ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Water.tex
Revision: 3028
Committed: Tue Sep 26 23:15:24 2006 UTC (17 years, 11 months ago) by chrisfen
Content type: application/x-tex
File size: 60972 byte(s)
Log Message:
down to proofing the appendix...

File Contents

# User Rev Content
1 chrisfen 2987 \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 chrisfen 3028 Thus, the motion and behavior of solute molecules in biochemical
9     simulations are highly dependent on the properties of the water model
10     that is chosen as the solvent.
11 chrisfen 2987 \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 chrisfen 3023 As discussed in the previous chapter, water is typically modeled with
23 chrisfen 2987 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 chrisfen 3023 main criteria for choosing a water model are:
35 chrisfen 2987
36 chrisfen 3023 \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 chrisfen 2987 \section{Soft Sticky Dipole Model for Water}
45    
46 chrisfen 3023 One recently-developed model that largely succeeds in retaining the
47 chrisfen 2987 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 chrisfen 3028 is also added. This sticky potential directs the molecules to assume
57 chrisfen 2987 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 chrisfen 3028 reaction field (RF) technique (and later, the {\sc sf} method), the
140     correction techniques discussed in the previous chapter, or even a
141     simple cutoff.\cite{Onsager36,Fennell06} This chapter addresses these
142     issues by looking at the structural and transport behavior of SSD over
143     a variety of temperatures with the purpose of utilizing the RF
144 chrisfen 3023 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 chrisfen 3028 the SSD water model have suggested adjustments to address abnormal
148     density behavior (also observed here), calling the corrected model
149     SSD1.\cite{Tan03} In the later sections of this chapter, we compare
150     our modified variants of SSD with both the original SSD and SSD1
151     models and discuss how our changes improve the depiction of water.
152 chrisfen 2987
153 chrisfen 3001 \section{Simulation Methods}{\label{sec:waterSimMethods}
154 chrisfen 2987
155     Most of the calculations in this particular study were performed using
156 chrisfen 3023 an internally developed simulation code that was one of the precursors
157 chrisfen 2987 of the {\sc oopse} molecular dynamics (MD) package.\cite{Meineke05}
158 chrisfen 3023 All of the capabilities of this code have been incorporated into {\sc
159     oopse}, and calculation results are consistent between the two
160     simulation packages. The later calculations involving the damped
161     shifted force ({\sc sf}) techniques were performed using {\sc oopse}.
162 chrisfen 2987
163     In the primary simulations of this study, long-range dipole-dipole
164 chrisfen 3023 corrections were accounted for by using either the reaction field
165     technique or a simple cubic switching function at the cutoff
166     radius. Interestingly, one of the early applications of a reaction
167     field was in Monte Carlo simulations of liquid water.\cite{Barker73}
168     In this method, the magnitude of the reaction field acting on dipole
169     $i$ is
170 chrisfen 2987 \begin{equation}
171     \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
172     \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
173     \label{eq:rfequation}
174     \end{equation}
175     where $\mathcal{R}$ is the cavity defined by the cutoff radius
176     ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
177     system, ${\bf\mu}_{j}$ is the dipole moment vector of particle $j$,
178     and $s(r_{ij})$ is a cubic switching function.\cite{Allen87} The
179 chrisfen 3023 reaction field contribution to the total energy from particle $i$ is
180 chrisfen 2987 given by $-\frac{1}{2}{\bf\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
181     on dipole $i$ by ${\bf\mu}_{i}\times\mathcal{E}_{i}$.\cite{Allen87} An
182     applied reaction field will alter the bulk orientational properties of
183     simulated water, and there is particular sensitivity of these
184     properties on changes in the length of the cutoff
185 chrisfen 3023 radius.\cite{vanderSpoel98} This behavior makes the reaction field a
186     less attractive method than the Ewald sum; however, for very large
187 chrisfen 3028 systems, the computational benefit of a reaction field is significant.
188 chrisfen 2987
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 chrisfen 3028 simulations in the isobaric-isothermal ($NPT$) ensemble. All
197 chrisfen 2987 dynamical properties for these models were then obtained from
198 chrisfen 3028 microcanonical ($NVE$) simulations done at densities matching the
199     $NPT$ density for a particular target temperature. The constant
200 chrisfen 2987 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 chrisfen 3028 this decision was that, in $NVE$ simulations, the energy drift
213 chrisfen 2987 when using quaternions was substantially greater than when using the
214 chrisfen 3028 {\sc dlm} method (figure \ref{fig:timeStepIntegration}). This steady
215 chrisfen 2987 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 $\Delta
224     t$ = 0.1~fs) for clarity.}
225     \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 chrisfen 3028 same degree of energy conservation.
237 chrisfen 2987
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     handle orientational motion. At time steps of 0.1 and 0.5~fs, both
243     methods for propagating the orientational degrees of freedom conserve
244     energy fairly well, with the quaternion method showing a slight energy
245     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     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     unnoticeable for time steps up to 3~fs. We observed a slight energy
253     drift on the order of 0.012~kcal/mol per nanosecond with a time step
254     of 4~fs. As expected, this drift increases dramatically with increasing
255 chrisfen 3028 time step. To insure accuracy in our $NVE$ simulations, time steps
256     were set at 2~fs and were also kept at this value for $NPT$
257 chrisfen 2987 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 chrisfen 3028 structure of ice I$_\textrm{h}$, the simulation boxes were
265 chrisfen 2987 orthorhombic in shape with an edge length ratio of approximately
266     1.00$\times$1.06$\times$1.23. We then allowed the particles to orient
267     freely about their fixed lattice positions with angular momenta values
268     randomly sampled at 400~K. The rotational temperature was then scaled
269     down in stages to slowly cool the crystals to 25~K. The particles were
270     then allowed to translate with fixed orientations at a constant
271 chrisfen 3023 pressure of 1~atm for 50~ps at 25~K. Finally, all constraints were
272 chrisfen 2987 removed and the ice crystals were allowed to equilibrate for 50~ps at
273 chrisfen 3023 25~K and a constant pressure of 1~atm. This procedure resulted in
274 chrisfen 3028 structurally stable ice I$_\textrm{h}$ crystals that obey the
275 chrisfen 2987 Bernal-Fowler rules.\cite{Bernal33,Rahman72} This method was also
276 chrisfen 3028 utilized in the making of diamond lattice ice I$_\textrm{c}$ crystals,
277 chrisfen 2987 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 chrisfen 3028 the $NPT$ ensemble. During melting simulations, the melting
286 chrisfen 2987 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     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     in that the fluctuations in the volume autocorrelation function damped
304     out in all of the simulations in under 20~ps.
305    
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     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     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     experiment.\cite{Jorgensen98b,Baez94,CRC80}. Note that using a
323     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 3023 sources.\cite{Jorgensen98b,Baez94,CRC80} Of the water models shown,
333     SSD has a temperature closest to the experimentally observed density
334     maximum. Of the {\it charge-based} models in figure
335 chrisfen 2987 \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 chrisfen 3028 \ref{sec:t5peDensity}).
340 chrisfen 2987
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     radius of 12~\AA\, complementing the 9~\AA\ $R_\textrm{c}$ used in the
346     previous SSD simulations. All of the resulting densities overlapped
347     within error and showed no significant trend toward lower or higher
348 chrisfen 3028 densities in simulations both with and without a reaction field.
349 chrisfen 2987
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 chrisfen 3023 models at the same pressure. This behavior is especially apparent at
354 chrisfen 2987 temperatures greater than 300~K. Lower than expected densities have
355     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 3028 density maximum location, down to 245~K, is also observed. This is a
368     more accurate comparison to the other listed water models, in that no
369     long-range corrections were applied in those
370 chrisfen 2987 simulations.\cite{Baez94,Jorgensen98b} However, even without the
371     reaction field, the density around 300~K is still significantly lower
372     than experiment and comparable water models. This anomalous behavior
373     was what lead Tan {\it et al.} to recently reparametrize
374 chrisfen 3028 SSD.\cite{Tan03} Throughout the remainder of this chapter our
375 chrisfen 2987 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 chrisfen 3028 biological membranes. In order to probe transport in bulk water,
383     $NVE$ simulations were performed at the average densities obtained
384     from the $NPT$ simulations at the target temperature. Simulations
385     started with randomized velocities and underwent 50~ps of temperature
386     scaling and 50~ps of constant energy equilibration before a 200~ps
387     data collection run. Diffusion constants were calculated via linear
388     fits to the long-time behavior of the mean-square displacement as a
389     function of time.\cite{Allen87} The average results from the five sets
390     of $NVE$ simulations are displayed in figure \ref{fig:ssdDiffuse},
391     alongside experimental, SPC/E, and TIP5P
392     results.\cite{Gillen72,Holz00,Baez94,Mahoney01}
393 chrisfen 2987
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     data.\cite{Baez94,Mahoney01,Gillen72,Holz00} Of the three water
400     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     reproducing values similar to experiment around 290~K; however, it
412     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 chrisfen 3028 water densities at these higher temperatures. When calculating the
418 chrisfen 2987 diffusion coefficients for SSD at experimental densities (instead of
419 chrisfen 3028 the densities from the $NPT$ simulations), the resulting values
420     fall more in line with experiment at the respective temperatures.
421 chrisfen 2987
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 chrisfen 3028 melting point, we can study the phase behavior of the liquid. The
427     constant pressure heat capacity ($C_\textrm{p}$) was monitored to
428     locate $T_\textrm{m}$ in each of the simulations. In the melting
429     simulations of the 1024 particle ice I$_\textrm{h}$ crystals, a large
430     spike in $C_\textrm{p}$ occurs at 245~K, indicating a first order
431     phase 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 2987 transition occurs at 235~K. These melting transitions are considerably
434     lower than the experimental value of 273~K, indicating that the solid
435     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     active reaction field. Note the large spike in $C_p$ around 245~K,
442     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     512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
452     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     \caption{ An illustration of angles involved in the correlations
462 chrisfen 3023 displayed in figure \ref{fig:contour}.}
463 chrisfen 2987 \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     \begin{equation}
470 chrisfen 3028 g_{\text{AB}}(r,\cos\theta) = \frac{V}{N_\text{A}N_\text{B}}\left\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\theta-\cos\theta_{ij})\delta(r-\left|{\bf r}_{ij}\right|)\right\rangle\ ,
471 chrisfen 2987 \end{equation}
472     \begin{equation}
473     g_{\text{AB}}(r,\cos\omega) =
474 chrisfen 3028 \frac{V}{N_\text{A}N_\text{B}}\left\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\omega-\cos\omega_{ij})\delta(r-\left|{\bf r}_{ij}\right|)\right\rangle\ ,
475 chrisfen 2987 \end{equation}
476     where $\theta$ and $\omega$ refer to the angles shown in figure
477     \ref{fig:corrAngle}. By binning over both distance and the cosine of the
478     desired angle between the two dipoles, the $g(r)$ can be analyzed to
479     determine the common dipole arrangements that constitute the peaks and
480     troughs in the standard one-dimensional $g(r)$ plots. Frames A and B
481     of figure \ref{fig:contour} show results from an ice I$_\textrm{c}$
482     simulation. The first peak in the $g(r)$ consists primarily of the
483     preferred hydrogen bonding arrangements as dictated by the tetrahedral
484     sticky potential - one peak for the hydrogen bond donor and the other
485     for the hydrogen bond acceptor. Due to the high degree of
486     crystallinity of the sample, the second and third solvation shells
487     show a repeated peak arrangement which decays at distances around the
488     fourth solvation shell, near the imposed cutoff for the Lennard-Jones
489     and dipole-dipole interactions. In the higher temperature simulation
490     shown in frames C and D, these long-range features deteriorate
491     rapidly. The first solvation shell still shows the strong effect of
492     the sticky-potential, although it covers a larger area, extending to
493     include a fraction of aligned dipole peaks within the first solvation
494 chrisfen 3028 shell. The latter peaks are lost due to thermal motion and as the
495 chrisfen 2987 competing dipole force overcomes the sticky potential's tight
496     tetrahedral structuring of the crystal.
497    
498     This complex interplay between dipole and sticky interactions was
499     remarked upon as a possible reason for the split second peak in the
500     oxygen-oxygen pair correlation function,
501     $g_\textrm{OO}(r)$.\cite{Liu96b} At low temperatures, the second
502     solvation shell peak appears to have two distinct components that
503     blend together to form one observable peak. At higher temperatures,
504     this split character alters to show the leading 4~\AA\ peak dominated
505     by equatorial anti-parallel dipole orientations. There is also a
506     tightly bunched group of axially arranged dipoles that most likely
507 chrisfen 3028 consist of a smaller fraction of aligned dipole pairs. The trailing
508 chrisfen 2987 component of the split peak at 5~\AA\ is dominated by aligned dipoles
509     that assume hydrogen bond arrangements similar to those seen in the
510     first solvation shell. This evidence indicates that the dipole pair
511     interaction begins to dominate outside of the range of the dipolar
512     repulsion term. The energetically favorable dipole arrangements
513     populate the region immediately outside this repulsion region (around
514     4~\AA), while arrangements that seek to satisfy both the sticky and
515     dipole forces locate themselves just beyond this initial buildup
516     (around 5~\AA).
517    
518     This analysis indicates that the split second peak is primarily the
519     product of the dipolar repulsion term of the sticky potential. In
520     fact, the inner peak can be pushed out and merged with the outer split
521     peak just by extending the switching function ($s^\prime(r_{ij})$)
522     from its normal 4~\AA\ cutoff to values of 4.5 or even 5~\AA. This
523     type of correction is not recommended for improving the liquid
524     structure, since the second solvation shell would still be shifted too
525     far out. In addition, this would have an even more detrimental effect
526     on the system densities, leading to a liquid with a more open
527     structure and a density considerably lower than the already low SSD
528     density. A better correction would be to include the
529     quadrupole-quadrupole interactions for the water particles outside of
530     the first solvation shell, but this would remove the simplicity and
531     speed advantage of SSD.
532    
533     \section{Adjusted Potentials: SSD/RF and SSD/E}
534    
535     The propensity of SSD to adopt lower than expected densities under
536     varying conditions is troubling, especially at higher temperatures. In
537     order to correct this model for use with a reaction field, it is
538     necessary to adjust the force field parameters for the primary
539     intermolecular interactions. In undergoing a reparametrization, it is
540     important not to focus on just one property and neglect the others. In
541     this case, it would be ideal to correct the densities while
542     maintaining the accurate transport behavior.
543    
544     The parameters available for tuning include the $\sigma$ and
545     $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
546     strength of the sticky potential ($\nu_0$), and the cutoff distances
547     for the sticky attractive and dipole repulsive cubic switching
548     function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
549     respectively). The results of the reparametrizations are shown in
550     table \ref{tab:ssdParams}. We are calling these reparametrizations the
551     Soft Sticky Dipole Reaction Field (SSD/RF - for use with a reaction
552     field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve
553     the liquid structure in simulations without a long-range correction).
554    
555     \begin{table}
556     \caption{PARAMETERS FOR THE ORIGINAL AND ADJUSTED SSD MODELS}
557    
558     \centering
559     \begin{tabular}{ llcccc }
560     \toprule
561     \toprule
562     Parameters & & SSD & SSD1 & SSD/E & SSD/RF \\
563     \midrule
564     $\sigma$ & (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
565     $\epsilon$ & (kcal mol$^{-1}$) & 0.152 & 0.152 & 0.152 & 0.152\\
566     $\mu$ & (D) & 2.35 & 2.35 & 2.42 & 2.48\\
567     $\nu_0$ & (kcal mol$^{-1}$) & 3.7284 & 3.6613 & 3.90 & 3.90\\
568     $\omega^\circ$ & & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
569     $r_l$ & (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
570     $r_u$ & (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
571     $r_l^\prime$ & (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
572     $r_u^\prime$ & (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
573     \bottomrule
574     \end{tabular}
575     \label{tab:ssdParams}
576     \end{table}
577    
578     \begin{figure}
579     \centering
580     \includegraphics[width=4.5in]{./figures/newGofRCompare.pdf}
581 chrisfen 3028 \caption{ Plots showing the experimental $g(r)$ (reference \cite{Hura00})
582 chrisfen 2987 with SSD/E and SSD1 without reaction field (top), as well as SSD/RF
583     and SSD1 with reaction field turned on (bottom). The changes in
584     parameters have lowered and broadened the first peak of SSD/E and
585     SSD/RF, resulting in a better fit to the first solvation shell.}
586     \label{fig:gofrCompare}
587     \end{figure}
588    
589     \begin{figure}
590     \centering
591     \includegraphics[width=\linewidth]{./figures/dualPotentials.pdf}
592     \caption{ Positive and negative isosurfaces of the sticky potential for
593     SSD and SSD1 (A) and SSD/E \& SSD/RF (B). Gold areas correspond to the
594     tetrahedral attractive component, and blue areas correspond to the
595     dipolar repulsive component.}
596     \label{fig:isosurface}
597     \end{figure}
598    
599     In the original paper detailing the development of SSD, Liu and Ichiye
600     placed particular emphasis on an accurate description of the first
601     solvation shell. This resulted in a somewhat tall and narrow first
602     peak in $g(r)$ that integrated to give similar coordination numbers to
603     the experimental data obtained by Soper and
604     Phillips.\cite{Liu96b,Soper86} New experimental x-ray scattering data
605     from Hura {\it et al.} indicates a slightly lower and shifted first
606     peak in the $g_\textrm{OO}(r)$, so our adjustments to SSD were made
607     after taking into consideration the new experimental
608     findings.\cite{Hura00} Figure \ref{fig:gofrCompare} shows the
609     relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
610     the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
611     experimental results. Both modified water models have shorter peaks
612     that match more closely to the experimental peak (as seen in the
613     insets of figure \ref{fig:gofrCompare}). This structural alteration
614     was accomplished by the combined reduction in the Lennard-Jones
615     $\sigma$ variable and adjustment of the sticky potential strength and
616     cutoffs. As can be seen in table \ref{tab:ssdParams}, the cutoffs for
617     the tetrahedral attractive and dipolar repulsive terms were nearly
618     swapped with each other. Isosurfaces of the original and modified
619     sticky potentials are shown in figure \ref{fig:isosurface}. In these
620     isosurfaces, it is easy to see how altering the cutoffs changes the
621     repulsive and attractive character of the particles. With a reduced
622     repulsive surface, the particles can move closer to one another,
623     increasing the density for the overall system. This change in
624     interaction cutoff also results in a more gradual orientational motion
625     by allowing the particles to maintain preferred dipolar arrangements
626     before they begin to feel the pull of the tetrahedral
627     restructuring. As the particles move closer together, the dipolar
628     repulsion term becomes active and excludes unphysical nearest-neighbor
629     arrangements. This compares with how SSD and SSD1 exclude preferred
630     dipole alignments before the particles feel the pull of the ``hydrogen
631     bonds''. Aside from improving the shape of the first peak in the
632     $g(r)$, this modification improves the densities considerably by
633     allowing the persistence of full dipolar character below the previous
634     4~\AA\ cutoff.
635    
636     While adjusting the location and shape of the first peak of $g(r)$
637     improves the densities, these changes alone are insufficient to bring
638 chrisfen 3028 the system densities up to the experimental values. To further
639     increase the densities, the dipole moments were increased in both of
640     our adjusted models. Since SSD is a dipole based model, the structure
641     and transport are very sensitive to changes in the dipole moment. The
642     original SSD used the dipole moment of the TIP3P water model, which at
643     2.35~D is significantly greater than the experimental gas phase value
644     of 1.84~D. The larger dipole moment is a more realistic value and
645     improves the dielectric properties of the fluid. Both theoretical and
646     experimental measurements indicate a liquid phase dipole moment
647     ranging from 2.4~D to values as high as 3.11~D, providing a
648     substantial range of reasonable values for a dipole
649     moment.\cite{Sprik91,Gubskaya02,Badyal00,Barriol64} Moderately
650 chrisfen 2987 increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF,
651     respectively, leads to significant changes in the density and
652     transport of the water models.
653    
654     \subsection{Density Behavior}
655    
656     In order to demonstrate the benefits of these reparametrizations, we
657 chrisfen 3028 performed a series of $NPT$ and $NVE$ simulations to probe the density
658     and transport properties of the adapted models and compared the
659 chrisfen 2987 results to the original SSD model. This comparison involved full {\it
660 chrisfen 3028 NPT} melting sequences for both SSD/E and SSD/RF, as well as $NVE$
661 chrisfen 2987 transport calculations at the calculated self-consistent
662     densities. Again, the results were obtained from five separate
663     simulations of 1024 particle systems, and the melting sequences were
664     started from different ice I$_\textrm{h}$ crystals constructed as
665 chrisfen 3028 described previously. Each $NPT$ simulation was equilibrated for
666 chrisfen 2987 100~ps before a 200~ps data collection run at each temperature step,
667     and the final configuration from the previous temperature simulation
668 chrisfen 3028 was used as a starting point. All $NVE$ simulations had the same
669 chrisfen 2987 thermalization, equilibration, and data collection times as stated
670     previously.
671    
672     \begin{figure}
673     \centering
674     \includegraphics[width=\linewidth]{./figures/ssdeDense.pdf}
675     \caption{ Comparison of densities calculated with SSD/E to
676     SSD1 without a reaction field, TIP3P, SPC/E, TIP5P, and
677     experiment.\cite{Jorgensen98b,Baez94,Mahoney00,CRC80} Both SSD1 and
678     SSD/E show good agreement with experiment when the long-range
679     correction is neglected.}
680     \label{fig:ssdeDense}
681     \end{figure}
682    
683     Figure \ref{fig:ssdeDense} shows the density profiles for SSD/E, SSD1,
684     TIP3P, TIP4P, and SPC/E alongside the experimental results. The
685     calculated densities for both SSD/E and SSD1 have increased
686     significantly over the original SSD model (see figure
687     \ref{fig:ssdDense}) and are in better agreement with the experimental
688     values. At 298~K, the densities of SSD/E and SSD1 without a long-range
689     correction are 0.996~g/cm$^3$ and 0.999~g/cm$^3$ respectively. These
690     both compare well with the experimental value of 0.997~g/cm$^3$, and
691     they are considerably better than the SSD value of 0.967~g/cm$^3$. The
692     changes to the dipole moment and sticky switching functions have
693     improved the structuring of the liquid (as seen in figure
694     \ref{fig:gofrCompare}), but they have shifted the density maximum to
695     much lower temperatures. This comes about via an increase in the
696     liquid disorder through the weakening of the sticky potential and
697     strengthening of the dipolar character. However, this increasing
698     disorder in the SSD/E model has little effect on the melting
699     transition. By monitoring $C_p$ throughout these simulations, we
700     observed a melting transition for SSD/E at 235~K, the same as SSD and
701     SSD1.
702    
703     \begin{figure}
704     \centering
705     \includegraphics[width=\linewidth]{./figures/ssdrfDense.pdf}
706     \caption{ Comparison of densities calculated with SSD/RF to
707     SSD1 with a reaction field, TIP3P, SPC/E, TIP5P, and
708     experiment.\cite{Jorgensen98b,Baez94,Mahoney00,CRC80} This plot
709     shows the benefit afforded by the reparametrization for use with a
710     reaction field correction - SSD/RF provides significantly more
711     accurate densities than SSD1 when performing room temperature
712     simulations.}
713     \label{fig:ssdrfDense}
714     \end{figure}
715    
716 chrisfen 3023 Including the reaction field long-range correction results is a more
717 chrisfen 2987 interesting comparison. A density profile including SSD/RF and SSD1
718     with an active reaction field is shown in figure \ref{fig:ssdrfDense}.
719     As observed in the simulations without a reaction field, the densities
720     of SSD/RF and SSD1 show a dramatic increase over normal SSD (see
721     figure \ref{fig:ssdDense}). At 298~K, SSD/RF has a density of 0.997
722     g/cm$^3$, directly in line with experiment and considerably better
723     than the original SSD value of 0.941~g/cm$^3$ and the SSD1 value of
724     0.972~g/cm$^3$. These results further emphasize the importance of
725     reparametrization in order to model the density properly under
726     different simulation conditions. Again, these changes have only a
727     minor effect on the melting point, which observed at 245~K for SSD/RF,
728     is identical to SSD and only 5~K lower than SSD1 with a reaction
729     field. Additionally, the difference in density maxima is not as
730     extreme, with SSD/RF showing a density maximum at 255~K, fairly close
731     to the density maxima of 260~K and 265~K, shown by SSD and SSD1
732     respectively.
733    
734     \subsection{Transport Behavior}
735    
736     \begin{figure}
737     \centering
738     \includegraphics[width=\linewidth]{./figures/ssdeDiffuse.pdf}
739     \caption{ The diffusion constants calculated from SSD/E and
740     SSD1 (both without a reaction field) along with experimental
741 chrisfen 3028 results.\cite{Gillen72,Holz00} The $NVE$ calculations were
742     performed at the average densities from the $NPT$ simulations for
743 chrisfen 2987 the respective models. SSD/E is slightly more mobile than experiment
744     at all of the temperatures, but it is closer to experiment at
745     biologically relevant temperatures than SSD1 without a long-range
746     correction.}
747     \label{fig:ssdeDiffuse}
748     \end{figure}
749    
750     The reparametrization of the SSD water model, both for use with and
751     without an applied long-range correction, brought the densities up to
752     what is expected for proper simulation of liquid water. In addition to
753     improving the densities, it is important that the diffusive behavior
754     of SSD be maintained or improved. Figure \ref{fig:ssdeDiffuse}
755     compares the temperature dependence of the diffusion constant of SSD/E
756     to SSD1 without an active reaction field at the densities calculated
757 chrisfen 3028 from their respective $NPT$ simulations at 1 atm. The diffusion
758 chrisfen 2987 constant for SSD/E is consistently higher than experiment, while SSD1
759     remains lower than experiment until relatively high temperatures
760     (around 360~K). Both models follow the shape of the experimental curve
761     below 300~K but tend to diffuse too rapidly at higher temperatures, as
762     seen in SSD1 crossing above 360~K. This increasing diffusion relative
763     to the experimental values is caused by the rapidly decreasing system
764     density with increasing temperature. Both SSD1 and SSD/E show this
765     deviation in particle mobility, but this trend has different
766     implications on the diffusive behavior of the models. While SSD1
767     shows more experimentally accurate diffusive behavior in the high
768     temperature regimes, SSD/E shows more accurate behavior in the
769     supercooled and biologically relevant temperature ranges. Thus, the
770     changes made to improve the liquid structure may have had an adverse
771     affect on the density maximum, but they improve the transport behavior
772     of SSD/E relative to SSD1 under the most commonly simulated
773     conditions.
774    
775     \begin{figure}
776     \centering
777     \includegraphics[width=\linewidth]{./figures/ssdrfDiffuse.pdf}
778     \caption{ The diffusion constants calculated from SSD/RF and
779     SSD1 (both with an active reaction field) along with experimental
780 chrisfen 3028 results.\cite{Gillen72,Holz00} The $NVE$ calculations were
781     performed at the average densities from the $NPT$ simulations for
782 chrisfen 2987 both of the models. SSD/RF captures the self-diffusion of water
783     throughout most of this temperature range. The increasing diffusion
784     constants at high temperatures for both models can be attributed to
785     lower calculated densities than those observed in experiment.}
786     \label{fig:ssdrfDiffuse}
787     \end{figure}
788    
789     In figure \ref{fig:ssdrfDiffuse}, the diffusion constants for SSD/RF are
790     compared to SSD1 with an active reaction field. Note that SSD/RF
791     tracks the experimental results quantitatively, identical within error
792     throughout most of the temperature range shown and exhibiting only a
793     slight increasing trend at higher temperatures. SSD1 tends to diffuse
794     more slowly at low temperatures and deviates to diffuse too rapidly at
795     temperatures greater than 330~K. As stated above, this deviation away
796     from the ideal trend is due to a rapid decrease in density at higher
797     temperatures. SSD/RF does not suffer from this problem as much as SSD1
798     because the calculated densities are closer to the experimental
799     values. These results again emphasize the importance of careful
800     reparametrization when using an altered long-range correction.
801    
802     \subsection{Summary of Liquid State Properties}
803    
804     \begin{table}
805     \caption{PROPERTIES OF THE SINGLE-POINT WATER MODELS COMPARED WITH
806     EXPERIMENTAL DATA AT AMBIENT CONDITIONS}
807     \footnotesize
808     \centering
809     \begin{tabular}{ llccccc }
810     \toprule
811     \toprule
812     & & SSD1 & SSD/E & SSD1 (RF) & SSD/RF & Experiment [Ref.] \\
813     \midrule
814     $\rho$ & (g cm$^{-3}$) & 0.999(1) & 0.996(1) & 0.972(2) & 0.997(1) & 0.997 \cite{CRC80}\\
815     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 28.80(11) & 25.45(9) & 28.28(6) & 23.83(16) & 18.005 \cite{Wagner02}\\
816     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 1.78(7) & 2.51(18) & 2.00(17) & 2.32(6) & 2.299 \cite{Mills73}\\
817     $n_C$ & & 3.9 & 4.3 & 3.8 & 4.4 & 4.7 \cite{Hura00}\\
818     $n_H$ & & 3.7 & 3.6 & 3.7 & 3.7 & 3.5 \cite{Soper86}\\
819     $\tau_1$ & (ps) & 10.9(6) & 7.3(4) & 7.5(7) & 7.2(4) & 5.7 \cite{Eisenberg69}\\
820     $\tau_2$ & (ps) & 4.7(4) & 3.1(2) & 3.5(3) & 3.2(2) & 2.3 \cite{Krynicki66}\\
821     \bottomrule
822     \end{tabular}
823     \label{tab:liquidProperties}
824     \end{table}
825    
826     Table \ref{tab:liquidProperties} gives a synopsis of the liquid state
827     properties of the water models compared in this study along with the
828     experimental values for liquid water at ambient conditions. The
829     coordination number ($n_C$) and number of hydrogen bonds per particle
830     ($n_H$) were calculated by integrating the following relations:
831     \begin{equation}
832     n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2g_{\textrm{OO}}(r)dr,
833     \end{equation}
834     \begin{equation}
835     n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2g_{\textrm{OH}}(r)dr,
836     \end{equation}
837     where $\rho$ is the number density of the specified pair interactions,
838 chrisfen 3028 and $a$ and $b$ are the radial locations of the minima following the
839     first peak in $g_\textrm{OO}(r)$ or $g_\textrm{OH}(r)$
840     respectively. The number of hydrogen bonds stays relatively constant
841     across all of the models, but the coordination numbers of SSD/E and
842     SSD/RF show an improvement over SSD1. This improvement is primarily
843     due to extension of the first solvation shell in the new parameter
844     sets. Because $n_H$ and $n_C$ are nearly identical in SSD1, it
845     appears that all molecules in the first solvation shell are involved
846     in hydrogen bonds. Since $n_H$ and $n_C$ differ in the newly
847     parameterized models, the orientations in the first solvation shell
848     are a bit more ``fluid''. Therefore SSD1 over-structures the first
849     solvation shell and our reparametrizations have returned this shell to
850     more realistic liquid-like behavior.
851 chrisfen 2987
852     The time constants for the orientational autocorrelation functions
853     are also displayed in Table \ref{tab:liquidProperties}. The dipolar
854     orientational time correlation functions ($C_{l}$) are described
855     by:
856     \begin{equation}
857     C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
858     \label{eq:reorientCorr}
859     \end{equation}
860     where $P_l$ are Legendre polynomials of order $l$ and
861     $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular
862     dipole.\cite{Rahman71} Note that this is identical to equation
863     (\ref{eq:OrientCorr}) were $\alpha$ is equal to $z$. From these
864     correlation functions, the orientational relaxation time of the dipole
865     vector can be calculated from an exponential fit in the long-time
866     regime ($t > \tau_l$).\cite{Rothschild84} Calculation of these time
867 chrisfen 3028 constants were averaged over five detailed $NVE$ simulations
868 chrisfen 2987 performed at the ambient conditions for each of the respective
869     models. It should be noted that the commonly cited value of 1.9~ps for
870 chrisfen 3028 $\tau_2$ was determined from the NMR data in reference \cite{Krynicki66} at
871 chrisfen 2987 a temperature near 34$^\circ$C.\cite{Rahman71} Because of the strong
872     temperature dependence of $\tau_2$, it is necessary to recalculate it
873     at 298~K to make proper comparisons. The value shown in Table
874     \ref{tab:liquidProperties} was calculated from the same NMR data in the
875 chrisfen 3028 fashion described in reference \cite{Krynicki66}. Similarly, $\tau_1$ was
876     recomputed for 298~K from the data in reference \cite{Eisenberg69}.
877 chrisfen 2987 Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
878     and without an active reaction field. Turning on the reaction field
879     leads to much improved time constants for SSD1; however, these results
880     also include a corresponding decrease in system density.
881     Orientational relaxation times published in the original SSD dynamics
882     paper are smaller than the values observed here, and this difference
883     can be attributed to the use of the Ewald sum.\cite{Chandra99}
884    
885     \subsection{SSD/RF and Damped Electrostatics}\label{sec:ssdrfDamped}
886    
887     In section \ref{sec:dampingMultipoles}, a method was described for
888 chrisfen 3028 applying the damped {\sc sf} or {\sc sp} techniques to systems
889 chrisfen 2987 containing point multipoles. The SSD family of water models is the
890     perfect test case because of the dipole-dipole (and
891     charge-dipole/quadrupole) interactions that are present. The {\sc sf}
892     and {\sc sp} techniques were presented as a pairwise replacement for
893     the Ewald summation. It has been suggested that models parametrized
894     for the Ewald summation (like TIP5P-E) would be appropriate for use
895     with a reaction field and vice versa.\cite{Rick04} Therefore, we
896     decided to test the SSD/RF water model with this damped electrostatic
897     technique in place of the reaction field to see how the calculated
898     properties change.
899    
900     \begin{table}
901     \caption{PROPERTIES OF SSD/RF WHEN USING DIFFERENT ELECTROSTATIC
902     CORRECTION METHODS}
903     \footnotesize
904     \centering
905     \begin{tabular}{ llccc }
906     \toprule
907     \toprule
908     & & Reaction Field & Damped Electrostatics & Experiment [Ref.] \\
909     & & $\epsilon = 80$ & $\alpha = 0.2125$~\AA$^{-1}$ & \\
910     \midrule
911     $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 \cite{CRC80}\\
912     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 \cite{Wagner02} \\
913     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 \cite{Mills73}\\
914     $n_C$ & & 4.4 & 4.2 & 4.7 \cite{Hura00}\\
915     $n_H$ & & 3.7 & 3.7 & 3.5 \cite{Soper86}\\
916     $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 \cite{Eisenberg69}\\
917     $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 \cite{Krynicki66}\\
918     \bottomrule
919     \end{tabular}
920     \label{tab:dampedSSDRF}
921     \end{table}
922    
923     The properties shown in table \ref{tab:dampedSSDRF} compare
924     surprisingly well. The average density shows a modest increase when
925     using damped electrostatics in place of the reaction field. This comes
926     about because we neglect the pressure effect due to the surroundings
927 chrisfen 3028 outside of the cutoff, instead relying on screening effects to
928 chrisfen 2987 neutralize electrostatic interactions at long distances. The $C_p$
929     also shows a slight increase, indicating greater fluctuation in the
930     enthalpy at constant pressure. The only other differences between the
931     damped and reaction field results are the dipole reorientational time
932     constants, $\tau_1$ and $\tau_2$. When using damped electrostatics,
933 chrisfen 3023 the water molecules relax more quickly and exhibit behavior very
934     similar to the experimental values. These results indicate that not
935     only is it reasonable to use damped electrostatics with SSD/RF, it is
936     recommended if capturing realistic dynamics is of primary
937     importance. This is an encouraging result because the damping methods
938     are more generally applicable than reaction field. Using damping,
939     SSD/RF can be used effectively with mixed systems, such as dissolved
940     ions, dissolved organic molecules, or even proteins.
941 chrisfen 2987
942     In addition to the properties tabulated in table
943     \ref{tab:dampedSSDRF}, we calculated the static dielectric constant
944     from a 5~ns simulation of SSD/RF using the damped electrostatics. The
945     resulting value of 82.6(6) compares very favorably with the
946     experimental value of 78.3.\cite{Malmberg56} This value is closer to
947     the experimental value than what was expected according to figure
948     \ref{fig:dielectricMap}, raising some questions as to the accuracy of
949     the visual contours in the figure. This highlights the qualitative
950     nature of contour plotting.
951    
952     \section{Tetrahedrally Restructured Elongated Dipole (TRED) Water Model}\label{sec:tredWater}
953    
954     The SSD/RF model works well with damped electrostatics, but because of
955     its point multipole character, there is no charge neutralization
956     correction at $R_\textrm{c}$. This has the effect of increasing the
957     density, since there is no consideration of the ``surroundings''. In
958     an attempt to optimize a water model for these conditions, we decided
959     to both simplify the parameters of the SSD type models and break the
960     point dipole into a charge pair so that it will gain some effect from
961     the shifting action in the {\sc sf} technique. This process resulted
962     in a two-point model that we are calling the Tetrahedrally
963     Restructured Elongated Dipole (TRED) water model.
964    
965     \begin{figure}
966     \centering
967     \includegraphics[width=2.75in]{./figures/tredLayout.pdf}
968     \caption{Geometry of TRED water. In this two-point model, the origin
969     is the center-of-mass of the water molecule with the same moments of
970     inertia as SSD/RF. The negatively charged site at the origin is also
971     both a Lennard-Jones and sticky interaction site.}
972     \label{fig:tredLayout}
973     \end{figure}
974     \begin{table}
975     \centering
976     \caption{PARAMETERS FOR THE TRED WATER MODEL}
977     \begin{tabular}{ llr }
978     \toprule
979     \toprule
980     $\sigma$ & (\AA) & 2.980 \\
981     $\epsilon$ & (kcal mol$^{-1}$) & 0.2045 \\
982     $q$ & ($e$) & 1.041 \\
983     $v_0$ & (kcal mol$^{-1}$) & 4.22 \\
984     $\omega^\circ$ & & 0.07715 \\
985     $r_l$ \& $r_l^\prime$ & (\AA) & 2.4 \\
986     $r_u$ \& $r_u^\prime$ & (\AA) & 4.0 \\
987     Q$_xx$ & (D \AA) & -1.682 \\
988     Q$_yy$ & (D \AA) & 1.762 \\
989     Q$_zz$ & (D \AA) & -0.080 \\
990     \bottomrule
991     \end{tabular}
992     \label{tab:tredParams}
993     \end{table}
994     \begin{figure}
995     \centering
996     \includegraphics[width=\linewidth]{./figures/tredGofR.pdf}
997     \caption{The $g_\textrm{OO}(r)$ for TRED water. The first peak has a
998     closer to the experimental plot; however, the second and third peaks
999     exhibit a more expanded structure similar to SSD/RF. The minimum
1000     following the first peak is located at 3.55~\AA , which is further out
1001     than both experiment and SSD/RF (3.42 and 3.3~\AA\ respectively),
1002     leading to a larger coordination number. If all the curves were
1003     integrated to the experimental minimum, the $n_C$ results would all be
1004     similar, with TRED having an $n_C$ only 0.2 larger than experiment.}
1005     \label{fig:tredGofR}
1006     \end{figure}
1007     The geometric layout of TRED water is shown in figure
1008     \ref{fig:tredLayout} and the electrostatic, Lennard-Jones, and sticky
1009     parameters are displayed in table \ref{tab:tredParams}. The negatively
1010     charged site is located at the center of mass of the molecule and is
1011     also home to the Lennard-Jones and sticky interactions. The charge
1012     separation distance of 0.5~\AA\ along the dipolar ($z$) axis combined
1013     with the charge magnitude ($q$) results in a dipole moment of
1014     2.5~D. This value is simply the value used for SSD/RF (2.48~D) rounded
1015     to the tenths place. We also unified the sticky parameters for the
1016     switching functions on the repulsive and attractive interactions in
1017     the interest of simplicity, and we left the quadrupole moment elements
1018     and $\omega^\circ$ unaltered. It should be noted that additional logic
1019     needs to be included into the electrostatic code when using TRED to
1020     insure that the charges of each water do not interact with the other
1021     water's quadrupole moment. Finally, the strength of the sticky
1022     interaction ($v_0$) was modified to improve the shape of the first
1023     peaks in $g_\textrm{OO}(r)$ and $g_\textrm{OH}(r)$, while the $\sigma$
1024     and $\epsilon$ values were varied to adjust the location of the first
1025     peak of $g_\textrm{OO}(r)$ (see figure \ref{fig:tredGofR}) and the
1026 chrisfen 3023 density. The $\sigma$ and $\epsilon$ optimization was carried out by
1027 chrisfen 2987 separating the Lennard-Jones potential into near entirely repulsive
1028     ($A$) and attractive ($C$) parts, such that
1029     \begin{equation}
1030     \sigma = \left(\frac{A}{C}\right)^\frac{1}{6},
1031     \end{equation}
1032     and
1033     \begin{equation}
1034     \epsilon = \frac{C^2}{4A}.
1035     \end{equation}
1036     The location of the peak is adjusted through $A$, while the
1037     interaction strength is modified through $C$. All of the above changes
1038     were made with the goal of capturing the experimental density and
1039     translational diffusion constant at 298~K and 1~atm.
1040    
1041 chrisfen 3023 Since TRED is a two-point water model, we expect its computational
1042     efficiency to fall some place in between the one-point and three-point
1043     water models. In comparative simulations, TRED was approximately 85\%
1044     slower than SSD/RF, while SPC/E was 225\% slower than SSD/RF. While
1045     TRED loses some of the performance advantage of SSD, it is still
1046     nearly twice as computationally efficient as SPC/E and TIP3P.
1047 chrisfen 2987
1048     \begin{table}
1049     \caption{PROPERTIES OF TRED COMPARED WITH SSD/RF AND EXPERIMENT}
1050     \footnotesize
1051     \centering
1052     \begin{tabular}{ llccc }
1053     \toprule
1054     \toprule
1055     & & SSD/RF & TRED & Experiment [Ref.]\\
1056     & & $\alpha = 0.2125$~\AA$^{-1}$ & $\alpha = 0.2125$~\AA$^{-1}$ & \\
1057     \midrule
1058     $\rho$ & (g cm$^{-3}$) & 1.004(4) & 0.995(5) & 0.997 \cite{CRC80}\\
1059     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 27(1) & 23(1) & 18.005 \cite{Wagner02} \\
1060     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.33(2) & 2.30(5) & 2.299 \cite{Mills73}\\
1061     $n_C$ & & 4.2 & 5.3 & 4.7 \cite{Hura00}\\
1062     $n_H$ & & 3.7 & 4.1 & 3.5 \cite{Soper86}\\
1063     $\tau_1$ & (ps) & 5.86(8) & 6.0(1) & 5.7 \cite{Eisenberg69}\\
1064     $\tau_2$ & (ps) & 2.45(7) & 2.49(5) & 2.3 \cite{Krynicki66}\\
1065     $\epsilon_0$ & & 82.6(6) & 83(1) & 78.3 \cite{Malmberg56}\\
1066     $\tau_D$ & (ps) & 9.1(2) & 10.6(3) & 8.2(4) \cite{Kindt96}\\
1067     \bottomrule
1068     \end{tabular}
1069     \label{tab:tredProperties}
1070     \end{table}
1071     We calculated thermodynamic and dynamic properties in the same manner
1072     described in section \ref{sec:ssdrfDamped} for SSD/RF, and the results
1073     are presented in table \ref{tab:tredProperties}. These results show that
1074     TRED improves upon the $\rho$ and $C_p$ properties of SSD/RF with
1075     damped electrostatics while maintaining the excellent dynamics
1076     behavior of both the translational self-diffusivity and the
1077     reorientational time constants, $\tau_1$ and $\tau_2$. The structural
1078     results show some differences between the two models. Despite the
1079     improved peak shape for the first solvation shell (see figure
1080     \ref{fig:tredGofR}), $n_C$ and $n_H$ counts increase because of the
1081     further first minimum distance locations. This results in the
1082     integration extending over a larger water volume. If we integrate to
1083     the first minimum value of the experimental $g_\textrm{OO}(r)$
1084     (3.42~\AA ) in both the SSD/RF and TRED plots, the $n_C$ values for
1085     both are much closer to experiment (4.7 for SSD/RF and 4.9 for TRED).
1086    
1087     \begin{figure}
1088     \centering
1089     \includegraphics[width=3in]{./figures/tredDielectric.pdf}
1090     \caption{Contour map of the dielectric constant for TRED as a function
1091     of damping parameter and cutoff radius. The dielectric behavior for TRED
1092     is very similar to SSD/RF (see figure \ref{fig:dielectricMap}D), which
1093     is to be expected due to the similar dipole moment and sticky interaction
1094     strength.}
1095     \label{fig:tredDielectric}
1096     \end{figure}
1097     A comparison of the dielectric behavior was also included at the
1098     bottom of table \ref{tab:tredProperties}. The static dielectric
1099     constant results for SSD/RF and TRED are identical within error. This
1100     is not surprising given the similar dipole moment, similar sticky
1101     interaction strength, and identical applied damping
1102     constant. Comparing the static dielectric constant contour map (figure
1103     \ref{fig:tredDielectric}) with the dielectric map for SSD/RF (figure
1104     \ref{fig:dielectricMap}D) highlights the similarities in how these
1105     models respond to dielectric damping and how the dipolar and monopolar
1106     electrostatic damping act in an equivalent fashion. Both these
1107 chrisfen 3028 dielectric maps span a larger range than the three-, four-, and
1108     five-point water models; however, the SSD/RF range is greater than
1109     TRED, indicating that multipoles are more sensitive to damping than
1110     monopoles.
1111 chrisfen 2987
1112     The final dielectric comparison comes through the Debye relaxation
1113     time ($\tau_D$) or the collective dipolar relaxation time when
1114     assuming a Debye treatment for the dielectric
1115     relaxation.\cite{Chandra99,Kindt96} This value is calculated through
1116     equation (\ref{eq:reorientCorr}) applied to the total system dipole
1117 chrisfen 3028 moment. The values for both of the models are slower than the
1118 chrisfen 2987 experimental relaxation; however, they compare compare very well to
1119     experiment considering the Debye relaxation times calculated for the
1120     original SSD (11.95~ps) and the SPC/E (6.95~ps) and TIP3P (6.1~ps)
1121     values. The $\tau_D$ for TRED is about 1.5~ps slower than the $\tau_D$
1122     for SSD/RF, most likely due to the slower decay of the charge-charge
1123     interaction, even when screened by the same damping constant.
1124    
1125     \section{Conclusions}
1126    
1127     In the above sections, the density maximum and temperature dependence
1128     of the self-diffusion constant were studied for the SSD water model,
1129 chrisfen 3028 both with and without the use of reaction field, via a series of $NPT$
1130     and $NVE$ simulations. The constant pressure simulations showed a
1131     density maximum near 260~K. In most cases, the calculated densities
1132     were significantly lower than the densities obtained from other water
1133     models (and experiment). Analysis of self-diffusion showed SSD to
1134     capture the transport properties of water well in both the liquid and
1135     supercooled liquid regimes.
1136 chrisfen 2987
1137     In order to correct the density behavior, we reparametrized the
1138     original SSD model for use both with and without a reaction field
1139     (SSD/RF and SSD/E), and made comparisons with SSD1, an alternate
1140     density corrected version of SSD. Both models improve the liquid
1141     structure, densities, and diffusive properties under their respective
1142     simulation conditions, indicating the necessity of reparametrization
1143     when changing the method of calculating long-range electrostatic
1144     interactions.
1145    
1146     We also showed that SSD/RF performs well under the alternative damped
1147     electrostatic conditions, validating the multipolar damping work in
1148 chrisfen 3028 the previous chapter. To improve the modeling of water with the {\sc
1149     sf} technique, the TRED water model was developed. This model improves
1150     upon the thermodynamic properties of SSD/RF using damped
1151     electrostatics while maintaining the exceptional depiction of water
1152     dynamics.
1153 chrisfen 2987
1154     The simple water models investigated here are excellent choices for
1155 chrisfen 3023 representing water in large scale simulations of biochemical
1156 chrisfen 2987 systems. They are more computationally efficient than the common
1157 chrisfen 3023 charge based water models, and often exhibit more realistic bulk phase
1158     fluid properties. These models are one of the few cases in which
1159     maximizing efficiency does not result in a loss in realistic
1160     representation.