dipole and a
50     Lennard-Jones core. However, since the normal aligned and
51     anti-aligned geometries favored by point dipoles are poor mimics of
52     local structure in liquid water, a short ranged ``sticky'' potential
53     is also added. The sticky potential directs the molecules to assume
54     the proper hydrogen bond orientation in the first solvation shell.
55    
56     The interaction between two SSD water molecules \emph{i} and \emph{j}
57     is given by the potential
58     \begin{equation}
59     u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
60     ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ +
61     u_{ij}^{sp}
62     ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j),
63     \end{equation}
64     where the ${\bf r}_{ij}$ is the position vector between molecules
65     \emph{i} and \emph{j} with magnitude $r_{ij}$, and
66     ${\bf \Omega}_i$ and ${\bf \Omega}_j$ represent the orientations of
67     the two molecules. The Lennard-Jones and dipole interactions are given
68     by the following familiar forms:
69     \begin{equation}
70     u_{ij}^{LJ}(r_{ij}) = 4\epsilon
71     \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right]
72     \ ,
73     \end{equation}
74     and
75     \begin{equation}
76     u_{ij}^{dp} = \frac{|\mu_i||\mu_j|}{4 \pi \epsilon_0 r_{ij}^3} \left(
77     \hat{\bf u}_i \cdot \hat{\bf u}_j - 3(\hat{\bf u}_i\cdot\hat{\bf
78     r}_{ij})(\hat{\bf u}_j\cdot\hat{\bf r}_{ij}) \right)\ ,
79     \end{equation}
80     where $\hat{\bf u}_i$ and $\hat{\bf u}_j$ are the unit vectors along
81     the dipoles of molecules $i$ and $j$ respectively. $|\mu_i|$ and
82     $|\mu_j|$ are the strengths of the dipole moments, and $\hat{\bf
83     r}_{ij}$ is the unit vector pointing from molecule $j$ to molecule
84     $i$.
85    
86     The sticky potential is somewhat less familiar:
87     \begin{equation}
88     u_{ij}^{sp}
89     ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
90     \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)
91     + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf
92     \Omega}_j)]\ .
93     \label{eq:stickyfunction}
94     \end{equation}
95     Here, $\nu_0$ is a strength parameter for the sticky potential, and
96     $s$ and $s^\prime$ are cubic switching functions which turn off the
97     sticky interaction beyond the first solvation shell. The $w$ function
98     can be thought of as an attractive potential with tetrahedral
99     geometry:
100     \begin{equation}
101     w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
102     \end{equation}
103     while the $w^\prime$ function counters the normal aligned and
104     anti-aligned structures favored by point dipoles:
105     \begin{equation}
106     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,
107     \end{equation}
108     It should be noted that $w$ is proportional to the sum of the $Y_3^2$
109     and $Y_3^{-2}$ spherical harmonics (a linear combination which
110     enhances the tetrahedral geometry for hydrogen bonded structures),
111     while $w^\prime$ is a purely empirical function. A more detailed
112     description of the functional parts and variables in this potential
113     can be found in the original SSD
114     articles.\cite{Liu96,Liu96b,Chandra99,Tan03}
115    
116     Since SSD is a single-point {\it dipolar} model, the force
117     calculations are simplified significantly relative to the standard
118     {\it charged} multi-point models. In the original Monte Carlo
119     simulations using this model, Liu and Ichiye reported that using SSD
120     decreased computer time by a factor of 6-7 compared to other
121     models.\cite{Liu96b} What is most impressive is that this savings did
122     not come at the expense of accurate depiction of the liquid state
123     properties. Indeed, SSD maintains reasonable agreement with the Soper
124     data for the structural features of liquid water.\cite{Soper86,Liu96b}
125     Additionally, the dynamical properties exhibited by SSD agree with
126     experiment better than those of more computationally expensive models
127     (like TIP3P and SPC/E).\cite{Chandra99} The combination of speed and
128     accurate depiction of solvent properties makes SSD a very attractive
129     model for the simulation of large scale biochemical simulations.
130    
131     It is important to note that the SSD model was originally developed
132     for use with the Ewald summation for handling long-range
133     electrostatics.\cite{Ewald21} In applying this water model in a
134     variety of molecular systems, it would be useful to know its
135     properties and behavior under the more computationally efficient
136     reaction field (RF) technique, the correction techniques discussed in
137     the previous chapter, or even a simple
138     cutoff.\cite{Onsager36,Fennell06} This study addresses these issues by
139     looking at the structural and transport behavior of SSD over a variety
140     of temperatures with the purpose of utilizing the RF correction
141     technique. We then suggest modifications to the parameters that
142     result in more realistic bulk phase behavior. It should be noted that
143     in a recent publication, some of the original investigators of the SSD
144     water model have suggested adjustments to the SSD water model to
145     address abnormal density behavior (also observed here), calling the
146     corrected model SSD1.\cite{Tan03} In the later sections of this
147     chapter, we compare our modified variants of SSD with both the
148     original SSD and SSD1 models and discuss how our changes improve the
149     depiction of water.
150    
151     \section{Simulation Methods}
152    
153     Most of the calculations in this particular study were performed using
154     a internally developed simulation code that was one of the precursors
155     of the {\sc oopse} molecular dynamics (MD) package.\cite{Meineke05}
156     All of the capabilities of this code have been efficiently
157     incorporated into {\sc oopse}, and calculation results are consistent
158     between the two simulation packages. The later calculations involving
159     the damped shifted force ({\sc sf}) techniques were performed using
160     {\sc oopse}.
161    
162     In the primary simulations of this study, long-range dipole-dipole
163     interaction corrections were accounted for by using either the
164     reaction field technique or a simple cubic switching function at the
165     cutoff radius. Interestingly, one of the early applications of a
166     reaction field was in Monte Carlo simulations of liquid
167     water.\cite{Barker73} In this method, the magnitude of the reaction
168     field acting on dipole $i$ is
169     \begin{equation}
170     \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
171     \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
172     \label{eq:rfequation}
173     \end{equation}
174     where $\mathcal{R}$ is the cavity defined by the cutoff radius
175     ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
176     system, ${\bf\mu}_{j}$ is the dipole moment vector of particle $j$,
177     and $s(r_{ij})$ is a cubic switching function.\cite{Allen87} The
178     reaction field contribution to the total energy by particle $i$ is
179     given by $-\frac{1}{2}{\bf\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
180     on dipole $i$ by ${\bf\mu}_{i}\times\mathcal{E}_{i}$.\cite{Allen87} An
181     applied reaction field will alter the bulk orientational properties of
182     simulated water, and there is particular sensitivity of these
183     properties on changes in the length of the cutoff
184     radius.\cite{vanderSpoel98} This variable behavior makes reaction
185     field a less attractive method than the Ewald sum; however, for very
186     large systems, the computational benefit of reaction field is is
187     significant.
188    
189     In contrast to the simulations with a reaction field, we have also
190     performed a companion set of simulations {\it without} a surrounding
191     dielectric (i.e. using a simple cubic switching function at the cutoff
192     radius). As a result, we have developed two reparametrizations of SSD
193     which can be used either with or without an active reaction field.
194    
195     To determine the preferred densities of the models, we performed
196     simulations in the isobaric-isothermal ({\it NPT}) ensemble. All
197     dynamical properties for these models were then obtained from
198     microcanonical ({\it NVE}) simulations done at densities matching the
199     {\it NPT} density for a particular target temperature. The constant
200     pressure simulations were implemented using an integral thermostat and
201     barostat as outlined by Hoover.\cite{Hoover85,Hoover86} All molecules
202     were treated as non-linear rigid bodies. Vibrational constraints are
203     not necessary in simulations of SSD, because there are no explicit
204     hydrogen atoms, and thus no molecular vibrational modes need to be
205     considered.
206    
207     The symplectic splitting method proposed by Dullweber, Leimkuhler, and
208     McLachlan ({\sc dlm}, see section \ref{sec:IntroIntegrate}) was used
209     to carry out the integration of the equations of motion in place of
210     the more prevalent quaternion
211     method.\cite{Dullweber97,Evans77,Evans77b,Allen87} The reason behind
212     this decision was that, in {\it NVE} simulations, the energy drift
213     when using quaternions was substantially greater than when using the
214     {\sc dlm} method (Fig. \ref{fig:timeStepIntegration}). This steady
215     drift in the total energy has also been observed in other
216     studies.\cite{Kol97}
217    
218     \begin{figure}
219     \centering
220     \includegraphics[width=\linewidth]{./figures/timeStepIntegration.pdf}
221     \caption{Energy conservation using both quaternion-based integration
222     and the {\sc dlm} method with increasing time step. The larger time
223     step plots are shifted from the true energy baseline (that of $\Delta
224 chrisfen 2981 t$ = 0.1~fs) for clarity.}
225 chrisfen 2977 \label{fig:timeStepIntegration}
226     \end{figure}
227    
228     The {\sc dlm} method allows for Verlet style integration orientational
229     motion of rigid bodies via a sequence of rotation matrix
230     operations. Because these matrix operations are more costly than the
231     simpler arithmetic operations for quaternion propagation, typical SSD
232     particle simulations using {\sc dlm} are 5-10\% slower than
233     simulations using the quaternion method and an identical time
234     step. This additional expense is justified because of the ability to
235     use time steps that are more that twice as long and still achieve the
236     same energy conservation.
237    
238     Figure \ref{fig:timeStepIntegration} shows the resulting energy drift
239     at various time steps for both {\sc dlm} and quaternion
240     integration. All of the 1000 SSD particle simulations started with the
241     same configuration, and the only difference was the method used to
242 chrisfen 2981 handle orientational motion. At time steps of 0.1 and 0.5~fs, both
243 chrisfen 2977 methods for propagating the orientational degrees of freedom conserve
244     energy fairly well, with the quaternion method showing a slight energy
245 chrisfen 2981 drift over time in the 0.5~fs time step simulation. Time steps of 1 and
246     2~fs clearly demonstrate the benefits in energy conservation that come
247 chrisfen 2977 with the {\sc dlm} method. Thus, while maintaining the same degree of
248     energy conservation, one can take considerably longer time steps,
249     leading to an overall reduction in computation time.
250    
251     Energy drifts in water simulations using {\sc dlm} integration were
252 chrisfen 2981 unnoticeable for time steps up to 3~fs. We observed a slight energy
253 chrisfen 2977 drift on the order of 0.012~kcal/mol per nanosecond with a time step
254 chrisfen 2981 of 4~fs. As expected, this drift increases dramatically with increasing
255 chrisfen 2977 time step. To insure accuracy in our {\it NVE} simulations, time steps
256 chrisfen 2981 were set at 2~fs and were also kept at this value for {\it NPT}
257 chrisfen 2977 simulations.
258    
259     Proton-disordered ice crystals in both the I$_\textrm{h}$ and
260     I$_\textrm{c}$ lattices were generated as starting points for all
261     simulations. The I$_\textrm{h}$ crystals were formed by first
262     arranging the centers of mass of the SSD particles into a
263     ``hexagonal'' ice lattice of 1024 particles. Because of the crystal
264     structure of I$_\textrm{h}$ ice, the simulation boxes were
265     orthorhombic in shape with an edge length ratio of approximately
266     1.00$\times$1.06$\times$1.23. We then allowed the particles to orient
267     freely about their fixed lattice positions with angular momenta values
268 chrisfen 2981 randomly sampled at 400~K. The rotational temperature was then scaled
269     down in stages to slowly cool the crystals to 25~K. The particles were
270 chrisfen 2977 then allowed to translate with fixed orientations at a constant
271 chrisfen 2981 pressure of 1atm for 50~ps at 25~K. Finally, all constraints were
272     removed and the ice crystals were allowed to equilibrate for 50~ps at
273     25~K and a constant pressure of 1atm. This procedure resulted in
274 chrisfen 2977 structurally stable I$_\textrm{h}$ ice crystals that obey the
275     Bernal-Fowler rules.\cite{Bernal33,Rahman72} This method was also
276     utilized in the making of diamond lattice I$_\textrm{c}$ ice crystals,
277     with each cubic simulation box consisting of either 512 or 1000
278     particles. Only isotropic volume fluctuations were performed under
279     constant pressure, so the ratio of edge lengths remained constant
280     throughout the simulations.
281    
282     \section{SSD Density Behavior}
283    
284     Melting studies were performed on the randomized ice crystals using
285     the {\it NPT} ensemble. During melting simulations, the melting
286     transition and the density maximum can both be observed, provided that
287     the density maximum occurs in the liquid and not the supercooled
288     regime. It should be noted that the calculated melting temperature
289     ($T_\textrm{m}$) will not be the true $T_\textrm{m}$ because of
290     super-heating due to the relatively short time scales in molecular
291     simulations. This behavior results in inflated $T_\textrm{m}$ values;
292     however, these values provide a reasonable initial estimate of
293     $T_\textrm{m}$.
294    
295     An ensemble average from five separate melting simulations was
296     acquired, each starting from different ice crystals generated as
297 chrisfen 2981 described previously. All simulations were equilibrated for 100~ps
298     prior to a 200~ps data collection run at each temperature setting. The
299     temperature range of study spanned from 25 to 400~K, with a maximum
300     degree increment of 25~K. For regions of interest along this stepwise
301     progression, the temperature increment was decreased from 25~K to 10
302     and 5~K. The above equilibration and production times were sufficient
303 chrisfen 2977 in that the fluctuations in the volume autocorrelation function damped
304 chrisfen 2981 out in all of the simulations in under 20~ps.
305 chrisfen 2977
306     Our initial simulations focused on the original SSD water model, and
307     an average density versus temperature plot is shown in figure
308     \ref{fig:ssdDense}. Note that the density maximum when using a
309 chrisfen 2981 reaction field appears between 255 and 265~K. There were smaller
310     fluctuations in the density at 260~K than at either 255 or 265~K, so we
311 chrisfen 2977 report this value as the location of the density maximum. Figure
312     \ref{fig:ssdDense} was constructed using ice I$_\textrm{h}$ crystals
313     for the initial configuration; though not pictured, the simulations
314     starting from ice I$_\textrm{c}$ crystal configurations showed similar
315     results, with a liquid-phase density maximum at the same temperature.
316    
317     \begin{figure}
318     \centering
319     \includegraphics[width=\linewidth]{./figures/ssdDense.pdf}
320     \caption{ Density versus temperature for TIP3P, SPC/E, TIP4P, SSD,
321     SSD with a reaction field, and
322 chrisfen 2979 experiment.\cite{Jorgensen98b,Baez94,CRC80}. Note that using a
323 chrisfen 2977 reaction field lowers the density more than the already lowered SSD
324     densities. The lower than expected densities for the SSD model
325     prompted the original reparametrization of SSD to SSD1.\cite{Tan03}}
326     \label{fig:ssdDense}
327     \end{figure}
328    
329     The density maximum for SSD compares quite favorably to other simple
330     water models. Figure \ref{fig:ssdDense} also shows calculated
331     densities of several other models and experiment obtained from other
332 chrisfen 2979 sources.\cite{Jorgensen98b,Baez94,CRC80} Of the listed simple water
333 chrisfen 2977 models, SSD has a temperature closest to the experimentally observed
334     density maximum. Of the {\it charge-based} models in figure
335     \ref{fig:ssdDense}, TIP4P has a density maximum behavior most like
336     that seen in SSD. Though not included in this plot, it is useful to
337     note that TIP5P has a density maximum nearly identical to the
338     experimentally measured temperature (see section
339     \ref{sec:t5peDensity}.
340    
341     Liquid state densities in water have been observed to be dependent on
342     the cutoff radius ($R_\textrm{c}$), both with and without the use of a
343     reaction field.\cite{vanderSpoel98} In order to address the possible
344     effect of $R_\textrm{c}$, simulations were performed with a cutoff
345 chrisfen 2981 radius of 12~\AA\, complementing the 9~\AA\ $R_\textrm{c}$ used in the
346 chrisfen 2977 previous SSD simulations. All of the resulting densities overlapped
347     within error and showed no significant trend toward lower or higher
348     densities in simulations both with and without reaction field.
349    
350     The key feature to recognize in figure \ref{fig:ssdDense} is the
351     density scaling of SSD relative to other common models at any given
352     temperature. SSD assumes a lower density than any of the other listed
353     models at the same pressure, behavior which is especially apparent at
354 chrisfen 2981 temperatures greater than 300~K. Lower than expected densities have
355 chrisfen 2977 been observed for other systems using a reaction field for long-range
356     electrostatic interactions, so the most likely reason for the reduced
357     densities is the presence of the reaction
358     field.\cite{vanderSpoel98,Nezbeda02} In order to test the effect of
359     the reaction field on the density of the systems, the simulations were
360     repeated without a reaction field present. The results of these
361     simulations are also displayed in figure \ref{fig:ssdDense}. Without
362     the reaction field, the densities increase to more experimentally
363     reasonable values, especially around the freezing point of liquid
364     water. The shape of the curve is similar to the curve produced from
365     SSD simulations using reaction field, specifically the rapidly
366     decreasing densities at higher temperatures; however, a shift in the
367 chrisfen 2981 density maximum location, down to 245~K, is observed. This is a more
368 chrisfen 2977 accurate comparison to the other listed water models, in that no long
369     range corrections were applied in those
370 chrisfen 2979 simulations.\cite{Baez94,Jorgensen98b} However, even without the
371 chrisfen 2981 reaction field, the density around 300~K is still significantly lower
372 chrisfen 2977 than experiment and comparable water models. This anomalous behavior
373     was what lead Tan {\it et al.} to recently reparametrize
374     SSD.\cite{Tan03} Throughout the remainder of the paper our
375     reparametrizations of SSD will be compared with their newer SSD1
376     model.
377    
378     \section{SSD Transport Behavior}
379    
380     Accurate dynamical properties of a water model are particularly
381     important when using the model to study permeation or transport across
382     biological membranes. In order to probe transport in bulk water, {\it
383     NVE} simulations were performed at the average densities obtained from
384     the {\it NPT} simulations at an identical target
385     temperature. Simulations started with randomized velocities and
386 chrisfen 2981 underwent 50~ps of temperature scaling and 50~ps of constant energy
387     equilibration before a 200~ps data collection run. Diffusion constants
388 chrisfen 2977 were calculated via linear fits to the long-time behavior of the
389     mean-square displacement as a function of time.\cite{Allen87} The
390     averaged results from five sets of {\it NVE} simulations are displayed
391     in figure \ref{fig:ssdDiffuse}, alongside experimental, SPC/E, and TIP5P
392 chrisfen 2979 results.\cite{Gillen72,Holz00,Baez94,Mahoney01}
393 chrisfen 2977
394     \begin{figure}
395     \centering
396     \includegraphics[width=\linewidth]{./figures/ssdDiffuse.pdf}
397     \caption{ Average self-diffusion constant as a function of temperature for
398     SSD, SPC/E, and TIP5P compared with experimental
399 chrisfen 2979 data.\cite{Baez94,Mahoney01,Gillen72,Holz00} Of the three water
400 chrisfen 2977 models shown, SSD has the least deviation from the experimental
401     values. The rapidly increasing diffusion constants for TIP5P and SSD
402     correspond to significant decreases in density at the higher
403     temperatures.}
404     \label{fig:ssdDiffuse}
405     \end{figure}
406    
407     The observed values for the diffusion constant point out one of the
408     strengths of the SSD model. Of the three models shown, the SSD model
409     has the most accurate depiction of self-diffusion in both the
410     supercooled and liquid regimes. SPC/E does a respectable job by
411 chrisfen 2981 reproducing values similar to experiment around 290~K; however, it
412 chrisfen 2977 deviates at both higher and lower temperatures, failing to predict the
413     correct thermal trend. TIP5P and SSD both start off low at colder
414     temperatures and tend to diffuse too rapidly at higher temperatures.
415     This behavior at higher temperatures is not particularly surprising
416     since the densities of both TIP5P and SSD are lower than experimental
417     water densities at higher temperatures. When calculating the
418     diffusion coefficients for SSD at experimental densities (instead of
419     the densities from the {\it NPT} simulations), the resulting values
420     fall more in line with experiment at these temperatures.
421    
422     \section{Structural Changes and Characterization}
423    
424     By starting the simulations from the crystalline state, we can get an
425     estimation of the $T_\textrm{m}$ of the ice structure, and beyond the
426     melting point, we study the phase behavior of the liquid. The constant
427     pressure heat capacity ($C_\textrm{p}$) was monitored to locate
428     $T_\textrm{m}$ in each of the simulations. In the melting simulations
429     of the 1024 particle ice I$_\textrm{h}$ simulations, a large spike in
430 chrisfen 2981 $C_\textrm{p}$ occurs at 245~K, indicating a first order phase
431 chrisfen 2977 transition for the melting of these ice crystals (see figure
432     \ref{fig:ssdCp}. When the reaction field is turned off, the melting
433 chrisfen 2981 transition occurs at 235~K. These melting transitions are considerably
434     lower than the experimental value of 273~K, indicating that the solid
435 chrisfen 2977 ice I$_\textrm{h}$ is not thermodynamically preferred relative to the
436     liquid state at these lower temperatures.
437     \begin{figure}
438     \centering
439     \includegraphics[width=\linewidth]{./figures/ssdCp.pdf}
440     \caption{Heat capacity versus temperature for the SSD model with an
441 chrisfen 2981 active reaction field. Note the large spike in $C_p$ around 245~K,
442 chrisfen 2977 indicating a phase transition from the ordered crystal to disordered
443     liquid.}
444     \label{fig:ssdCp}
445     \end{figure}
446    
447     \begin{figure}
448     \centering
449     \includegraphics[width=\linewidth]{./figures/fullContour.pdf}
450     \caption{ Contour plots of 2D angular pair correlation functions for
451 chrisfen 2981 512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
452 chrisfen 2977 signify regions of enhanced density while light areas signify
453     depletion relative to the bulk density. White areas have pair
454     correlation values below 0.5 and black areas have values above 1.5.}
455     \label{fig:contour}
456     \end{figure}
457    
458     \begin{figure}
459     \centering
460     \includegraphics[width=2.5in]{./figures/corrDiag.pdf}
461 chrisfen 2980 \caption{ An illustration of angles involved in the correlations
462     observed in figure \ref{fig:contour}.}
463 chrisfen 2977 \label{fig:corrAngle}
464     \end{figure}
465    
466     Additional analysis of the melting process was performed using
467     two-dimensional structure and dipole angle correlations. Expressions
468     for these correlations are as follows:
469    
470     \begin{equation}
471     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\ ,
472     \end{equation}
473     \begin{equation}
474     g_{\text{AB}}(r,\cos\omega) =
475     \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\ ,
476     \end{equation}
477     where $\theta$ and $\omega$ refer to the angles shown in figure
478     \ref{fig:corrAngle}. By binning over both distance and the cosine of the
479     desired angle between the two dipoles, the $g(r)$ can be analyzed to
480     determine the common dipole arrangements that constitute the peaks and
481     troughs in the standard one-dimensional $g(r)$ plots. Frames A and B
482     of figure \ref{fig:contour} show results from an ice I$_\textrm{c}$
483     simulation. The first peak in the $g(r)$ consists primarily of the
484     preferred hydrogen bonding arrangements as dictated by the tetrahedral
485     sticky potential - one peak for the hydrogen bond donor and the other
486     for the hydrogen bond acceptor. Due to the high degree of
487     crystallinity of the sample, the second and third solvation shells
488     show a repeated peak arrangement which decays at distances around the
489     fourth solvation shell, near the imposed cutoff for the Lennard-Jones
490     and dipole-dipole interactions. In the higher temperature simulation
491     shown in frames C and D, these long-range features deteriorate
492     rapidly. The first solvation shell still shows the strong effect of
493     the sticky-potential, although it covers a larger area, extending to
494     include a fraction of aligned dipole peaks within the first solvation
495     shell. The latter peaks lose due to thermal motion and as the
496     competing dipole force overcomes the sticky potential's tight
497     tetrahedral structuring of the crystal.
498    
499     This complex interplay between dipole and sticky interactions was
500     remarked upon as a possible reason for the split second peak in the
501     oxygen-oxygen pair correlation function,
502     $g_\textrm{OO}(r)$.\cite{Liu96b} At low temperatures, the second
503     solvation shell peak appears to have two distinct components that
504     blend together to form one observable peak. At higher temperatures,
505 chrisfen 2981 this split character alters to show the leading 4~\AA\ peak dominated
506 chrisfen 2977 by equatorial anti-parallel dipole orientations. There is also a
507     tightly bunched group of axially arranged dipoles that most likely
508     consist of the smaller fraction of aligned dipole pairs. The trailing
509 chrisfen 2981 component of the split peak at 5~\AA\ is dominated by aligned dipoles
510 chrisfen 2977 that assume hydrogen bond arrangements similar to those seen in the
511     first solvation shell. This evidence indicates that the dipole pair
512     interaction begins to dominate outside of the range of the dipolar
513     repulsion term. The energetically favorable dipole arrangements
514     populate the region immediately outside this repulsion region (around
515 chrisfen 2981 4~\AA), while arrangements that seek to satisfy both the sticky and
516 chrisfen 2977 dipole forces locate themselves just beyond this initial buildup
517 chrisfen 2981 (around 5~\AA).
518 chrisfen 2977
519     This analysis indicates that the split second peak is primarily the
520     product of the dipolar repulsion term of the sticky potential. In
521     fact, the inner peak can be pushed out and merged with the outer split
522     peak just by extending the switching function ($s^\prime(r_{ij})$)
523 chrisfen 2981 from its normal 4~\AA\ cutoff to values of 4.5 or even 5~\AA. This
524 chrisfen 2977 type of correction is not recommended for improving the liquid
525     structure, since the second solvation shell would still be shifted too
526     far out. In addition, this would have an even more detrimental effect
527     on the system densities, leading to a liquid with a more open
528     structure and a density considerably lower than the already low SSD
529     density. A better correction would be to include the
530     quadrupole-quadrupole interactions for the water particles outside of
531     the first solvation shell, but this would remove the simplicity and
532     speed advantage of SSD.
533    
534     \section{Adjusted Potentials: SSD/RF and SSD/E}
535    
536     The propensity of SSD to adopt lower than expected densities under
537     varying conditions is troubling, especially at higher temperatures. In
538     order to correct this model for use with a reaction field, it is
539     necessary to adjust the force field parameters for the primary
540     intermolecular interactions. In undergoing a reparametrization, it is
541     important not to focus on just one property and neglect the others. In
542     this case, it would be ideal to correct the densities while
543     maintaining the accurate transport behavior.
544    
545     The parameters available for tuning include the $\sigma$ and
546     $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
547     strength of the sticky potential ($\nu_0$), and the cutoff distances
548     for the sticky attractive and dipole repulsive cubic switching
549     function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
550     respectively). The results of the reparametrizations are shown in
551     table \ref{tab:ssdParams}. We are calling these reparametrizations the
552     Soft Sticky Dipole Reaction Field (SSD/RF - for use with a reaction
553     field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve
554     the liquid structure in simulations without a long-range correction).
555    
556     \begin{table}
557     \caption{PARAMETERS FOR THE ORIGINAL AND ADJUSTED SSD MODELS}
558    
559     \centering
560 chrisfen 2981 \begin{tabular}{ llcccc }
561 chrisfen 2977 \toprule
562     \toprule
563 chrisfen 2981 Parameters & & SSD & SSD1 & SSD/E & SSD/RF \\
564 chrisfen 2977 \midrule
565 chrisfen 2981 $\sigma$ & (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
566     $\epsilon$ & (kcal mol$^{-1}$) & 0.152 & 0.152 & 0.152 & 0.152\\
567     $\mu$ & (D) & 2.35 & 2.35 & 2.42 & 2.48\\
568     $\nu_0$ & (kcal mol$^{-1}$) & 3.7284 & 3.6613 & 3.90 & 3.90\\
569     $\omega^\circ$ & & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
570     $r_l$ & (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
571     $r_u$ & (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
572     $r_l^\prime$ & (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
573     $r_u^\prime$ & (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
574 chrisfen 2977 \bottomrule
575     \end{tabular}
576     \label{tab:ssdParams}
577     \end{table}
578    
579     \begin{figure}
580     \centering
581     \includegraphics[width=4.5in]{./figures/newGofRCompare.pdf}
582     \caption{ Plots showing the experimental $g(r)$ (Ref. \cite{Hura00})
583     with SSD/E and SSD1 without reaction field (top), as well as SSD/RF
584     and SSD1 with reaction field turned on (bottom). The changes in
585     parameters have lowered and broadened the first peak of SSD/E and
586     SSD/RF, resulting in a better fit to the first solvation shell.}
587     \label{fig:gofrCompare}
588     \end{figure}
589    
590     \begin{figure}
591     \centering
592     \includegraphics[width=\linewidth]{./figures/dualPotentials.pdf}
593     \caption{ Positive and negative isosurfaces of the sticky potential for
594     SSD and SSD1 (A) and SSD/E \& SSD/RF (B). Gold areas correspond to the
595     tetrahedral attractive component, and blue areas correspond to the
596     dipolar repulsive component.}
597     \label{fig:isosurface}
598     \end{figure}
599    
600     In the original paper detailing the development of SSD, Liu and Ichiye
601     placed particular emphasis on an accurate description of the first
602     solvation shell. This resulted in a somewhat tall and narrow first
603     peak in $g(r)$ that integrated to give similar coordination numbers to
604     the experimental data obtained by Soper and
605     Phillips.\cite{Liu96b,Soper86} New experimental x-ray scattering data
606     from Hura {\it et al.} indicates a slightly lower and shifted first
607     peak in the $g_\textrm{OO}(r)$, so our adjustments to SSD were made
608     after taking into consideration the new experimental
609     findings.\cite{Hura00} Figure \ref{fig:gofrCompare} shows the
610     relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
611     the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
612     experimental results. Both modified water models have shorter peaks
613     that match more closely to the experimental peak (as seen in the
614     insets of figure \ref{fig:gofrCompare}). This structural alteration
615     was accomplished by the combined reduction in the Lennard-Jones
616     $\sigma$ variable and adjustment of the sticky potential strength and
617     cutoffs. As can be seen in table \ref{tab:ssdParams}, the cutoffs for
618     the tetrahedral attractive and dipolar repulsive terms were nearly
619     swapped with each other. Isosurfaces of the original and modified
620     sticky potentials are shown in figure \ref{fig:isosurface}. In these
621     isosurfaces, it is easy to see how altering the cutoffs changes the
622     repulsive and attractive character of the particles. With a reduced
623     repulsive surface, the particles can move closer to one another,
624     increasing the density for the overall system. This change in
625     interaction cutoff also results in a more gradual orientational motion
626     by allowing the particles to maintain preferred dipolar arrangements
627     before they begin to feel the pull of the tetrahedral
628     restructuring. As the particles move closer together, the dipolar
629     repulsion term becomes active and excludes unphysical nearest-neighbor
630     arrangements. This compares with how SSD and SSD1 exclude preferred
631     dipole alignments before the particles feel the pull of the ``hydrogen
632     bonds''. Aside from improving the shape of the first peak in the
633     $g(r)$, this modification improves the densities considerably by
634     allowing the persistence of full dipolar character below the previous
635 chrisfen 2981 4~\AA\ cutoff.
636 chrisfen 2977
637     While adjusting the location and shape of the first peak of $g(r)$
638     improves the densities, these changes alone are insufficient to bring
639     the system densities up to the values observed experimentally. To
640     further increase the densities, the dipole moments were increased in
641     both of our adjusted models. Since SSD is a dipole based model, the
642     structure and transport are very sensitive to changes in the dipole
643     moment. The original SSD simply used the dipole moment calculated from
644     the TIP3P water model, which at 2.35~D is significantly greater than
645     the experimental gas phase value of 1.84~D. The larger dipole moment
646     is a more realistic value and improves the dielectric properties of
647     the fluid. Both theoretical and experimental measurements indicate a
648     liquid phase dipole moment ranging from 2.4~D to values as high as
649     3.11~D, providing a substantial range of reasonable values for a
650     dipole moment.\cite{Sprik91,Gubskaya02,Badyal00,Barriol64} Moderately
651     increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF,
652     respectively, leads to significant changes in the density and
653     transport of the water models.
654    
655     \subsection{Density Behavior}
656    
657     In order to demonstrate the benefits of these reparametrizations, we
658     performed a series of {\it NPT} and {\it NVE} simulations to probe the
659     density and transport properties of the adapted models and compare the
660     results to the original SSD model. This comparison involved full {\it
661     NPT} melting sequences for both SSD/E and SSD/RF, as well as {\it NVE}
662     transport calculations at the calculated self-consistent
663     densities. Again, the results were obtained from five separate
664     simulations of 1024 particle systems, and the melting sequences were
665     started from different ice I$_\textrm{h}$ crystals constructed as
666     described previously. Each {\it NPT} simulation was equilibrated for
667 chrisfen 2981 100~ps before a 200~ps data collection run at each temperature step,
668 chrisfen 2977 and the final configuration from the previous temperature simulation
669     was used as a starting point. All {\it NVE} simulations had the same
670     thermalization, equilibration, and data collection times as stated
671     previously.
672    
673     \begin{figure}
674     \centering
675     \includegraphics[width=\linewidth]{./figures/ssdeDense.pdf}
676     \caption{ Comparison of densities calculated with SSD/E to
677     SSD1 without a reaction field, TIP3P, SPC/E, TIP5P, and
678 chrisfen 2979 experiment.\cite{Jorgensen98b,Baez94,Mahoney00,CRC80} Both SSD1 and
679 chrisfen 2977 SSD/E show good agreement with experiment when the long-range
680     correction is neglected.}
681     \label{fig:ssdeDense}
682     \end{figure}
683    
684     Figure \ref{fig:ssdeDense} shows the density profiles for SSD/E, SSD1,
685     TIP3P, TIP4P, and SPC/E alongside the experimental results. The
686     calculated densities for both SSD/E and SSD1 have increased
687     significantly over the original SSD model (see figure
688     \ref{fig:ssdDense}) and are in better agreement with the experimental
689 chrisfen 2981 values. At 298~K, the densities of SSD/E and SSD1 without a long-range
690     correction are 0.996~g/cm$^3$ and 0.999~g/cm$^3$ respectively. These
691     both compare well with the experimental value of 0.997~g/cm$^3$, and
692     they are considerably better than the SSD value of 0.967~g/cm$^3$. The
693 chrisfen 2977 changes to the dipole moment and sticky switching functions have
694     improved the structuring of the liquid (as seen in figure
695     \ref{fig:gofrCompare}), but they have shifted the density maximum to
696     much lower temperatures. This comes about via an increase in the
697     liquid disorder through the weakening of the sticky potential and
698     strengthening of the dipolar character. However, this increasing
699     disorder in the SSD/E model has little effect on the melting
700     transition. By monitoring $C_p$ throughout these simulations, we
701 chrisfen 2981 observed a melting transition for SSD/E at 235~K, the same as SSD and
702 chrisfen 2977 SSD1.
703    
704     \begin{figure}
705     \centering
706     \includegraphics[width=\linewidth]{./figures/ssdrfDense.pdf}
707     \caption{ Comparison of densities calculated with SSD/RF to
708     SSD1 with a reaction field, TIP3P, SPC/E, TIP5P, and
709 chrisfen 2979 experiment.\cite{Jorgensen98b,Baez94,Mahoney00,CRC80} This plot
710 chrisfen 2977 shows the benefit afforded by the reparametrization for use with a
711     reaction field correction - SSD/RF provides significantly more
712     accurate densities than SSD1 when performing room temperature
713     simulations.}
714     \label{fig:ssdrfDense}
715     \end{figure}
716    
717     Including the reaction field long-range correction results in a more
718     interesting comparison. A density profile including SSD/RF and SSD1
719     with an active reaction field is shown in figure \ref{fig:ssdrfDense}.
720     As observed in the simulations without a reaction field, the densities
721     of SSD/RF and SSD1 show a dramatic increase over normal SSD (see
722 chrisfen 2981 figure \ref{fig:ssdDense}). At 298~K, SSD/RF has a density of 0.997
723 chrisfen 2977 g/cm$^3$, directly in line with experiment and considerably better
724 chrisfen 2981 than the original SSD value of 0.941~g/cm$^3$ and the SSD1 value of
725     0.972~g/cm$^3$. These results further emphasize the importance of
726 chrisfen 2977 reparametrization in order to model the density properly under
727     different simulation conditions. Again, these changes have only a
728 chrisfen 2981 minor effect on the melting point, which observed at 245~K for SSD/RF,
729     is identical to SSD and only 5~K lower than SSD1 with a reaction
730 chrisfen 2977 field. Additionally, the difference in density maxima is not as
731 chrisfen 2981 extreme, with SSD/RF showing a density maximum at 255~K, fairly close
732     to the density maxima of 260~K and 265~K, shown by SSD and SSD1
733 chrisfen 2977 respectively.
734    
735     \subsection{Transport Behavior}
736    
737     \begin{figure}
738     \centering
739     \includegraphics[width=\linewidth]{./figures/ssdeDiffuse.pdf}
740     \caption{ The diffusion constants calculated from SSD/E and
741     SSD1 (both without a reaction field) along with experimental
742     results.\cite{Gillen72,Holz00} The {\it NVE} calculations were
743     performed at the average densities from the {\it NPT} simulations for
744     the respective models. SSD/E is slightly more mobile than experiment
745     at all of the temperatures, but it is closer to experiment at
746     biologically relevant temperatures than SSD1 without a long-range
747     correction.}
748     \label{fig:ssdeDiffuse}
749     \end{figure}
750    
751     The reparametrization of the SSD water model, both for use with and
752     without an applied long-range correction, brought the densities up to
753     what is expected for proper simulation of liquid water. In addition to
754     improving the densities, it is important that the diffusive behavior
755     of SSD be maintained or improved. Figure \ref{fig:ssdeDiffuse}
756     compares the temperature dependence of the diffusion constant of SSD/E
757     to SSD1 without an active reaction field at the densities calculated
758     from their respective {\it NPT} simulations at 1 atm. The diffusion
759     constant for SSD/E is consistently higher than experiment, while SSD1
760     remains lower than experiment until relatively high temperatures
761 chrisfen 2981 (around 360~K). Both models follow the shape of the experimental curve
762     below 300~K but tend to diffuse too rapidly at higher temperatures, as
763     seen in SSD1 crossing above 360~K. This increasing diffusion relative
764 chrisfen 2977 to the experimental values is caused by the rapidly decreasing system
765     density with increasing temperature. Both SSD1 and SSD/E show this
766     deviation in particle mobility, but this trend has different
767     implications on the diffusive behavior of the models. While SSD1
768     shows more experimentally accurate diffusive behavior in the high
769     temperature regimes, SSD/E shows more accurate behavior in the
770     supercooled and biologically relevant temperature ranges. Thus, the
771     changes made to improve the liquid structure may have had an adverse
772     affect on the density maximum, but they improve the transport behavior
773     of SSD/E relative to SSD1 under the most commonly simulated
774     conditions.
775    
776     \begin{figure}
777     \centering
778     \includegraphics[width=\linewidth]{./figures/ssdrfDiffuse.pdf}
779     \caption{ The diffusion constants calculated from SSD/RF and
780     SSD1 (both with an active reaction field) along with experimental
781     results.\cite{Gillen72,Holz00} The {\it NVE} calculations were
782     performed at the average densities from the {\it NPT} simulations for
783     both of the models. SSD/RF captures the self-diffusion of water
784     throughout most of this temperature range. The increasing diffusion
785     constants at high temperatures for both models can be attributed to
786     lower calculated densities than those observed in experiment.}
787     \label{fig:ssdrfDiffuse}
788     \end{figure}
789    
790     In figure \ref{fig:ssdrfDiffuse}, the diffusion constants for SSD/RF are
791     compared to SSD1 with an active reaction field. Note that SSD/RF
792     tracks the experimental results quantitatively, identical within error
793     throughout most of the temperature range shown and exhibiting only a
794     slight increasing trend at higher temperatures. SSD1 tends to diffuse
795     more slowly at low temperatures and deviates to diffuse too rapidly at
796 chrisfen 2981 temperatures greater than 330~K. As stated above, this deviation away
797 chrisfen 2977 from the ideal trend is due to a rapid decrease in density at higher
798     temperatures. SSD/RF does not suffer from this problem as much as SSD1
799     because the calculated densities are closer to the experimental
800     values. These results again emphasize the importance of careful
801     reparametrization when using an altered long-range correction.
802    
803     \subsection{Summary of Liquid State Properties}
804    
805     \begin{table}
806     \caption{PROPERTIES OF THE SINGLE-POINT WATER MODELS COMPARED WITH
807     EXPERIMENTAL DATA AT AMBIENT CONDITIONS}
808     \footnotesize
809     \centering
810 chrisfen 2978 \begin{tabular}{ llccccc }
811 chrisfen 2977 \toprule
812     \toprule
813 chrisfen 2978 & & SSD1 & SSD/E & SSD1 (RF) & SSD/RF & Experiment [Ref.] \\
814 chrisfen 2977 \midrule
815 chrisfen 2978 $\rho$ & (g cm$^{-3}$) & 0.999(1) & 0.996(1) & 0.972(2) & 0.997(1) & 0.997 \cite{CRC80}\\
816     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 28.80(11) & 25.45(9) & 28.28(6) & 23.83(16) & 18.005 \cite{Wagner02}\\
817     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 1.78(7) & 2.51(18) & 2.00(17) & 2.32(6) & 2.299 \cite{Mills73}\\
818     $n_C$ & & 3.9 & 4.3 & 3.8 & 4.4 & 4.7 \cite{Hura00}\\
819     $n_H$ & & 3.7 & 3.6 & 3.7 & 3.7 & 3.5 \cite{Soper86}\\
820     $\tau_1$ & (ps) & 10.9(6) & 7.3(4) & 7.5(7) & 7.2(4) & 5.7 \cite{Eisenberg69}\\
821     $\tau_2$ & (ps) & 4.7(4) & 3.1(2) & 3.5(3) & 3.2(2) & 2.3 \cite{Krynicki66}\\
822 chrisfen 2977 \bottomrule
823     \end{tabular}
824     \label{tab:liquidProperties}
825     \end{table}
826    
827     Table \ref{tab:liquidProperties} gives a synopsis of the liquid state
828     properties of the water models compared in this study along with the
829     experimental values for liquid water at ambient conditions. The
830     coordination number ($n_C$) and number of hydrogen bonds per particle
831     ($n_H$) were calculated by integrating the following relations:
832     \begin{equation}
833     n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2g_{\textrm{OO}}(r)dr,
834     \end{equation}
835     \begin{equation}
836     n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2g_{\textrm{OH}}(r)dr,
837     \end{equation}
838     where $\rho$ is the number density of the specified pair interactions,
839     $a$ and $b$ are the radial locations of the minima following the first
840     peak in $g_\textrm{OO}(r)$ or $g_\textrm{OH}(r)$ respectively. The
841     number of hydrogen bonds stays relatively constant across all of the
842     models, but the coordination numbers of SSD/E and SSD/RF show an
843     improvement over SSD1. This improvement is primarily due to extension
844     of the first solvation shell in the new parameter sets. Because $n_H$
845     and $n_C$ are nearly identical in SSD1, it appears that all molecules
846     in the first solvation shell are involved in hydrogen bonds. Since
847     $n_H$ and $n_C$ differ in the newly parameterized models, the
848     orientations in the first solvation shell are a bit more ``fluid''.
849     Therefore SSD1 over-structures the first solvation shell and our
850     reparametrizations have returned this shell to more realistic
851     liquid-like behavior.
852    
853     The time constants for the orientational autocorrelation functions
854     are also displayed in Table \ref{tab:liquidProperties}. The dipolar
855     orientational time correlation functions ($C_{l}$) are described
856     by:
857     \begin{equation}
858     C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
859 chrisfen 2986 \label{eq:reorientCorr}
860 chrisfen 2977 \end{equation}
861     where $P_l$ are Legendre polynomials of order $l$ and
862     $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular
863     dipole.\cite{Rahman71} Note that this is identical to equation
864     (\ref{eq:OrientCorr}) were $\alpha$ is equal to $z$. From these
865     correlation functions, the orientational relaxation time of the dipole
866     vector can be calculated from an exponential fit in the long-time
867 chrisfen 2978 regime ($t > \tau_l$).\cite{Rothschild84} Calculation of these time
868     constants were averaged over five detailed {\it NVE} simulations
869     performed at the ambient conditions for each of the respective
870 chrisfen 2981 models. It should be noted that the commonly cited value of 1.9~ps for
871 chrisfen 2978 $\tau_2$ was determined from the NMR data in Ref. \cite{Krynicki66} at
872     a temperature near 34$^\circ$C.\cite{Rahman71} Because of the strong
873     temperature dependence of $\tau_2$, it is necessary to recalculate it
874 chrisfen 2981 at 298~K to make proper comparisons. The value shown in Table
875 chrisfen 2977 \ref{tab:liquidProperties} was calculated from the same NMR data in the
876     fashion described in Ref. \cite{Krynicki66}. Similarly, $\tau_1$ was
877 chrisfen 2981 recomputed for 298~K from the data in Ref. \cite{Eisenberg69}.
878 chrisfen 2977 Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
879     and without an active reaction field. Turning on the reaction field
880     leads to much improved time constants for SSD1; however, these results
881     also include a corresponding decrease in system density.
882     Orientational relaxation times published in the original SSD dynamics
883     paper are smaller than the values observed here, and this difference
884     can be attributed to the use of the Ewald sum.\cite{Chandra99}
885    
886 chrisfen 2981 \subsection{SSD/RF and Damped Electrostatics}\label{sec:ssdrfDamped}
887 chrisfen 2977
888 chrisfen 2978 In section \ref{sec:dampingMultipoles}, a method was described for
889     applying the damped {\sc sf} or {\sc sp} techniques to for systems
890     containing point multipoles. The SSD family of water models is the
891     perfect test case because of the dipole-dipole (and
892     charge-dipole/quadrupole) interactions that are present. The {\sc sf}
893     and {\sc sp} techniques were presented as a pairwise replacement for
894     the Ewald summation. It has been suggested that models parametrized
895     for the Ewald summation (like TIP5P-E) would be appropriate for use
896     with a reaction field and vice versa.\cite{Rick04} Therefore, we
897     decided to test the SSD/RF water model with this damped electrostatic
898     technique in place of the reaction field to see how the calculated
899     properties change.
900    
901 chrisfen 2977 \begin{table}
902 chrisfen 2980 \caption{PROPERTIES OF SSD/RF WHEN USING DIFFERENT ELECTROSTATIC
903     CORRECTION METHODS}
904 chrisfen 2977 \footnotesize
905     \centering
906 chrisfen 2978 \begin{tabular}{ llccc }
907 chrisfen 2977 \toprule
908     \toprule
909 chrisfen 2978 & & Reaction Field & Damped Electrostatics & Experiment [Ref.] \\
910 chrisfen 2981 & & $\epsilon = 80$ & $\alpha = 0.2125$~\AA$^{-1}$ & \\
911 chrisfen 2977 \midrule
912 chrisfen 2978 $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 \cite{CRC80}\\
913     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 \cite{Wagner02} \\
914     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 \cite{Mills73}\\
915 chrisfen 2981 $n_C$ & & 4.4 & 4.2 & 4.7 \cite{Hura00}\\
916 chrisfen 2978 $n_H$ & & 3.7 & 3.7 & 3.5 \cite{Soper86}\\
917     $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 \cite{Eisenberg69}\\
918     $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 \cite{Krynicki66}\\
919 chrisfen 2977 \bottomrule
920     \end{tabular}
921     \label{tab:dampedSSDRF}
922     \end{table}
923    
924 chrisfen 2980 The properties shown in table \ref{tab:dampedSSDRF} compare
925     surprisingly well. The average density shows a modest increase when
926     using damped electrostatics in place of the reaction field. This comes
927     about because we neglect the pressure effect due to the surroundings
928     outside of the cuttoff, instead relying on screening effects to
929     neutralize electrostatic interactions at long distances. The $C_p$
930     also shows a slight increase, indicating greater fluctuation in the
931     enthalpy at constant pressure. The only other differences between the
932     damped and reaction field results are the dipole reorientational time
933     constants, $\tau_1$ and $\tau_2$. When using damped electrostatics,
934     the water molecules relax more quickly and are almost identical to the
935     experimental values. These results indicate that not only is it
936     reasonable to use damped electrostatics with SSD/RF, it is recommended
937     if capturing realistic dynamics is of primary importance. This is an
938     encouraging result because of the more varied applicability of damping
939     over the reaction field technique. Rather than be limited to
940     homogeneous systems, SSD/RF can be used effectively with mixed
941 chrisfen 2981 systems, such as dissolved ions, dissolved organic molecules, or even
942 chrisfen 2980 proteins.
943    
944 chrisfen 2977 In addition to the properties tabulated in table
945 chrisfen 2980 \ref{tab:dampedSSDRF}, we calculated the static dielectric constant
946 chrisfen 2981 from a 5~ns simulation of SSD/RF using the damped electrostatics. The
947 chrisfen 2977 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 chrisfen 2980 the visual contours in the figure. This highlights the qualitative
952     nature of contour plotting.
953 chrisfen 2977
954 chrisfen 2986 \section{Tetrahedrally Restructured Elongated Dipole (TRED) Water Model}\label{sec:tredWater}
955 chrisfen 2977
956 chrisfen 2981 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 chrisfen 2980
967 chrisfen 2981 \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 chrisfen 2978 \begin{table}
977 chrisfen 2981 \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 chrisfen 2986 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 chrisfen 2981 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     Being that TRED is a two-point water model, we expect its
1044     computational efficiency to fall some place in between the single and
1045     three-point water models. In comparative simulations, TRED was
1046     approximately 85\% slower than SSD/RF, while SPC/E was 225\% slower
1047     than SSD/RF. While TRED loses some of the performance advantage of
1048     SSD, it is still nearly twice as computationally efficient as SPC/E
1049     and TIP3P.
1050    
1051     \begin{table}
1052 chrisfen 2978 \caption{PROPERTIES OF TRED COMPARED WITH SSD/RF AND EXPERIMENT}
1053     \footnotesize
1054     \centering
1055     \begin{tabular}{ llccc }
1056     \toprule
1057     \toprule
1058     & & SSD/RF & TRED & Experiment [Ref.]\\
1059 chrisfen 2981 & & $\alpha = 0.2125$~\AA$^{-1}$ & $\alpha = 0.2125$~\AA$^{-1}$ & \\
1060 chrisfen 2978 \midrule
1061 chrisfen 2980 $\rho$ & (g cm$^{-3}$) & 1.004(4) & 0.995(5) & 0.997 \cite{CRC80}\\
1062     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 27(1) & 23(1) & 18.005 \cite{Wagner02} \\
1063 chrisfen 2978 $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.33(2) & 2.30(5) & 2.299 \cite{Mills73}\\
1064 chrisfen 2981 $n_C$ & & 4.2 & 5.3 & 4.7 \cite{Hura00}\\
1065 chrisfen 2978 $n_H$ & & 3.7 & 4.1 & 3.5 \cite{Soper86}\\
1066     $\tau_1$ & (ps) & 5.86(8) & 6.0(1) & 5.7 \cite{Eisenberg69}\\
1067     $\tau_2$ & (ps) & 2.45(7) & 2.49(5) & 2.3 \cite{Krynicki66}\\
1068 chrisfen 2980 $\epsilon_0$ & & 82.6(6) & 83(1) & 78.3 \cite{Malmberg56}\\
1069     $\tau_D$ & (ps) & 9.1(2) & 10.6(3) & 8.2(4) \cite{Kindt96}\\
1070 chrisfen 2978 \bottomrule
1071     \end{tabular}
1072 chrisfen 2981 \label{tab:tredProperties}
1073 chrisfen 2978 \end{table}
1074 chrisfen 2981 We calculated thermodynamic and dynamic properties in the same manner
1075     described in section \ref{sec:ssdrfDamped} for SSD/RF, and the results
1076     are presented in table \ref{tab:tredProperties}. These results show that
1077     TRED improves upon the $\rho$ and $C_p$ properties of SSD/RF with
1078     damped electrostatics while maintaining the excellent dynamics
1079     behavior of both the translational self-diffusivity and the
1080     reorientational time constants, $\tau_1$ and $\tau_2$. The structural
1081     results show some differences between the two models. Despite the
1082     improved peak shape for the first solvation shell (see figure
1083     \ref{fig:tredGofR}), $n_C$ and $n_H$ counts increase because of the
1084     further first minimum distance locations. This results in the
1085     integration extending over a larger water volume. If we integrate to
1086 chrisfen 2986 the first minimum value of the experimental $g_\textrm{OO}(r)$
1087     (3.42~\AA ) in both the SSD/RF and TRED plots, the $n_C$ values for
1088     both are much closer to experiment (4.7 for SSD/RF and 4.9 for TRED).
1089 chrisfen 2978
1090 chrisfen 2981 \begin{figure}
1091     \centering
1092     \includegraphics[width=3in]{./figures/tredDielectric.pdf}
1093     \caption{Contour map of the dielectric constant for TRED as a function
1094     of damping parameter and cutoff radius. The dielectric behavior for TRED
1095     is very similar to SSD/RF (see figure \ref{fig:dielectricMap}D), which
1096     is to be expected due to the similar dipole moment and sticky interaction
1097     strength.}
1098     \label{fig:tredDielectric}
1099     \end{figure}
1100     A comparison of the dielectric behavior was also included at the
1101     bottom of table \ref{tab:tredProperties}. The static dielectric
1102     constant results for SSD/RF and TRED are identical within error. This
1103     is not surprising given the similar dipole moment, similar sticky
1104 chrisfen 2986 interaction strength, and identical applied damping
1105     constant. Comparing the static dielectric constant contour map (figure
1106     \ref{fig:tredDielectric}) with the dielectric map for SSD/RF (figure
1107     \ref{fig:dielectricMap}D) highlights the similarities in how these
1108     models respond to dielectric damping and how the dipolar and monopolar
1109     electrostatic damping act in an equivalent fashion. Both these
1110     dielectric maps span a larger range than the 3, 4, and 5 point-charge
1111     water models; however, the SSD/RF range is greater than TRED,
1112     indicating that multipoles are a little more sensitive to damping than
1113     monopoles.
1114 chrisfen 2981
1115 chrisfen 2986 The final dielectric comparison comes through the Debye relaxation
1116     time ($\tau_D$) or the collective dipolar relaxation time when
1117     assuming a Debye treatment for the dielectric
1118     relaxation.\cite{Chandra99,Kindt96} This value is calculated through
1119     equation (\ref{eq:reorientCorr}) applied to the total system dipole
1120     moment. The values for both of the models are a slower than the
1121     experimental relaxation; however, they compare compare very well to
1122     experiment considering the Debye relaxation times calculated for the
1123     original SSD (11.95~ps) and the SPC/E (6.95~ps) and TIP3P (6.1~ps)
1124     values. The $\tau_D$ for TRED is about 1.5~ps slower than the $\tau_D$
1125     for SSD/RF, most likely due to the slower decay of the charge-charge
1126     interaction, even when screened by the same damping constant.
1127    
1128 chrisfen 2977 \section{Conclusions}
1129    
1130     In the above sections, the density maximum and temperature dependence
1131     of the self-diffusion constant were studied for the SSD water model,
1132     both with and without the use of reaction field, via a series of {\it
1133     NPT} and {\it NVE} simulations. The constant pressure simulations
1134 chrisfen 2981 showed a density maximum near 260~K. In most cases, the calculated
1135 chrisfen 2977 densities were significantly lower than the densities obtained from
1136     other water models (and experiment). Analysis of self-diffusion showed
1137     SSD to capture the transport properties of water well in both the
1138     liquid and supercooled liquid regimes.
1139    
1140     In order to correct the density behavior, we reparametrized the
1141     original SSD model for use both with and without a reaction field
1142     (SSD/RF and SSD/E), and made comparisons with SSD1, an alternate
1143     density corrected version of SSD. Both models improve the liquid
1144     structure, densities, and diffusive properties under their respective
1145     simulation conditions, indicating the necessity of reparametrization
1146     when changing the method of calculating long-range electrostatic
1147     interactions.
1148    
1149 chrisfen 2986 We also showed that SSD/RF performs well under the alternative damped
1150     electrostatic conditions, validating the multipolar damping work in
1151     the previous chapter. To improve the modeling of water when {\sc sf},
1152     the TRED water model was developed. This model maintains improves upon
1153     the thermodynamic properties of SSD/RF using damped electrostatics
1154     while maintaining the exceptional depiction of water dynamics.
1155    
1156 chrisfen 2981 The simple water models investigated here are excellent choices for
1157     representing explicit water in large scale simulations of biochemical
1158     systems. They are more computationally efficient than the common
1159     charge based water models, and, in many cases, exhibit more realistic
1160     bulk phase fluid properties. These models are one of the few cases in
1161     which maximizing efficiency does not result in a loss in realistic
1162 chrisfen 2986 liquid water representation.