1 |
\chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION \\ TECHNIQUES} |
2 |
|
3 |
In molecular simulations, proper accumulation of electrostatic |
4 |
interactions is essential and one of the most |
5 |
computationally-demanding tasks. The common molecular mechanics force |
6 |
fields represent atomic sites with full or partial charges protected |
7 |
by repulsive Lennard-Jones interactions. This means that nearly every |
8 |
pair interaction involves a calculation of charge-charge forces. |
9 |
Coupled with relatively long-ranged $r^{-1}$ decay, the monopole |
10 |
interactions quickly become the most expensive part of molecular |
11 |
simulations. Historically, the electrostatic pair interaction would |
12 |
not have decayed appreciably within the typical box lengths that could |
13 |
be feasibly simulated. In the larger systems that are more typical of |
14 |
modern simulations, large cutoffs should be used to incorporate |
15 |
electrostatics correctly. |
16 |
|
17 |
There have been many efforts to address the proper and practical |
18 |
handling of electrostatic interactions, and these have resulted in a |
19 |
variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are |
20 |
typically classified as implicit methods (i.e., continuum dielectrics, |
21 |
static dipolar fields),\cite{Born20,Grossfield00} explicit methods |
22 |
(i.e., Ewald summations, interaction shifting or |
23 |
truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e., |
24 |
reaction field type methods, fast multipole |
25 |
methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are |
26 |
often preferred because they physically incorporate solvent molecules |
27 |
in the system of interest, but these methods are sometimes difficult |
28 |
to utilize because of their high computational cost.\cite{Roux99} In |
29 |
addition to the computational cost, there have been some questions |
30 |
regarding possible artifacts caused by the inherent periodicity of the |
31 |
explicit Ewald summation.\cite{Tobias01} |
32 |
|
33 |
In this chapter, we focus on a new set of pairwise methods devised by |
34 |
Wolf {\it et al.},\cite{Wolf99} which we further extend. These |
35 |
methods, along with a few other mixed methods (i.e. reaction field), |
36 |
are compared with the smooth particle mesh Ewald |
37 |
sum,\cite{Onsager36,Essmann99} which is our reference method for |
38 |
handling long-range electrostatic interactions. The new methods for |
39 |
handling electrostatics have the potential to scale linearly with |
40 |
increasing system size, since they involve only a simple modification |
41 |
to the direct pairwise sum. They also lack the added periodicity of |
42 |
the Ewald sum, so they can be used for systems which are non-periodic |
43 |
or which have one- or two-dimensional periodicity. Below, these |
44 |
methods are evaluated using a variety of model systems to establish |
45 |
their usability in molecular simulations. |
46 |
|
47 |
\section{The Ewald Sum} |
48 |
|
49 |
The complete accumulation of the electrostatic interactions in a system with |
50 |
periodic boundary conditions (PBC) requires the consideration of the |
51 |
effect of all charges within a (cubic) simulation box as well as those |
52 |
in the periodic replicas, |
53 |
\begin{equation} |
54 |
V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime |
55 |
\left[ \sum_{i=1}^N\sum_{j=1}^N \phi |
56 |
\left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right) |
57 |
\right], |
58 |
\label{eq:PBCSum} |
59 |
\end{equation} |
60 |
where the sum over $\mathbf{n}$ is a sum over all periodic box |
61 |
replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the |
62 |
prime indicates $i = j$ are neglected for $\mathbf{n} = |
63 |
0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic |
64 |
particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is |
65 |
the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and |
66 |
$j$, and $\phi$ is the solution to Poisson's equation |
67 |
($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for |
68 |
charge-charge interactions). In the case of monopole electrostatics, |
69 |
equation (\ref{eq:PBCSum}) is conditionally convergent and is divergent for |
70 |
non-neutral systems. |
71 |
|
72 |
The electrostatic summation problem was originally studied by Ewald |
73 |
for the case of an infinite crystal.\cite{Ewald21}. The approach he |
74 |
took was to convert this conditionally convergent sum into two |
75 |
absolutely convergent summations: a short-ranged real-space summation |
76 |
and a long-ranged reciprocal-space summation, |
77 |
\begin{equation} |
78 |
\begin{split} |
79 |
V_\textrm{elec} = \frac{1}{2}& |
80 |
\sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime |
81 |
\frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)} |
82 |
{|\mathbf{r}_{ij}+\mathbf{n}|} \\ |
83 |
&+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2} |
84 |
\exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right) |
85 |
\cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\ |
86 |
&- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2 |
87 |
+ \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3} |
88 |
\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2, |
89 |
\end{split} |
90 |
\label{eq:EwaldSum} |
91 |
\end{equation} |
92 |
where $\alpha$ is the damping or convergence parameter with units of |
93 |
\AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to |
94 |
$2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric |
95 |
constant of the surrounding medium. The final two terms of equation |
96 |
(\ref{eq:EwaldSum}) are a particle-self term and a dipolar term for |
97 |
interacting with a surrounding dielectric.\cite{Allen87} This dipolar |
98 |
term was neglected in early applications of this technique in |
99 |
molecular simulations,\cite{Brush66,Woodcock71} until it was |
100 |
introduced by de Leeuw {\it et al.} to address situations where the |
101 |
unit cell has a dipole moment which is magnified through replication |
102 |
of the periodic images.\cite{deLeeuw80,Smith81} If this term is taken |
103 |
to be zero, the system is said to be using conducting (or |
104 |
``tin-foil'') boundary conditions, $\epsilon_{\rm S} = \infty$. |
105 |
|
106 |
\begin{figure} |
107 |
\includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf} |
108 |
\caption{The change in the need for the Ewald sum with |
109 |
increasing computational power. A:~Initially, only small systems |
110 |
could be studied, and the Ewald sum replicated the simulation box to |
111 |
convergence. B:~Now, radial cutoff methods should be able to reach |
112 |
convergence for the larger systems of charges that are common today.} |
113 |
\label{fig:ewaldTime} |
114 |
\end{figure} |
115 |
Figure \ref{fig:ewaldTime} shows how the Ewald sum has been applied |
116 |
over time. Initially, due to the small system sizes that could be |
117 |
simulated feasibly, the entire simulation box was replicated to |
118 |
convergence. In more modern simulations, the systems have grown large |
119 |
enough that a real-space cutoff could potentially give convergent |
120 |
behavior. Indeed, it has been observed that with the choice of a |
121 |
small $\alpha$, the reciprocal-space portion of the Ewald sum can be |
122 |
rapidly convergent and small relative to the real-space |
123 |
portion.\cite{Karasawa89,Kolafa92} |
124 |
|
125 |
The original Ewald summation is an $\mathcal{O}(N^2)$ algorithm. The |
126 |
convergence parameter $(\alpha)$ plays an important role in balancing |
127 |
the computational cost between the direct and reciprocal-space |
128 |
portions of the summation. The choice of this value allows one to |
129 |
select whether the real-space or reciprocal space portion of the |
130 |
summation is an $\mathcal{O}(N^2)$ calculation (with the other being |
131 |
$\mathcal{O}(N)$).\cite{Sagui99} With the appropriate choice of |
132 |
$\alpha$ and thoughtful algorithm development, this cost can be |
133 |
reduced to $\mathcal{O}(N^{3/2})$.\cite{Perram88} The typical route |
134 |
taken to reduce the cost of the Ewald summation even further is to set |
135 |
$\alpha$ such that the real-space interactions decay rapidly, allowing |
136 |
for a short spherical cutoff. Then the reciprocal space summation is |
137 |
optimized. These optimizations usually involve utilization of the |
138 |
fast Fourier transform (FFT),\cite{Hockney81} leading to the |
139 |
particle-particle particle-mesh (P3M) and particle mesh Ewald (PME) |
140 |
methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these |
141 |
methods, the cost of the reciprocal-space portion of the Ewald |
142 |
summation is reduced from $\mathcal{O}(N^2)$ down to $\mathcal{O}(N |
143 |
\log N)$. |
144 |
|
145 |
These developments and optimizations have made the use of the Ewald |
146 |
summation routine in simulations with periodic boundary |
147 |
conditions. However, in certain systems, such as vapor-liquid |
148 |
interfaces and membranes, the intrinsic three-dimensional periodicity |
149 |
can prove problematic. The Ewald sum has been reformulated to handle |
150 |
2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these |
151 |
methods are computationally expensive.\cite{Spohr97,Yeh99} More |
152 |
recently, there have been several successful efforts toward reducing |
153 |
the computational cost of 2-D lattice |
154 |
summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} |
155 |
bringing them more in line with the cost of the full 3-D summation. |
156 |
|
157 |
Several studies have recognized that the inherent periodicity in the |
158 |
Ewald sum can have an effect not just on reduced dimensionality |
159 |
system, but on three-dimensional systems as |
160 |
well.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00} |
161 |
As an example, solvated proteins are essentially kept at high |
162 |
concentration due to the periodicity of the electrostatic summation |
163 |
method. In these systems, the more compact folded states of a protein |
164 |
can be artificially stabilized by the periodic replicas introduced by |
165 |
the Ewald summation.\cite{Weber00} Thus, care must be taken when |
166 |
considering the use of the Ewald summation where the assumed |
167 |
periodicity would introduce spurious effects. |
168 |
|
169 |
|
170 |
\section{The Wolf and Zahn Methods} |
171 |
|
172 |
In a recent paper by Wolf \textit{et al.}, a procedure was outlined |
173 |
for the accurate accumulation of electrostatic interactions in an |
174 |
efficient pairwise fashion. This procedure lacks the inherent |
175 |
periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.} |
176 |
observed that the electrostatic interaction is effectively |
177 |
short-ranged in condensed phase systems and that neutralization of the |
178 |
charges contained within the cutoff radius is crucial for potential |
179 |
stability. They devised a pairwise summation method that ensures |
180 |
charge neutrality and gives results similar to those obtained with the |
181 |
Ewald summation. The resulting shifted Coulomb potential includes |
182 |
image-charges subtracted out through placement on the cutoff sphere |
183 |
and a distance-dependent damping function (identical to that seen in |
184 |
the real-space portion of the Ewald sum) to aid convergence: |
185 |
\begin{equation} |
186 |
V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}} |
187 |
- \lim_{r_{ij}\rightarrow R_\textrm{c}} |
188 |
\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}. |
189 |
\label{eq:WolfPot} |
190 |
\end{equation} |
191 |
Equation (\ref{eq:WolfPot}) is essentially the common form of a shifted |
192 |
potential. However, neutralizing the charge contained within each |
193 |
cutoff sphere requires the placement of a self-image charge on the |
194 |
surface of the cutoff sphere. This additional self-term in the total |
195 |
potential enabled Wolf {\it et al.} to obtain excellent estimates of |
196 |
Madelung energies for many crystals. |
197 |
|
198 |
In order to use their charge-neutralized potential in molecular |
199 |
dynamics simulations, Wolf \textit{et al.} suggested taking the |
200 |
derivative of this potential prior to evaluation of the limit. This |
201 |
procedure gives an expression for the forces, |
202 |
\begin{equation} |
203 |
\begin{split} |
204 |
F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{& |
205 |
\Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}} |
206 |
+ \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}} |
207 |
\Biggr]\\ |
208 |
&-\Biggr[ |
209 |
\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2} |
210 |
+ \frac{2\alpha}{\pi^{1/2}} |
211 |
\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}} |
212 |
\Biggr]\Biggr\}, |
213 |
\end{split} |
214 |
\label{eq:WolfForces} |
215 |
\end{equation} |
216 |
that incorporates both image charges and damping of the electrostatic |
217 |
interaction. |
218 |
|
219 |
More recently, Zahn \textit{et al.} investigated these potential and |
220 |
force expressions for use in simulations involving water.\cite{Zahn02} |
221 |
In their work, they pointed out that the forces and derivative of |
222 |
the potential are not commensurate. Attempts to use both |
223 |
equations (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead |
224 |
to poor energy conservation. They correctly observed that taking the |
225 |
limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the |
226 |
derivatives gives forces for a different potential energy function |
227 |
than the one shown in equation (\ref{eq:WolfPot}). |
228 |
|
229 |
Zahn \textit{et al.} introduced a modified form of this summation |
230 |
method as a way to use the technique in Molecular Dynamics |
231 |
simulations. They proposed a new damped Coulomb potential, |
232 |
\begin{equation} |
233 |
\begin{split} |
234 |
V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{& |
235 |
\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\ |
236 |
&- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2} |
237 |
+ \frac{2\alpha}{\pi^{1/2}} |
238 |
\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}} |
239 |
\right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\}, |
240 |
\end{split} |
241 |
\label{eq:ZahnPot} |
242 |
\end{equation} |
243 |
and showed that this potential does fairly well at capturing the |
244 |
structural and dynamic properties of water compared with the same |
245 |
properties obtained using the Ewald sum. |
246 |
|
247 |
\section{Simple Forms for Pairwise Electrostatics}\label{sec:PairwiseDerivation} |
248 |
|
249 |
The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et |
250 |
al.} are constructed using two different (and separable) computational |
251 |
tricks: |
252 |
|
253 |
\begin{enumerate}[itemsep=0pt] |
254 |
\item shifting through the use of image charges, and |
255 |
\item damping the electrostatic interaction. |
256 |
\end{enumerate} |
257 |
Wolf \textit{et al.} treated the development of their summation method |
258 |
as a progressive application of these techniques,\cite{Wolf99} while |
259 |
Zahn \textit{et al.} founded their damped Coulomb modification |
260 |
(Eq. (\ref{eq:ZahnPot})) on the post-limit forces |
261 |
(Eq. (\ref{eq:WolfForces})) which were derived using both techniques. |
262 |
It is possible, however, to separate these tricks and study their |
263 |
effects independently. |
264 |
|
265 |
Starting with the original observation that the effective range of the |
266 |
electrostatic interaction in condensed phases is considerably less |
267 |
than $r^{-1}$, either the cutoff sphere neutralization or the |
268 |
distance-dependent damping technique could be used as a foundation for |
269 |
a new pairwise summation method. Wolf \textit{et al.} made the |
270 |
observation that charge neutralization within the cutoff sphere plays |
271 |
a significant role in energy convergence; therefore we will begin our |
272 |
analysis with the various shifted forms that maintain this charge |
273 |
neutralization. We can evaluate the methods of Wolf {\it et al.} and |
274 |
Zahn {\it et al.} by considering the standard shifted potential, |
275 |
\begin{equation} |
276 |
V_\textrm{SP}(r) = \begin{cases} |
277 |
v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > |
278 |
R_\textrm{c} |
279 |
\end{cases}, |
280 |
\label{eq:shiftingPotForm} |
281 |
\end{equation} |
282 |
and shifted force, |
283 |
\begin{equation} |
284 |
V_\textrm{SF}(r) = \begin{cases} |
285 |
v(r) - v_\textrm{c} |
286 |
- \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c}) |
287 |
&\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c} |
288 |
\end{cases}, |
289 |
\label{eq:shiftingForm} |
290 |
\end{equation} |
291 |
functions where $v(r)$ is the unshifted form of the potential, and |
292 |
$v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures |
293 |
that both the potential and the forces goes to zero at the cutoff |
294 |
radius, while the Shifted Potential ({\sc sp}) form only ensures the |
295 |
potential is smooth at the cutoff radius |
296 |
($R_\textrm{c}$).\cite{Allen87} |
297 |
|
298 |
The forces associated with the shifted potential are simply the forces |
299 |
of the unshifted potential itself (when inside the cutoff sphere), |
300 |
\begin{equation} |
301 |
F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right), |
302 |
\end{equation} |
303 |
and are zero outside. Inside the cutoff sphere, the forces associated |
304 |
with the shifted force form can be written, |
305 |
\begin{equation} |
306 |
F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d |
307 |
v(r)}{dr} \right)_{r=R_\textrm{c}}. |
308 |
\end{equation} |
309 |
|
310 |
If the potential, $v(r)$, is taken to be the normal Coulomb potential, |
311 |
\begin{equation} |
312 |
v(r) = \frac{q_i q_j}{r}, |
313 |
\label{eq:Coulomb} |
314 |
\end{equation} |
315 |
then the {\sc sp} form will give Wolf {\it et al.}'s undamped |
316 |
prescription: |
317 |
\begin{equation} |
318 |
V_\textrm{SP}(r) = |
319 |
q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad |
320 |
r\leqslant R_\textrm{c}, |
321 |
\label{eq:SPPot} |
322 |
\end{equation} |
323 |
with associated forces, |
324 |
\begin{equation} |
325 |
F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) |
326 |
\quad r\leqslant R_\textrm{c}. |
327 |
\label{eq:SPForces} |
328 |
\end{equation} |
329 |
These forces are identical to the forces of the standard Coulomb |
330 |
interaction, and cutting these off at $R_c$ was addressed by Wolf |
331 |
\textit{et al.} as undesirable. They pointed out that the effect of |
332 |
the image charges is neglected in the forces when this form is |
333 |
used,\cite{Wolf99} thereby eliminating any benefit from the method in |
334 |
molecular dynamics. Additionally, there is a discontinuity in the |
335 |
forces at the cutoff radius which results in energy drift during MD |
336 |
simulations. |
337 |
|
338 |
The {\sc sf} form using the normal Coulomb potential will give, |
339 |
\begin{equation} |
340 |
V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}} |
341 |
+ \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] |
342 |
\quad r\leqslant R_\textrm{c}. |
343 |
\label{eq:SFPot} |
344 |
\end{equation} |
345 |
with associated forces, |
346 |
\begin{equation} |
347 |
F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) |
348 |
\quad r\leqslant R_\textrm{c}. |
349 |
\label{eq:SFForces} |
350 |
\end{equation} |
351 |
This formulation has the benefits that there are no discontinuities at |
352 |
the cutoff radius and the neutralizing image charges are present in |
353 |
both the energy and force expressions. It would be simple to add the |
354 |
self-neutralizing term back when computing the total energy of the |
355 |
system, thereby maintaining the agreement with the Madelung energies. |
356 |
A side effect of this treatment is the alteration in the shape of the |
357 |
potential that comes from the derivative term. Thus, a degree of |
358 |
clarity about agreement with the empirical potential is lost in order |
359 |
to gain functionality in dynamics simulations. |
360 |
|
361 |
Wolf \textit{et al.} originally discussed the energetics of the |
362 |
shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was |
363 |
insufficient for accurate determination of the energy with reasonable |
364 |
cutoff distances. The calculated Madelung energies fluctuated wildly |
365 |
around the expected value, but as the cutoff radius was increased, the |
366 |
oscillations converged toward the correct value.\cite{Wolf99} A |
367 |
damping function was incorporated to accelerate this convergence; and |
368 |
though alternative forms for the damping function could be |
369 |
used,\cite{Jones56,Heyes81} the complimentary error function was |
370 |
chosen to mirror the effective screening used in the Ewald summation. |
371 |
Incorporating this error function damping into the simple Coulomb |
372 |
potential, |
373 |
\begin{equation} |
374 |
v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r}, |
375 |
\label{eq:dampCoulomb} |
376 |
\end{equation} |
377 |
the {\sc sp} potential function (Eq. (\ref{eq:SPPot})) becomes |
378 |
\begin{equation} |
379 |
V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r} |
380 |
- \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) |
381 |
\quad r\leqslant R_\textrm{c}, |
382 |
\label{eq:DSPPot} |
383 |
\end{equation} |
384 |
with associated forces, |
385 |
\begin{equation} |
386 |
F_{\textrm{DSP}}(r) = q_iq_j |
387 |
\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2} |
388 |
+ \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) |
389 |
\quad r\leqslant R_\textrm{c}. |
390 |
\label{eq:DSPForces} |
391 |
\end{equation} |
392 |
Again, this damped shifted potential suffers from a discontinuity in |
393 |
the forces at the cutoff radius, and the image charges play no role in |
394 |
the forces. To remedy these concerns, one may derive a {\sc sf} |
395 |
variant by including the derivative term present in |
396 |
equation~(\ref{eq:shiftingForm}), |
397 |
\begin{equation} |
398 |
\begin{split} |
399 |
V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}& |
400 |
\frac{\mathrm{erfc}\left(\alpha r\right)}{r} |
401 |
- \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ |
402 |
&+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2} |
403 |
+ \frac{2\alpha}{\pi^{1/2}} |
404 |
\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}} |
405 |
\right)\left(r-R_\mathrm{c}\right)\ \Biggr{]} |
406 |
\quad r\leqslant R_\textrm{c}. |
407 |
\label{eq:DSFPot} |
408 |
\end{split} |
409 |
\end{equation} |
410 |
The derivative of the above potential will lead to the following forces, |
411 |
\begin{equation} |
412 |
\begin{split} |
413 |
F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}& |
414 |
\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2} |
415 |
+ \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ |
416 |
&- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)} |
417 |
{R_{\textrm{c}}^2} |
418 |
+ \frac{2\alpha}{\pi^{1/2}} |
419 |
\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}} |
420 |
\right)\Biggr{]} |
421 |
\quad r\leqslant R_\textrm{c}. |
422 |
\label{eq:DSFForces} |
423 |
\end{split} |
424 |
\end{equation} |
425 |
If the damping parameter $(\alpha)$ is set to zero, the undamped case, |
426 |
equations (\ref{eq:SPPot}) through (\ref{eq:SFForces}) are correctly |
427 |
recovered from equations (\ref{eq:DSPPot}) through (\ref{eq:DSFForces}). |
428 |
|
429 |
This new {\sc sf} potential is similar to equation (\ref{eq:ZahnPot}) |
430 |
derived by Zahn \textit{et al.}; however, there are two important |
431 |
differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from equation |
432 |
(\ref{eq:shiftingForm}) is equal to equation (\ref{eq:dampCoulomb}) |
433 |
with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present |
434 |
in the Zahn potential, resulting in a potential discontinuity as |
435 |
particles cross $R_\textrm{c}$. Second, the sign of the derivative |
436 |
portion is different. The missing $v_\textrm{c}$ term would not |
437 |
affect molecular dynamics simulations (although the computed energy |
438 |
would be expected to have sudden jumps as particle distances crossed |
439 |
$R_c$); however, the sign problem is a potential source of errors. In |
440 |
fact, equation~(\ref{eq:ZahnPot}) introduces a discontinuity in the |
441 |
forces at the cutoff, because the force function is shifted in the |
442 |
wrong direction and does not cross zero at $R_\textrm{c}$. |
443 |
|
444 |
Equations (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an |
445 |
electrostatic summation method in which the potential and forces are |
446 |
continuous at the cutoff radius and which incorporates the damping |
447 |
function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of |
448 |
this chapter, we will evaluate exactly how good these methods ({\sc |
449 |
sp}, {\sc sf}, damping) are at reproducing the correct electrostatic |
450 |
summation performed by the Ewald sum. |
451 |
|
452 |
|
453 |
\section{Evaluating Pairwise Summation Techniques} |
454 |
|
455 |
As mentioned in the introduction, there are two primary techniques |
456 |
utilized to obtain information about the system of interest in |
457 |
classical molecular mechanics simulations: Monte Carlo (MC) and |
458 |
molecular dynamics (MD). Both of these techniques utilize pairwise |
459 |
summations of interactions between particle sites, but they use these |
460 |
summations in different ways. |
461 |
|
462 |
In MC, the potential energy difference between configurations dictates |
463 |
the progression of MC sampling. Going back to the origins of this |
464 |
method, the acceptance criterion for the canonical ensemble laid out |
465 |
by Metropolis \textit{et al.} states that a subsequent configuration |
466 |
is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where |
467 |
$\xi$ is a random number between 0 and 1.\cite{Metropolis53} |
468 |
Maintaining the correct $\Delta E$ when using an alternate method for |
469 |
handling the long-range electrostatics will ensure proper sampling |
470 |
from the ensemble. |
471 |
|
472 |
In MD, the derivative of the potential governs how the system will |
473 |
progress in time. Consequently, the force and torque vectors on each |
474 |
body in the system dictate how the system evolves. If the magnitude |
475 |
and direction of these vectors are similar when using alternate |
476 |
electrostatic summation techniques, the dynamics in the short term |
477 |
will be indistinguishable. Because error in MD calculations is |
478 |
cumulative, one should expect greater deviation at longer times, |
479 |
and methods which have large differences in the force and torque |
480 |
vectors will diverge from each other more rapidly. |
481 |
|
482 |
\subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods} |
483 |
|
484 |
\begin{figure} |
485 |
\centering |
486 |
\includegraphics[width = 3.5in]{./figures/dualLinear.pdf} |
487 |
\caption{Example least squares regressions of the configuration energy |
488 |
differences for SPC/E water systems. The upper plot shows a data set |
489 |
with a poor correlation coefficient ($R^2$), while the lower plot |
490 |
shows a data set with a good correlation coefficient.} |
491 |
\label{fig:linearFit} |
492 |
\end{figure} |
493 |
The pairwise summation techniques (outlined in section |
494 |
\ref{sec:ESMethods}) were evaluated for use in MC simulations by |
495 |
studying the energy differences between conformations. We took the |
496 |
{\sc spme}-computed energy difference between two conformations to be the |
497 |
correct behavior. An ideal performance by an alternative method would |
498 |
reproduce these energy differences exactly (even if the absolute |
499 |
energies calculated by the methods are different). Since none of the |
500 |
methods provide exact energy differences, we used linear least squares |
501 |
regressions of energy gap data to evaluate how closely the methods |
502 |
mimicked the Ewald energy gaps. Unitary results for both the |
503 |
correlation (slope) and correlation coefficient for these regressions |
504 |
indicate perfect agreement between the alternative method and {\sc spme}. |
505 |
Sample correlation plots for two alternate methods are shown in |
506 |
figure \ref{fig:linearFit}. |
507 |
|
508 |
Each of the seven system types (detailed in section \ref{sec:RepSims}) |
509 |
were represented using 500 independent configurations. Thus, each of |
510 |
the alternative (non-Ewald) electrostatic summation methods was |
511 |
evaluated using an accumulated 873,250 configurational energy |
512 |
differences. Results for and discussions regarding the individual |
513 |
analysis of each of the system types appear in appendix |
514 |
\ref{app:IndividualResults}, while the cumulative results over all the |
515 |
investigated systems appear below in section~\ref{sec:EnergyResults}. |
516 |
|
517 |
\subsection{Molecular Dynamics and the Force and Torque |
518 |
Vectors}\label{sec:MDMethods} We evaluated the pairwise methods |
519 |
(outlined in section \ref{sec:ESMethods}) for use in MD simulations by |
520 |
comparing the force and torque vectors with those obtained using the |
521 |
reference Ewald summation ({\sc spme}). Both the magnitude and the |
522 |
direction of these vectors on each of the bodies in the system were |
523 |
analyzed. For the magnitude of these vectors, linear least squares |
524 |
regression analyses were performed as described previously for |
525 |
comparing $\Delta E$ values. Instead of a single energy difference |
526 |
between two system configurations, we compared the magnitudes of the |
527 |
forces (and torques) on each molecule in each configuration. For a |
528 |
system of 1000 water molecules and 40 ions, there are 1040 force and |
529 |
1000 torque vectors. With 500 configurations, this results in 520,000 |
530 |
force and 500,000 torque vector comparisons. Additionally, data from |
531 |
seven different system types was aggregated before comparisons were |
532 |
made. |
533 |
|
534 |
The {\it directionality} of the force and torque vectors was |
535 |
investigated through measurement of the angle ($\theta$) formed |
536 |
between those computed from the particular method and those from {\sc spme}, |
537 |
\begin{equation} |
538 |
\theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} |
539 |
\cdot \hat{F}_\textrm{M}\right), |
540 |
\end{equation} |
541 |
where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force |
542 |
vector computed using method M. Each of these $\theta$ values was |
543 |
accumulated in a distribution function and weighted by the area on the |
544 |
unit sphere. Since this distribution is a measure of angular error |
545 |
between two different electrostatic summation methods, there is no |
546 |
{\it a priori} reason for the profile to adhere to any specific |
547 |
shape. Thus, Gaussian fits were used to measure the width of the |
548 |
resulting distributions. The variance ($\sigma^2$) was extracted from |
549 |
each of these fits and was used to compare distribution widths. |
550 |
Values of $\sigma^2$ near zero indicate vector directions |
551 |
indistinguishable from those calculated when using the reference |
552 |
method ({\sc spme}). |
553 |
|
554 |
\subsection{Short-time Dynamics} |
555 |
|
556 |
The effects of the alternative electrostatic summation methods on the |
557 |
short-time dynamics of charged systems were evaluated by considering a |
558 |
NaCl crystal at a temperature of 1000~K. A subset of the best |
559 |
performing pairwise methods was used in this comparison. The NaCl |
560 |
crystal was chosen to avoid possible complications from the treatment |
561 |
of orientational motion in molecular systems. All systems were |
562 |
started with the same initial positions and velocities. Simulations |
563 |
were performed under the microcanonical ensemble, and velocity |
564 |
autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each |
565 |
of the trajectories, |
566 |
\begin{equation} |
567 |
C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}. |
568 |
\label{eq:vCorr} |
569 |
\end{equation} |
570 |
Velocity autocorrelation functions require detailed short time data, |
571 |
thus velocity information was saved every 2~fs over 10~ps |
572 |
trajectories. Because the NaCl crystal is composed of two different |
573 |
atom types, the average of the two resulting velocity autocorrelation |
574 |
functions was used for comparisons. |
575 |
|
576 |
\subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods} |
577 |
|
578 |
The effects of the same subset of alternative electrostatic methods on |
579 |
the {\it long-time} dynamics of charged systems were evaluated using |
580 |
the same model system (NaCl crystals at 1000K). The power spectrum |
581 |
($I(\omega)$) was obtained via Fourier transform of the velocity |
582 |
autocorrelation function, |
583 |
\begin{equation} I(\omega) = |
584 |
\frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt, |
585 |
\label{eq:powerSpec} |
586 |
\end{equation} |
587 |
where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the |
588 |
NaCl crystal is composed of two different atom types, the average of |
589 |
the two resulting power spectra was used for comparisons. Simulations |
590 |
were performed under the microcanonical ensemble, and velocity |
591 |
information was saved every 5~fs over 100~ps trajectories. |
592 |
|
593 |
\subsection{Representative Simulations}\label{sec:RepSims} |
594 |
A variety of representative molecular simulations were analyzed to |
595 |
determine the relative effectiveness of the pairwise summation |
596 |
techniques in reproducing the energetics and dynamics exhibited by |
597 |
{\sc spme}. We wanted to span the space of typical molecular |
598 |
simulations (i.e. from liquids of neutral molecules to ionic |
599 |
crystals), so the systems studied were: |
600 |
|
601 |
\begin{enumerate}[itemsep=0pt] |
602 |
\item liquid water (SPC/E),\cite{Berendsen87} |
603 |
\item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E), |
604 |
\item NaCl crystals, |
605 |
\item NaCl melts, |
606 |
\item a low ionic strength solution of NaCl in water (0.11 M), |
607 |
\item a high ionic strength solution of NaCl in water (1.1 M), and |
608 |
\item a 6~\AA\ radius sphere of Argon in water. |
609 |
\end{enumerate} |
610 |
By utilizing the pairwise techniques (outlined in section |
611 |
\ref{sec:ESMethods}) in systems composed entirely of neutral groups, |
612 |
charged particles, and mixtures of the two, we hope to discern under |
613 |
which conditions it will be possible to use one of the alternative |
614 |
summation methodologies instead of the Ewald sum. |
615 |
|
616 |
For the solid and liquid water configurations, configurations were |
617 |
taken at regular intervals from high temperature trajectories of 1000 |
618 |
SPC/E water molecules. Each configuration was equilibrated |
619 |
independently at a lower temperature (300~K for the liquid, 200~K for |
620 |
the crystal). The solid and liquid NaCl systems consisted of 500 |
621 |
$\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for |
622 |
these systems were selected and equilibrated in the same manner as the |
623 |
water systems. In order to introduce measurable fluctuations in the |
624 |
configuration energy differences, the crystalline simulations were |
625 |
equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid |
626 |
NaCl configurations needed to represent a fully disordered array of |
627 |
point charges, so the high temperature of 7000~K was selected for |
628 |
equilibration. The ionic solutions were made by solvating 4 (or 40) |
629 |
ions in a periodic box containing 1000 SPC/E water molecules. Ion and |
630 |
water positions were then randomly swapped, and the resulting |
631 |
configurations were again equilibrated individually. Finally, for the |
632 |
Argon / Water ``charge void'' systems, the identities of all the SPC/E |
633 |
waters within 6~\AA\ of the center of the equilibrated water |
634 |
configurations were converted to argon. |
635 |
|
636 |
These procedures guaranteed us a set of representative configurations |
637 |
from chemically-relevant systems sampled from appropriate |
638 |
ensembles. Force field parameters for the ions and Argon were taken |
639 |
from the force field utilized by {\sc oopse}.\cite{Meineke05} |
640 |
|
641 |
\subsection{Comparison of Summation Methods}\label{sec:ESMethods} |
642 |
We compared the following alternative summation methods with results |
643 |
from the reference method ({\sc spme}): |
644 |
|
645 |
\begin{enumerate}[itemsep=0pt] |
646 |
\item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2, |
647 |
and 0.3~\AA$^{-1}$, |
648 |
\item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2, |
649 |
and 0.3~\AA$^{-1}$, |
650 |
\item reaction field with an infinite dielectric constant, and |
651 |
\item an unmodified cutoff. |
652 |
\end{enumerate} |
653 |
|
654 |
Group-based cutoffs with a fifth-order polynomial switching function |
655 |
were utilized for the reaction field simulations. Additionally, we |
656 |
investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure |
657 |
cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker} |
658 |
implementation of {\sc spme},\cite{Ponder87} while all other calculations |
659 |
were performed using the {\sc oopse} molecular mechanics |
660 |
package.\cite{Meineke05} All other portions of the energy calculation |
661 |
(i.e. Lennard-Jones interactions) were handled in exactly the same |
662 |
manner across all systems and configurations. |
663 |
|
664 |
The alternative methods were also evaluated with three different |
665 |
cutoff radii (9, 12, and 15~\AA). As noted previously, the |
666 |
convergence parameter ($\alpha$) plays a role in the balance of the |
667 |
real-space and reciprocal-space portions of the Ewald calculation. |
668 |
Typical molecular mechanics packages set this to a value dependent on |
669 |
the cutoff radius and a tolerance (typically less than $1 \times |
670 |
10^{-4}$~kcal/mol). Smaller tolerances are typically associated with |
671 |
increasing accuracy at the expense of computational time spent on the |
672 |
reciprocal-space portion of the summation.\cite{Perram88,Essmann95} |
673 |
The default {\sc tinker} tolerance of $1 \times 10^{-8}$~kcal/mol was used |
674 |
in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200, |
675 |
0.3119, and 0.2476~\AA$^{-1}$ for cutoff radii of 9, 12, and 15~\AA\ |
676 |
respectively. |
677 |
|
678 |
\section{Configuration Energy Difference Results}\label{sec:EnergyResults} |
679 |
In order to evaluate the performance of the pairwise electrostatic |
680 |
summation methods for Monte Carlo (MC) simulations, the energy |
681 |
differences between configurations were compared to the values |
682 |
obtained when using {\sc spme}. The results for the combined |
683 |
regression analysis of all of the systems are shown in figure |
684 |
\ref{fig:delE}. |
685 |
|
686 |
\begin{figure} |
687 |
\centering |
688 |
\includegraphics[width=4.75in]{./figures/delEplot.pdf} |
689 |
\caption{Statistical analysis of the quality of configurational energy |
690 |
differences for a given electrostatic method compared with the |
691 |
reference Ewald sum. Results with a value equal to 1 (dashed line) |
692 |
indicate $\Delta E$ values indistinguishable from those obtained using |
693 |
{\sc spme}. Different values of the cutoff radius are indicated with |
694 |
different symbols (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ = |
695 |
inverted triangles).} |
696 |
\label{fig:delE} |
697 |
\end{figure} |
698 |
The most striking feature of this plot is how well the Shifted Force |
699 |
({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy |
700 |
differences. For the undamped {\sc sf} method, and the |
701 |
moderately-damped {\sc sp} methods, the results are nearly |
702 |
indistinguishable from the Ewald results. The other common methods do |
703 |
significantly less well. |
704 |
|
705 |
The unmodified cutoff method is essentially unusable. This is not |
706 |
surprising since hard cutoffs give large energy fluctuations as atoms |
707 |
or molecules move in and out of the cutoff |
708 |
radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to |
709 |
some degree by using group based cutoffs with a switching |
710 |
function.\cite{Adams79,Steinbach94,Leach01} However, we do not see |
711 |
significant improvement using the group-switched cutoff because the |
712 |
salt and salt solution systems contain non-neutral groups. Appendix |
713 |
\ref{app:IndividualResults} includes results for systems comprised |
714 |
entirely of neutral groups. |
715 |
|
716 |
For the {\sc sp} method, inclusion of electrostatic damping improves |
717 |
the agreement with Ewald, and using an $\alpha$ of 0.2~\AA $^{-1}$ |
718 |
shows an excellent correlation and quality of fit with the {\sc spme} |
719 |
results, particularly with a cutoff radius greater than 12~\AA . Use |
720 |
of a larger damping parameter is more helpful for the shortest cutoff |
721 |
shown, but it has a detrimental effect on simulations with larger |
722 |
cutoffs. |
723 |
|
724 |
In the {\sc sf} sets, increasing damping results in progressively {\it |
725 |
worse} correlation with Ewald. Overall, the undamped case is the best |
726 |
performing set, as the correlation and quality of fits are |
727 |
consistently superior regardless of the cutoff distance. The undamped |
728 |
case is also less computationally demanding (because no evaluation of |
729 |
the complementary error function is required). |
730 |
|
731 |
The reaction field results illustrates some of that method's |
732 |
limitations, primarily that it was developed for use in homogeneous |
733 |
systems. It does, however, provide results that are an improvement |
734 |
over those from an unmodified cutoff. |
735 |
|
736 |
\section{Magnitude of the Force and Torque Vector Results}\label{sec:FTMagResults} |
737 |
|
738 |
Evaluation of pairwise methods for use in Molecular Dynamics |
739 |
simulations requires consideration of effects on the forces and |
740 |
torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the |
741 |
regression results for the force and torque vector magnitudes, |
742 |
respectively. The data in these figures was generated from an |
743 |
accumulation of the statistics from all of the system types. |
744 |
|
745 |
\begin{figure} |
746 |
\centering |
747 |
\includegraphics[width=4.75in]{./figures/frcMagplot.pdf} |
748 |
\caption{Statistical analysis of the quality of the force vector |
749 |
magnitudes for a given electrostatic method compared with the |
750 |
reference Ewald sum. Results with a value equal to 1 (dashed line) |
751 |
indicate force magnitude values indistinguishable from those obtained |
752 |
using {\sc spme}. Different values of the cutoff radius are indicated with |
753 |
different symbols (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ = |
754 |
inverted triangles).} |
755 |
\label{fig:frcMag} |
756 |
\end{figure} |
757 |
Again, it is striking how well the {\sc sp} and {\sc sf} methods |
758 |
reproduce the {\sc spme} forces. The undamped and weakly-damped {\sc |
759 |
sf} method gives the best agreement with Ewald. This is perhaps |
760 |
expected because this method explicitly incorporates a smooth |
761 |
transition in the forces at the cutoff radius as well as the |
762 |
neutralizing image charges. |
763 |
|
764 |
Figure \ref{fig:frcMag}, for the most part, parallels the results seen |
765 |
in the previous $\Delta E$ section. The unmodified cutoff results are |
766 |
poor, but using group based cutoffs and a switching function provides |
767 |
an improvement much more significant than what was seen with $\Delta |
768 |
E$. |
769 |
|
770 |
With moderate damping and a large enough cutoff radius, the {\sc sp} |
771 |
method is generating usable forces. Further increases in damping, |
772 |
while beneficial for simulations with a cutoff radius of 9~\AA\ , are |
773 |
detrimental to simulations with larger cutoff radii. |
774 |
|
775 |
The reaction field results are surprisingly good, considering the poor |
776 |
quality of the fits for the $\Delta E$ results. There is still a |
777 |
considerable degree of scatter in the data, but in general, the forces |
778 |
correlate well with the Ewald forces. We note that the pure NaCl |
779 |
systems were not included in the system set used in the reaction field |
780 |
calculations, so these results are partly biased towards conditions in |
781 |
which the method performs more favorably. |
782 |
|
783 |
\begin{figure} |
784 |
\centering |
785 |
\includegraphics[width=4.75in]{./figures/trqMagplot.pdf} |
786 |
\caption{Statistical analysis of the quality of the torque vector |
787 |
magnitudes for a given electrostatic method compared with the |
788 |
reference Ewald sum. Results with a value equal to 1 (dashed line) |
789 |
indicate torque magnitude values indistinguishable from those obtained |
790 |
using {\sc spme}. Different values of the cutoff radius are indicated with |
791 |
different symbols (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ = |
792 |
inverted triangles).} |
793 |
\label{fig:trqMag} |
794 |
\end{figure} |
795 |
Molecular torques were only available from the systems which contained |
796 |
rigid molecules (i.e. the systems containing water). The data in |
797 |
figure \ref{fig:trqMag} is taken from this smaller sampling pool. |
798 |
|
799 |
Torques appear to be much more sensitive to charge interactions at |
800 |
longer distances. The most noticeable feature in comparing the new |
801 |
electrostatic methods with {\sc spme} is how much the agreement |
802 |
improves with increasing cutoff radius. Again, the weakly damped and |
803 |
undamped {\sc sf} method appears to reproduce the {\sc spme} torques |
804 |
most accurately. |
805 |
|
806 |
Water molecules are dipolar, and the reaction field method reproduces |
807 |
the effect of the surrounding polarized medium on each of the |
808 |
molecular bodies. Therefore it is not surprising that reaction field |
809 |
performs best of all of the methods on molecular torques. |
810 |
|
811 |
\section{Directionality of the Force and Torque Vector Results}\label{sec:FTDirResults} |
812 |
|
813 |
It is clearly important that a new electrostatic method should be able |
814 |
to reproduce the magnitudes of the force and torque vectors obtained |
815 |
via the Ewald sum. However, the {\it directionality} of these vectors |
816 |
will also be vital in calculating dynamical quantities accurately. |
817 |
Force and torque directionalities were investigated by measuring the |
818 |
angles formed between these vectors and the same vectors calculated |
819 |
using {\sc spme}. The results (figure \ref{fig:frcTrqAng}) are compared |
820 |
through the variance ($\sigma^2$) of the Gaussian fits of the angle |
821 |
error distributions of the combined set over all system types. |
822 |
|
823 |
\begin{figure} |
824 |
\centering |
825 |
\includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf} |
826 |
\caption{Statistical analysis of the width of the angular distribution |
827 |
that the force and torque vectors from a given electrostatic method |
828 |
make with their counterparts obtained using the reference Ewald sum. |
829 |
Results with a variance ($\sigma^2$) equal to zero (dashed line) |
830 |
indicate force and torque directions indistinguishable from those |
831 |
obtained using {\sc spme}. Different values of the cutoff radius are |
832 |
indicated with different symbols (9~\AA\ = circles, 12~\AA\ = squares, |
833 |
and 15~\AA\ = inverted triangles).} |
834 |
\label{fig:frcTrqAng} |
835 |
\end{figure} |
836 |
Both the force and torque $\sigma^2$ results from the analysis of the |
837 |
total accumulated system data are tabulated in figure |
838 |
\ref{fig:frcTrqAng}. Here it is clear that the {\sc sp} method would |
839 |
be essentially unusable for molecular dynamics unless the damping |
840 |
function is added. The {\sc sf} method, however, is generating force |
841 |
and torque vectors which are within a few degrees of the Ewald results |
842 |
even with weak (or no) damping. |
843 |
|
844 |
All of the sets (aside from the over-damped case) show the improvement |
845 |
afforded by choosing a larger cutoff radius. Increasing the cutoff |
846 |
from 9 to 12~\AA\ typically results in a halving of the width of the |
847 |
distribution, with a similar improvement when going from 12 to |
848 |
15~\AA . |
849 |
|
850 |
The undamped {\sc sf}, group-based cutoff, and reaction field methods |
851 |
all do equivalently well at capturing the direction of both the force |
852 |
and torque vectors. Using the electrostatic damping improves the |
853 |
angular behavior significantly for the {\sc sp} and moderately for the |
854 |
{\sc sf} methods. Over-damping is detrimental to both methods. Again |
855 |
it is important to recognize that the force vectors cover all |
856 |
particles in all seven systems, while torque vectors are only |
857 |
available for neutral molecular groups. Damping is more beneficial to |
858 |
charged bodies, and this observation is investigated further in |
859 |
appendix \ref{app:IndividualResults}. |
860 |
|
861 |
Although not discussed previously, group based cutoffs can be applied |
862 |
to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs |
863 |
will reintroduce small discontinuities at the cutoff radius, but the |
864 |
effects of these can be minimized by utilizing a switching function. |
865 |
Though there are no significant benefits or drawbacks observed in |
866 |
$\Delta E$ and the force and torque magnitudes when doing this, there |
867 |
is a measurable improvement in the directionality of the forces and |
868 |
torques. Table \ref{tab:groupAngle} shows the angular variances |
869 |
obtained both without (N) and with (Y) group based cutoffs and a |
870 |
switching function. Note that the $\alpha$ values have units of |
871 |
\AA$^{-1}$ and the variance values have units of degrees$^2$. The |
872 |
{\sc sp} (with an $\alpha$ of 0.2~\AA$^{-1}$ or smaller) shows much |
873 |
narrower angular distributions when using group-based cutoffs. The |
874 |
{\sc sf} method likewise shows improvement in the undamped and lightly |
875 |
damped cases. |
876 |
|
877 |
\begin{table} |
878 |
\caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$) |
879 |
THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A |
880 |
GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING |
881 |
THE REFERENCE EWALD SUMMATION} |
882 |
|
883 |
\footnotesize |
884 |
\begin{center} |
885 |
\begin{tabular}{@{} ccrrrrrrrr @{}} |
886 |
\toprule |
887 |
\toprule |
888 |
& &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted |
889 |
Force} \\ |
890 |
\cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups & |
891 |
$\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & |
892 |
$\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\ |
893 |
|
894 |
\midrule |
895 |
|
896 |
9~\AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\ |
897 |
& \textbf{Y} & \textbf{2.486} & \textbf{2.160} & \textbf{0.667} & \textbf{0.608} & \textbf{1.768} & \textbf{1.766} & \textbf{0.676} & \textbf{0.609} \\ |
898 |
12~\AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\ |
899 |
& \textbf{Y} & \textbf{0.515} & \textbf{0.288} & \textbf{0.127} & \textbf{0.586} & \textbf{0.308} & \textbf{0.249} & \textbf{0.127} & \textbf{0.586} \\ |
900 |
15~\AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\ |
901 |
& \textbf{Y} & \textbf{0.228} & \textbf{0.099} & \textbf{0.121} & \textbf{0.598} & \textbf{0.144} & \textbf{0.090} & \textbf{0.121} & \textbf{0.598} \\ |
902 |
|
903 |
\midrule |
904 |
|
905 |
9~\AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\ |
906 |
& \textbf{Y} & \textbf{2.115} & \textbf{1.914} & \textbf{1.878} & \textbf{5.142} & \textbf{2.076} & \textbf{2.039} & \textbf{1.972} & \textbf{5.146} \\ |
907 |
12~\AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\ |
908 |
& \textbf{Y} & \textbf{0.810} & \textbf{0.685} & \textbf{1.352} & \textbf{5.082} & \textbf{0.765} & \textbf{0.714} & \textbf{1.360} & \textbf{5.082} \\ |
909 |
15~\AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\ |
910 |
& \textbf{Y} & \textbf{0.282} & \textbf{0.294} & \textbf{1.272} & \textbf{4.999} & \textbf{0.324} & \textbf{0.318} & \textbf{1.272} & \textbf{4.999} \\ |
911 |
|
912 |
\bottomrule |
913 |
\end{tabular} |
914 |
\end{center} |
915 |
\label{tab:groupAngle} |
916 |
\end{table} |
917 |
|
918 |
One additional trend in table \ref{tab:groupAngle} is that the |
919 |
$\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$ |
920 |
increases, something that is more obvious with group-based cutoffs. |
921 |
The complimentary error function inserted into the potential weakens |
922 |
the electrostatic interaction as the value of $\alpha$ is increased. |
923 |
However, at larger values of $\alpha$, it is possible to over-damp the |
924 |
electrostatic interaction and remove it completely. Kast |
925 |
\textit{et al.} developed a method for choosing appropriate $\alpha$ |
926 |
values for these types of electrostatic summation methods by fitting |
927 |
to $g(r)$ data, and their methods indicate optimal values of 0.34, |
928 |
0.25, and 0.16~\AA$^{-1}$ for cutoff values of 9, 12, and 15~\AA\ |
929 |
respectively.\cite{Kast03} These appear to be reasonable choices to |
930 |
obtain proper MC behavior (figure \ref{fig:delE}); however, based on |
931 |
these findings, choices this high would introduce error in the |
932 |
molecular torques, particularly for the shorter cutoffs. Based on the |
933 |
above observations, empirical damping up to 0.2~\AA$^{-1}$ is |
934 |
beneficial, but damping may be unnecessary when using the {\sc sf} |
935 |
method. |
936 |
|
937 |
|
938 |
\section{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}\label{sec:ShortTimeDynamics} |
939 |
|
940 |
Zahn {\it et al.} investigated the structure and dynamics of water |
941 |
using equations (\ref{eq:ZahnPot}) and |
942 |
(\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated |
943 |
that a method similar (but not identical with) the damped {\sc sf} |
944 |
method resulted in properties very close to those obtained when |
945 |
using the Ewald summation. The properties they studied (pair |
946 |
distribution functions, diffusion constants, and velocity and |
947 |
orientational correlation functions) may not be particularly sensitive |
948 |
to the long-range and collective behavior that governs the |
949 |
low-frequency behavior in crystalline systems. Additionally, the |
950 |
ionic crystals are a worst case scenario for the pairwise methods |
951 |
because they lack the reciprocal space contribution contained in the |
952 |
Ewald summation. |
953 |
|
954 |
We used two separate measures to probe the effects of these |
955 |
alternative electrostatic methods on the dynamics in crystalline |
956 |
materials. For short- and intermediate-time dynamics, we computed the |
957 |
velocity autocorrelation function, and for long-time and large |
958 |
length-scale collective motions, we looked at the low-frequency |
959 |
portion of the power spectrum. |
960 |
|
961 |
\begin{figure} |
962 |
\centering |
963 |
\includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf} |
964 |
\caption{Velocity autocorrelation functions of NaCl crystals at |
965 |
1000~K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& |
966 |
0.2~\AA$^{-1}$), and {\sc sp} ($\alpha$ = 0.2~\AA$^{-1}$). The inset is |
967 |
a magnification of the area around the first minimum. The times to |
968 |
first collision are nearly identical, but differences can be seen in |
969 |
the peaks and troughs, where the undamped and weakly damped methods |
970 |
are stiffer than the moderately damped and {\sc spme} methods.} |
971 |
\label{fig:vCorrPlot} |
972 |
\end{figure} |
973 |
The short-time decay of the velocity autocorrelation functions through |
974 |
the first collision are nearly identical in figure |
975 |
\ref{fig:vCorrPlot}, but the peaks and troughs of the functions show |
976 |
how the methods differ. The undamped {\sc sf} method has deeper |
977 |
troughs (see inset in figure \ref{fig:vCorrPlot}) and higher peaks than |
978 |
any of the other methods. As the damping parameter ($\alpha$) is |
979 |
increased, these peaks are smoothed out, and the {\sc sf} method |
980 |
approaches the {\sc spme} results. With $\alpha$ values of 0.2~\AA$^{-1}$, |
981 |
the {\sc sf} and {\sc sp} functions are nearly identical and track the |
982 |
{\sc spme} features quite well. This is not surprising because the {\sc sf} |
983 |
and {\sc sp} potentials become nearly identical with increased |
984 |
damping. However, this appears to indicate that once damping is |
985 |
utilized, the details of the form of the potential (and forces) |
986 |
constructed out of the damped electrostatic interaction are less |
987 |
important. |
988 |
|
989 |
\section{Collective Motion: Power Spectra of NaCl Crystals}\label{sec:LongTimeDynamics} |
990 |
|
991 |
\begin{figure} |
992 |
\centering |
993 |
\includegraphics[width = \linewidth]{./figures/spectraSquare.pdf} |
994 |
\caption{Power spectra obtained from the velocity auto-correlation |
995 |
functions of NaCl crystals at 1000~K while using {\sc spme}, {\sc sf} |
996 |
($\alpha$ = 0, 0.1, \& 0.2~\AA$^{-1}$), and {\sc sp} ($\alpha$ = |
997 |
0.2~\AA$^{-1}$). The inset shows the frequency region below |
998 |
100~cm$^{-1}$ to highlight where the spectra differ.} |
999 |
\label{fig:methodPS} |
1000 |
\end{figure} |
1001 |
To evaluate how the differences between the methods affect the |
1002 |
collective long-time motion, we computed power spectra from long-time |
1003 |
traces of the velocity autocorrelation function. The power spectra for |
1004 |
the best-performing alternative methods are shown in |
1005 |
figure \ref{fig:methodPS}. Apodization of the correlation functions via |
1006 |
a cubic switching function between 40 and 50~ps was used to reduce the |
1007 |
ringing resulting from data truncation. This procedure had no |
1008 |
noticeable effect on peak location or magnitude. |
1009 |
|
1010 |
While the high frequency regions of the power spectra for the |
1011 |
alternative methods are quantitatively identical with Ewald spectrum, |
1012 |
the low frequency region shows how the summation methods differ. |
1013 |
Considering the low-frequency inset (expanded in the upper frame of |
1014 |
figure \ref{fig:methodPS}), at frequencies below 100~cm$^{-1}$, the |
1015 |
correlated motions are blue-shifted when using undamped or weakly |
1016 |
damped {\sc sf}. When using moderate damping ($\alpha = |
1017 |
0.2$~\AA$^{-1}$), both the {\sc sf} and {\sc sp} methods produce |
1018 |
correlated motions nearly identical to the Ewald method (which has a |
1019 |
convergence parameter of 0.3119~\AA$^{-1}$). This weakening of the |
1020 |
electrostatic interaction with increased damping explains why the |
1021 |
long-ranged correlated motions are at lower frequencies for the |
1022 |
moderately damped methods than for undamped or weakly damped methods. |
1023 |
|
1024 |
\begin{figure} |
1025 |
\centering |
1026 |
\includegraphics[width = \linewidth]{./figures/increasedDamping.pdf} |
1027 |
\caption{Effect of damping on the two lowest-frequency phonon modes in |
1028 |
the NaCl crystal at 1000~K. The undamped shifted force ({\sc sf}) |
1029 |
method is off by less than 10~cm$^{-1}$, and increasing the |
1030 |
electrostatic damping to 0.25~\AA$^{-1}$ gives quantitative agreement |
1031 |
with the power spectrum obtained using the Ewald sum. Over-damping can |
1032 |
result in underestimates of frequencies of the long-wavelength |
1033 |
motions.} |
1034 |
\label{fig:dampInc} |
1035 |
\end{figure} |
1036 |
To isolate the role of the damping constant, we have computed the |
1037 |
spectra for a single method ({\sc sf}) with a range of damping |
1038 |
constants and compared this with the {\sc spme} spectrum. Figure |
1039 |
\ref{fig:dampInc} shows more clearly that increasing the electrostatic |
1040 |
damping red-shifts the lowest frequency phonon modes. However, even |
1041 |
without any electrostatic damping, the {\sc sf} method has at most a |
1042 |
10 cm$^{-1}$ error in the lowest frequency phonon mode. Without the |
1043 |
{\sc sf} modifications, an undamped (pure cutoff) method would predict |
1044 |
the lowest frequency peak near 325~cm$^{-1}$, an error significantly |
1045 |
larger than that of the undamped {\sc sf} technique. This indicates |
1046 |
that {\it most} of the collective behavior in the crystal is |
1047 |
accurately captured using the {\sc sf} method. Quantitative agreement |
1048 |
with Ewald can be obtained using moderate damping in addition to the |
1049 |
shifting at the cutoff distance. |
1050 |
|
1051 |
\section{An Application: TIP5P-E Water}\label{sec:t5peApplied} |
1052 |
|
1053 |
The above sections focused on the energetics and dynamics of a variety |
1054 |
of systems when utilizing the {\sc sp} and {\sc sf} pairwise |
1055 |
techniques. A unitary correlation with results obtained using the |
1056 |
Ewald summation should result in a successful reproduction of both the |
1057 |
static and dynamic properties of any selected system. To test this, |
1058 |
we decided to calculate a series of properties for the TIP5P-E water |
1059 |
model when using the {\sc sf} technique. |
1060 |
|
1061 |
The TIP5P-E water model is a variant of Mahoney and Jorgensen's |
1062 |
five-point transferable intermolecular potential (TIP5P) model for |
1063 |
water.\cite{Mahoney00} TIP5P was developed to reproduce the density |
1064 |
maximum anomaly present in liquid water near 4$^\circ$C. As with many |
1065 |
previous point charge water models (such as ST2, TIP3P, TIP4P, SPC, |
1066 |
and SPC/E), TIP5P was parametrized using a simple cutoff with no |
1067 |
long-range electrostatic |
1068 |
correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87} |
1069 |
Without this correction, the pressure term on the central particle |
1070 |
from the surroundings is missing. Because they expand to compensate |
1071 |
for this added pressure term when this correction is included, systems |
1072 |
composed of these particles tend to under-predict the density of water |
1073 |
under standard conditions. When using any form of long-range |
1074 |
electrostatic correction, it has become common practice to develop or |
1075 |
utilize a reparametrized water model that corrects for this |
1076 |
effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows |
1077 |
this practice and was optimized specifically for use with the Ewald |
1078 |
summation.\cite{Rick04} In his publication, Rick preserved the |
1079 |
geometry and point charge magnitudes in TIP5P and focused on altering |
1080 |
the Lennard-Jones parameters to correct the density at |
1081 |
298K.\cite{Rick04} With the density corrected, he compared common |
1082 |
water properties for TIP5P-E using the Ewald sum with TIP5P using a |
1083 |
9~\AA\ cutoff. |
1084 |
|
1085 |
In the following sections, we compared these same water properties |
1086 |
calculated from TIP5P-E using the Ewald sum with TIP5P-E using the |
1087 |
{\sc sf} technique. In the above evaluation of the pairwise |
1088 |
techniques, we observed some flexibility in the choice of parameters. |
1089 |
Because of this, the following comparisons include the {\sc sf} |
1090 |
technique with a 12~\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and |
1091 |
0.2~\AA$^{-1}$, as well as a 9~\AA\ cutoff with an $\alpha$ = |
1092 |
0.2~\AA$^{-1}$. |
1093 |
|
1094 |
\subsection{Density}\label{sec:t5peDensity} |
1095 |
|
1096 |
As stated previously, the property that prompted the development of |
1097 |
TIP5P-E was the density at 1 atm. The density depends upon the |
1098 |
internal pressure of the system in the $NPT$ ensemble, and the |
1099 |
calculation of the pressure includes a components from both the |
1100 |
kinetic energy and the virial. More specifically, the instantaneous |
1101 |
molecular pressure ($p(t)$) is given by |
1102 |
\begin{equation} |
1103 |
p(t) = \frac{1}{\textrm{d}V}\sum_\mu |
1104 |
\left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}} |
1105 |
+ \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right], |
1106 |
\label{eq:MolecularPressure} |
1107 |
\end{equation} |
1108 |
where d is the dimensionality of the system, $V$ is the volume, |
1109 |
$\mathbf{P}_{\mu}$ is the momentum of molecule $\mu$, $\mathbf{R}_\mu$ |
1110 |
is the position of the center of mass ($M_\mu$) of molecule $\mu$, and |
1111 |
$\mathbf{F}_{\mu i}$ is the force on atom $i$ of molecule |
1112 |
$\mu$.\cite{Melchionna93} The virial term (the right term in the |
1113 |
brackets of equation |
1114 |
\ref{eq:MolecularPressure}) is directly dependent on the interatomic |
1115 |
forces. Since the {\sc sp} method does not modify the forces (see |
1116 |
section. \ref{sec:PairwiseDerivation}), the pressure using {\sc sp} |
1117 |
will be identical to that obtained without an electrostatic |
1118 |
correction. The {\sc sf} method does alter the virial component and, |
1119 |
by way of the modified pressures, should provide densities more in |
1120 |
line with those obtained using the Ewald summation. |
1121 |
|
1122 |
To compare densities, $NPT$ simulations were performed with the same |
1123 |
temperatures as those selected by Rick in his Ewald summation |
1124 |
simulations.\cite{Rick04} In order to improve statistics around the |
1125 |
density maximum, 3~ns trajectories were accumulated at 0, 12.5, and |
1126 |
25$^\circ$C, while 2~ns trajectories were obtained at all other |
1127 |
temperatures. The average densities were calculated from the later |
1128 |
three-fourths of each trajectory. Similar to Mahoney and Jorgensen's |
1129 |
method for accumulating statistics, these sequences were spliced into |
1130 |
200 segments to calculate the average density and standard deviation |
1131 |
at each temperature.\cite{Mahoney00} |
1132 |
|
1133 |
\begin{figure} |
1134 |
\includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf} |
1135 |
\caption{Density versus temperature for the TIP5P-E water model when |
1136 |
using the Ewald summation (Ref. \cite{Rick04}) and the {\sc sf} method |
1137 |
with various parameters. The pressure term from the image-charge shell |
1138 |
is larger than that provided by the reciprocal-space portion of the |
1139 |
Ewald summation, leading to slightly lower densities. This effect is |
1140 |
more visible with the 9~\AA\ cutoff, where the image charges exert a |
1141 |
greater force on the central particle. The error bars for the {\sc sf} |
1142 |
methods show the average one-sigma uncertainty of the density |
1143 |
measurement, and this uncertainty is the same for all the {\sc sf} |
1144 |
curves.} |
1145 |
\label{fig:t5peDensities} |
1146 |
\end{figure} |
1147 |
Figure \ref{fig:t5peDensities} shows the densities calculated for |
1148 |
TIP5P-E using differing electrostatic corrections overlaid on the |
1149 |
experimental values.\cite{CRC80} The densities when using the {\sc sf} |
1150 |
technique are close to, though typically lower than, those calculated |
1151 |
while using the Ewald summation. These slightly reduced densities |
1152 |
indicate that the pressure component from the image charges at |
1153 |
R$_\textrm{c}$ is larger than that exerted by the reciprocal-space |
1154 |
portion of the Ewald summation. Bringing the image charges closer to |
1155 |
the central particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than |
1156 |
the preferred 12~\AA\ R$_\textrm{c}$) increases the strength of their |
1157 |
interactions, resulting in a further reduction of the densities. |
1158 |
|
1159 |
Because the strength of the image charge interactions has a noticeable |
1160 |
effect on the density, we would expect the use of electrostatic |
1161 |
damping to also play a role in these calculations. Larger values of |
1162 |
$\alpha$ weaken the pair-interactions; and since electrostatic damping |
1163 |
is distance-dependent, force components from the image charges will be |
1164 |
reduced more than those from particles close the the central |
1165 |
charge. This effect is visible in figure \ref{fig:t5peDensities} with |
1166 |
the damped {\sc sf} sums showing slightly higher densities; however, |
1167 |
it is apparent that the choice of cutoff radius plays a much more |
1168 |
important role in the resulting densities. |
1169 |
|
1170 |
As a final note, all of the above density calculations were performed |
1171 |
with systems of 512 water molecules. Rick observed a system size |
1172 |
dependence of the computed densities when using the Ewald summation, |
1173 |
most likely due to his tying of the convergence parameter to the box |
1174 |
dimensions.\cite{Rick04} For systems of 256 water molecules, the |
1175 |
calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A |
1176 |
system size of 256 molecules would force the use of a shorter |
1177 |
R$_\textrm{c}$ when using the {\sc sf} technique, and this would also |
1178 |
lower the densities. Moving to larger systems, as long as the |
1179 |
R$_\textrm{c}$ remains at a fixed value, we would expect the densities |
1180 |
to remain constant. |
1181 |
|
1182 |
\subsection{Liquid Structure}\label{sec:t5peLiqStructure} |
1183 |
|
1184 |
A common function considered when developing and comparing water |
1185 |
models is the oxygen-oxygen radial distribution function |
1186 |
($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of |
1187 |
finding a pair of oxygen atoms some distance ($r$) apart relative to a |
1188 |
random distribution at the same density.\cite{Allen87} It is |
1189 |
calculated via |
1190 |
\begin{equation} |
1191 |
g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i} |
1192 |
\delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle, |
1193 |
\label{eq:GOOofR} |
1194 |
\end{equation} |
1195 |
where the double sum is over all $i$ and $j$ pairs of $N$ oxygen |
1196 |
atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or |
1197 |
neutron scattering experiments through the oxygen-oxygen structure |
1198 |
factor ($S_\textrm{OO}(k)$) by the following relationship: |
1199 |
\begin{equation} |
1200 |
S_\textrm{OO}(k) = 1 + 4\pi\rho |
1201 |
\int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r. |
1202 |
\label{eq:SOOofK} |
1203 |
\end{equation} |
1204 |
Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform |
1205 |
of $g_\textrm{OO}(r)$. |
1206 |
|
1207 |
The experimentally determined $g_\textrm{OO}(r)$ for liquid water has |
1208 |
been compared in great detail with the various common water models, |
1209 |
and TIP5P was found to be in better agreement than other rigid, |
1210 |
non-polarizable models.\cite{Sorenson00} This excellent agreement with |
1211 |
experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To |
1212 |
check whether the choice of using the Ewald summation or the {\sc sf} |
1213 |
technique alters the liquid structure, the $g_\textrm{OO}(r)$s at 298K |
1214 |
and 1atm were determined for the systems compared in the previous |
1215 |
section. |
1216 |
|
1217 |
\begin{figure} |
1218 |
\includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf} |
1219 |
\caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298~K and |
1220 |
1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc |
1221 |
sf} technique with varying parameters. Even with the reduced densities |
1222 |
using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially |
1223 |
identical.} |
1224 |
\label{fig:t5peGofRs} |
1225 |
\end{figure} |
1226 |
The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc |
1227 |
sf} technique with a various parameters are overlaid on the |
1228 |
$g_\textrm{OO}(r)$ while using the Ewald summation in figure |
1229 |
\ref{fig:t5peGofRs}. The differences in density do not appear to have |
1230 |
any effect on the liquid structure as the $g_\textrm{OO}(r)$s are |
1231 |
indistinguishable. These results indicate that the $g_\textrm{OO}(r)$ |
1232 |
is insensitive to the choice of electrostatic correction. |
1233 |
|
1234 |
\subsection{Thermodynamic Properties}\label{sec:t5peThermo} |
1235 |
|
1236 |
In addition to the density, there are a variety of thermodynamic |
1237 |
quantities that can be calculated for water and compared directly to |
1238 |
experimental values. Some of these additional quantities include the |
1239 |
latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant |
1240 |
pressure heat capacity ($C_p$), the isothermal compressibility |
1241 |
($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static |
1242 |
dielectric constant ($\epsilon$). All of these properties were |
1243 |
calculated for TIP5P-E with the Ewald summation, so they provide a |
1244 |
good set for comparisons involving the {\sc sf} technique. |
1245 |
|
1246 |
The $\Delta H_\textrm{vap}$ is the enthalpy change required to |
1247 |
transform one mol of substance from the liquid phase to the gas |
1248 |
phase.\cite{Berry00} In molecular simulations, this quantity can be |
1249 |
determined via |
1250 |
\begin{equation} |
1251 |
\begin{split} |
1252 |
\Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\ |
1253 |
&= E_\textrm{gas} - E_\textrm{liq.} |
1254 |
+ p(V_\textrm{gas} - V_\textrm{liq.}) \\ |
1255 |
&\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT, |
1256 |
\end{split} |
1257 |
\label{eq:DeltaHVap} |
1258 |
\end{equation} |
1259 |
where $E$ is the total energy, $U$ is the potential energy, $p$ is the |
1260 |
pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is |
1261 |
the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As |
1262 |
seen in the last line of equation (\ref{eq:DeltaHVap}), we can |
1263 |
approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas |
1264 |
state. This allows us to cancel the kinetic energy terms, leaving only |
1265 |
the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is |
1266 |
several orders of magnitude larger than that of the liquid, so we can |
1267 |
neglect the liquid $pV$ term. |
1268 |
|
1269 |
The remaining thermodynamic properties can all be calculated from |
1270 |
fluctuations of the enthalpy, volume, and system dipole |
1271 |
moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the |
1272 |
enthalpy in constant pressure simulations via |
1273 |
\begin{equation} |
1274 |
\begin{split} |
1275 |
C_p = \left(\frac{\partial H}{\partial T}\right)_{N,p} |
1276 |
= \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2}, |
1277 |
\end{split} |
1278 |
\label{eq:Cp} |
1279 |
\end{equation} |
1280 |
where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via |
1281 |
\begin{equation} |
1282 |
\begin{split} |
1283 |
\kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T} |
1284 |
= \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})} |
1285 |
{k_BT\langle V\rangle_{N,P,T}}, |
1286 |
\end{split} |
1287 |
\label{eq:kappa} |
1288 |
\end{equation} |
1289 |
and $\alpha_p$ can be calculated via |
1290 |
\begin{equation} |
1291 |
\begin{split} |
1292 |
\alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P} |
1293 |
= \frac{(\langle VH\rangle_{N,P,T} |
1294 |
- \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})} |
1295 |
{k_BT^2\langle V\rangle_{N,P,T}}. |
1296 |
\end{split} |
1297 |
\label{eq:alpha} |
1298 |
\end{equation} |
1299 |
Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can |
1300 |
be calculated for systems of non-polarizable substances via |
1301 |
\begin{equation} |
1302 |
\epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV}, |
1303 |
\label{eq:staticDielectric} |
1304 |
\end{equation} |
1305 |
where $\epsilon_0$ is the permittivity of free space and $\langle |
1306 |
M^2\rangle$ is the fluctuation of the system dipole |
1307 |
moment.\cite{Allen87} The numerator in the fractional term in equation |
1308 |
(\ref{eq:staticDielectric}) is the fluctuation of the simulation-box |
1309 |
dipole moment, identical to the quantity calculated in the |
1310 |
finite-system Kirkwood $g$ factor ($G_k$): |
1311 |
\begin{equation} |
1312 |
G_k = \frac{\langle M^2\rangle}{N\mu^2}, |
1313 |
\label{eq:KirkwoodFactor} |
1314 |
\end{equation} |
1315 |
where $\mu$ is the dipole moment of a single molecule of the |
1316 |
homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The |
1317 |
fluctuation term in both equation (\ref{eq:staticDielectric}) and |
1318 |
\ref{eq:KirkwoodFactor} is calculated as follows, |
1319 |
\begin{equation} |
1320 |
\begin{split} |
1321 |
\langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle |
1322 |
- \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\ |
1323 |
&= \langle M_x^2+M_y^2+M_z^2\rangle |
1324 |
- (\langle M_x\rangle^2 + \langle M_x\rangle^2 |
1325 |
+ \langle M_x\rangle^2). |
1326 |
\end{split} |
1327 |
\label{eq:fluctBoxDipole} |
1328 |
\end{equation} |
1329 |
This fluctuation term can be accumulated during the simulation; |
1330 |
however, it converges rather slowly, thus requiring multi-nanosecond |
1331 |
simulation times.\cite{Horn04} In the case of tin-foil boundary |
1332 |
conditions, the dielectric/surface term of equation (\ref{eq:EwaldSum}) |
1333 |
is equal to zero. Since the {\sc sf} method also lacks this |
1334 |
dielectric/surface term, equation (\ref{eq:staticDielectric}) is still |
1335 |
valid for determining static dielectric constants. |
1336 |
|
1337 |
All of the above properties were calculated from the same trajectories |
1338 |
used to determine the densities in section \ref{sec:t5peDensity} |
1339 |
except for the static dielectric constants. The $\epsilon$ values were |
1340 |
accumulated from 2~ns $NVE$ ensemble trajectories with system densities |
1341 |
fixed at the average values from the $NPT$ simulations at each of the |
1342 |
temperatures. The resulting values are displayed in figure |
1343 |
\ref{fig:t5peThermo}. |
1344 |
\begin{figure} |
1345 |
\centering |
1346 |
\includegraphics[width=4.5in]{./figures/t5peThermo.pdf} |
1347 |
\caption{Thermodynamic properties for TIP5P-E using the Ewald summation |
1348 |
and the {\sc sf} techniques along with the experimental values. Units |
1349 |
for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$, |
1350 |
cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$, |
1351 |
and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from |
1352 |
reference \cite{Rick04}. Experimental values for $\Delta |
1353 |
H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference |
1354 |
\cite{Kell75}. Experimental values for $C_p$ are from reference |
1355 |
\cite{Wagner02}. Experimental values for $\epsilon$ are from reference |
1356 |
\cite{Malmberg56}.} |
1357 |
\label{fig:t5peThermo} |
1358 |
\end{figure} |
1359 |
|
1360 |
As observed for the density in section \ref{sec:t5peDensity}, the |
1361 |
property trends with temperature seen when using the Ewald summation |
1362 |
are reproduced with the {\sc sf} technique. One noticable difference |
1363 |
between the properties calculated using the two methods are the lower |
1364 |
$\Delta H_\textrm{vap}$ values when using {\sc sf}. This is to be |
1365 |
expected due to the direct weakening of the electrostatic interaction |
1366 |
through forced neutralization. This results in an increase of the |
1367 |
intermolecular potential producing lower values from equation |
1368 |
(\ref{eq:DeltaHVap}). The slopes of these values with temperature are |
1369 |
similar to that seen using the Ewald summation; however, they are both |
1370 |
steeper than the experimental trend, indirectly resulting in the |
1371 |
inflated $C_p$ values at all temperatures. |
1372 |
|
1373 |
Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ |
1374 |
values all overlap within error. As indicated for the $\Delta |
1375 |
H_\textrm{vap}$ and $C_p$ results discussed in the previous paragraph, |
1376 |
the deviations between experiment and simulation in this region are |
1377 |
not the fault of the electrostatic summation methods but are due to |
1378 |
the TIP5P class model itself. Like most rigid, non-polarizable, |
1379 |
point-charge water models, the density decreases with temperature at a |
1380 |
much faster rate than experiment (see figure |
1381 |
\ref{fig:t5peDensities}). The reduced density leads to the inflated |
1382 |
compressibility and expansivity values at higher temperatures seen |
1383 |
here in figure \ref{fig:t5peThermo}. Incorporation of polarizability |
1384 |
and many-body effects are required in order for simulation to overcome |
1385 |
these differences with experiment.\cite{Laasonen93,Donchev06} |
1386 |
|
1387 |
At temperatures below the freezing point for experimental water, the |
1388 |
differences between {\sc sf} and the Ewald summation results are more |
1389 |
apparent. The larger $C_p$ and lower $\alpha_p$ values in this region |
1390 |
indicate a more pronounced transition in the supercooled regime, |
1391 |
particularly in the case of {\sc sf} without damping. This points to |
1392 |
the onset of a more frustrated or glassy behavior for TIP5P-E at |
1393 |
temperatures below 250~K in the {\sc sf} simulations, indicating that |
1394 |
disorder in the reciprical-space term of the Ewald summation might act |
1395 |
to loosen up the local structure more than the image-charges in {\sc |
1396 |
sf}. Because the systems are locked in different regions of |
1397 |
phase-space, comparisons between properties at these temperatures are |
1398 |
not exactly fair. This observation is explored in more detail in |
1399 |
section \ref{sec:t5peDynamics}. |
1400 |
|
1401 |
The final thermodynamic property displayed in figure |
1402 |
\ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy |
1403 |
between the Ewald summation and the {\sc sf} technique (and experiment |
1404 |
for that matter). It is known that the dielectric constant is |
1405 |
dependent upon and quite sensitive to the imposed boundary |
1406 |
conditions.\cite{Neumann80,Neumann83} This is readily apparent in the |
1407 |
converged $\epsilon$ values accumulated for the {\sc sf} |
1408 |
simulations. Lack of a damping function results in dielectric |
1409 |
constants significantly smaller than that obtained using the Ewald |
1410 |
sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the |
1411 |
agreement considerably. It should be noted that the choice of the |
1412 |
``Ewald coefficient'' value also has a significant effect on the |
1413 |
calculated value when using the Ewald summation. In the simulations of |
1414 |
TIP5P-E with the Ewald sum, this screening parameter was tethered to |
1415 |
the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04} In |
1416 |
general, systems with larger screening parameters reported larger |
1417 |
dielectric constant values, the same behavior we see here with {\sc |
1418 |
sf}; however, the choice of cutoff radius also plays an important |
1419 |
role. In section \ref{sec:dampingDielectric}, this connection is |
1420 |
further explored as optimal damping coefficients for different choices |
1421 |
of $R_\textrm{c}$ are determined for {\sc sf} for capturing the |
1422 |
dielectric behavior. |
1423 |
|
1424 |
\subsection{Dynamic Properties}\label{sec:t5peDynamics} |
1425 |
|
1426 |
To look at the dynamic properties of TIP5P-E when using the {\sc sf} |
1427 |
method, 200~ps $NVE$ simulations were performed for each temperature at |
1428 |
the average density reported by the $NPT$ simulations. The |
1429 |
self-diffusion constants ($D$) were calculated with the Einstein |
1430 |
relation using the mean square displacement (MSD), |
1431 |
\begin{equation} |
1432 |
D = \lim_{t\rightarrow\infty} |
1433 |
\frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t}, |
1434 |
\label{eq:MSD} |
1435 |
\end{equation} |
1436 |
where $t$ is time, and $\mathbf{r}_i$ is the position of particle |
1437 |
$i$.\cite{Allen87} Figure \ref{fig:ExampleMSD} shows an example MSD |
1438 |
plot. As labeled in the figure, MSD plots consist of three distinct |
1439 |
regions: |
1440 |
|
1441 |
\begin{enumerate}[itemsep=0pt] |
1442 |
\item parabolic short-time ballistic motion, |
1443 |
\item linear diffusive regime, and |
1444 |
\item a region with poor statistics. |
1445 |
\end{enumerate} |
1446 |
The slope from the linear region (region 2) is used to calculate $D$. |
1447 |
\begin{figure} |
1448 |
\centering |
1449 |
\includegraphics[width=3.5in]{./figures/ExampleMSD.pdf} |
1450 |
\caption{Example plot of mean square displacement verses time. The |
1451 |
left red region is the ballistic motion regime, the middle green |
1452 |
region is the linear diffusive regime, and the right blue region is |
1453 |
the region with poor statistics.} |
1454 |
\label{fig:ExampleMSD} |
1455 |
\end{figure} |
1456 |
|
1457 |
\begin{figure} |
1458 |
\centering |
1459 |
\includegraphics[width=3.5in]{./figures/waterFrame.pdf} |
1460 |
\caption{Body-fixed coordinate frame for a water molecule. The |
1461 |
respective molecular principle axes point in the direction of the |
1462 |
labeled frame axes.} |
1463 |
\label{fig:waterFrame} |
1464 |
\end{figure} |
1465 |
In addition to translational diffusion, reorientational time constants |
1466 |
were calculated for comparisons with the Ewald simulations and with |
1467 |
experiments. These values were determined from 25~ps $NVE$ trajectories |
1468 |
through calculation of the orientational time correlation function, |
1469 |
\begin{equation} |
1470 |
C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t) |
1471 |
\cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle, |
1472 |
\label{eq:OrientCorr} |
1473 |
\end{equation} |
1474 |
where $P_l$ is the Legendre polynomial of order $l$ and |
1475 |
$\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along |
1476 |
principle axis $\alpha$. The principle axis frame for these water |
1477 |
molecules is shown in figure \ref{fig:waterFrame}. As an example, |
1478 |
$C_l^y$ is calculated from the time evolution of the unit vector |
1479 |
connecting the two hydrogen atoms. |
1480 |
|
1481 |
\begin{figure} |
1482 |
\centering |
1483 |
\includegraphics[width=3.5in]{./figures/ExampleOrientCorr.pdf} |
1484 |
\caption{Example plots of the orientational autocorrelation functions |
1485 |
for the first and second Legendre polynomials. These curves show the |
1486 |
time decay of the unit vector along the $y$ principle axis.} |
1487 |
\label{fig:OrientCorr} |
1488 |
\end{figure} |
1489 |
From the orientation autocorrelation functions, we can obtain time |
1490 |
constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows |
1491 |
some example plots of orientational autocorrelation functions for the |
1492 |
first and second Legendre polynomials. The relatively short time |
1493 |
portions (between 1 and 3~ps for water) of these curves can be fit to |
1494 |
an exponential decay to obtain these constants, and they are directly |
1495 |
comparable to water orientational relaxation times from nuclear |
1496 |
magnetic resonance (NMR). The relaxation constant obtained from |
1497 |
$C_2^y(t)$ is of particular interest because it describes the |
1498 |
relaxation of the principle axis connecting the hydrogen atoms. Thus, |
1499 |
$C_2^y(t)$ can be compared to the intermolecular portion of the |
1500 |
dipole-dipole relaxation from a proton NMR signal and should provide |
1501 |
the best estimate of the NMR relaxation time constant.\cite{Impey82} |
1502 |
|
1503 |
\begin{figure} |
1504 |
\centering |
1505 |
\includegraphics[width=3.5in]{./figures/t5peDynamics.pdf} |
1506 |
\caption{Diffusion constants ({\it upper}) and reorientational time |
1507 |
constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf} |
1508 |
technique compared with experiment. Data at temperatures less that |
1509 |
0$^\circ$C were not included in the $\tau_2^y$ plot to allow for |
1510 |
easier comparisons in the more relevant temperature regime.} |
1511 |
\label{fig:t5peDynamics} |
1512 |
\end{figure} |
1513 |
Results for the diffusion constants and orientational relaxation times |
1514 |
are shown in figure \ref{fig:t5peDynamics}. From this figure, it is |
1515 |
apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using |
1516 |
the Ewald sum are reproduced with the {\sc sf} technique. The enhanced |
1517 |
diffusion at high temperatures are again the product of the lower |
1518 |
densities in comparison with experiment and do not provide any special |
1519 |
insight into differences between the electrostatic summation |
1520 |
techniques. With the undamped {\sc sf} technique, TIP5P-E tends to |
1521 |
diffuse a little faster than with the Ewald sum; however, use of light |
1522 |
to moderate damping results in indistinguishable $D$ values. Though |
1523 |
not apparent in this figure, {\sc sf} values at the lowest temperature |
1524 |
are approximately an order of magnitude lower than with Ewald. These |
1525 |
values support the observation from section \ref{sec:t5peThermo} that |
1526 |
there appeared to be a change to a more glassy-like phase with the |
1527 |
{\sc sf} technique at these lower temperatures. |
1528 |
|
1529 |
The $\tau_2^y$ results in the lower frame of figure |
1530 |
\ref{fig:t5peDynamics} show a much greater difference between the {\sc |
1531 |
sf} results and the Ewald results. At all temperatures shown, TIP5P-E |
1532 |
relaxes faster than experiment with the Ewald sum while tracking |
1533 |
experiment fairly well when using the {\sc sf} technique, independent |
1534 |
of the choice of damping constant. Their are several possible reasons |
1535 |
for this deviation between techniques. The Ewald results were taken |
1536 |
shorter (10ps) trajectories than the {\sc sf} results (25ps). A quick |
1537 |
calculation from a 10~ps trajectory with {\sc sf} with an $\alpha$ of |
1538 |
0.2~\AA$^{-1}$ at 25$^\circ$C showed a 0.4~ps drop in $\tau_2^y$, |
1539 |
placing the result more in line with that obtained using the Ewald |
1540 |
sum. These results support this explanation; however, recomputing the |
1541 |
results to meet a poorer statistical standard is |
1542 |
counter-productive. Assuming the Ewald results are not the product of |
1543 |
poor statistics, differences in techniques to integrate the |
1544 |
orientational motion could also play a role. {\sc shake} is the most |
1545 |
commonly used technique for approximating rigid-body orientational |
1546 |
motion,\cite{Ryckaert77} where as in {\sc oopse}, we maintain and |
1547 |
integrate the entire rotation matrix using the {\sc dlm} |
1548 |
method.\cite{Meineke05} Since {\sc shake} is an iterative constraint |
1549 |
technique, if the convergence tolerances are raised for increased |
1550 |
performance, error will accumulate in the orientational |
1551 |
motion. Finally, the Ewald results were calculated using the $NVT$ |
1552 |
ensemble, while the $NVE$ ensemble was used for {\sc sf} |
1553 |
calculations. The additional mode of motion due to the thermostat will |
1554 |
alter the dynamics, resulting in differences between $NVT$ and $NVE$ |
1555 |
results. These differences are increasingly noticeable as the |
1556 |
thermostat time constant decreases. |
1557 |
|
1558 |
\section{Damping of Point Multipoles}\label{sec:dampingMultipoles} |
1559 |
|
1560 |
As discussed above, the {\sc sp} and {\sc sf} methods operate by |
1561 |
neutralizing the cutoff sphere with charge-charge interaction shifting |
1562 |
and by damping the electrostatic interactions. Now we would like to |
1563 |
consider an extension of these techniques to include point multipole |
1564 |
interactions. How will the shifting and damping need to be modified in |
1565 |
order to accommodate point multipoles? |
1566 |
|
1567 |
Of the two techniques, the easiest to adapt is shifting. Shifting is |
1568 |
employed to neutralize the cutoff sphere; however, in a system |
1569 |
composed purely of point multipoles, the cutoff sphere is already |
1570 |
neutralized. This means that shifting is not necessary between point |
1571 |
multipoles. In a mixed system of monopoles and multipoles, the |
1572 |
undamped {\sc sf} potential needs only to shift the force terms of the |
1573 |
monopole (and use the monopole potential of equation (\ref{eq:SFPot})) |
1574 |
and smoothly cutoff the multipole interactions with a switching |
1575 |
function. The switching function is required in order to conserve |
1576 |
energy, because a discontinuity will exist at $R_\textrm{c}$ in the |
1577 |
absence of shifting terms. |
1578 |
|
1579 |
If we consider damping the {\sc sf} potential (Eq. (\ref{eq:DSFPot})), |
1580 |
then we need to incorporate the complimentary error function term into |
1581 |
the multipole potentials. The most direct way to do this is by |
1582 |
replacing $r^{-1}$ with erfc$(\alpha r)\cdot r^{-1}$ in the multipole |
1583 |
expansion.\cite{Hirschfelder67} In the multipole expansion, rather |
1584 |
than considering only the interactions between single point charges, |
1585 |
the electrostatic interaction is reformulated such that it describes |
1586 |
the interaction between charge distributions about central sites of |
1587 |
the respective sets of charges. This procedure is what leads to the |
1588 |
familiar charge-dipole, |
1589 |
\begin{equation} |
1590 |
V_\textrm{cd} = -q_i\frac{\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}{r^3_{ij}} |
1591 |
= -q_i\mu_j\frac{\cos\theta}{r^2_{ij}}, |
1592 |
\label{eq:chargeDipole} |
1593 |
\end{equation} |
1594 |
and dipole-dipole, |
1595 |
\begin{equation} |
1596 |
V_\textrm{dd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}) |
1597 |
(\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}} - |
1598 |
\frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}, |
1599 |
\label{eq:dipoleDipole} |
1600 |
\end{equation} |
1601 |
interaction potentials. |
1602 |
|
1603 |
Using the charge-dipole interaction as an example, if we insert |
1604 |
erfc$(\alpha r)\cdot r^{-1}$ in place of $r^{-1}$, a damped |
1605 |
charge-dipole results, |
1606 |
\begin{equation} |
1607 |
V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}), |
1608 |
\label{eq:dChargeDipole} |
1609 |
\end{equation} |
1610 |
where $c_1(r_{ij})$ is |
1611 |
\begin{equation} |
1612 |
c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}} |
1613 |
+ \textrm{erfc}(\alpha r_{ij}). |
1614 |
\label{eq:c1Func} |
1615 |
\end{equation} |
1616 |
Thus, $c_1(r_{ij})$ is the resulting damping term that modifies the |
1617 |
standard charge-dipole potential (Eq. (\ref{eq:chargeDipole})). Note |
1618 |
that this damping term is dependent upon distance and not upon |
1619 |
orientation, and that it is acting on what was originally an |
1620 |
$r^{-3}$ function. By writing the damped form in this manner, we |
1621 |
can collect the damping into one function and apply it to the original |
1622 |
potential when damping is desired. This works well for potentials that |
1623 |
have only one $r^{-n}$ term (where $n$ is an odd positive integer); |
1624 |
but in the case of the dipole-dipole potential, there is one part |
1625 |
dependent on $r^{-3}$ and another dependent on $r^{-5}$. In order to |
1626 |
properly damping this potential, each of these parts is dampened with |
1627 |
separate damping functions. We can determine the necessary damping |
1628 |
functions by continuing with the multipole expansion; however, it |
1629 |
quickly becomes more complex with ``two-center'' systems, like the |
1630 |
dipole-dipole potential, and is typically approached with a spherical |
1631 |
harmonic formalism.\cite{Hirschfelder67} A simpler method for |
1632 |
determining these functions arises from adopting the tensor formalism |
1633 |
for expressing the electrostatic interactions.\cite{Stone02} |
1634 |
|
1635 |
The tensor formalism for electrostatic interactions involves obtaining |
1636 |
the multipole interactions from successive gradients of the monopole |
1637 |
potential. Thus, tensors of rank one through three are |
1638 |
\begin{equation} |
1639 |
T = \frac{1}{4\pi\epsilon_0r_{ij}}, |
1640 |
\label{eq:tensorRank1} |
1641 |
\end{equation} |
1642 |
\begin{equation} |
1643 |
T_\alpha = \frac{1}{4\pi\epsilon_0}\nabla_\alpha \frac{1}{r_{ij}}, |
1644 |
\label{eq:tensorRank2} |
1645 |
\end{equation} |
1646 |
\begin{equation} |
1647 |
T_{\alpha\beta} = \frac{1}{4\pi\epsilon_0} |
1648 |
\nabla_\alpha\nabla_\beta \frac{1}{r_{ij}}, |
1649 |
\label{eq:tensorRank3} |
1650 |
\end{equation} |
1651 |
where the form of the first tensor gives the monopole-monopole |
1652 |
potential, the second gives the monopole-dipole potential, and the |
1653 |
third gives the monopole-quadrupole and dipole-dipole |
1654 |
potentials.\cite{Stone02} Since the force is $-\nabla V$, the forces |
1655 |
for each potential come from the next higher tensor. |
1656 |
|
1657 |
To obtain the damped electrostatic forms, we replace $r^{-1}$ with |
1658 |
erfc$(\alpha r)\cdot r^{-1}$. Equation \ref{eq:tensorRank2} generates |
1659 |
$c_1(r_{ij})$, just like the multipole expansion, while equation |
1660 |
\ref{eq:tensorRank3} generates $c_2(r_{ij})$, where |
1661 |
\begin{equation} |
1662 |
c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}} |
1663 |
+ \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}} |
1664 |
+ \textrm{erfc}(\alpha r_{ij}). |
1665 |
\end{equation} |
1666 |
Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional |
1667 |
term. Continuing with higher rank tensors, we can obtain the damping |
1668 |
functions for higher multipoles as well as the forces. Each subsequent |
1669 |
damping function includes one additional term, and we can simplify the |
1670 |
procedure for obtaining these terms by writing out the following |
1671 |
generating function, |
1672 |
\begin{equation} |
1673 |
c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}} |
1674 |
{(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}), |
1675 |
\label{eq:dampingGeneratingFunc} |
1676 |
\end{equation} |
1677 |
where, |
1678 |
\begin{equation} |
1679 |
m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l} |
1680 |
m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\ |
1681 |
m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\ |
1682 |
1 & m = -1\textrm{ or }0, |
1683 |
\end{array}\right. |
1684 |
\label{eq:doubleFactorial} |
1685 |
\end{equation} |
1686 |
and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function |
1687 |
is similar in form to those obtained by researchers for the |
1688 |
application of the Ewald sum to |
1689 |
multipoles.\cite{Smith82,Smith98,Aguado03} |
1690 |
|
1691 |
Returning to the dipole-dipole example, the potential consists of a |
1692 |
portion dependent upon $r^{-5}$ and another dependent upon |
1693 |
$r^{-3}$. In the damped dipole-dipole potential, |
1694 |
\begin{equation} |
1695 |
V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}) |
1696 |
(\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}} |
1697 |
c_2(r_{ij}) - |
1698 |
\frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}} |
1699 |
c_1(r_{ij}), |
1700 |
\label{eq:dampDipoleDipole} |
1701 |
\end{equation} |
1702 |
$c_2(r_{ij})$ and $c_1(r_{ij})$ respectively dampen these two |
1703 |
parts. The forces for the damped dipole-dipole interaction, |
1704 |
\begin{equation} |
1705 |
\begin{split} |
1706 |
F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}) |
1707 |
(\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}} |
1708 |
c_3(r_{ij})\\ &- |
1709 |
3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j + |
1710 |
\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i + |
1711 |
\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}} |
1712 |
{r^5_{ij}} c_2(r_{ij}), |
1713 |
\end{split} |
1714 |
\label{eq:dampDipoleDipoleForces} |
1715 |
\end{equation} |
1716 |
rely on higher order damping functions because we perform another |
1717 |
gradient operation. In this manner, we can dampen higher order |
1718 |
multipolar interactions along with the monopole interactions, allowing |
1719 |
us to include multipoles in simulations involving damped electrostatic |
1720 |
interactions. |
1721 |
|
1722 |
|
1723 |
\section{Damping and Dielectric Constants}\label{sec:dampingDielectric} |
1724 |
|
1725 |
In section \ref{sec:t5peThermo}, we observed that the choice of |
1726 |
damping coefficient plays a major role in the calculated dielectric |
1727 |
constant. This is not too surprising given the results for damping |
1728 |
parameter influence on the long-time correlated motions of the NaCl |
1729 |
crystal in section \ref{sec:LongTimeDynamics}. The static dielectric |
1730 |
constant is calculated from the long-time fluctuations of the system's |
1731 |
accumulated dipole moment (Eq. (\ref{eq:staticDielectric})), so it is |
1732 |
going to be quite sensitive to the choice of damping parameter. We |
1733 |
would like to choose an optimal damping constant for any particular |
1734 |
cutoff radius choice that would properly capture the dielectric |
1735 |
behavior of the liquid. |
1736 |
|
1737 |
In order to find these optimal values, we mapped out the static |
1738 |
dielectric constant as a function of both the damping parameter and |
1739 |
cutoff radius for several different water models. To calculate the |
1740 |
static dielectric constant, we performed 5~ns $NPT$ calculations on |
1741 |
systems of 512 water molecules, using the TIP5P-E, TIP4P-Ew, SPC/E, |
1742 |
and SSD/RF water models. TIP4P-Ew is a reparametrized version of the |
1743 |
four-point transferable intermolecular potential (TIP4P) for water |
1744 |
targeted for use with the Ewald summation.\cite{Horn04} SSD/RF is the |
1745 |
reaction field modified variant of the soft sticky dipole (SSD) model |
1746 |
for water\cite{Fennell04} This model is discussed in more detail in |
1747 |
the next chapter. One thing to note about it, electrostatic |
1748 |
interactions are handled via dipole-dipole interactions rather than |
1749 |
charge-charge interactions like the other three models. Damping of the |
1750 |
dipole-dipole interaction was handled as described in section |
1751 |
\ref{sec:dampingMultipoles}. Each of these systems were studied with |
1752 |
cutoff radii of 9, 10, 11, and 12~\AA\ and with damping parameter values |
1753 |
ranging from 0 to 0.35~\AA$^{-1}$. |
1754 |
|
1755 |
\begin{figure} |
1756 |
\centering |
1757 |
\includegraphics[width=\linewidth]{./figures/dielectricMap.pdf} |
1758 |
\caption{The static dielectric constant for the TIP5P-E (A), TIP4P-Ew |
1759 |
(B), SPC/E (C), and SSD/RF (D) water models as a function of cutoff |
1760 |
radius ($R_\textrm{c}$) and damping coefficient ($\alpha$).} |
1761 |
\label{fig:dielectricMap} |
1762 |
\end{figure} |
1763 |
The results of these calculations are displayed in figure |
1764 |
\ref{fig:dielectricMap} in the form of shaded contour plots. An |
1765 |
interesting aspect of all four contour plots is that the dielectric |
1766 |
constant is effectively linear with respect to $\alpha$ and |
1767 |
$R_\textrm{c}$ in the low to moderate damping regions, and the slope |
1768 |
is 0.025~\AA$^{-1}\cdot R_\textrm{c}^{-1}$. Another point to note is |
1769 |
that choosing $\alpha$ and $R_\textrm{c}$ identical to those used in |
1770 |
studies with the Ewald summation results in the same calculated |
1771 |
dielectric constant. As an example, in the paper outlining the |
1772 |
development of TIP5P-E, the real-space cutoff and Ewald coefficient |
1773 |
were tethered to the system size, and for a 512 molecule system are |
1774 |
approximately 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04} |
1775 |
These parameters resulted in a dielectric constant of 92$\pm$14, while |
1776 |
with {\sc sf} these parameters give a dielectric constant of |
1777 |
90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where |
1778 |
$\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and |
1779 |
0.35~\AA$^{-1}$, and these parameters resulted in a $\epsilon_0$ equal |
1780 |
to 63$\pm$1.\cite{Horn04} We did not perform calculations with these |
1781 |
exact parameters, but interpolating between surrounding values gives a |
1782 |
$\epsilon_0$ of 61$\pm$1. Seeing a dependence of the dielectric |
1783 |
constant on $\alpha$ and $R_\textrm{c}$ with the {\sc sf} technique, |
1784 |
it might be interesting to investigate the dielectric dependence of |
1785 |
the real-space Ewald parameters. |
1786 |
|
1787 |
Although it is tempting to choose damping parameters equivalent to |
1788 |
these Ewald examples, the results discussed in sections |
1789 |
\ref{sec:EnergyResults} through \ref{sec:FTDirResults} and appendix |
1790 |
\ref{app:IndividualResults} indicate that values this high are |
1791 |
destructive to both the energetics and dynamics. Ideally, $\alpha$ |
1792 |
should not exceed 0.3~\AA$^{-1}$ for any of the cutoff values in this |
1793 |
range. If the optimal damping parameter is chosen to be midway between |
1794 |
0.275 and 0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff, |
1795 |
then the linear trend with $R_\textrm{c}$ will always keep $\alpha$ |
1796 |
below 0.3~\AA$^{-1}$. This linear progression would give values of |
1797 |
0.2875, 0.2625, 0.2375, and 0.2125~\AA$^{-1}$ for cutoff radii of 9, |
1798 |
10, 11, and 12~\AA. Setting this to be the default behavior for the |
1799 |
damped {\sc sf} technique will result in consistent dielectric |
1800 |
behavior for these and other condensed molecular systems, regardless |
1801 |
of the chosen cutoff radius. The static dielectric constants for |
1802 |
TIP5P-E, TIP4P-Ew, SPC/E, and SSD/RF will be fixed at approximately |
1803 |
74, 52, 58, and 89 respectively. These values are generally lower than |
1804 |
the values reported in the literature; however, the relative |
1805 |
dielectric behavior scales as expected when comparing the models to |
1806 |
one another. |
1807 |
|
1808 |
\section{Conclusions}\label{sec:PairwiseConclusions} |
1809 |
|
1810 |
The above investigation of pairwise electrostatic summation techniques |
1811 |
shows that there are viable and computationally efficient alternatives |
1812 |
to the Ewald summation. These methods are derived from the damped and |
1813 |
cutoff-neutralized Coulombic sum originally proposed by Wolf |
1814 |
\textit{et al.}\cite{Wolf99} In particular, the {\sc sf} |
1815 |
method, reformulated above as equations (\ref{eq:DSFPot}) and |
1816 |
(\ref{eq:DSFForces}), shows a remarkable ability to reproduce the |
1817 |
energetic and dynamic characteristics exhibited by simulations |
1818 |
employing lattice summation techniques. The cumulative energy |
1819 |
difference results showed the undamped {\sc sf} and moderately damped |
1820 |
{\sc sp} methods produced results nearly identical to the Ewald |
1821 |
summation. Similarly for the dynamic features, the undamped or |
1822 |
moderately damped {\sc sf} and moderately damped {\sc sp} methods |
1823 |
produce force and torque vector magnitude and directions very similar |
1824 |
to the expected values. These results translate into long-time |
1825 |
dynamic behavior equivalent to that produced in simulations using the |
1826 |
Ewald summation. A detailed study of water simulations showed that |
1827 |
liquid properties calculated when using {\sc sf} will also be |
1828 |
equivalent to those obtained using the Ewald summation. |
1829 |
|
1830 |
As in all purely-pairwise cutoff methods, these methods are expected |
1831 |
to scale approximately {\it linearly} with system size, and they are |
1832 |
easily parallelizable. This should result in substantial reductions |
1833 |
in the computational cost of performing large simulations. |
1834 |
|
1835 |
Aside from the computational cost benefit, these techniques have |
1836 |
applicability in situations where the use of the Ewald sum can prove |
1837 |
problematic. Of greatest interest is their potential use in |
1838 |
interfacial systems, where the unmodified lattice sum techniques |
1839 |
artificially accentuate the periodicity of the system in an |
1840 |
undesirable manner. There have been alterations to the standard Ewald |
1841 |
techniques, via corrections and reformulations, to compensate for |
1842 |
these systems; but the pairwise techniques discussed here require no |
1843 |
modifications, making them natural tools to tackle these problems. |
1844 |
Additionally, this transferability gives them benefits over other |
1845 |
pairwise methods, like reaction field, because estimations of physical |
1846 |
properties (e.g. the dielectric constant) are unnecessary. |
1847 |
|
1848 |
If a researcher is using Monte Carlo simulations of large chemical |
1849 |
systems containing point charges, most structural features will be |
1850 |
accurately captured using the undamped {\sc sf} method or the {\sc sp} |
1851 |
method with an electrostatic damping of 0.2~\AA$^{-1}$. These methods |
1852 |
would also be appropriate for molecular dynamics simulations where the |
1853 |
data of interest is either structural or short-time dynamical |
1854 |
quantities. For long-time dynamics and collective motions, the safest |
1855 |
pairwise method we have evaluated is the {\sc sf} method with an |
1856 |
electrostatic damping between 0.2 and 0.25~\AA$^{-1}$. It is also |
1857 |
important to note that the static dielectric constant in water |
1858 |
simulations is highly dependent on both $\alpha$ and |
1859 |
$R_\textrm{c}$. For consistent dielectric behavior, the damped {\sc |
1860 |
sf} method should use an $\alpha$ of 0.2175~\AA$^{-1}$ for an |
1861 |
$R_\textrm{c}$ of 12~\AA, and $\alpha$ should decrease by |
1862 |
0.025~\AA$^{-1}$ for every 1~\AA\ increase in cutoff radius. |
1863 |
|
1864 |
We are not suggesting that there is any flaw with the Ewald sum; in |
1865 |
fact, it is the standard by which these simple pairwise sums have been |
1866 |
judged. However, these results do suggest that in the typical |
1867 |
simulations performed today, the Ewald summation may no longer be |
1868 |
required to obtain the level of accuracy most researchers have come to |
1869 |
expect. |