ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
(Generate patch)

Comparing trunk/electrostaticMethodsPaper/electrostaticMethods.tex (file contents):
Revision 2641 by chrisfen, Mon Mar 20 15:43:13 2006 UTC vs.
Revision 2643 by gezelter, Mon Mar 20 17:32:33 2006 UTC

# Line 65 | Line 65 | interactions is considered one of the most essential a
65   \section{Introduction}
66  
67   In molecular simulations, proper accumulation of the electrostatic
68 < interactions is considered one of the most essential and
69 < computationally demanding tasks.  The common molecular mechanics force
70 < fields are founded on representation of the atomic sites centered on
71 < full or partial charges shielded by Lennard-Jones type interactions.
72 < This means that nearly every pair interaction involves an
73 < charge-charge calculation.  Coupled with $r^{-1}$ decay, the monopole
74 < interactions quickly become a burden for molecular systems of all
75 < sizes.  For example, in small systems, the electrostatic pair
76 < interaction may not have decayed appreciably within the box length
77 < leading to an effect excluded from the pair interactions within a unit
78 < box.  In large systems, excessively large cutoffs need to be used to
79 < accurately incorporate their effect, and since the computational cost
80 < increases proportionally with the cutoff sphere, it quickly becomes
81 < very time-consuming to perform these calculations.
68 > interactions is essential and is one of the most
69 > computationally-demanding tasks.  The common molecular mechanics force
70 > fields represent atomic sites with full or partial charges protected
71 > by Lennard-Jones (short range) interactions.  This means that nearly
72 > every pair interaction involves a calculation of charge-charge forces.
73 > Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
74 > interactions quickly become the most expensive part of molecular
75 > simulations.  Historically, the electrostatic pair interaction would
76 > not have decayed appreciably within the typical box lengths that could
77 > be feasibly simulated.  In the larger systems that are more typical of
78 > modern simulations, large cutoffs should be used to incorporate
79 > electrostatics correctly.
80  
81 < There have been many efforts to address this issue of both proper and
82 < practical handling of electrostatic interactions, and these have
83 < resulted in the availability of a variety of
84 < techniques.\cite{Roux99,Sagui99,Tobias01} These are typically
85 < classified as implicit methods (i.e., continuum dielectrics, static
86 < dipolar fields),\cite{Born20,Grossfield00} explicit methods (i.e.,
89 < Ewald summations, interaction shifting or
81 > There have been many efforts to address the proper and practical
82 > handling of electrostatic interactions, and these have resulted in a
83 > variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
84 > typically classified as implicit methods (i.e., continuum dielectrics,
85 > static dipolar fields),\cite{Born20,Grossfield00} explicit methods
86 > (i.e., Ewald summations, interaction shifting or
87   truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
88   reaction field type methods, fast multipole
89   methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
90 < often preferred because they incorporate dynamic solvent molecules in
91 < the system of interest, but these methods are sometimes difficult to
92 < utilize because of their high computational cost.\cite{Roux99} In
93 < addition to this cost, there has been some question of the inherent
94 < periodicity of the explicit Ewald summation artificially influencing
95 < systems dynamics.\cite{Tobias01}
90 > often preferred because they physically incorporate solvent molecules
91 > in the system of interest, but these methods are sometimes difficult
92 > to utilize because of their high computational cost.\cite{Roux99} In
93 > addition to the computational cost, there have been some questions
94 > regarding possible artifacts caused by the inherent periodicity of the
95 > explicit Ewald summation.\cite{Tobias01}
96  
97 < In this paper, we focus on the common mixed and explicit methods of
98 < reaction filed and smooth particle mesh
99 < Ewald\cite{Onsager36,Essmann99} and a new set of shifted methods
100 < devised by Wolf {\it et al.} which we further extend.\cite{Wolf99}
101 < These new methods for handling electrostatics are quite
102 < computationally efficient, since they involve only a simple
103 < modification to the direct pairwise sum, and they lack the added
104 < periodicity of the Ewald sum. Below, these methods are evaluated using
105 < a variety of model systems and comparison methodologies to establish
97 > In this paper, we focus on a new set of shifted methods devised by
98 > Wolf {\it et al.},\cite{Wolf99} which we further extend.  These
99 > methods along with a few other mixed methods (i.e. reaction field) are
100 > compared with the smooth particle mesh Ewald
101 > sum,\cite{Onsager36,Essmann99} which is our reference method for
102 > handling long-range electrostatic interactions. The new methods for
103 > handling electrostatics have the potential to scale linearly with
104 > increasing system size since they involve only a simple modification
105 > to the direct pairwise sum.  They also lack the added periodicity of
106 > the Ewald sum, so they can be used for systems which are non-periodic
107 > or which have one- or two-dimensional periodicity.  Below, these
108 > methods are evaluated using a variety of model systems to establish
109   their usability in molecular simulations.
110  
111   \subsection{The Ewald Sum}
112   The complete accumulation electrostatic interactions in a system with
113   periodic boundary conditions (PBC) requires the consideration of the
114 < effect of all charges within a simulation box, as well as those in the
115 < periodic replicas,
114 > effect of all charges within a (cubic) simulation box as well as those
115 > in the periodic replicas,
116   \begin{equation}
117   V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime \left[ \sum_{i=1}^N\sum_{j=1}^N \phi\left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right) \right],
118   \label{eq:PBCSum}
# Line 123 | Line 123 | $j$, and $\phi$ is Poisson's equation ($\phi(\mathbf{r
123   0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
124   particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
125   the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
126 < $j$, and $\phi$ is Poisson's equation ($\phi(\mathbf{r}_{ij}) = q_i
127 < q_j |\mathbf{r}_{ij}|^{-1}$ for charge-charge interactions). In the
128 < case of monopole electrostatics, eq. (\ref{eq:PBCSum}) is
129 < conditionally convergent and is discontinuous for non-neutral systems.
126 > $j$, and $\phi$ is the solution to Poisson's equation
127 > ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
128 > charge-charge interactions). In the case of monopole electrostatics,
129 > eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
130 > non-neutral systems.
131  
132 < This electrostatic summation problem was originally studied by Ewald
132 > The electrostatic summation problem was originally studied by Ewald
133   for the case of an infinite crystal.\cite{Ewald21}. The approach he
134   took was to convert this conditionally convergent sum into two
135   absolutely convergent summations: a short-ranged real-space summation
# Line 140 | Line 141 | units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal v
141   \label{eq:EwaldSum}
142   \end{equation}
143   where $\alpha$ is a damping parameter, or separation constant, with
144 < units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and equal
145 < $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
146 < constant of the encompassing medium. The final two terms of
144 > units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are
145 > equal to $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the
146 > dielectric constant of the surrounding medium. The final two terms of
147   eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
148   for interacting with a surrounding dielectric.\cite{Allen87} This
149   dipolar term was neglected in early applications in molecular
150   simulations,\cite{Brush66,Woodcock71} until it was introduced by de
151   Leeuw {\it et al.} to address situations where the unit cell has a
152 < dipole moment and this dipole moment gets magnified through
153 < replication of the periodic images.\cite{deLeeuw80,Smith81} If this
154 < term is taken to be zero, the system is using conducting boundary
152 > dipole moment which is magnified through replication of the periodic
153 > images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
154 > system is said to be using conducting (or ``tin-foil'') boundary
155   conditions, $\epsilon_{\rm S} = \infty$. Figure
156   \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
157 < time.  Initially, due to the small size of systems, the entire
158 < simulation box was replicated to convergence.  Currently, we balance a
159 < spherical real-space cutoff with the reciprocal sum and consider the
160 < surrounding dielectric.
157 > time.  Initially, due to the small sizes of the systems that could be
158 > feasibly simulated, the entire simulation box was replicated to
159 > convergence.  In more modern simulations, the simulation boxes have
160 > grown large enough that a real-space cutoff could potentially give
161 > convergent behavior.  Indeed, it has often been observed that the
162 > reciprocal-space portion of the Ewald sum can be vanishingly
163 > small compared to the real-space portion.\cite{XXX}
164 >
165   \begin{figure}
166   \centering
167   \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
# Line 170 | Line 175 | The Ewald summation in the straight-forward form is an
175   \label{fig:ewaldTime}
176   \end{figure}
177  
178 < The Ewald summation in the straight-forward form is an
179 < $\mathscr{O}(N^2)$ algorithm.  The separation constant $(\alpha)$
180 < plays an important role in the computational cost balance between the
181 < direct and reciprocal-space portions of the summation.  The choice of
182 < the magnitude of this value allows one to select whether the
183 < real-space or reciprocal space portion of the summation is an
184 < $\mathscr{O}(N^2)$ calculation (with the other being
185 < $\mathscr{O}(N)$).\cite{Sagui99} With appropriate choice of $\alpha$
186 < and thoughtful algorithm development, this cost can be brought down to
187 < $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route taken to
188 < reduce the cost of the Ewald summation further is to set $\alpha$ such
189 < that the real-space interactions decay rapidly, allowing for a short
190 < spherical cutoff, and then optimize the reciprocal space summation.
191 < These optimizations usually involve the utilization of the fast
187 < Fourier transform (FFT),\cite{Hockney81} leading to the
178 > The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm.  The
179 > separation constant $(\alpha)$ plays an important role in balancing
180 > the computational cost between the direct and reciprocal-space
181 > portions of the summation.  The choice of this value allows one to
182 > select whether the real-space or reciprocal space portion of the
183 > summation is an $\mathscr{O}(N^2)$ calculation (with the other being
184 > $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
185 > $\alpha$ and thoughtful algorithm development, this cost can be
186 > reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
187 > taken to reduce the cost of the Ewald summation even further is to set
188 > $\alpha$ such that the real-space interactions decay rapidly, allowing
189 > for a short spherical cutoff. Then the reciprocal space summation is
190 > optimized.  These optimizations usually involve utilization of the
191 > fast Fourier transform (FFT),\cite{Hockney81} leading to the
192   particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
193   methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
194   methods, the cost of the reciprocal-space portion of the Ewald
195 < summation is from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N \log N)$.
195 > summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
196 > \log N)$.
197  
198 < These developments and optimizations have led the use of the Ewald
199 < summation to become routine in simulations with periodic boundary
200 < conditions. However, in certain systems the intrinsic three
201 < dimensional periodicity can prove to be problematic, such as two
202 < dimensional surfaces and membranes.  The Ewald sum has been
203 < reformulated to handle 2D
204 < systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the new
205 < methods have been found to be computationally
206 < expensive.\cite{Spohr97,Yeh99} Inclusion of a correction term in the
207 < full Ewald summation is a possible direction for enabling the handling
203 < of 2D systems and the inclusion of the optimizations described
204 < previously.\cite{Yeh99}
198 > These developments and optimizations have made the use of the Ewald
199 > summation routine in simulations with periodic boundary
200 > conditions. However, in certain systems, such as vapor-liquid
201 > interfaces and membranes, the intrinsic three-dimensional periodicity
202 > can prove problematic.  The Ewald sum has been reformulated to handle
203 > 2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
204 > new methods are computationally expensive.\cite{Spohr97,Yeh99}
205 > Inclusion of a correction term in the Ewald summation is a possible
206 > direction for handling 2D systems while still enabling the use of the
207 > modern optimizations.\cite{Yeh99}
208  
209   Several studies have recognized that the inherent periodicity in the
210 < Ewald sum can also have an effect on systems that have the same
211 < dimensionality.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
212 < Good examples are solvated proteins kept at high relative
213 < concentration due to the periodicity of the electrostatics.  In these
210 > Ewald sum can also have an effect on three-dimensional
211 > systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
212 > Solvated proteins are essentially kept at high concentration due to
213 > the periodicity of the electrostatic summation method.  In these
214   systems, the more compact folded states of a protein can be
215   artificially stabilized by the periodic replicas introduced by the
216 < Ewald summation.\cite{Weber00} Thus, care ought to be taken when
217 < considering the use of the Ewald summation where the intrinsic
218 < periodicity may negatively affect the system dynamics.
216 > Ewald summation.\cite{Weber00} Thus, care must be taken when
217 > considering the use of the Ewald summation where the assumed
218 > periodicity would introduce spurious effects in the system dynamics.
219  
217
220   \subsection{The Wolf and Zahn Methods}
221   In a recent paper by Wolf \textit{et al.}, a procedure was outlined
222   for the accurate accumulation of electrostatic interactions in an
223 < efficient pairwise fashion and lacks the inherent periodicity of the
224 < Ewald summation.\cite{Wolf99} Wolf \textit{et al.} observed that the
225 < electrostatic interaction is effectively short-ranged in condensed
226 < phase systems and that neutralization of the charge contained within
227 < the cutoff radius is crucial for potential stability. They devised a
228 < pairwise summation method that ensures charge neutrality and gives
229 < results similar to those obtained with the Ewald summation.  The
230 < resulting shifted Coulomb potential (Eq. \ref{eq:WolfPot}) includes
231 < image-charges subtracted out through placement on the cutoff sphere
232 < and a distance-dependent damping function (identical to that seen in
233 < the real-space portion of the Ewald sum) to aid convergence
223 > efficient pairwise fashion.  This procedure lacks the inherent
224 > periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
225 > observed that the electrostatic interaction is effectively
226 > short-ranged in condensed phase systems and that neutralization of the
227 > charge contained within the cutoff radius is crucial for potential
228 > stability. They devised a pairwise summation method that ensures
229 > charge neutrality and gives results similar to those obtained with the
230 > Ewald summation.  The resulting shifted Coulomb potential
231 > (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
232 > placement on the cutoff sphere and a distance-dependent damping
233 > function (identical to that seen in the real-space portion of the
234 > Ewald sum) to aid convergence
235   \begin{equation}
236   V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
237   \label{eq:WolfPot}
# Line 255 | Line 258 | Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) toge
258   force expressions for use in simulations involving water.\cite{Zahn02}
259   In their work, they pointed out that the forces and derivative of
260   the potential are not commensurate.  Attempts to use both
261 < Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
261 > eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
262   to poor energy conservation.  They correctly observed that taking the
263   limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
264   derivatives gives forces for a different potential energy function
265 < than the one shown in Eq. (\ref{eq:WolfPot}).
265 > than the one shown in eq. (\ref{eq:WolfPot}).
266  
267 < Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
268 < method'' as a way to use this technique in Molecular Dynamics
269 < simulations.  Taking the integral of the forces shown in equation
267 < \ref{eq:WolfForces}, they proposed a new damped Coulomb
268 < potential,
267 > Zahn \textit{et al.} introduced a modified form of this summation
268 > method as a way to use the technique in Molecular Dynamics
269 > simulations.  They proposed a new damped Coulomb potential,
270   \begin{equation}
271 < V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\}.
271 > V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\},
272   \label{eq:ZahnPot}
273   \end{equation}
274 < They showed that this potential does fairly well at capturing the
274 > and showed that this potential does fairly well at capturing the
275   structural and dynamic properties of water compared the same
276   properties obtained using the Ewald sum.
277  
# Line 301 | Line 302 | v_\textrm{SP}(r) =     \begin{cases}
302   \textit{et al.}  and Zahn \textit{et al.} by considering the standard
303   shifted potential,
304   \begin{equation}
305 < v_\textrm{SP}(r) =      \begin{cases}
305 > V_\textrm{SP}(r) =      \begin{cases}
306   v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
307   R_\textrm{c}  
308   \end{cases},
# Line 309 | Line 310 | v_\textrm{SF}(r) =     \begin{cases}
310   \end{equation}
311   and shifted force,
312   \begin{equation}
313 < v_\textrm{SF}(r) =      \begin{cases}
313 > V_\textrm{SF}(r) =      \begin{cases}
314   v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
315   &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
316                                                  \end{cases},
# Line 325 | Line 326 | f_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
326   The forces associated with the shifted potential are simply the forces
327   of the unshifted potential itself (when inside the cutoff sphere),
328   \begin{equation}
329 < f_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
329 > F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
330   \end{equation}
331   and are zero outside.  Inside the cutoff sphere, the forces associated
332   with the shifted force form can be written,
333   \begin{equation}
334 < f_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
334 > F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
335   v(r)}{dr} \right)_{r=R_\textrm{c}}.
336   \end{equation}
337  
338 < If the potential ($v(r)$) is taken to be the normal Coulomb potential,
338 > If the potential, $v(r)$, is taken to be the normal Coulomb potential,
339   \begin{equation}
340   v(r) = \frac{q_i q_j}{r},
341   \label{eq:Coulomb}
# Line 342 | Line 343 | v_\textrm{SP}(r) =
343   then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
344   al.}'s undamped prescription:
345   \begin{equation}
346 < v_\textrm{SP}(r) =
346 > V_\textrm{SP}(r) =
347   q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
348   r\leqslant R_\textrm{c},
349   \label{eq:SPPot}
350   \end{equation}
351   with associated forces,
352   \begin{equation}
353 < f_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
353 > F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
354   \label{eq:SPForces}
355   \end{equation}
356   These forces are identical to the forces of the standard Coulomb
# Line 364 | Line 365 | v_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_
365   The shifted force ({\sc sf}) form using the normal Coulomb potential
366   will give,
367   \begin{equation}
368 < v_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}.
368 > V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}.
369   \label{eq:SFPot}
370   \end{equation}
371   with associated forces,
372   \begin{equation}
373 < f_\textrm{SF}(r) =  q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}.
373 > F_\textrm{SF}(r) =  q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}.
374   \label{eq:SFForces}
375   \end{equation}
376   This formulation has the benefits that there are no discontinuities at
377 < the cutoff distance, while the neutralizing image charges are present
378 < in both the energy and force expressions.  It would be simple to add
379 < the self-neutralizing term back when computing the total energy of the
377 > the cutoff radius, while the neutralizing image charges are present in
378 > both the energy and force expressions.  It would be simple to add the
379 > self-neutralizing term back when computing the total energy of the
380   system, thereby maintaining the agreement with the Madelung energies.
381   A side effect of this treatment is the alteration in the shape of the
382   potential that comes from the derivative term.  Thus, a degree of
# Line 383 | Line 384 | shifted Coulomb potential (Eq. \ref{eq:SPPot}), and th
384   to gain functionality in dynamics simulations.
385  
386   Wolf \textit{et al.} originally discussed the energetics of the
387 < shifted Coulomb potential (Eq. \ref{eq:SPPot}), and they found that
388 < it was still insufficient for accurate determination of the energy
389 < with reasonable cutoff distances.  The calculated Madelung energies
390 < fluctuate around the expected value with increasing cutoff radius, but
391 < the oscillations converge toward the correct value.\cite{Wolf99} A
387 > shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
388 > insufficient for accurate determination of the energy with reasonable
389 > cutoff distances.  The calculated Madelung energies fluctuated around
390 > the expected value as the cutoff radius was increased, but the
391 > oscillations converged toward the correct value.\cite{Wolf99} A
392   damping function was incorporated to accelerate the convergence; and
393 < though alternative functional forms could be
393 > though alternative forms for the damping function could be
394   used,\cite{Jones56,Heyes81} the complimentary error function was
395   chosen to mirror the effective screening used in the Ewald summation.
396   Incorporating this error function damping into the simple Coulomb
# Line 398 | Line 399 | the shifted potential (Eq. (\ref{eq:SPPot})) can be re
399   v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
400   \label{eq:dampCoulomb}
401   \end{equation}
402 < the shifted potential (Eq. (\ref{eq:SPPot})) can be reacquired using
402 < eq. (\ref{eq:shiftingForm}),
402 > the shifted potential (eq. (\ref{eq:SPPot})) becomes
403   \begin{equation}
404 < v_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c},
404 > V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c},
405   \label{eq:DSPPot}
406   \end{equation}
407   with associated forces,
408   \begin{equation}
409 < f_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}.
409 > F_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}.
410   \label{eq:DSPForces}
411   \end{equation}
412 < Again, this damped shifted potential suffers from a discontinuity and
413 < a lack of the image charges in the forces.  To remedy these concerns,
414 < one may derive a {\sc sf} variant by including  the derivative
415 < term in eq. (\ref{eq:shiftingForm}),
412 > Again, this damped shifted potential suffers from a
413 > force-discontinuity at the cutoff radius, and the image charges play
414 > no role in the forces.  To remedy these concerns, one may derive a
415 > {\sc sf} variant by including the derivative term in
416 > eq. (\ref{eq:shiftingForm}),
417   \begin{equation}
418   \begin{split}
419 < v_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r\right)}{r}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ &\left.+\left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right)\left(r-R_\mathrm{c}\right)\ \right] \quad r\leqslant R_\textrm{c}.
419 > V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r\right)}{r}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ &\left.+\left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right)\left(r-R_\mathrm{c}\right)\ \right] \quad r\leqslant R_\textrm{c}.
420   \label{eq:DSFPot}
421   \end{split}
422   \end{equation}
423   The derivative of the above potential will lead to the following forces,
424   \begin{equation}
425   \begin{split}
426 < f_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
426 > F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
427   \label{eq:DSFForces}
428   \end{split}
429   \end{equation}
430 < If the damping parameter $(\alpha)$ is chosen to be zero, the undamped
431 < case, eqs. (\ref{eq:SPPot}-\ref{eq:SFForces}) are correctly recovered
432 < from eqs. (\ref{eq:DSPPot}-\ref{eq:DSFForces}).
430 > If the damping parameter $(\alpha)$ is set to zero, the undamped case,
431 > eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
432 > recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
433  
434   This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
435   derived by Zahn \textit{et al.}; however, there are two important
# Line 440 | Line 441 | $R_c$).  The sign problem would be a potential source
441   portion is different.  The missing $v_\textrm{c}$ term would not
442   affect molecular dynamics simulations (although the computed energy
443   would be expected to have sudden jumps as particle distances crossed
444 < $R_c$).  The sign problem would be a potential source of errors,
445 < however.  In fact, it introduces a discontinuity in the forces at the
446 < cutoff, because the force function is shifted in the wrong direction
447 < and doesn't cross zero at $R_\textrm{c}$.
444 > $R_c$).  The sign problem is a potential source of errors, however.
445 > In fact, it introduces a discontinuity in the forces at the cutoff,
446 > because the force function is shifted in the wrong direction and
447 > doesn't cross zero at $R_\textrm{c}$.
448  
449   Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
450 < electrostatic summation method that is continuous in both the
451 < potential and forces and which incorporates the damping function
452 < proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
453 < paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
454 < sf}, damping) are at reproducing the correct electrostatic summation
455 < performed by the Ewald sum.
450 > electrostatic summation method in which the potential and forces are
451 > continuous at the cutoff radius and which incorporates the damping
452 > function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
453 > this paper, we will evaluate exactly how good these methods ({\sc sp},
454 > {\sc sf}, damping) are at reproducing the correct electrostatic
455 > summation performed by the Ewald sum.
456  
457   \subsection{Other alternatives}
458 < In addition to the methods described above, we will consider some
459 < other techniques that commonly get used in molecular simulations.  The
458 > In addition to the methods described above, we considered some other
459 > techniques that are commonly used in molecular simulations.  The
460   simplest of these is group-based cutoffs.  Though of little use for
461 < non-neutral molecules, collecting atoms into neutral groups takes
461 > charged molecules, collecting atoms into neutral groups takes
462   advantage of the observation that the electrostatic interactions decay
463   faster than those for monopolar pairs.\cite{Steinbach94} When
464 < considering these molecules as groups, an orientational aspect is
465 < introduced to the interactions.  Consequently, as these molecular
466 < particles move through $R_\textrm{c}$, the energy will drift upward
467 < due to the anisotropy of the net molecular dipole
468 < interactions.\cite{Rahman71} To maintain good energy conservation,
469 < both the potential and derivative need to be smoothly switched to zero
470 < at $R_\textrm{c}$.\cite{Adams79} This is accomplished using a
471 < switching function,
472 < \begin{equation}
473 < S(r) = \begin{cases} 1 &\quad r\leqslant r_\textrm{sw} \\
473 < \frac{(R_\textrm{c}+2r-3r_\textrm{sw})(R_\textrm{c}-r)^2}{(R_\textrm{c}-r_\textrm{sw})^3} &\quad r_\textrm{sw}<r\leqslant R_\textrm{c} \\
474 < 0 &\quad r>R_\textrm{c}
475 < \end{cases},
476 < \end{equation}
477 < where the above form is for a cubic function.  If a smooth second
478 < derivative is desired, a fifth (or higher) order polynomial can be
479 < used.\cite{Andrea83}
464 > considering these molecules as neutral groups, the relative
465 > orientations of the molecules control the strength of the interactions
466 > at the cutoff radius.  Consequently, as these molecular particles move
467 > through $R_\textrm{c}$, the energy will drift upward due to the
468 > anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
469 > maintain good energy conservation, both the potential and derivative
470 > need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
471 > This is accomplished using a standard switching function.  If a smooth
472 > second derivative is desired, a fifth (or higher) order polynomial can
473 > be used.\cite{Andrea83}
474  
475   Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
476 < and to incorporate their effect, a method like Reaction Field ({\sc
477 < rf}) can be used.  The original theory for {\sc rf} was originally
478 < developed by Onsager,\cite{Onsager36} and it was applied in
479 < simulations for the study of water by Barker and Watts.\cite{Barker73}
480 < In application, it is simply an extension of the group-based cutoff
481 < method where the net dipole within the cutoff sphere polarizes an
482 < external dielectric, which reacts back on the central dipole.  The
483 < same switching function considerations for group-based cutoffs need to
484 < made for {\sc rf}, with the additional pre-specification of a
485 < dielectric constant.
476 > and to incorporate the effects of the surroundings, a method like
477 > Reaction Field ({\sc rf}) can be used.  The original theory for {\sc
478 > rf} was originally developed by Onsager,\cite{Onsager36} and it was
479 > applied in simulations for the study of water by Barker and
480 > Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
481 > an extension of the group-based cutoff method where the net dipole
482 > within the cutoff sphere polarizes an external dielectric, which
483 > reacts back on the central dipole.  The same switching function
484 > considerations for group-based cutoffs need to made for {\sc rf}, with
485 > the additional pre-specification of a dielectric constant.
486  
487   \section{Methods}
488  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines