Import cmake Modules/FindCUDA.cmake
[gromacs.git] / docs / manual / forcefield.tex
blob8574bc409459fe2b117bfe8df3b7c4d1ee615ba4
2 % This file is part of the GROMACS molecular simulation package.
4 % Copyright (c) 2013,2014,2015,2016,2017, by the GROMACS development team, led by
5 % Mark Abraham, David van der Spoel, Berk Hess, and Erik Lindahl,
6 % and including many others, as listed in the AUTHORS file in the
7 % top-level source directory and at http://www.gromacs.org.
9 % GROMACS is free software; you can redistribute it and/or
10 % modify it under the terms of the GNU Lesser General Public License
11 % as published by the Free Software Foundation; either version 2.1
12 % of the License, or (at your option) any later version.
14 % GROMACS is distributed in the hope that it will be useful,
15 % but WITHOUT ANY WARRANTY; without even the implied warranty of
16 % MERCHANTABILITY or FITNESS FOR A PARTICULAR PURPOSE. See the GNU
17 % Lesser General Public License for more details.
19 % You should have received a copy of the GNU Lesser General Public
20 % License along with GROMACS; if not, see
21 % http://www.gnu.org/licenses, or write to the Free Software Foundation,
22 % Inc., 51 Franklin Street, Fifth Floor, Boston, MA 02110-1301 USA.
24 % If you want to redistribute modifications to GROMACS, please
25 % consider that scientific software is very special. Version
26 % control is crucial - bugs must be traceable. We will be happy to
27 % consider code for inclusion in the official distribution, but
28 % derived work must not be called official GROMACS. Details are found
29 % in the README & COPYING files - if they are missing, get the
30 % official version at http://www.gromacs.org.
32 % To help us fund GROMACS development, we humbly ask that you cite
33 % the research papers on the package. Check out http://www.gromacs.org.
35 \chapter{Interaction function and force fields\index{force field}}
36 \label{ch:ff}
37 To accommodate the potential functions used
38 in some popular force fields (see \ref{sec:ff}), {\gromacs} offers a choice of functions,
39 both for non-bonded interaction and for dihedral interactions. They
40 are described in the appropriate subsections.
42 The potential functions can be subdivided into three parts
43 \begin{enumerate}
44 \item {\em Non-bonded}: Lennard-Jones or Buckingham, and Coulomb or
45 modified Coulomb. The non-bonded interactions are computed on the
46 basis of a neighbor list (a list of non-bonded atoms within a certain
47 radius), in which exclusions are already removed.
48 \item {\em Bonded}: covalent bond-stretching, angle-bending,
49 improper dihedrals, and proper dihedrals. These are computed on the
50 basis of fixed lists.
51 \item {\em Restraints}: position restraints, angle restraints,
52 distance restraints, orientation restraints and dihedral restraints, all
53 based on fixed lists.
54 \item {\em Applied Forces}:
55 externally applied forces, see \chref{special}.
56 \end{enumerate}
58 \section{Non-bonded interactions}
59 Non-bonded interactions in {\gromacs} are pair-additive and centro-symmetric:
60 \beq
61 V(\ve{r}_1,\ldots \ve{r}_N) = \sum_{i<j}V_{ij}(\rvij);
62 \eeq
63 \beq
64 \ve{F}_i = -\sum_j \frac{dV_{ij}(r_{ij})}{dr_{ij}} \frac{\rvij}{r_{ij}} = -\ve{F}_j
65 \eeq
66 The non-bonded interactions contain a \normindex{repulsion} term,
67 a \normindex{dispersion}
68 term, and a Coulomb term. The repulsion and dispersion term are
69 combined in either the Lennard-Jones (or 6-12 interaction), or the
70 Buckingham (or exp-6 potential). In addition, (partially) charged atoms
71 act through the Coulomb term.
73 \subsection{The Lennard-Jones interaction}
74 \label{sec:lj}
75 The \normindex{Lennard-Jones} potential $V_{LJ}$ between two atoms equals:
76 \beq
77 V_{LJ}(\rij) = \frac{C_{ij}^{(12)}}{\rij^{12}} -
78 \frac{C_{ij}^{(6)}}{\rij^6}
79 \eeq
80 See also \figref{lj}
81 The parameters $C^{(12)}_{ij}$ and $C^{(6)}_{ij}$ depend on pairs of
82 {\em atom types}; consequently they are taken from a matrix of
83 LJ-parameters. In the Verlet cut-off scheme, the potential is shifted
84 by a constant such that it is zero at the cut-off distance.
86 \begin{figure}
87 \centerline{\includegraphics[width=8cm]{plots/f-lj}}
88 \caption {The Lennard-Jones interaction.}
89 \label{fig:lj}
90 \end{figure}
92 The force derived from this potential is:
93 \beq
94 \ve{F}_i(\rvij) = \left( 12~\frac{C_{ij}^{(12)}}{\rij^{13}} -
95 6~\frac{C_{ij}^{(6)}}{\rij^7} \right) \rnorm
96 \eeq
98 The LJ potential may also be written in the following form:
99 \beq
100 V_{LJ}(\rvij) = 4\epsilon_{ij}\left(\left(\frac{\sigma_{ij}} {\rij}\right)^{12}
101 - \left(\frac{\sigma_{ij}}{\rij}\right)^{6} \right)
102 \label{eqn:sigeps}
103 \eeq
105 In constructing the parameter matrix for the non-bonded LJ-parameters,
106 two types of \normindex{combination rule}s can be used within {\gromacs},
107 only geometric averages (type 1 in the input section of the force-field file):
108 \beq
109 \begin{array}{rcl}
110 C_{ij}^{(6)} &=& \left({C_{ii}^{(6)} \, C_{jj}^{(6)}}\right)^{1/2} \\
111 C_{ij}^{(12)} &=& \left({C_{ii}^{(12)} \, C_{jj}^{(12)}}\right)^{1/2}
112 \label{eqn:comb}
113 \end{array}
114 \eeq
115 or, alternatively the Lorentz-Berthelot rules can be used. An arithmetic average is used to calculate $\sigma_{ij}$, while a geometric average is used to calculate $\epsilon_{ij}$ (type 2):
116 \beq
117 \begin{array}{rcl}
118 \sigma_{ij} &=& \frac{1}{ 2}(\sigma_{ii} + \sigma_{jj}) \\
119 \epsilon_{ij} &=& \left({\epsilon_{ii} \, \epsilon_{jj}}\right)^{1/2}
120 \label{eqn:lorentzberthelot}
121 \end{array}
122 \eeq
123 finally an geometric average for both parameters can be used (type 3):
124 \beq
125 \begin{array}{rcl}
126 \sigma_{ij} &=& \left({\sigma_{ii} \, \sigma_{jj}}\right)^{1/2} \\
127 \epsilon_{ij} &=& \left({\epsilon_{ii} \, \epsilon_{jj}}\right)^{1/2}
128 \end{array}
129 \eeq
130 This last rule is used by the OPLS force field.
133 \subsection{\normindex{Buckingham potential}}
134 The Buckingham
135 potential has a more flexible and realistic repulsion term
136 than the Lennard-Jones interaction, but is also more expensive to
137 compute. The potential form is:
138 \beq
139 V_{bh}(\rij) = A_{ij} \exp(-B_{ij} \rij) -
140 \frac{C_{ij}}{\rij^6}
141 \eeq
142 \begin{figure}
143 \centerline{\includegraphics[width=8cm]{plots/f-bham}}
144 \caption {The Buckingham interaction.}
145 \label{fig:bham}
146 \end{figure}
148 See also \figref{bham}. The force derived from this is:
149 \beq
150 \ve{F}_i(\rij) = \left[ A_{ij}B_{ij}\exp(-B_{ij} \rij) -
151 6\frac{C_{ij}}{\rij^7} \right] \rnorm
152 \eeq
155 \subsection{Coulomb interaction}
156 \label{sec:coul}
157 \newcommand{\epsr}{\varepsilon_r}
158 \newcommand{\epsrf}{\varepsilon_{rf}}
159 The \normindex{Coulomb} interaction between two charge particles is given by:
160 \beq
161 V_c(\rij) = f \frac{q_i q_j}{\epsr \rij}
162 \label{eqn:vcoul}
163 \eeq
164 See also \figref{coul}, where $f = \frac{1}{4\pi \varepsilon_0} =
165 \electricConvFactorValue$ (see \chref{defunits})
167 \begin{figure}
168 \centerline{\includegraphics[width=8cm]{plots/vcrf}}
169 \caption[The Coulomb interaction with and without reaction field.]{The
170 Coulomb interaction (for particles with equal signed charge) with and
171 without reaction field. In the latter case $\epsr$ was 1, $\epsrf$ was 78,
172 and $r_c$ was 0.9 nm.
173 The dot-dashed line is the same as the dashed line, except for a constant.}
174 \label{fig:coul}
175 \end{figure}
177 The force derived from this potential is:
178 \beq
179 \ve{F}_i(\rvij) = f \frac{q_i q_j}{\epsr\rij^2}\rnorm
180 \eeq
182 A plain Coulomb interaction should only be used without cut-off or when all pairs fall within the cut-off, since there is an abrupt, large change in the force at the cut-off. In case you do want to use a cut-off, the potential can be shifted by a constant to make the potential the integral of the force. With the group cut-off scheme, this shift is only applied to non-excluded pairs. With the Verlet cut-off scheme, the shift is also applied to excluded pairs and self interactions, which makes the potential equivalent to a reaction field with $\epsrf=1$ (see below).
184 In {\gromacs} the relative \swapindex{dielectric}{constant}
185 \normindex{$\epsr$}
186 may be set in the in the input for {\tt grompp}.
188 \subsection{Coulomb interaction with \normindex{reaction field}}
189 \label{sec:coulrf}
190 The Coulomb interaction can be modified for homogeneous systems by
191 assuming a constant dielectric environment beyond the cut-off $r_c$
192 with a dielectric constant of {$\epsrf$}. The interaction then reads:
193 \beq
194 V_{crf} ~=~
195 f \frac{q_i q_j}{\epsr\rij}\left[1+\frac{\epsrf-\epsr}{2\epsrf+\epsr}
196 \,\frac{\rij^3}{r_c^3}\right]
197 - f\frac{q_i q_j}{\epsr r_c}\,\frac{3\epsrf}{2\epsrf+\epsr}
198 \label{eqn:vcrf}
199 \eeq
200 in which the constant expression on the right makes the potential
201 zero at the cut-off $r_c$. For charged cut-off spheres this corresponds
202 to neutralization with a homogeneous background charge.
203 We can rewrite \eqnref{vcrf} for simplicity as
204 \beq
205 V_{crf} ~=~ f \frac{q_i q_j}{\epsr}\left[\frac{1}{\rij} + k_{rf}~ \rij^2 -c_{rf}\right]
206 \eeq
207 with
208 \bea
209 k_{rf} &=& \frac{1}{r_c^3}\,\frac{\epsrf-\epsr}{(2\epsrf+\epsr)} \label{eqn:krf}\\
210 c_{rf} &=& \frac{1}{r_c}+k_{rf}\,r_c^2 ~=~ \frac{1}{r_c}\,\frac{3\epsrf}{(2\epsrf+\epsr)}
211 \label{eqn:crf}
212 \eea
213 For large $\epsrf$ the $k_{rf}$ goes to $r_c^{-3}/2$,
214 while for $\epsrf$ = $\epsr$ the correction vanishes.
215 In \figref{coul}
216 the modified interaction is plotted, and it is clear that the derivative
217 with respect to $\rij$ (= -force) goes to zero at the cut-off distance.
218 The force derived from this potential reads:
219 \beq
220 \ve{F}_i(\rvij) = f \frac{q_i q_j}{\epsr}\left[\frac{1}{\rij^2} - 2 k_{rf}\rij\right]\rnorm
221 \label{eqn:fcrf}
222 \eeq
223 The reaction-field correction should also be applied to all excluded
224 atoms pairs, including self pairs, in which case the normal Coulomb
225 term in \eqnsref{vcrf}{fcrf} is absent.
227 Tironi {\etal} have introduced a generalized reaction field in which
228 the dielectric continuum beyond the cut-off $r_c$ also has an ionic strength
229 $I$~\cite{Tironi95}. In this case we can rewrite the constants $k_{rf}$ and
230 $c_{rf}$ using the inverse Debye screening length $\kappa$:
231 \bea
232 \kappa^2 &=&
233 \frac{2 I \,F^2}{\varepsilon_0 \epsrf RT}
234 = \frac{F^2}{\varepsilon_0 \epsrf RT}\sum_{i=1}^{K} c_i z_i^2 \\
235 k_{rf} &=& \frac{1}{r_c^3}\,
236 \frac{(\epsrf-\epsr)(1 + \kappa r_c) + \half\epsrf(\kappa r_c)^2}
237 {(2\epsrf + \epsr)(1 + \kappa r_c) + \epsrf(\kappa r_c)^2}
238 \label{eqn:kgrf}\\
239 c_{rf} &=& \frac{1}{r_c}\,
240 \frac{3\epsrf(1 + \kappa r_c + \half(\kappa r_c)^2)}
241 {(2\epsrf+\epsr)(1 + \kappa r_c) + \epsrf(\kappa r_c)^2}
242 \label{eqn:cgrf}
243 \eea
244 where $F$ is Faraday's constant, $R$ is the ideal gas constant, $T$
245 the absolute temperature, $c_i$ the molar concentration for species
246 $i$ and $z_i$ the charge number of species $i$ where we have $K$
247 different species. In the limit of zero ionic strength ($\kappa=0$)
248 \eqnsref{kgrf}{cgrf} reduce to the simple forms of \eqnsref{krf}{crf}
249 respectively.
251 \subsection{Modified non-bonded interactions}
252 \label{sec:mod_nb_int}
253 In {\gromacs}, the non-bonded potentials can be
254 modified by a shift function, also called a force-switch function,
255 since it switches the force to zero at the cut-off.
256 The purpose of this is to replace the
257 truncated forces by forces that are continuous and have continuous
258 derivatives at the \normindex{cut-off} radius. With such forces the
259 time integration produces smaller errors. But note that for
260 Lennard-Jones interactions these errors are usually smaller than other errors,
261 such as integration errors at the repulsive part of the potential.
262 For Coulomb interactions we advise against using a shifted potential
263 and for use of a reaction field or a proper long-range method such as PME.
265 There is {\em no} fundamental difference between a switch function
266 (which multiplies the potential with a function) and a shift function
267 (which adds a function to the force or potential)~\cite{Spoel2006a}. The switch
268 function is a special case of the shift function, which we apply to
269 the {\em force function} $F(r)$, related to the electrostatic or
270 van der Waals force acting on particle $i$ by particle $j$ as:
271 \beq
272 \ve{F}_i = c \, F(r_{ij}) \frac{\rvij}{r_{ij}}
273 \eeq
274 For pure Coulomb or Lennard-Jones interactions
275 $F(r) = F_\alpha(r) = \alpha \, r^{-(\alpha+1)}$.
276 The switched force $F_s(r)$ can generally be written as:
277 \beq
278 \begin{array}{rcl}
279 \vspace{2mm}
280 F_s(r)~=&~F_\alpha(r) & r < r_1 \\
281 \vspace{2mm}
282 F_s(r)~=&~F_\alpha(r)+S(r) & r_1 \le r < r_c \\
283 F_s(r)~=&~0 & r_c \le r
284 \end{array}
285 \eeq
286 When $r_1=0$ this is a traditional shift function, otherwise it acts as a
287 switch function. The corresponding shifted potential function then reads:
288 \beq
289 V_s(r) = \int^{\infty}_r~F_s(x)\, dx
290 \eeq
292 The {\gromacs} {\bf force switch} function $S_F(r)$ should be smooth at the boundaries, therefore
293 the following boundary conditions are imposed on the switch function:
294 \beq
295 \begin{array}{rcl}
296 S_F(r_1) &=&0 \\
297 S_F'(r_1) &=&0 \\
298 S_F(r_c) &=&-F_\alpha(r_c) \\
299 S_F'(r_c) &=&-F_\alpha'(r_c)
300 \end{array}
301 \eeq
302 A 3$^{rd}$ degree polynomial of the form
303 \beq
304 S_F(r) = A(r-r_1)^2 + B(r-r_1)^3
305 \eeq
306 fulfills these requirements. The constants A and B are given by the
307 boundary condition at $r_c$:
308 \beq
309 \begin{array}{rcl}
310 \vspace{2mm}
311 A &~=~& -\alpha \, \displaystyle
312 \frac{(\alpha+4)r_c~-~(\alpha+1)r_1} {r_c^{\alpha+2}~(r_c-r_1)^2} \\
313 B &~=~& \alpha \, \displaystyle
314 \frac{(\alpha+3)r_c~-~(\alpha+1)r_1}{r_c^{\alpha+2}~(r_c-r_1)^3}
315 \end{array}
316 \eeq
317 Thus the total force function is:
318 \beq
319 F_s(r) = \frac{\alpha}{r^{\alpha+1}} + A(r-r_1)^2 + B(r-r_1)^3
320 \eeq
321 and the potential function reads:
322 \beq
323 V_s(r) = \frac{1}{r^\alpha} - \frac{A}{3} (r-r_1)^3 - \frac{B}{4} (r-r_1)^4 - C
324 \eeq
325 where
326 \beq
327 C = \frac{1}{r_c^\alpha} - \frac{A}{3} (r_c-r_1)^3 - \frac{B}{4} (r_c-r_1)^4
328 \eeq
330 The {\gromacs} {\bf potential-switch} function $S_V(r)$ scales the potential between
331 $r_1$ and $r_c$, and has similar boundary conditions, intended to produce
332 smoothly-varying potential and forces:
333 \beq
334 \begin{array}{rcl}
335 S_V(r_1) &=&1 \\
336 S_V'(r_1) &=&0 \\
337 S_V''(r_1) &=&0 \\
338 S_V(r_c) &=&0 \\
339 S_V'(r_c) &=&0 \\
340 S_V''(r_c) &=&0
341 \end{array}
342 \eeq
344 The fifth-degree polynomial that has these properties is
345 \beq
346 S_V(r; r_1, r_c) = \frac{1 - 10(r-r_1)^3(r_c-r_1)^2 + 15(r-r_1)^4(r_c-r_1) - 6(r-r_1)}{(r_c-r_1)^5}
347 \eeq
349 This implementation is found in several other simulation
350 packages,\cite{Ohmine1988,Kitchen1990,Guenot1993} but differs from
351 that in CHARMM.\cite{Steinbach1994} Switching the potential leads to
352 artificially large forces in the switching region, therefore it is not
353 recommended to switch Coulomb interactions using this
354 function,\cite{Spoel2006a} but switching Lennard-Jones interactions
355 using this function produces acceptable results.
357 \subsection{Modified short-range interactions with Ewald summation}
358 When Ewald summation\index{Ewald sum} or \seeindex{particle-mesh
359 Ewald}{PME}\index{Ewald, particle-mesh} is used to calculate the
360 long-range interactions, the
361 short-range Coulomb potential must also be modified. Here the potential
362 is switched to (nearly) zero at the cut-off, instead of the force.
363 In this case the short range potential is given by:
364 \beq
365 V(r) = f \frac{\mbox{erfc}(\beta r_{ij})}{r_{ij}} q_i q_j,
366 \eeq
367 where $\beta$ is a parameter that determines the relative weight
368 between the direct space sum and the reciprocal space sum and erfc$(x)$ is
369 the complementary error function. For further
370 details on long-range electrostatics, see \secref{lr_elstat}.
373 \section{Bonded interactions}
374 Bonded interactions are based on a fixed list of atoms. They are not
375 exclusively pair interactions, but include 3- and 4-body interactions
376 as well. There are {\em bond stretching} (2-body), {\em bond angle}
377 (3-body), and {\em dihedral angle} (4-body) interactions. A special
378 type of dihedral interaction (called {\em improper dihedral}) is used
379 to force atoms to remain in a plane or to prevent transition to a
380 configuration of opposite chirality (a mirror image).
382 \subsection{Bond stretching}
383 \label{sec:bondpot}
384 \subsubsection{Harmonic potential}
385 \label{subsec:harmonicbond}
386 The \swapindex{bond}{stretching} between two covalently bonded atoms
387 $i$ and $j$ is represented by a harmonic potential:
389 \begin{figure}
390 \centerline{\raisebox{2cm}{\includegraphics[width=5cm]{plots/bstretch}}\includegraphics[width=7cm]{plots/f-bond}}
391 \caption[Bond stretching.]{Principle of bond stretching (left), and the bond
392 stretching potential (right).}
393 \label{fig:bstretch1}
394 \end{figure}
396 \beq
397 V_b~(\rij) = \half k^b_{ij}(\rij-b_{ij})^2
398 \eeq
399 See also \figref{bstretch1}, with the force given by:
400 \beq
401 \ve{F}_i(\rvij) = k^b_{ij}(\rij-b_{ij}) \rnorm
402 \eeq
404 \subsubsection{Fourth power potential}
405 \label{subsec:G96bond}
406 In the \gromosv{96} force field~\cite{gromos96}, the covalent bond potential
407 is, for reasons of computational efficiency, written as:
408 \beq
409 V_b~(\rij) = \frac{1}{4}k^b_{ij}\left(\rij^2-b_{ij}^2\right)^2
410 \eeq
411 The corresponding force is:
412 \beq
413 \ve{F}_i(\rvij) = k^b_{ij}(\rij^2-b_{ij}^2)~\rvij
414 \eeq
415 The force constants for this form of the potential are related to the usual
416 harmonic force constant $k^{b,\mathrm{harm}}$ (\secref{bondpot}) as
417 \beq
418 2 k^b b_{ij}^2 = k^{b,\mathrm{harm}}
419 \eeq
420 The force constants are mostly derived from the harmonic ones used in
421 \gromosv{87}~\cite{biomos}. Although this form is computationally more
422 efficient
423 (because no square root has to be evaluated), it is conceptually more
424 complex. One particular disadvantage is that since the form is not harmonic,
425 the average energy of a single bond is not equal to $\half kT$ as it is for
426 the normal harmonic potential.
428 \subsection{\normindex{Morse potential} bond stretching}
429 \label{subsec:Morsebond}
430 %\author{F.P.X. Everdij}
432 For some systems that require an anharmonic bond stretching potential,
433 the Morse potential~\cite{Morse29}
434 between two atoms {\it i} and {\it j} is available
435 in {\gromacs}. This potential differs from the harmonic potential in
436 that it has an asymmetric potential well and a zero force at infinite
437 distance. The functional form is:
438 \beq
439 \displaystyle V_{morse} (r_{ij}) = D_{ij} [1 - \exp(-\beta_{ij}(r_{ij}-b_{ij}))]^2,
440 \eeq
441 See also \figref{morse}, and the corresponding force is:
442 \beq
443 \begin{array}{rcl}
444 \displaystyle {\bf F}_{morse} ({\bf r}_{ij})&=&2 D_{ij} \beta_{ij} \exp(-\beta_{ij}(r_{ij}-b_{ij})) * \\
445 \displaystyle \: & \: &[1 - \exp(-\beta_{ij}(r_{ij}-b_{ij}))] \frac{\displaystyle {\bf r}_{ij}}{\displaystyle r_{ij}},
446 \end{array}
447 \eeq
448 where \( \displaystyle D_{ij} \) is the depth of the well in kJ/mol,
449 \( \displaystyle \beta_{ij} \) defines the steepness of the well (in
450 nm\(^{-1} \)), and \( \displaystyle b_{ij} \) is the equilibrium
451 distance in nm. The steepness parameter \( \displaystyle \beta_{ij}
452 \) can be expressed in terms of the reduced mass of the atoms {\it i}
453 and {\it j}, the fundamental vibration frequency \( \displaystyle
454 \omega_{ij} \) and the well depth \( \displaystyle D_{ij} \):
455 \beq
456 \displaystyle \beta_{ij}= \omega_{ij} \sqrt{\frac{\mu_{ij}}{2 D_{ij}}}
457 \eeq
458 and because \( \displaystyle \omega = \sqrt{k/\mu} \), one can rewrite \( \displaystyle \beta_{ij} \) in terms of the harmonic force constant \( \displaystyle k_{ij} \):
459 \beq
460 \displaystyle \beta_{ij}= \sqrt{\frac{k_{ij}}{2 D_{ij}}}
461 \label{eqn:betaij}
462 \eeq
463 For small deviations \( \displaystyle (r_{ij}-b_{ij}) \), one can
464 approximate the \( \displaystyle \exp \)-term to first-order using a
465 Taylor expansion:
466 \beq
467 \displaystyle \exp(-x) \approx 1-x
468 \label{eqn:expminx}
469 \eeq
470 and substituting \eqnref{betaij} and \eqnref{expminx} in the functional form:
471 \beq
472 \begin{array}{rcl}
473 \displaystyle V_{morse} (r_{ij})&=&D_{ij} [1 - \exp(-\beta_{ij}(r_{ij}-b_{ij}))]^2\\
474 \displaystyle \:&=&D_{ij} [1 - (1 -\sqrt{\frac{k_{ij}}{2 D_{ij}}}(r_{ij}-b_{ij}))]^2\\
475 \displaystyle \:&=&\frac{1}{2} k_{ij} (r_{ij}-b_{ij}))^2
476 \end{array}
477 \eeq
478 we recover the harmonic bond stretching potential.
480 \begin{figure}
481 \centerline{\includegraphics[width=7cm]{plots/f-morse}}
482 \caption{The Morse potential well, with bond length 0.15 nm.}
483 \label{fig:morse}
484 \end{figure}
486 \subsection{Cubic bond stretching potential}
487 \label{subsec:cubicbond}
488 Another anharmonic bond stretching potential that is slightly simpler
489 than the Morse potential adds a cubic term in the distance to the
490 simple harmonic form:
491 \beq
492 V_b~(\rij) = k^b_{ij}(\rij-b_{ij})^2 + k^b_{ij}k^{cub}_{ij}(\rij-b_{ij})^3
493 \eeq
494 A flexible \normindex{water} model (based on
495 the SPC water model~\cite{Berendsen81}) including
496 a cubic bond stretching potential for the O-H bond
497 was developed by Ferguson~\cite{Ferguson95}. This model was found
498 to yield a reasonable infrared spectrum. The Ferguson water model is
499 available in the {\gromacs} library ({\tt flexwat-ferguson.itp}).
500 It should be noted that the potential is asymmetric: overstretching leads to
501 infinitely low energies. The \swapindex{integration}{timestep} is therefore
502 limited to 1 fs.
504 The force corresponding to this potential is:
505 \beq
506 \ve{F}_i(\rvij) = 2k^b_{ij}(\rij-b_{ij})~\rnorm + 3k^b_{ij}k^{cub}_{ij}(\rij-b_{ij})^2~\rnorm
507 \eeq
509 \subsection{FENE bond stretching potential\index{FENE potential}}
510 \label{subsec:FENEbond}
511 In coarse-grained polymer simulations the beads are often connected
512 by a FENE (finitely extensible nonlinear elastic) potential~\cite{Warner72}:
513 \beq
514 V_{\mbox{\small FENE}}(\rij) =
515 -\half k^b_{ij} b^2_{ij} \log\left(1 - \frac{\rij^2}{b^2_{ij}}\right)
516 \eeq
517 The potential looks complicated, but the expression for the force is simpler:
518 \beq
519 F_{\mbox{\small FENE}}(\rvij) =
520 -k^b_{ij} \left(1 - \frac{\rij^2}{b^2_{ij}}\right)^{-1} \rvij
521 \eeq
522 At short distances the potential asymptotically goes to a harmonic
523 potential with force constant $k^b$, while it diverges at distance $b$.
525 \subsection{Harmonic angle potential}
526 \label{subsec:harmonicangle}
527 \newcommand{\tijk}{\theta_{ijk}}
528 The bond-\swapindex{angle}{vibration} between a triplet of atoms $i$ - $j$ - $k$
529 is also represented by a harmonic potential on the angle $\tijk$
531 \begin{figure}
532 \centerline{\raisebox{1cm}{\includegraphics[width=5cm]{plots/angle}}\includegraphics[width=7cm]{plots/f-angle}}
533 \caption[Angle vibration.]{Principle of angle vibration (left) and the
534 bond angle potential (right).}
535 \label{fig:angle}
536 \end{figure}
538 \beq
539 V_a(\tijk) = \half k^{\theta}_{ijk}(\tijk-\tijk^0)^2
540 \eeq
541 As the bond-angle vibration is represented by a harmonic potential, the
542 form is the same as the bond stretching (\figref{bstretch1}).
544 The force equations are given by the chain rule:
545 \beq
546 \begin{array}{l}
547 \Fvi ~=~ -\displaystyle\frac{d V_a(\tijk)}{d \rvi} \\
548 \Fvk ~=~ -\displaystyle\frac{d V_a(\tijk)}{d \rvk} \\
549 \Fvj ~=~ -\Fvi-\Fvk
550 \end{array}
551 ~ \mbox{ ~ where ~ } ~
552 \tijk = \arccos \frac{(\rvij \cdot \ve{r}_{kj})}{r_{ij}r_{kj}}
553 \eeq
554 The numbering $i,j,k$ is in sequence of covalently bonded atoms. Atom
555 $j$ is in the middle; atoms $i$ and $k$ are at the ends (see \figref{angle}).
556 {\bf Note} that in the input in topology files, angles are given in degrees and
557 force constants in kJ/mol/rad$^2$.
559 \subsection{Cosine based angle potential}
560 \label{subsec:G96angle}
561 In the \gromosv{96} force field a simplified function is used to represent angle
562 vibrations:
563 \beq
564 V_a(\tijk) = \half k^{\theta}_{ijk}\left(\cos(\tijk) - \cos(\tijk^0)\right)^2
565 \label{eq:G96angle}
566 \eeq
567 where
568 \beq
569 \cos(\tijk) = \frac{\rvij\cdot\ve{r}_{kj}}{\rij r_{kj}}
570 \eeq
571 The corresponding force can be derived by partial differentiation with respect
572 to the atomic positions. The force constants in this function are related
573 to the force constants in the harmonic form $k^{\theta,\mathrm{harm}}$
574 (\ssecref{harmonicangle}) by:
575 \beq
576 k^{\theta} \sin^2(\tijk^0) = k^{\theta,\mathrm{harm}}
577 \eeq
578 In the \gromosv{96} manual there is a much more complicated conversion formula
579 which is temperature dependent. The formulas are equivalent at 0 K
580 and the differences at 300 K are on the order of 0.1 to 0.2\%.
581 {\bf Note} that in the input in topology files, angles are given in degrees and
582 force constants in kJ/mol.
584 \subsection{Restricted bending potential}
585 \label{subsec:ReB}
586 The restricted bending (ReB) potential~\cite{MonicaGoga2013} prevents the bending angle $\theta$
587 from reaching the $180^{\circ}$ value. In this way, the numerical instabilities
588 due to the calculation of the torsion angle and potential are eliminated when
589 performing coarse-grained molecular dynamics simulations.
591 To systematically hinder the bending angles from reaching the $180^{\circ}$ value,
592 the bending potential \ref{eq:G96angle} is divided by a $\sin^2\theta$ factor:
594 \beq
595 V_{\rm ReB}(\theta_i) = \frac{1}{2} k_{\theta} \frac{(\cos\theta_i - \cos\theta_0)^2}{\sin^2\theta_i}.
596 \label{eq:ReB}
597 \eeq
599 Figure ~\figref{ReB} shows the comparison between the ReB potential, \ref{eq:ReB},
600 and the standard one \ref{eq:G96angle}.
602 \begin{figure}
603 \centerline{\includegraphics[width=10cm]{plots/fig-02}}
604 \vspace*{8pt}
605 \caption{Bending angle potentials: cosine harmonic (solid black line), angle harmonic
606 (dashed black line) and restricted bending (red) with the same bending constant
607 $k_{\theta}=85$ kJ mol$^{-1}$ and equilibrium angle $\theta_0=130^{\circ}$.
608 The orange line represents the sum of a cosine harmonic ($k =50$ kJ mol$^{-1}$)
609 with a restricted bending ($k =25$ kJ mol$^{-1}$) potential, both with $\theta_0=130^{\circ}$.}
610 \label{fig:ReB}
611 \end{figure}
613 The wall of the ReB potential is very repulsive in the region close to $180^{\circ}$ and,
614 as a result, the bending angles are kept within a safe interval, far from instabilities.
615 The power $2$ of $\sin\theta_i$ in the denominator has been chosen to guarantee this behavior
616 and allows an elegant differentiation:
618 \beq
619 F_{\rm ReB}(\theta_i) = \frac{2k_{\theta}}{\sin^4\theta_i}(\cos\theta_i - \cos\theta_0) (1 - \cos\theta_i\cos\theta_0) \frac{\partial \cos\theta_i}{\partial \vec r_{k}}.
620 \label{eq:diff_ReB}
621 \eeq
623 Due to its construction, the restricted bending potential cannot be used for equilibrium
624 $\theta_0$ values too close to $0^{\circ}$ or $180^{\circ}$ (from experience, at least $10^{\circ}$
625 difference is recommended). It is very important that, in the starting configuration,
626 all the bending angles have to be in the safe interval to avoid initial instabilities.
627 This bending potential can be used in combination with any form of torsion potential.
628 It will always prevent three consecutive particles from becoming collinear and,
629 as a result, any torsion potential will remain free of singularities.
630 It can be also added to a standard bending potential to affect the angle around $180^{\circ}$,
631 but to keep its original form around the minimum (see the orange curve in \figref{ReB}).
634 \subsection{Urey-Bradley potential}
635 \label{subsec:Urey-Bradley}
636 The \swapindex{Urey-Bradley bond-angle}{vibration} between a triplet
637 of atoms $i$ - $j$ - $k$ is represented by a harmonic potential on the
638 angle $\tijk$ and a harmonic correction term on the distance between
639 the atoms $i$ and $k$. Although this can be easily written as a simple
640 sum of two terms, it is convenient to have it as a single entry in the
641 topology file and in the output as a separate energy term. It is used mainly
642 in the CHARMm force field~\cite{BBrooks83}. The energy is given by:
644 \beq
645 V_a(\tijk) = \half k^{\theta}_{ijk}(\tijk-\tijk^0)^2 + \half k^{UB}_{ijk}(r_{ik}-r_{ik}^0)^2
646 \eeq
648 The force equations can be deduced from sections~\ssecref{harmonicbond}
649 and~\ssecref{harmonicangle}.
651 \subsection{Bond-Bond cross term}
652 \label{subsec:bondbondcross}
653 The bond-bond cross term for three particles $i, j, k$ forming bonds
654 $i-j$ and $k-j$ is given by~\cite{Lawrence2003b}:
655 \begin{equation}
656 V_{rr'} ~=~ k_{rr'} \left(\left|\ve{r}_{i}-\ve{r}_j\right|-r_{1e}\right) \left(\left|\ve{r}_{k}-\ve{r}_j\right|-r_{2e}\right)
657 \label{crossbb}
658 \end{equation}
659 where $k_{rr'}$ is the force constant, and $r_{1e}$ and $r_{2e}$ are the
660 equilibrium bond lengths of the $i-j$ and $k-j$ bonds respectively. The force
661 associated with this potential on particle $i$ is:
662 \begin{equation}
663 \ve{F}_{i} = -k_{rr'}\left(\left|\ve{r}_{k}-\ve{r}_j\right|-r_{2e}\right)\frac{\ve{r}_i-\ve{r}_j}{\left|\ve{r}_{i}-\ve{r}_j\right|}
664 \end{equation}
665 The force on atom $k$ can be obtained by swapping $i$ and $k$ in the above
666 equation. Finally, the force on atom $j$ follows from the fact that the sum
667 of internal forces should be zero: $\ve{F}_j = -\ve{F}_i-\ve{F}_k$.
669 \subsection{Bond-Angle cross term}
670 \label{subsec:bondanglecross}
671 The bond-angle cross term for three particles $i, j, k$ forming bonds
672 $i-j$ and $k-j$ is given by~\cite{Lawrence2003b}:
673 \begin{equation}
674 V_{r\theta} ~=~ k_{r\theta} \left(\left|\ve{r}_{i}-\ve{r}_k\right|-r_{3e} \right) \left(\left|\ve{r}_{i}-\ve{r}_j\right|-r_{1e} + \left|\ve{r}_{k}-\ve{r}_j\right|-r_{2e}\right)
675 \end{equation}
676 where $k_{r\theta}$ is the force constant, $r_{3e}$ is the $i-k$ distance,
677 and the other constants are the same as in Equation~\ref{crossbb}. The force
678 associated with the potential on atom $i$ is:
679 \begin{equation}
680 \ve{F}_{i} ~=~ -k_{r\theta}\left[\left(\left|\ve{r}_{i}-\ve{r}_{k}\right|-r_{3e}\right)\frac{\ve{r}_i-\ve{r}_j}{\left|\ve{r}_{i}-\ve{r}_j\right|} \\
681 + \left(\left|\ve{r}_{i}-\ve{r}_j\right|-r_{1e} + \left|\ve{r}_{k}-\ve{r}_j\right|-r_{2e}\right)\frac{\ve{r}_i-\ve{r}_k}{\left|\ve{r}_{i}-\ve{r}_k\right|}\right]
682 \end{equation}
684 \subsection{Quartic angle potential}
685 \label{subsec:quarticangle}
686 For special purposes there is an angle potential
687 that uses a fourth order polynomial:
688 \beq
689 V_q(\tijk) ~=~ \sum_{n=0}^5 C_n (\tijk-\tijk^0)^n
690 \eeq
692 %% new commands %%%%%%%%%%%%%%%%%%%%%%
693 \newcommand{\rvkj}{{\bf r}_{kj}}
694 \newcommand{\rkj}{r_{kj}}
695 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
697 \subsection{Improper dihedrals\swapindexquiet{improper}{dihedral}}
698 \label{sec:imp}
699 Improper dihedrals are meant to keep \swapindex{planar}{group}s ({\eg}
700 aromatic rings) planar, or to prevent molecules from flipping over to their
701 \normindex{mirror image}s, see \figref{imp}.
703 \begin {figure}
704 \centerline{\includegraphics[width=4cm]{plots/ring-imp}\hspace{1cm}
705 \includegraphics[width=3cm]{plots/subst-im}\hspace{1cm}\includegraphics[width=3cm]{plots/tetra-im}}
706 \caption[Improper dihedral angles.]{Principle of improper
707 dihedral angles. Out of plane bending for rings (left), substituents
708 of rings (middle), out of tetrahedral (right). The improper dihedral
709 angle $\xi$ is defined as the angle between planes (i,j,k) and (j,k,l)
710 in all cases.}
711 \label{fig:imp}
712 \end {figure}
714 \subsubsection{Improper dihedrals: harmonic type}
715 \label{subsec:harmonicimproperdihedral}
716 The simplest improper dihedral potential is a harmonic potential; it is plotted in
717 \figref{imps}.
718 \beq
719 V_{id}(\xi_{ijkl}) = \half k_{\xi}(\xi_{ijkl}-\xi_0)^2
720 \eeq
721 Since the potential is harmonic it is discontinuous,
722 but since the discontinuity is chosen at 180$^\circ$ distance from $\xi_0$
723 this will never cause problems.
724 {\bf Note} that in the input in topology files, angles are given in degrees and
725 force constants in kJ/mol/rad$^2$.
727 \begin{figure}
728 \centerline{\includegraphics[width=10cm]{plots/f-imps.pdf}}
729 \caption{Improper dihedral potential.}
730 \label{fig:imps}
731 \end{figure}
733 \subsubsection{Improper dihedrals: periodic type}
734 \label{subsec:periodicimproperdihedral}
735 This potential is identical to the periodic proper dihedral (see below).
736 There is a separate dihedral type for this (type 4) only to be able
737 to distinguish improper from proper dihedrals in the parameter section
738 and the output.
740 \subsection{Proper dihedrals\swapindexquiet{proper}{dihedral}}
741 For the normal \normindex{dihedral} interaction there is a choice of
742 either the {\gromos} periodic function or a function based on
743 expansion in powers of $\cos \phi$ (the so-called Ryckaert-Bellemans
744 potential). This choice has consequences for the inclusion of special
745 interactions between the first and the fourth atom of the dihedral
746 quadruple. With the periodic {\gromos} potential a special 1-4
747 LJ-interaction must be included; with the Ryckaert-Bellemans potential
748 {\em for alkanes} the \normindex{1-4 interaction}s must be excluded
749 from the non-bonded list. {\bf Note:} Ryckaert-Bellemans potentials
750 are also used in {\eg} the OPLS force field in combination with 1-4
751 interactions. You should therefore not modify topologies generated by
752 {\tt \normindex{pdb2gmx}} in this case.
754 \subsubsection{Proper dihedrals: periodic type}
755 \label{subsec:properdihedral}
756 Proper dihedral angles are defined according to the IUPAC/IUB
757 convention, where $\phi$ is the angle between the $ijk$ and the $jkl$
758 planes, with {\bf zero} corresponding to the {\em cis} configuration
759 ($i$ and $l$ on the same side). There are two dihedral function types
760 in {\gromacs} topology files. There is the standard type 1 which behaves
761 like any other bonded interactions. For certain force fields, type 9
762 is useful. Type 9 allows multiple potential functions to be applied
763 automatically to a single dihedral in the {\tt [ dihedral ]} section
764 when multiple parameters are defined for the same atomtypes
765 in the {\tt [ dihedraltypes ]} section.
767 \begin{figure}
768 \centerline{\raisebox{1cm}{\includegraphics[width=5cm]{plots/dih}}\includegraphics[width=7cm]{plots/f-dih}}
769 \caption[Proper dihedral angle.]{Principle of proper dihedral angle
770 (left, in {\em trans} form) and the dihedral angle potential (right).}
771 \label{fig:pdihf}
772 \end{figure}
773 \beq
774 V_d(\phi_{ijkl}) = k_{\phi}(1 + \cos(n \phi - \phi_s))
775 \eeq
777 \subsubsection{Proper dihedrals: Ryckaert-Bellemans function}
778 \label{subsec:RBdihedral}
779 For alkanes, the following proper dihedral potential is often used
780 (see \figref{rbdih}):
781 \beq
782 V_{rb}(\phi_{ijkl}) = \sum_{n=0}^5 C_n( \cos(\psi ))^n,
783 \eeq
784 where $\psi = \phi - 180^\circ$. \\
785 {\bf Note:} A conversion from one convention to another can be achieved by
786 multiplying every coefficient \( \displaystyle C_n \)
787 by \( \displaystyle (-1)^n \).
789 An example of constants for $C$ is given in \tabref{crb}.
791 \begin{table}
792 \centerline{
793 \begin{tabular}{|lr|lr|lr|}
794 \dline
795 $C_0$ & 9.28 & $C_2$ & -13.12 & $C_4$ & 26.24 \\
796 $C_1$ & 12.16 & $C_3$ & -3.06 & $C_5$ & -31.5 \\
797 \dline
798 \end{tabular}
800 \caption{Constants for Ryckaert-Bellemans potential (kJ mol$^{-1}$).}
801 \label{tab:crb}
802 \end{table}
804 \begin{figure}
805 \centerline{\includegraphics[width=8cm]{plots/f-rbs}}
806 \caption{Ryckaert-Bellemans dihedral potential.}
807 \label{fig:rbdih}
808 \end{figure}
810 ({\bf Note:} The use of this potential implies exclusion of LJ interactions
811 between the first and the last atom of the dihedral, and $\psi$ is defined
812 according to the ``polymer convention'' ($\psi_{trans}=0$).)
814 The RB dihedral function can also be used to include Fourier dihedrals
815 (see below):
816 \beq
817 V_{rb} (\phi_{ijkl}) ~=~ \frac{1}{2} \left[F_1(1+\cos(\phi)) + F_2(
818 1-\cos(2\phi)) + F_3(1+\cos(3\phi)) + F_4(1-\cos(4\phi))\right]
819 \eeq
820 Because of the equalities \( \cos(2\phi) = 2\cos^2(\phi) - 1 \),
821 \( \cos(3\phi) = 4\cos^3(\phi) - 3\cos(\phi) \) and
822 \( \cos(4\phi) = 8\cos^4(\phi) - 8\cos^2(\phi) + 1 \)
823 one can translate the OPLS parameters to
824 Ryckaert-Bellemans parameters as follows:
825 \beq
826 \displaystyle
827 \begin{array}{rcl}
828 \displaystyle C_0&=&F_2 + \frac{1}{2} (F_1 + F_3)\\
829 \displaystyle C_1&=&\frac{1}{2} (- F_1 + 3 \, F_3)\\
830 \displaystyle C_2&=& -F_2 + 4 \, F_4\\
831 \displaystyle C_3&=&-2 \, F_3\\
832 \displaystyle C_4&=&-4 \, F_4\\
833 \displaystyle C_5&=&0
834 \end{array}
835 \eeq
836 with OPLS parameters in protein convention and RB parameters in
837 polymer convention (this yields a minus sign for the odd powers of
838 cos$(\phi)$).\\
839 \noindent{\bf Note:} Mind the conversion from {\bf kcal mol$^{-1}$} for
840 literature OPLS and RB parameters to {\bf kJ mol$^{-1}$} in {\gromacs}.\\
842 \subsubsection{Proper dihedrals: Fourier function}
843 \label{subsec:Fourierdihedral}
844 The OPLS potential function is given as the first three
845 ~\cite{Jorgensen1996} or four~\cite{Robertson2015a} cosine terms of a Fourier series.
846 In {\gromacs} the four term function is implemented:
847 \beq
848 V_{F} (\phi_{ijkl}) ~=~ \frac{1}{2} \left[C_1(1+\cos(\phi)) + C_2(
849 1-\cos(2\phi)) + C_3(1+\cos(3\phi)) + C_4(1-\cos(4\phi))\right],
850 \eeq
851 Internally, {\gromacs}
852 uses the Ryckaert-Bellemans code
853 to compute Fourier dihedrals (see above), because this is more efficient.\\
854 \noindent{\bf Note:} Mind the conversion from {\emph kcal mol$^{-1}$} for
855 literature OPLS parameters to {\bf kJ mol$^{-1}$} in {\gromacs}.\\
857 \subsubsection{Proper dihedrals: Restricted torsion potential}
858 \label{subsec:ReT}
859 In a manner very similar to the restricted bending potential (see \ref{subsec:ReB}),
860 a restricted torsion/dihedral potential is introduced:
862 \beq
863 V_{\rm ReT}(\phi_i) = \frac{1}{2} k_{\phi} \frac{(\cos\phi_i - \cos\phi_0)^2}{\sin^2\phi_i}
864 \label{eq:ReT}
865 \eeq
867 with the advantages of being a function of $\cos\phi$ (no problems taking the derivative of $\sin\phi$)
868 and of keeping the torsion angle at only one minimum value. In this case, the factor $\sin^2\phi$ does
869 not allow the dihedral angle to move from the [$-180^{\circ}$:0] to [0:$180^{\circ}$] interval, i.e. it cannot have maxima both at $-\phi_0$ and $+\phi_0$ maxima, but only one of them.
870 For this reason, all the dihedral angles of the starting configuration should have their values in the
871 desired angles interval and the the equilibrium $\phi_0$ value should not be too close to the interval limits
872 (as for the restricted bending potential, described in \ref{subsec:ReB}, at least $10^{\circ}$ difference is recommended).
874 \subsubsection{Proper dihedrals: Combined bending-torsion potential}
875 \label{subsec:CBT}
876 When the four particles forming the dihedral angle become collinear (this situation will never happen in
877 atomistic simulations, but it can occur in coarse-grained simulations) the calculation of the
878 torsion angle and potential leads to numerical instabilities.
879 One way to avoid this is to use the restricted bending potential (see \ref{subsec:ReB})
880 that prevents the dihedral
881 from reaching the $180^{\circ}$ value.
883 Another way is to disregard any effects of the dihedral becoming ill-defined,
884 keeping the dihedral force and potential calculation continuous in entire angle range
885 by coupling the torsion potential (in a cosine form) with the bending potentials of the
886 adjacent bending angles in a unique expression:
888 \beq
889 V_{\rm CBT}(\theta_{i-1}, \theta_i, \phi_i) = k_{\phi} \sin^3\theta_{i-1} \sin^3\theta_{i} \sum_{n=0}^4 { a_n \cos^n\phi_i}.
890 \label{eq:CBT}
891 \eeq
893 This combined bending-torsion (CBT) potential has been proposed by~\cite{BulacuGiessen2005}
894 for polymer melt simulations and is extensively described in~\cite{MonicaGoga2013}.
896 This potential has two main advantages:
897 \begin{itemize}
898 \item
899 it does not only depend on the dihedral angle $\phi_i$ (between the $i-2$, $i-1$, $i$ and $i+1$ beads)
900 but also on the bending angles $\theta_{i-1}$ and $\theta_i$ defined from three adjacent beads
901 ($i-2$, $i-1$ and $i$, and $i-1$, $i$ and $i+1$, respectively).
902 The two $\sin^3\theta$ pre-factors, tentatively suggested by~\cite{ScottScheragator1966} and theoretically
903 discussed by~\cite{PaulingBond}, cancel the torsion potential and force when either of the two bending angles
904 approaches the value of $180^\circ$.
905 \item
906 its dependence on $\phi_i$ is expressed through a polynomial in $\cos\phi_i$ that avoids the singularities in
907 $\phi=0^\circ$ or $180^\circ$ in calculating the torsional force.
908 \end{itemize}
910 These two properties make the CBT potential well-behaved for MD simulations with weak constraints
911 on the bending angles or even for steered / non-equilibrium MD in which the bending and torsion angles suffer major
912 modifications.
913 When using the CBT potential, the bending potentials for the adjacent $\theta_{i-1}$ and $\theta_i$ may have any form.
914 It is also possible to leave out the two angle bending terms ($\theta_{i-1}$ and $\theta_{i}$) completely.
915 \figref{CBT} illustrates the difference between a torsion potential with and without the $\sin^{3}\theta$ factors
916 (blue and gray curves, respectively).
918 \begin{figure}
919 \centerline{\includegraphics[width=10cm]{plots/fig-04}}
920 \caption{Blue: surface plot of the combined bending-torsion potential
921 (\ref{eq:CBT} with $k = 10$ kJ mol$^{-1}$, $a_0=2.41$, $a_1=-2.95$, $a_2=0.36$, $a_3=1.33$)
922 when, for simplicity, the bending angles behave the same ($\theta_1=\theta_2=\theta$).
923 Gray: the same torsion potential without the $\sin^{3}\theta$ terms (Ryckaert-Bellemans type).
924 $\phi$ is the dihedral angle.}
925 \label{fig:CBT}
926 \end{figure}
928 Additionally, the derivative of $V_{CBT}$ with respect to the Cartesian variables is straightforward:
930 \begin{equation}
931 \frac{\partial V_{\rm CBT}(\theta_{i-1},\theta_i,\phi_i)} {\partial \vec r_{l}} = \frac{\partial V_{\rm CBT}}{\partial \theta_{i-1}} \frac{\partial \theta_{i-1}}{\partial \vec r_{l}} +
932 \frac{\partial V_{\rm CBT}}{\partial \theta_{i }} \frac{\partial \theta_{i }}{\partial \vec r_{l}} +
933 \frac{\partial V_{\rm CBT}}{\partial \phi_{i }} \frac{\partial \phi_{i }}{\partial \vec r_{l}}
934 \label{eq:force_cbt}
935 \end{equation}
937 The CBT is based on a cosine form without multiplicity, so it can only be symmetrical around $0^{\circ}$.
938 To obtain an asymmetrical dihedral angle distribution (e.g. only one maximum in [$-180^{\circ}$:$180^{\circ}$] interval),
939 a standard torsion potential such as harmonic angle or periodic cosine potentials should be used instead of a CBT potential.
940 However, these two forms have the inconveniences of the force derivation ($1/\sin\phi$) and of the alignment of beads
941 ($\theta_i$ or $\theta_{i-1} = 0^{\circ}, 180^{\circ}$).
942 Coupling such non-$\cos\phi$ potentials with $\sin^3\theta$ factors does not improve simulation stability since there are
943 cases in which $\theta$ and $\phi$ are simultaneously $180^{\circ}$. The integration at this step would be possible
944 (due to the cancelling of the torsion potential) but the next step would be singular
945 ($\theta$ is not $180^{\circ}$ and $\phi$ is very close to $180^{\circ}$).
947 \subsection{Tabulated bonded interaction functions\index{tabulated bonded interaction function}}
948 \label{subsec:tabulatedinteraction}
949 For full flexibility, any functional shape can be used for
950 bonds, angles and dihedrals through user-supplied tabulated functions.
951 The functional shapes are:
952 \bea
953 V_b(r_{ij}) &=& k \, f^b_n(r_{ij}) \\
954 V_a(\tijk) &=& k \, f^a_n(\tijk) \\
955 V_d(\phi_{ijkl}) &=& k \, f^d_n(\phi_{ijkl})
956 \eea
957 where $k$ is a force constant in units of energy
958 and $f$ is a cubic spline function; for details see \ssecref{cubicspline}.
959 For each interaction, the force constant $k$ and the table number $n$
960 are specified in the topology.
961 There are two different types of bonds, one that generates exclusions (type 8)
962 and one that does not (type 9).
963 For details see \tabref{topfile2}.
964 The table files are supplied to the {\tt mdrun} program.
965 After the table file name an underscore, the letter ``b'' for bonds,
966 ``a'' for angles or ``d'' for dihedrals and the table number must be appended.
967 For example, a tabulated bond with $n=0$ can be read from the file {\tt table_b0.xvg}.
968 Multiple tables can be
969 supplied simply by adding files with different values of $n$, and are applied to the appropriate
970 bonds, as specified in the topology (\tabref{topfile2}).
971 The format for the table files is three fixed-format columns of any suitable width. These columns must contain $x$, $f(x)$, $-f'(x)$,
972 and the values of $x$ should be uniformly spaced. Requirements for entries in the topology
973 are given in~\tabref{topfile2}.
974 The setup of the tables is as follows:
975 \\{\bf bonds}:
976 $x$ is the distance in nm. For distances beyond the table length,
977 {\tt mdrun} will quit with an error message.
978 \\{\bf angles}:
979 $x$ is the angle in degrees. The table should go from
980 0 up to and including 180 degrees; the derivative is taken in degrees.
981 \\{\bf dihedrals}:
982 $x$ is the dihedral angle in degrees. The table should go from
983 -180 up to and including 180 degrees;
984 the IUPAC/IUB convention is used, {\ie} zero is cis,
985 the derivative is taken in degrees.
987 \section{Restraints}
988 Special potentials are used for imposing restraints on the motion of
989 the system, either to avoid disastrous deviations, or to include
990 knowledge from experimental data. In either case they are not really
991 part of the force field and the reliability of the parameters is not
992 important. The potential forms, as implemented in {\gromacs}, are
993 mentioned just for the sake of completeness. Restraints and constraints
994 refer to quite different algorithms in {\gromacs}.
996 \subsection{Position restraints\swapindexquiet{position}{restraint}}
997 \label{subsec:positionrestraint}
998 These are used to restrain particles to fixed reference positions
999 $\ve{R}_i$. They can be used during equilibration in order to avoid
1000 drastic rearrangements of critical parts ({\eg} to restrain motion
1001 in a protein that is subjected to large solvent forces when the
1002 solvent is not yet equilibrated). Another application is the
1003 restraining of particles in a shell around a region that is simulated
1004 in detail, while the shell is only approximated because it lacks
1005 proper interaction from missing particles outside the
1006 shell. Restraining will then maintain the integrity of the inner
1007 part. For spherical shells, it is a wise procedure to make the force
1008 constant depend on the radius, increasing from zero at the inner
1009 boundary to a large value at the outer boundary. This feature has
1010 not, however, been implemented in {\gromacs}.
1011 \newcommand{\unitv}[1]{\hat{\bf #1}}
1012 \newcommand{\halfje}[1]{\frac{#1}{2}}
1014 The following form is used:
1015 \beq
1016 V_{pr}(\ve{r}_i) = \halfje{1}k_{pr}|\rvi-\ve{R}_i|^2
1017 \eeq
1018 The potential is plotted in \figref{positionrestraint}.
1020 \begin{figure}
1021 \centerline{\includegraphics[width=8cm]{plots/f-pr}}
1022 \caption{Position restraint potential.}
1023 \label{fig:positionrestraint}
1024 \end{figure}
1026 The potential form can be rewritten without loss of generality as:
1027 \beq
1028 V_{pr}(\ve{r}_i) = \halfje{1} \left[ k_{pr}^x (x_i-X_i)^2 ~\unitv{x} + k_{pr}^y (y_i-Y_i)^2 ~\unitv{y} + k_{pr}^z (z_i-Z_i)^2 ~\unitv{z}\right]
1029 \eeq
1031 Now the forces are:
1032 \beq
1033 \begin{array}{rcl}
1034 F_i^x &=& -k_{pr}^x~(x_i - X_i) \\
1035 F_i^y &=& -k_{pr}^y~(y_i - Y_i) \\
1036 F_i^z &=& -k_{pr}^z~(z_i - Z_i)
1037 \end{array}
1038 \eeq
1039 Using three different force constants the position
1040 restraints can be turned on or off
1041 in each spatial dimension; this means that atoms can be harmonically
1042 restrained to a plane or a line.
1043 Position restraints are applied to a special fixed list of atoms. Such a
1044 list is usually generated by the {\tt \normindex{pdb2gmx}} program.
1046 \subsection{\swapindex{Flat-bottomed}{position restraint}s}
1047 \label{subsec:fbpositionrestraint}
1048 Flat-bottomed position restraints can be used to restrain particles to
1049 part of the simulation volume. No force acts on the restrained
1050 particle within the flat-bottomed region of the potential, however a
1051 harmonic force acts to move the particle to the flat-bottomed region
1052 if it is outside it. It is possible to apply normal and
1053 flat-bottomed position restraints on the same particle (however, only
1054 with the same reference position $\ve{R}_i$). The following general potential
1055 is used (Figure~\ref{fig:fbposres}A):
1056 \beq
1057 V_\mathrm{fb}(\ve{r}_i) = \frac{1}{2}k_\mathrm{fb} [d_g(\ve{r}_i;\ve{R}_i) - r_\mathrm{fb}]^2\,H[d_g(\ve{r}_i;\ve{R}_i) - r_\mathrm{fb}],
1058 \eeq
1059 where $\ve{R}_i$ is the reference position, $r_\mathrm{fb}$ is the distance
1060 from the center with a flat potential, $k_\mathrm{fb}$ the force constant, and $H$ is the Heaviside step
1061 function. The distance $d_g(\ve{r}_i;\ve{R}_i)$ from the reference
1062 position depends on the geometry $g$ of the flat-bottomed potential.
1064 \begin{figure}
1065 \centerline{\includegraphics[width=10cm]{plots/fbposres}}
1066 \caption{Flat-bottomed position restraint potential. (A) Not
1067 inverted, (B) inverted.}
1068 \label{fig:fbposres}
1069 \end{figure}
1071 The following geometries for the flat-bottomed potential are supported:\newline
1072 {\bfseries Sphere} ($g =1$): The particle is kept in a sphere of given
1073 radius. The force acts towards the center of the sphere. The following distance calculation is used:
1074 \beq
1075 d_g(\ve{r}_i;\ve{R}_i) = |\ve{r}_i-\ve{R}_i|
1076 \eeq
1077 {\bfseries Cylinder} ($g=6,7,8$): The particle is kept in a cylinder of given radius
1078 parallel to the $x$ ($g=6$), $y$ ($g=7$), or $z$-axis ($g=8$). For backwards compatibility, setting
1079 $g=2$ is mapped to $g=8$ in the code so that old {\tt .tpr} files and topologies work.
1080 The force from the flat-bottomed potential acts towards the axis of the cylinder.
1081 The component of the force parallel to the cylinder axis is zero.
1082 For a cylinder aligned along the $z$-axis:
1083 \beq
1084 d_g(\ve{r}_i;\ve{R}_i) = \sqrt{ (x_i-X_i)^2 + (y_i - Y_i)^2 }
1085 \eeq
1086 {\bfseries Layer} ($g=3,4,5$): The particle is kept in a layer defined by the
1087 thickness and the normal of the layer. The layer normal can be parallel to the $x$, $y$, or
1088 $z$-axis. The force acts parallel to the layer normal.\\
1089 \beq
1090 d_g(\ve{r}_i;\ve{R}_i) = |x_i-X_i|, \;\;\;\mbox{or}\;\;\;
1091 d_g(\ve{r}_i;\ve{R}_i) = |y_i-Y_i|, \;\;\;\mbox{or}\;\;\;
1092 d_g(\ve{r}_i;\ve{R}_i) = |z_i-Z_i|.
1093 \eeq
1095 It is possible to apply multiple independent flat-bottomed position
1096 restraints of different geometry on one particle. For example, applying
1097 a cylinder and a layer in $z$ keeps a particle within a
1098 disk. Applying three layers in $x$, $y$, and $z$ keeps the particle within a cuboid.
1100 In addition, it is possible to invert the restrained region with the
1101 unrestrained region, leading to a potential that acts to keep the particle {\it outside} of the volume
1102 defined by $\ve{R}_i$, $g$, and $r_\mathrm{fb}$. That feature is
1103 switched on by defining a negative $r_\mathrm{fb}$ in the
1104 topology. The following potential is used (Figure~\ref{fig:fbposres}B):
1105 \beq
1106 V_\mathrm{fb}^{\mathrm{inv}}(\ve{r}_i) = \frac{1}{2}k_\mathrm{fb}
1107 [d_g(\ve{r}_i;\ve{R}_i) - |r_\mathrm{fb}|]^2\,
1108 H[ -(d_g(\ve{r}_i;\ve{R}_i) - |r_\mathrm{fb}|)].
1109 \eeq
1113 \subsection{Angle restraints\swapindexquiet{angle}{restraint}}
1114 \label{subsec:anglerestraint}
1115 These are used to restrain the angle between two pairs of particles
1116 or between one pair of particles and the $z$-axis.
1117 The functional form is similar to that of a proper dihedral.
1118 For two pairs of atoms:
1119 \beq
1120 V_{ar}(\ve{r}_i,\ve{r}_j,\ve{r}_k,\ve{r}_l)
1121 = k_{ar}(1 - \cos(n (\theta - \theta_0))
1123 ,~~~~\mbox{where}~~
1124 \theta = \arccos\left(\frac{\ve{r}_j -\ve{r}_i}{\|\ve{r}_j -\ve{r}_i\|}
1125 \cdot \frac{\ve{r}_l -\ve{r}_k}{\|\ve{r}_l -\ve{r}_k\|} \right)
1126 \eeq
1127 For one pair of atoms and the $z$-axis:
1128 \beq
1129 V_{ar}(\ve{r}_i,\ve{r}_j) = k_{ar}(1 - \cos(n (\theta - \theta_0))
1131 ,~~~~\mbox{where}~~
1132 \theta = \arccos\left(\frac{\ve{r}_j -\ve{r}_i}{\|\ve{r}_j -\ve{r}_i\|}
1133 \cdot \left( \begin{array}{c} 0 \\ 0 \\ 1 \\ \end{array} \right) \right)
1134 \eeq
1135 A multiplicity ($n$) of 2 is useful when you do not want to distinguish
1136 between parallel and anti-parallel vectors.
1137 The equilibrium angle $\theta$ should be between 0 and 180 degrees
1138 for multiplicity 1 and between 0 and 90 degrees for multiplicity 2.
1141 \subsection{Dihedral restraints\swapindexquiet{dihedral}{restraint}}
1142 \label{subsec:dihedralrestraint}
1143 These are used to restrain the dihedral angle $\phi$ defined by four particles
1144 as in an improper dihedral (sec.~\ref{sec:imp}) but with a slightly
1145 modified potential. Using:
1146 \beq
1147 \phi' = \left(\phi-\phi_0\right) ~{\rm MOD}~ 2\pi
1148 \label{eqn:dphi}
1149 \eeq
1150 where $\phi_0$ is the reference angle, the potential is defined as:
1151 \beq
1152 V_{dihr}(\phi') ~=~ \left\{
1153 \begin{array}{lcllll}
1154 \half k_{dihr}(\phi'-\phi_0-\Delta\phi)^2
1155 &\mbox{for}& \phi' & > & \Delta\phi \\[1.5ex]
1156 0 &\mbox{for}& \phi' & \le & \Delta\phi \\[1.5ex]
1157 \end{array}\right.
1158 \label{eqn:dihre}
1159 \eeq
1160 where $\Delta\phi$ is a user defined angle and $k_{dihr}$ is the force
1161 constant.
1162 {\bf Note} that in the input in topology files, angles are given in degrees and
1163 force constants in kJ/mol/rad$^2$.
1165 \subsection{Distance restraints\swapindexquiet{distance}{restraint}}
1166 \label{subsec:distancerestraint}
1167 Distance restraints
1168 add a penalty to the potential when the distance between specified
1169 pairs of atoms exceeds a threshold value. They are normally used to
1170 impose experimental restraints from, for instance, experiments in nuclear
1171 magnetic resonance (NMR), on the motion of the system. Thus, MD can be
1172 used for structure refinement using NMR data\index{nmr
1173 refinement}\index{refinement,nmr}.
1174 In {\gromacs} there are three ways to impose restraints on pairs of atoms:
1175 \begin{itemize}
1176 \item Simple harmonic restraints: use {\tt [ bonds ]} type 6
1177 (see \secref{excl}).
1178 \item\label{subsec:harmonicrestraint}Piecewise linear/harmonic restraints: {\tt [ bonds ]} type 10.
1179 \item Complex NMR distance restraints, optionally with pair, time and/or
1180 ensemble averaging.
1181 \end{itemize}
1182 The last two options will be detailed now.
1184 The potential form for distance restraints is quadratic below a specified
1185 lower bound and between two specified upper bounds, and linear beyond the
1186 largest bound (see \figref{dist}).
1187 \beq
1188 V_{dr}(r_{ij}) ~=~ \left\{
1189 \begin{array}{lcllllll}
1190 \half k_{dr}(r_{ij}-r_0)^2
1191 &\mbox{for}& & & r_{ij} & < & r_0 \\[1.5ex]
1192 0 &\mbox{for}& r_0 & \le & r_{ij} & < & r_1 \\[1.5ex]
1193 \half k_{dr}(r_{ij}-r_1)^2
1194 &\mbox{for}& r_1 & \le & r_{ij} & < & r_2 \\[1.5ex]
1195 \half k_{dr}(r_2-r_1)(2r_{ij}-r_2-r_1)
1196 &\mbox{for}& r_2 & \le & r_{ij} & &
1197 \end{array}\right.
1198 \label{eqn:disre}
1199 \eeq
1201 \begin{figure}
1202 \centerline{\includegraphics[width=8cm]{plots/f-dr}}
1203 \caption{Distance Restraint potential.}
1204 \label{fig:dist}
1205 \end{figure}
1207 The forces are
1208 \beq
1209 \ve{F}_i~=~ \left\{
1210 \begin{array}{lcllllll}
1211 -k_{dr}(r_{ij}-r_0)\frac{\rvij}{r_{ij}}
1212 &\mbox{for}& & & r_{ij} & < & r_0 \\[1.5ex]
1213 0 &\mbox{for}& r_0 & \le & r_{ij} & < & r_1 \\[1.5ex]
1214 -k_{dr}(r_{ij}-r_1)\frac{\rvij}{r_{ij}}
1215 &\mbox{for}& r_1 & \le & r_{ij} & < & r_2 \\[1.5ex]
1216 -k_{dr}(r_2-r_1)\frac{\rvij}{r_{ij}}
1217 &\mbox{for}& r_2 & \le & r_{ij} & &
1218 \end{array} \right.
1219 \eeq
1221 For restraints not derived from NMR data, this functionality
1222 will usually suffice and a section of {\tt [ bonds ]} type 10
1223 can be used to apply individual restraints between pairs of
1224 atoms, see \ssecref{topfile}.
1225 For applying restraints derived from NMR measurements, more complex
1226 functionality might be required, which is provided through
1227 the {\tt [~distance_restraints~]} section and is described below.
1229 \subsubsection{Time averaging\swapindexquiet{time-averaged}{distance restraint}}
1230 Distance restraints based on instantaneous distances can potentially reduce
1231 the fluctuations in a molecule significantly. This problem can be overcome by restraining
1232 to a {\em time averaged} distance~\cite{Torda89}.
1233 The forces with time averaging are:
1234 \beq
1235 \ve{F}_i~=~ \left\{
1236 \begin{array}{lcllllll}
1237 -k^a_{dr}(\bar{r}_{ij}-r_0)\frac{\rvij}{r_{ij}}
1238 &\mbox{for}& & & \bar{r}_{ij} & < & r_0 \\[1.5ex]
1239 0 &\mbox{for}& r_0 & \le & \bar{r}_{ij} & < & r_1 \\[1.5ex]
1240 -k^a_{dr}(\bar{r}_{ij}-r_1)\frac{\rvij}{r_{ij}}
1241 &\mbox{for}& r_1 & \le & \bar{r}_{ij} & < & r_2 \\[1.5ex]
1242 -k^a_{dr}(r_2-r_1)\frac{\rvij}{r_{ij}}
1243 &\mbox{for}& r_2 & \le & \bar{r}_{ij} & &
1244 \end{array} \right.
1245 \eeq
1246 where $\bar{r}_{ij}$ is given by an exponential running average with decay time $\tau$:
1247 \beq
1248 \bar{r}_{ij} ~=~ < r_{ij}^{-3} >^{-1/3}
1249 \label{eqn:rav}
1250 \eeq
1251 The force constant $k^a_{dr}$ is switched on slowly to compensate for
1252 the lack of history at the beginning of the simulation:
1253 \beq
1254 k^a_{dr} = k_{dr} \left(1-\exp\left(-\frac{t}{\tau}\right)\right)
1255 \eeq
1256 Because of the time averaging, we can no longer speak of a distance restraint
1257 potential.
1259 This way an atom can satisfy two incompatible distance restraints
1260 {\em on average} by moving between two positions.
1261 An example would be an amino acid side-chain that is rotating around
1262 its $\chi$ dihedral angle, thereby coming close to various other groups.
1263 Such a mobile side chain can give rise to multiple NOEs that can not be
1264 fulfilled by a single structure.
1266 The computation of the time
1267 averaged distance in the {\tt mdrun} program is done in the following fashion:
1268 \beq
1269 \begin{array}{rcl}
1270 \overline{r^{-3}}_{ij}(0) &=& r_{ij}(0)^{-3} \\
1271 \overline{r^{-3}}_{ij}(t) &=& \overline{r^{-3}}_{ij}(t-\Delta t)~\exp{\left(-\frac{\Delta t}{\tau}\right)} + r_{ij}(t)^{-3}\left[1-\exp{\left(-\frac{\Delta t}{\tau}\right)}\right]
1272 \label{eqn:ravdisre}
1273 \end{array}
1274 \eeq
1276 When a pair is within the bounds, it can still feel a force
1277 because the time averaged distance can still be beyond a bound.
1278 To prevent the protons from being pulled too close together, a mixed
1279 approach can be used. In this approach, the penalty is zero when the
1280 instantaneous distance is within the bounds, otherwise the violation is
1281 the square root of the product of the instantaneous violation and the
1282 time averaged violation:
1283 \beq
1284 \ve{F}_i~=~ \left\{
1285 \begin{array}{lclll}
1286 k^a_{dr}\sqrt{(r_{ij}-r_0)(\bar{r}_{ij}-r_0)}\frac{\rvij}{r_{ij}}
1287 & \mbox{for} & r_{ij} < r_0 & \mbox{and} & \bar{r}_{ij} < r_0 \\[1.5ex]
1288 -k^a _{dr} \,
1289 \mbox{min}\left(\sqrt{(r_{ij}-r_1)(\bar{r}_{ij}-r_1)},r_2-r_1\right)
1290 \frac{\rvij}{r_{ij}}
1291 & \mbox{for} & r_{ij} > r_1 & \mbox{and} & \bar{r}_{ij} > r_1 \\[1.5ex]
1292 0 &\mbox{otherwise}
1293 \end{array} \right.
1294 \eeq
1296 \subsubsection{Averaging over multiple pairs\swapindexquiet{ensemble-averaged}{distance restraint}}
1298 Sometimes it is unclear from experimental data which atom pair
1299 gives rise to a single NOE, in other occasions it can be obvious that
1300 more than one pair contributes due to the symmetry of the system, {\eg} a
1301 methyl group with three protons. For such a group, it is not possible
1302 to distinguish between the protons, therefore they should all be taken into
1303 account when calculating the distance between this methyl group and another
1304 proton (or group of protons).
1305 Due to the physical nature of magnetic resonance, the intensity of the
1306 NOE signal is inversely proportional to the sixth power of the inter-atomic
1307 distance.
1308 Thus, when combining atom pairs,
1309 a fixed list of $N$ restraints may be taken together,
1310 where the apparent ``distance'' is given by:
1311 \beq
1312 r_N(t) = \left [\sum_{n=1}^{N} \bar{r}_{n}(t)^{-6} \right]^{-1/6}
1313 \label{eqn:rsix}
1314 \eeq
1315 where we use $r_{ij}$ or \eqnref{rav} for the $\bar{r}_{n}$.
1316 The $r_N$ of the instantaneous and time-averaged distances
1317 can be combined to do a mixed restraining, as indicated above.
1318 As more pairs of protons contribute to the same NOE signal, the intensity
1319 will increase, and the summed ``distance'' will be shorter than any of
1320 its components due to the reciprocal summation.
1322 There are two options for distributing the forces over the atom pairs.
1323 In the conservative option, the force is defined as the derivative of the
1324 restraint potential with respect to the coordinates. This results in
1325 a conservative potential when time averaging is not used.
1326 The force distribution over the pairs is proportional to $r^{-6}$.
1327 This means that a close pair feels a much larger force than a distant pair,
1328 which might lead to a molecule that is ``too rigid.''
1329 The other option is an equal force distribution. In this case each pair
1330 feels $1/N$ of the derivative of the restraint potential with respect to
1331 $r_N$. The advantage of this method is that more conformations might be
1332 sampled, but the non-conservative nature of the forces can lead to
1333 local heating of the protons.
1335 It is also possible to use {\em ensemble averaging} using multiple
1336 (protein) molecules. In this case the bounds should be lowered as in:
1337 \beq
1338 \begin{array}{rcl}
1339 r_1 &~=~& r_1 * M^{-1/6} \\
1340 r_2 &~=~& r_2 * M^{-1/6}
1341 \end{array}
1342 \eeq
1343 where $M$ is the number of molecules. The {\gromacs} preprocessor {\tt grompp}
1344 can do this automatically when the appropriate option is given.
1345 The resulting ``distance'' is
1346 then used to calculate the scalar force according to:
1347 \beq
1348 \ve{F}_i~=~\left\{
1349 \begin{array}{rcl}
1350 ~& 0 \hspace{4cm} & r_{N} < r_1 \\
1351 & k_{dr}(r_{N}-r_1)\frac{\rvij}{r_{ij}} & r_1 \le r_{N} < r_2 \\
1352 & k_{dr}(r_2-r_1)\frac{\rvij}{r_{ij}} & r_{N} \ge r_2
1353 \end{array} \right.
1354 \eeq
1355 where $i$ and $j$ denote the atoms of all the
1356 pairs that contribute to the NOE signal.
1358 \subsubsection{Using distance restraints}
1360 A list of distance restrains based on NOE data can be added to a molecule
1361 definition in your topology file, like in the following example:
1363 \begin{verbatim}
1364 [ distance_restraints ]
1365 ; ai aj type index type' low up1 up2 fac
1366 10 16 1 0 1 0.0 0.3 0.4 1.0
1367 10 28 1 1 1 0.0 0.3 0.4 1.0
1368 10 46 1 1 1 0.0 0.3 0.4 1.0
1369 16 22 1 2 1 0.0 0.3 0.4 2.5
1370 16 34 1 3 1 0.0 0.5 0.6 1.0
1371 \end{verbatim}
1373 In this example a number of features can be found. In columns {\tt
1374 ai} and {\tt aj} you find the atom numbers of the particles to be
1375 restrained. The {\tt type} column should always be 1. As explained in
1376 ~\ssecref{distancerestraint}, multiple distances can contribute to a single NOE
1377 signal. In the topology this can be set using the {\tt index}
1378 column. In our example, the restraints 10-28 and 10-46 both have index
1379 1, therefore they are treated simultaneously. An extra requirement
1380 for treating restraints together is that the restraints must be on
1381 successive lines, without any other intervening restraint. The {\tt
1382 type'} column will usually be 1, but can be set to 2 to obtain a
1383 distance restraint that will never be time- and ensemble-averaged;
1384 this can be useful for restraining hydrogen bonds. The columns {\tt
1385 low}, {\tt up1}, and {\tt up2} hold the values of $r_0$, $r_1$, and
1386 $r_2$ from ~\eqnref{disre}. In some cases it can be useful to have
1387 different force constants for some restraints; this is controlled by
1388 the column {\tt fac}. The force constant in the parameter file is
1389 multiplied by the value in the column {\tt fac} for each restraint.
1390 Information for each restraint is stored in the energy file and can
1391 be processed and plotted with {\tt gmx nmr}.
1393 \newcommand{\SSS}{{\mathbf S}}
1394 \newcommand{\DD}{{\mathbf D}}
1395 \newcommand{\RR}{{\mathbf R}}
1397 \subsection{Orientation restraints\swapindexquiet{orientation}{restraint}}
1398 \label{subsec:orientationrestraint}
1399 This section describes how orientations between vectors,
1400 as measured in certain NMR experiments, can be calculated
1401 and restrained in MD simulations.
1402 The presented refinement methodology and a comparison of results
1403 with and without time and ensemble averaging have been
1404 published~\cite{Hess2003}.
1405 \subsubsection{Theory}
1406 In an NMR experiment, orientations of vectors can be measured when a
1407 molecule does not tumble completely isotropically in the solvent.
1408 Two examples of such orientation measurements are
1409 residual \normindex{dipolar couplings}
1410 (between two nuclei) or chemical shift anisotropies.
1411 An observable for a vector $\ve{r}_i$ can be written as follows:
1412 \beq
1413 \delta_i = \frac{2}{3} \mbox{tr}(\SSS\DD_i)
1414 \eeq
1415 where $\SSS$ is the dimensionless order tensor of the molecule.
1416 The tensor $\DD_i$ is given by:
1417 \beq
1418 \label{orient_def}
1419 \DD_i = \frac{c_i}{\|\ve{r}_i\|^\alpha} \left(
1420 %\begin{array}{lll}
1421 %3 r_x r_x - \ve{r}\cdot\ve{r} & 3 r_x r_y & 3 r_x r_z \\
1422 %3 r_x r_y & 3 r_y r_y - \ve{r}\cdot\ve{r} & 3yz \\
1423 %3 r_x r_z & 3 r_y r_z & 3 r_z r_z - \ve{r}\cdot\ve{r}
1424 %\end{array} \right)
1425 \begin{array}{lll}
1426 3 x x - 1 & 3 x y & 3 x z \\
1427 3 x y & 3 y y - 1 & 3 y z \\
1428 3 x z & 3 y z & 3 z z - 1 \\
1429 \end{array} \right)
1430 \eeq
1431 \beq
1432 \mbox{with:} \quad
1433 x=\frac{r_{i,x}}{\|\ve{r}_i\|}, \quad
1434 y=\frac{r_{i,y}}{\|\ve{r}_i\|}, \quad
1435 z=\frac{r_{i,z}}{\|\ve{r}_i\|}
1436 \eeq
1437 For a dipolar coupling $\ve{r}_i$ is the vector connecting the two
1438 nuclei, $\alpha=3$ and the constant $c_i$ is given by:
1439 \beq
1440 c_i = \frac{\mu_0}{4\pi} \gamma_1^i \gamma_2^i \frac{\hbar}{4\pi}
1441 \eeq
1442 where $\gamma_1^i$ and $\gamma_2^i$ are the gyromagnetic ratios of the
1443 two nuclei.
1445 The order tensor is symmetric and has trace zero. Using a rotation matrix
1446 ${\mathbf T}$ it can be transformed into the following form:
1447 \beq
1448 {\mathbf T}^T \SSS {\mathbf T} = s \left( \begin{array}{ccc}
1449 -\frac{1}{2}(1-\eta) & 0 & 0 \\
1450 0 & -\frac{1}{2}(1+\eta) & 0 \\
1451 0 & 0 & 1
1452 \end{array} \right)
1453 \eeq
1454 where $-1 \leq s \leq 1$ and $0 \leq \eta \leq 1$.
1455 $s$ is called the order parameter and $\eta$ the asymmetry of the
1456 order tensor $\SSS$. When the molecule tumbles isotropically in the
1457 solvent, $s$ is zero, and no orientational effects can be observed
1458 because all $\delta_i$ are zero.
1460 %\newpage
1462 \subsubsection{Calculating orientations in a simulation}
1463 For reasons which are explained below, the $\DD$ matrices are calculated
1464 which respect to a reference orientation of the molecule. The orientation
1465 is defined by a rotation matrix $\RR$, which is needed to least-squares fit
1466 the current coordinates of a selected set of atoms onto
1467 a reference conformation. The reference conformation is the starting
1468 conformation of the simulation. In case of ensemble averaging, which will
1469 be treated later, the structure is taken from the first subsystem.
1470 The calculated $\DD_i^c$ matrix is given by:
1471 \begin{equation}
1472 \label{D_rot}
1473 \DD_i^c(t) = \RR(t) \DD_i(t) \RR^T(t)
1474 \end{equation}
1475 The calculated orientation for vector $i$ is given by:
1476 \beq
1477 \delta^c_i(t) = \frac{2}{3} \mbox{tr}(\SSS(t)\DD_i^c(t))
1478 \eeq
1479 The order tensor $\SSS(t)$ is usually unknown.
1480 A reasonable choice for the order tensor is the tensor
1481 which minimizes the (weighted) mean square difference between the calculated
1482 and the observed orientations:
1483 \begin{equation}
1484 \label{S_msd}
1485 MSD(t) = \left(\sum_{i=1}^N w_i\right)^{-1} \sum_{i=1}^N w_i (\delta_i^c (t) -\delta_i^{exp})^2
1486 \end{equation}
1487 To properly combine different types of measurements, the unit of $w_i$ should
1488 be such that all terms are dimensionless. This means the unit of $w_i$
1489 is the unit of $\delta_i$ to the power $-2$.
1490 {\bf Note} that scaling all $w_i$ with a constant factor does not influence
1491 the order tensor.
1493 \subsubsection{Time averaging}
1494 Since the tensors $\DD_i$ fluctuate rapidly in time, much faster than can
1495 be observed in an experiment, they should be averaged over time in the simulation.
1496 However, in a simulation the time and the number of copies of
1497 a molecule are limited. Usually one can not obtain a converged average
1498 of the $\DD_i$ tensors over all orientations of the molecule.
1499 If one assumes that the average orientations of the $\ve{r}_i$ vectors
1500 within the molecule converge much faster than the tumbling time of
1501 the molecule, the tensor can be averaged in an axis system that
1502 rotates with the molecule, as expressed by equation~(\ref{D_rot}).
1503 The time-averaged tensors are calculated
1504 using an exponentially decaying memory function:
1505 \beq
1506 \DD^a_i(t) = \frac{\displaystyle
1507 \int_{u=t_0}^t \DD^c_i(u) \exp\left(-\frac{t-u}{\tau}\right)\mbox{d} u
1508 }{\displaystyle
1509 \int_{u=t_0}^t \exp\left(-\frac{t-u}{\tau}\right)\mbox{d} u
1511 \eeq
1512 Assuming that the order tensor $\SSS$ fluctuates slower than the
1513 $\DD_i$, the time-averaged orientation can be calculated as:
1514 \beq
1515 \delta_i^a(t) = \frac{2}{3} \mbox{tr}(\SSS(t) \DD_i^a(t))
1516 \eeq
1517 where the order tensor $\SSS(t)$ is calculated using expression~(\ref{S_msd})
1518 with $\delta_i^c(t)$ replaced by $\delta_i^a(t)$.
1520 \subsubsection{Restraining}
1521 The simulated structure can be restrained by applying a force proportional
1522 to the difference between the calculated and the experimental orientations.
1523 When no time averaging is applied, a proper potential can be defined as:
1524 \beq
1525 V = \frac{1}{2} k \sum_{i=1}^N w_i (\delta_i^c (t) -\delta_i^{exp})^2
1526 \eeq
1527 where the unit of $k$ is the unit of energy.
1528 Thus the effective force constant for restraint $i$ is $k w_i$.
1529 The forces are given by minus the gradient of $V$.
1530 The force $\ve{F}\!_i$ working on vector $\ve{r}_i$ is:
1531 \begin{eqnarray*}
1532 \ve{F}\!_i(t)
1533 & = & - \frac{\mbox{d} V}{\mbox{d}\ve{r}_i} \\
1534 & = & -k w_i (\delta_i^c (t) -\delta_i^{exp}) \frac{\mbox{d} \delta_i (t)}{\mbox{d}\ve{r}_i} \\
1535 & = & -k w_i (\delta_i^c (t) -\delta_i^{exp})
1536 \frac{2 c_i}{\|\ve{r}\|^{2+\alpha}} \left(2 \RR^T \SSS \RR \ve{r}_i - \frac{2+\alpha}{\|\ve{r}\|^2} \mbox{tr}(\RR^T \SSS \RR \ve{r}_i \ve{r}_i^T) \ve{r}_i \right)
1537 \end{eqnarray*}
1539 \subsubsection{Ensemble averaging}
1540 Ensemble averaging can be applied by simulating a system of $M$ subsystems
1541 that each contain an identical set of orientation restraints. The systems only
1542 interact via the orientation restraint potential which is defined as:
1543 \beq
1544 V = M \frac{1}{2} k \sum_{i=1}^N w_i
1545 \langle \delta_i^c (t) -\delta_i^{exp} \rangle^2
1546 \eeq
1547 The force on vector $\ve{r}_{i,m}$ in subsystem $m$ is given by:
1548 \beq
1549 \ve{F}\!_{i,m}(t) = - \frac{\mbox{d} V}{\mbox{d}\ve{r}_{i,m}} =
1550 -k w_i \langle \delta_i^c (t) -\delta_i^{exp} \rangle \frac{\mbox{d} \delta_{i,m}^c (t)}{\mbox{d}\ve{r}_{i,m}} \\
1551 \eeq
1553 \subsubsection{Time averaging}
1554 When using time averaging it is not possible to define a potential.
1555 We can still define a quantity that gives a rough idea of the energy
1556 stored in the restraints:
1557 \beq
1558 V = M \frac{1}{2} k^a \sum_{i=1}^N w_i
1559 \langle \delta_i^a (t) -\delta_i^{exp} \rangle^2
1560 \eeq
1561 The force constant $k_a$ is switched on slowly to compensate for the
1562 lack of history at times close to $t_0$. It is exactly proportional
1563 to the amount of average that has been accumulated:
1564 \beq
1565 k^a =
1566 k \, \frac{1}{\tau}\int_{u=t_0}^t \exp\left(-\frac{t-u}{\tau}\right)\mbox{d} u
1567 \eeq
1568 What really matters is the definition of the force. It is chosen to
1569 be proportional to the square root of the product of the time-averaged
1570 and the instantaneous deviation.
1571 Using only the time-averaged deviation induces large oscillations.
1572 The force is given by:
1573 \beq
1574 \ve{F}\!_{i,m}(t) =
1575 %\left\{ \begin{array}{ll}
1576 %0 & \mbox{for} \quad \langle \delta_i^a (t) -\delta_i^{exp} \rangle \langle \delta_i (t) -\delta_i^{exp} \rangle \leq 0 \\
1577 %... & \mbox{for} \quad \langle \delta_i^a (t) -\delta_i^{exp} \rangle \langle \delta_i (t) -\delta_i^{exp} \rangle > 0
1578 %\end{array}
1579 %\right.
1580 \left\{ \begin{array}{ll}
1581 0 & \quad \mbox{for} \quad a\, b \leq 0 \\
1582 \displaystyle
1583 k^a w_i \frac{a}{|a|} \sqrt{a\, b} \, \frac{\mbox{d} \delta_{i,m}^c (t)}{\mbox{d}\ve{r}_{i,m}}
1584 & \quad \mbox{for} \quad a\, b > 0
1585 \end{array}
1586 \right.
1587 \eeq
1588 \begin{eqnarray*}
1589 a &=& \langle \delta_i^a (t) -\delta_i^{exp} \rangle \\
1590 b &=& \langle \delta_i^c (t) -\delta_i^{exp} \rangle
1591 \end{eqnarray*}
1593 \subsubsection{Using orientation restraints}
1594 Orientation restraints can be added to a molecule definition in
1595 the topology file in the section {\tt [~orientation_restraints~]}.
1596 Here we give an example section containing five N-H residual dipolar
1597 coupling restraints:
1599 \begin{verbatim}
1600 [ orientation_restraints ]
1601 ; ai aj type exp. label alpha const. obs. weight
1602 ; Hz nm^3 Hz Hz^-2
1603 31 32 1 1 3 3 6.083 -6.73 1.0
1604 43 44 1 1 4 3 6.083 -7.87 1.0
1605 55 56 1 1 5 3 6.083 -7.13 1.0
1606 65 66 1 1 6 3 6.083 -2.57 1.0
1607 73 74 1 1 7 3 6.083 -2.10 1.0
1608 \end{verbatim}
1610 The unit of the observable is Hz, but one can choose any other unit.
1611 In columns {\tt
1612 ai} and {\tt aj} you find the atom numbers of the particles to be
1613 restrained. The {\tt type} column should always be 1.
1614 The {\tt exp.} column denotes the experiment number, starting
1615 at 1. For each experiment a separate order tensor $\SSS$
1616 is optimized. The label should be a unique number larger than zero
1617 for each restraint. The {\tt alpha} column contains the power $\alpha$
1618 that is used in equation~(\ref{orient_def}) to calculate the orientation.
1619 The {\tt const.} column contains the constant $c_i$ used in the same
1620 equation. The constant should have the unit of the observable times
1621 nm$^\alpha$. The column {\tt obs.} contains the observable, in any
1622 unit you like. The last column contains the weights $w_i$; the unit
1623 should be the inverse of the square of the unit of the observable.
1625 Some parameters for orientation restraints can be specified in the
1626 {\tt grompp.mdp} file, for a study of the effect of different
1627 force constants and averaging times and ensemble averaging see~\cite{Hess2003}.
1628 Information for each restraint is stored in the energy file and can
1629 be processed and plotted with {\tt gmx nmr}.
1631 \section{Polarization}
1632 Polarization can be treated by {\gromacs} by attaching
1633 \normindex{shell} (\normindex{Drude}) particles to atoms and/or
1634 virtual sites. The energy of the shell particle is then minimized at
1635 each time step in order to remain on the Born-Oppenheimer surface.
1637 \subsection{Simple polarization}
1638 This is implemented as a harmonic potential with equilibrium distance
1640 The input given in the topology file is the polarizability $\alpha$ (in
1641 {\gromacs} units) as follows:
1642 \begin{verbatim}
1643 [ polarization ]
1644 ; Atom i j type alpha
1645 1 2 1 0.001
1646 \end{verbatim}
1647 in this case the polarizability volume is 0.001 nm$^3$ (or 1
1648 {\AA$^3$}). In order to compute the harmonic force constant $k_{cs}$
1649 (where $cs$ stands for core-shell), the
1650 following is used~\cite{Maaren2001a}:
1651 \begin{equation}
1652 k_{cs} ~=~ \frac{q_s^2}{\alpha}
1653 \end{equation}
1654 where $q_s$ is the charge on the shell particle.
1656 \subsection{Anharmonic polarization}
1657 For the development of the Drude force field by Roux and McKerell~\cite{Lopes2013a}
1658 it was found
1659 that some particles can overpolarize and this was fixed by introducing
1660 a higher order term in the polarization energy:
1661 \begin{eqnarray}
1662 V_{pol} ~=& \frac{k_{cs}}{2} r_{cs}^2 & r_{cs} \le \delta \\
1663 =& \frac{k_{cs}}{2} r_{cs}^2 + k_{hyp} (r_{cs}-\delta)^4 & r_{cs} > \delta
1664 \end{eqnarray}
1665 where $\delta$ is a user-defined constant that is set to 0.02 nm for
1666 anions in the Drude force field~\cite{HYu2010}. Since this original introduction it
1667 has also been used in other atom types~\cite{Lopes2013a}.
1668 \begin{verbatim}
1669 [ polarization ]
1670 ;Atom i j type alpha (nm^3) delta khyp
1671 1 2 2 0.001786 0.02 16.736e8
1672 \end{verbatim}
1673 The above force constant $k_{hyp}$ corresponds to 4$\cdot$10$^8$
1674 kcal/mol/nm$^4$, hence the strange number.
1676 \subsection{Water polarization}
1677 A special potential for water that allows anisotropic polarization of
1678 a single shell particle~\cite{Maaren2001a}.
1680 \subsection{Thole polarization}
1681 Based on early work by \normindex{Thole}~\cite{Thole81}, Roux and
1682 coworkers have implemented potentials for molecules like
1683 ethanol~\cite{Lamoureux2003a,Lamoureux2003b,Noskov2005a}. Within such
1684 molecules, there are intra-molecular interactions between shell
1685 particles, however these must be screened because full Coulomb would
1686 be too strong. The potential between two shell particles $i$ and $j$ is:
1687 \newcommand{\rbij}{\bar{r}_{ij}}
1688 \beq
1689 V_{thole} ~=~ \frac{q_i q_j}{r_{ij}}\left[1-\left(1+\frac{\rbij}{2}\right){\rm exp}^{-\rbij}\right]
1690 \eeq
1691 {\bf Note} that there is a sign error in Equation~1 of Noskov {\em et al.}~\cite{Noskov2005a}:
1692 \beq
1693 \rbij ~=~ a\frac{r_{ij}}{(\alpha_i \alpha_j)^{1/6}}
1694 \eeq
1695 where $a$ is a magic (dimensionless) constant, usually chosen to be
1696 2.6~\cite{Noskov2005a}; $\alpha_i$ and $\alpha_j$ are the polarizabilities
1697 of the respective shell particles.
1700 \section{Free energy interactions}
1701 \label{sec:feia}
1702 \index{free energy interactions}
1703 \newcommand{\LAM}{\lambda}
1704 \newcommand{\LL}{(1-\LAM)}
1705 \newcommand{\dvdl}[1]{\frac{\partial #1}{\partial \LAM}}
1706 This section describes the $\lambda$-dependence of the potentials
1707 used for free energy calculations (see \secref{fecalc}).
1708 All common types of potentials and constraints can be
1709 interpolated smoothly from state A ($\lambda=0$) to state B
1710 ($\lambda=1$) and vice versa.
1711 All bonded interactions are interpolated by linear interpolation
1712 of the interaction parameters. Non-bonded interactions can be
1713 interpolated linearly or via soft-core interactions.
1715 Starting in {\gromacs} 4.6, $\lambda$ is a vector, allowing different
1716 components of the free energy transformation to be carried out at
1717 different rates. Coulomb, Lennard-Jones, bonded, and restraint terms
1718 can all be controlled independently, as described in the {\tt .mdp}
1719 options.
1721 \subsubsection{Harmonic potentials}
1722 The example given here is for the bond potential, which is harmonic
1723 in {\gromacs}. However, these equations apply to the angle potential
1724 and the improper dihedral potential as well.
1725 \bea
1726 V_b &=&\half\left[\LL k_b^A +
1727 \LAM k_b^B\right] \left[b - \LL b_0^A - \LAM b_0^B\right]^2 \\
1728 \dvdl{V_b}&=&\half(k_b^B-k_b^A)
1729 \left[b - \LL b_0^A + \LAM b_0^B\right]^2 +
1730 \nonumber\\
1731 & & \phantom{\half}(b_0^A-b_0^B) \left[b - \LL b_0^A -\LAM b_0^B\right]
1732 \left[\LL k_b^A + \LAM k_b^B \right]
1733 \eea
1735 \subsubsection{\gromosv{96} bonds and angles}
1736 Fourth-power bond stretching and cosine-based angle potentials
1737 are interpolated by linear interpolation of the force constant
1738 and the equilibrium position. Formulas are not given here.
1740 \subsubsection{Proper dihedrals}
1741 For the proper dihedrals, the equations are somewhat more complicated:
1742 \bea
1743 V_d &=&\left[\LL k_d^A + \LAM k_d^B \right]
1744 \left( 1+ \cos\left[n_{\phi} \phi -
1745 \LL \phi_s^A - \LAM \phi_s^B
1746 \right]\right)\\
1747 \dvdl{V_d}&=&(k_d^B-k_d^A)
1748 \left( 1+ \cos
1749 \left[
1750 n_{\phi} \phi- \LL \phi_s^A - \LAM \phi_s^B
1751 \right]
1752 \right) +
1753 \nonumber\\
1754 &&(\phi_s^B - \phi_s^A) \left[\LL k_d^A - \LAM k_d^B\right]
1755 \sin\left[ n_{\phi}\phi - \LL \phi_s^A - \LAM \phi_s^B \right]
1756 \eea
1757 {\bf Note:} that the multiplicity $n_{\phi}$ can not be parameterized
1758 because the function should remain periodic on the interval $[0,2\pi]$.
1760 \subsubsection{Tabulated bonded interactions}
1761 For tabulated bonded interactions only the force constant can interpolated:
1762 \bea
1763 V &=& (\LL k^A + \LAM k^B) \, f \\
1764 \dvdl{V} &=& (k^B - k^A) \, f
1765 \eea
1767 \subsubsection{Coulomb interaction}
1768 The \normindex{Coulomb} interaction between two particles
1769 of which the charge varies with $\LAM$ is:
1770 \bea
1771 V_c &=& \frac{f}{\epsrf \rij}\left[\LL q_i^A q_j^A + \LAM\, q_i^B q_j^B\right] \\
1772 \dvdl{V_c}&=& \frac{f}{\epsrf \rij}\left[- q_i^A q_j^A + q_i^B q_j^B\right]
1773 \eea
1774 where $f = \frac{1}{4\pi \varepsilon_0} = \electricConvFactorValue$ (see \chref{defunits}).
1776 \subsubsection{Coulomb interaction with \normindex{reaction field}}
1777 The Coulomb interaction including a reaction field, between two particles
1778 of which the charge varies with $\LAM$ is:
1779 \bea
1780 V_c &=& f\left[\frac{1}{\rij} + k_{rf}~ \rij^2 -c_{rf}\right]
1781 \left[\LL q_i^A q_j^A + \LAM\, q_i^B q_j^B\right] \\
1782 \dvdl{V_c}&=& f\left[\frac{1}{\rij} + k_{rf}~ \rij^2 -c_{rf}\right]
1783 \left[- q_i^A q_j^A + q_i^B q_j^B\right]
1784 \label{eq:dVcoulombdlambda}
1785 \eea
1786 {\bf Note} that the constants $k_{rf}$ and $c_{rf}$ are
1787 defined using the dielectric
1788 constant $\epsrf$ of the medium (see \secref{coulrf}).
1790 \subsubsection{Lennard-Jones interaction}
1791 For the \normindex{Lennard-Jones} interaction between two particles
1792 of which the {\em atom type} varies with $\LAM$ we can write:
1793 \bea
1794 V_{LJ} &=& \frac{\LL C_{12}^A + \LAM\, C_{12}^B}{\rij^{12}} -
1795 \frac{\LL C_6^A + \LAM\, C_6^B}{\rij^6} \\
1796 \dvdl{V_{LJ}}&=&\frac{C_{12}^B - C_{12}^A}{\rij^{12}} -
1797 \frac{C_6^B - C_6^A}{\rij^6}
1798 \label{eq:dVljdlambda}
1799 \eea
1800 It should be noted that it is also possible to express a pathway from
1801 state A to state B using $\sigma$ and $\epsilon$ (see \eqnref{sigeps}).
1802 It may seem to make sense physically to vary the force field parameters
1803 $\sigma$ and $\epsilon$ rather
1804 than the derived parameters $C_{12}$ and $C_{6}$.
1805 However, the difference between the pathways in parameter space
1806 is not large, and the free energy itself
1807 does not depend on the pathway, so we use the simple formulation
1808 presented above.
1810 \subsubsection{Kinetic Energy}
1811 When the mass of a particle changes, there is also a contribution of
1812 the kinetic energy to the free energy (note that we can not write
1813 the momentum \ve{p} as m\ve{v}, since that would result
1814 in the sign of $\dvdl{E_k}$ being incorrect~\cite{Gunsteren98a}):
1816 \bea
1817 E_k &=& \half\frac{\ve{p}^2}{\LL m^A + \LAM m^B} \\
1818 \dvdl{E_k}&=& -\half\frac{\ve{p}^2(m^B-m^A)}{(\LL m^A + \LAM m^B)^2}
1819 \eea
1820 after taking the derivative, we {\em can} insert \ve{p} = m\ve{v}, such that:
1821 \beq
1822 \dvdl{E_k}~=~ -\half\ve{v}^2(m^B-m^A)
1823 \eeq
1825 \subsubsection{Constraints}
1826 \label{subsubsec:constraints}
1827 The constraints are formally part of the Hamiltonian, and therefore
1828 they give a contribution to the free energy. In {\gromacs} this can be
1829 calculated using the \normindex{LINCS} or the \normindex{SHAKE} algorithm.
1830 If we have $k = 1 \ldots K$ constraint equations $g_k$ for LINCS, then
1831 \beq
1832 g_k = |\ve{r}_{k}| - d_{k}
1833 \eeq
1834 where $\ve{r}_k$ is the displacement vector between two particles and
1835 $d_k$ is the constraint distance between the two particles. We can express
1836 the fact that the constraint distance has a $\LAM$ dependency by
1837 \beq
1838 d_k = \LL d_{k}^A + \LAM d_k^B
1839 \eeq
1841 Thus the $\LAM$-dependent constraint equation is
1842 \beq
1843 g_k = |\ve{r}_{k}| - \left(\LL d_{k}^A + \LAM d_k^B\right).
1844 \eeq
1846 The (zero) contribution $G$ to the Hamiltonian from the constraints
1847 (using Lagrange multipliers $\lambda_k$, which are logically distinct
1848 from the free-energy $\LAM$) is
1849 \bea
1850 G &=& \sum^K_k \lambda_k g_k \\
1851 \dvdl{G} &=& \frac{\partial G}{\partial d_k} \dvdl{d_k} \\
1852 &=& - \sum^K_k \lambda_k \left(d_k^B-d_k^A\right)
1853 \eea
1855 For SHAKE, the constraint equations are
1856 \beq
1857 g_k = \ve{r}_{k}^2 - d_{k}^2
1858 \eeq
1859 with $d_k$ as before, so
1860 \bea
1861 \dvdl{G} &=& -2 \sum^K_k \lambda_k \left(d_k^B-d_k^A\right)
1862 \eea
1864 \subsection{Soft-core interactions\index{soft-core interactions}}
1865 \begin{figure}
1866 \centerline{\includegraphics[height=6cm]{plots/softcore}}
1867 \caption{Soft-core interactions at $\LAM=0.5$, with $p=2$ and
1868 $C_6^A=C_{12}^A=C_6^B=C_{12}^B=1$.}
1869 \label{fig:softcore}
1870 \end{figure}
1871 In a free-energy calculation where particles grow out of nothing, or
1872 particles disappear, using the the simple linear interpolation of the
1873 Lennard-Jones and Coulomb potentials as described in Equations~\ref{eq:dVljdlambda}
1874 and \ref{eq:dVcoulombdlambda} may lead to poor convergence. When the particles have nearly disappeared, or are close to appearing (at $\LAM$ close to 0 or 1), the interaction energy will be weak enough for particles to get very
1875 close to each other, leading to large fluctuations in the measured values of
1876 $\partial V/\partial \LAM$ (which, because of the simple linear
1877 interpolation, depends on the potentials at both the endpoints of $\LAM$).
1879 To circumvent these problems, the singularities in the potentials need to be removed. This can be done by modifying the regular Lennard-Jones and Coulomb potentials with ``soft-core'' potentials that limit the energies and forces
1880 involved at $\LAM$ values between 0 and 1, but not \emph{at} $\LAM=0$
1881 or 1.
1883 In {\gromacs} the soft-core potentials $V_{sc}$ are shifted versions of the
1884 regular potentials, so that the singularity in the potential and its
1885 derivatives at $r=0$ is never reached:
1886 \bea
1887 V_{sc}(r) &=& \LL V^A(r_A) + \LAM V^B(r_B)
1889 r_A &=& \left(\alpha \sigma_A^6 \LAM^p + r^6 \right)^\frac{1}{6}
1891 r_B &=& \left(\alpha \sigma_B^6 \LL^p + r^6 \right)^\frac{1}{6}
1892 \eea
1893 where $V^A$ and $V^B$ are the normal ``hard core'' Van der Waals or
1894 electrostatic potentials in state A ($\LAM=0$) and state B ($\LAM=1$)
1895 respectively, $\alpha$ is the soft-core parameter (set with {\tt sc_alpha}
1896 in the {\tt .mdp} file), $p$ is the soft-core $\LAM$ power (set with
1897 {\tt sc_power}), $\sigma$ is the radius of the interaction, which is
1898 $(C_{12}/C_6)^{1/6}$ or an input parameter ({\tt sc_sigma}) when $C_6$
1899 or $C_{12}$ is zero.
1901 For intermediate $\LAM$, $r_A$ and $r_B$ alter the interactions very little
1902 for $r > \alpha^{1/6} \sigma$ and quickly switch the soft-core
1903 interaction to an almost constant value for smaller $r$ (\figref{softcore}).
1904 The force is:
1905 \beq
1906 F_{sc}(r) = -\frac{\partial V_{sc}(r)}{\partial r} =
1907 \LL F^A(r_A) \left(\frac{r}{r_A}\right)^5 +
1908 \LAM F^B(r_B) \left(\frac{r}{r_B}\right)^5
1909 \eeq
1910 where $F^A$ and $F^B$ are the ``hard core'' forces.
1911 The contribution to the derivative of the free energy is:
1912 \bea
1913 \dvdl{V_{sc}(r)} & = &
1914 V^B(r_B) -V^A(r_A) +
1915 \LL \frac{\partial V^A(r_A)}{\partial r_A}
1916 \frac{\partial r_A}{\partial \LAM} +
1917 \LAM\frac{\partial V^B(r_B)}{\partial r_B}
1918 \frac{\partial r_B}{\partial \LAM}
1919 \nonumber\\
1921 V^B(r_B) -V^A(r_A) + \nonumber \\
1923 \frac{p \alpha}{6}
1924 \left[ \LAM F^B(r_B) r^{-5}_B \sigma_B^6 \LL^{p-1} -
1925 \LL F^A(r_A) r^{-5}_A \sigma_A^6 \LAM^{p-1} \right]
1926 \eea
1928 The original GROMOS Lennard-Jones soft-core function~\cite{Beutler94}
1929 uses $p=2$, but $p=1$ gives a smoother $\partial H/\partial\LAM$ curve.
1930 %When the changes between the two states involve both the disappearing
1931 %and appearing of atoms, it is important that the overlapping of atoms
1932 %happens around $\LAM=0.5$. This can usually be achieved with
1933 %$\alpha$$\approx0.7$ for $p=1$ and $\alpha$$\approx1.5$ for $p=2$.
1934 %MRS: this is now eliminated as of 4.6, since changes between atoms are done linearly.
1936 Another issue that should be considered is the soft-core effect of hydrogens
1937 without Lennard-Jones interaction. Their soft-core $\sigma$ is
1938 set with {\tt sc-sigma} in the {\tt .mdp} file. These hydrogens
1939 produce peaks in $\partial H/\partial\LAM$ at $\LAM$ is 0 and/or 1 for $p=1$
1940 and close to 0 and/or 1 with $p=2$. Lowering {\tt\mbox{sc-sigma}} will decrease
1941 this effect, but it will also increase the interactions with hydrogens
1942 relative to the other interactions in the soft-core state.
1944 When soft-core potentials are selected (by setting {\tt sc-alpha} \textgreater
1945 0), and the Coulomb and Lennard-Jones potentials are turned on or off
1946 sequentially, then the Coulombic interaction is turned off linearly,
1947 rather than using soft-core interactions, which should be less
1948 statistically noisy in most cases. This behavior can be overwritten
1949 by using the mdp option {\tt sc-coul} to {\tt yes}. Note that the {\tt sc-coul}
1950 is only taken into account when lambda states are used, not with
1951 {\tt couple-lambda0}~/ {\tt couple-lambda1}, and you can still turn off soft-core
1952 interactions by setting {\tt sc-alpha=0}. Additionally, the soft-core
1953 interaction potential is only applied when either the A or B
1954 state has zero interaction potential. If both A and B states have
1955 nonzero interaction potential, default linear scaling described above
1956 is used. When both Coulombic and Lennard-Jones interactions are turned
1957 off simultaneously, a soft-core potential is used, and a hydrogen is
1958 being introduced or deleted, the sigma is set to {\tt sc-sigma-min},
1959 which itself defaults to {\tt sc-sigma-default}.
1961 Recently, a new formulation of the soft-core approach has been derived
1962 that in most cases gives lower and more even statistical variance than
1963 the standard soft-core path described above.~\cite{Pham2011,Pham2012}
1964 Specifically, we have:
1965 \bea
1966 V_{sc}(r) &=& \LL V^A(r_A) + \LAM V^B(r_B)
1968 r_A &=& \left(\alpha \sigma_A^{48} \LAM^p + r^{48} \right)^\frac{1}{48}
1970 r_B &=& \left(\alpha \sigma_B^{48} \LL^p + r^{48} \right)^\frac{1}{48}
1971 \eea
1972 This ``1-1-48'' path is also implemented in {\gromacs}. Note that for this path the soft core $\alpha$
1973 should satisfy $0.001 < \alpha < 0.003$, rather than $\alpha \approx
1974 0.5$.
1977 \section{Methods}
1978 \subsection{Exclusions and 1-4 Interactions.}
1979 Atoms within a molecule that are close by in the chain,
1980 {\ie} atoms that are covalently bonded, or linked by one or two
1981 atoms are called {\em first neighbors, second neighbors} and
1982 {\em \swapindex{third}{neighbor}s}, respectively (see \figref{chain}).
1983 Since the interactions of atom {\bf i} with atoms {\bf i+1} and {\bf i+2}
1984 are mainly quantum mechanical, they can not be modeled by a Lennard-Jones potential.
1985 Instead it is assumed that these interactions are adequately modeled
1986 by a harmonic bond term or constraint ({\bf i, i+1}) and a harmonic angle term
1987 ({\bf i, i+2}). The first and second neighbors (atoms {\bf i+1} and {\bf i+2})
1988 are therefore
1989 {\em excluded} from the Lennard-Jones \swapindex{interaction}{list}
1990 of atom {\bf i};
1991 atoms {\bf i+1} and {\bf i+2} are called {\em \normindex{exclusions}} of atom {\bf i}.
1993 \begin{figure}
1994 \centerline{\includegraphics[width=8cm]{plots/chain}}
1995 \caption{Atoms along an alkane chain.}
1996 \label{fig:chain}
1997 \end{figure}
1999 For third neighbors, the normal Lennard-Jones repulsion is sometimes
2000 still too strong, which means that when applied to a molecule, the
2001 molecule would deform or break due to the internal strain. This is
2002 especially the case for carbon-carbon interactions in a {\em
2003 cis}-conformation ({\eg} {\em cis}-butane). Therefore, for some of these
2004 interactions, the Lennard-Jones repulsion has been reduced in the
2005 {\gromos} force field, which is implemented by keeping a separate list of
2006 1-4 and normal Lennard-Jones parameters. In other force fields, such
2007 as OPLS~\cite{Jorgensen88}, the standard Lennard-Jones parameters are reduced
2008 by a factor of two, but in that case also the dispersion (r$^{-6}$)
2009 and the Coulomb interaction are scaled.
2010 {\gromacs} can use either of these methods.
2012 \subsection{Charge Groups\index{charge group}}
2013 \label{sec:cg}
2014 In principle, the force calculation in MD is an $O(N^2)$ problem.
2015 Therefore, we apply a \normindex{cut-off} for non-bonded force (NBF)
2016 calculations; only the particles within a certain distance of each
2017 other are interacting. This reduces the cost to $O(N)$ (typically
2018 $100N$ to $200N$) of the NBF. It also introduces an error, which is,
2019 in most cases, acceptable, except when applying the cut-off implies
2020 the creation of charges, in which case you should consider using the
2021 lattice sum methods provided by {\gromacs}.
2023 Consider a water molecule interacting with another atom. If we would apply
2024 a plain cut-off on an atom-atom basis we might include the atom-oxygen
2025 interaction (with a charge of $-0.82$) without the compensating charge
2026 of the protons, and as a result, induce a large dipole moment over the system.
2027 Therefore, we have to keep groups of atoms with total charge
2028 0 together. These groups are called {\em charge groups}. Note that with
2029 a proper treatment of long-range electrostatics (e.g. particle-mesh Ewald
2030 (\secref{pme}), keeping charge groups together is not required.
2032 \subsection{Treatment of Cut-offs in the group scheme\index{cut-off}}
2033 \newcommand{\rs}{$r_{short}$}
2034 \newcommand{\rl}{$r_{long}$}
2035 {\gromacs} is quite flexible in treating cut-offs, which implies
2036 there can be quite a number of parameters to set. These parameters are
2037 set in the input file for {\tt grompp}. There are two sort of parameters
2038 that affect the cut-off interactions; you can select which type
2039 of interaction to use in each case, and which cut-offs should be
2040 used in the neighbor searching.
2042 For both Coulomb and van der Waals interactions there are interaction
2043 type selectors (termed {\tt vdwtype} and {\tt coulombtype}) and two
2044 parameters, for a total of six non-bonded interaction parameters. See
2045 the User Guide for a complete description of these parameters.
2047 In the group cut-off scheme, all of the interaction functions in \tabref{funcparm}
2048 require that neighbor searching be done with a radius at least as large as the $r_c$
2049 specified for the functional form, because of the use of charge groups.
2050 The extra radius is typically of the order of 0.25 nm (roughly the
2051 largest distance between two atoms in a charge group plus the distance a
2052 charge group can diffuse within neighbor list updates).
2054 \begin{table}[ht]
2055 \centering
2056 \begin{tabular}{|ll|l|}
2057 \dline
2058 \multicolumn{2}{|c|}{Type} & Parameters \\
2059 \hline
2060 Coulomb&Plain cut-off & $r_c$, $\epsr$ \\
2061 &Reaction field & $r_c$, $\epsrf$ \\
2062 &Shift function & $r_1$, $r_c$, $\epsr$ \\
2063 &Switch function & $r_1$, $r_c$, $\epsr$ \\
2064 \hline
2065 VdW&Plain cut-off & $r_c$ \\
2066 &Shift function & $r_1$, $r_c$ \\
2067 &Switch function & $r_1$, $r_c$ \\
2068 \dline
2069 \end{tabular}
2070 \caption[Parameters for the different functional forms of the
2071 non-bonded interactions.]{Parameters for the different functional
2072 forms of the non-bonded interactions.}
2073 \label{tab:funcparm}
2074 \end{table}
2077 \newcommand{\vvis}{\ve{r}_s}
2078 \newcommand{\Fi}{\ve{F}_i'}
2079 \newcommand{\Fj}{\ve{F}_j'}
2080 \newcommand{\Fk}{\ve{F}_k'}
2081 \newcommand{\Fl}{\ve{F}_l'}
2082 \newcommand{\Fvis}{\ve{F}_{s}}
2083 \newcommand{\rvik}{\ve{r}_{ik}}
2084 \newcommand{\rvis}{\ve{r}_{is}}
2085 \newcommand{\rvjk}{\ve{r}_{jk}}
2086 \newcommand{\rvjl}{\ve{r}_{jl}}
2089 \section{Virtual interaction sites\index{virtual interaction sites}}
2090 \label{sec:virtual_sites}
2091 Virtual interaction sites (called \seeindex{dummy atoms}{virtual interaction sites} in {\gromacs} versions before 3.3)
2092 can be used in {\gromacs} in a number of ways.
2093 We write the position of the virtual site $\ve{r}_s$ as a function of
2094 the positions of other particles \ve{r}$_i$: $\ve{r}_s =
2095 f(\ve{r}_1..\ve{r}_n)$. The virtual site, which may carry charge or be
2096 involved in other interactions, can now be used in the force
2097 calculation. The force acting on the virtual site must be
2098 redistributed over the particles with mass in a consistent way.
2099 A good way to do this can be found in ref.~\cite{Berendsen84b}.
2100 We can write the potential energy as:
2101 \beq
2102 V = V(\vvis,\ve{r}_1,\ldots,\ve{r}_n) = V^*(\ve{r}_1,\ldots,\ve{r}_n)
2103 \eeq
2104 The force on the particle $i$ is then:
2105 \beq
2106 \ve{F}_i = -\frac{\partial V^*}{\partial \ve{r}_i}
2107 = -\frac{\partial V}{\partial \ve{r}_i} -
2108 \frac{\partial V}{\partial \vvis}
2109 \frac{\partial \vvis}{\partial \ve{r}_i}
2110 = \ve{F}_i^{direct} + \Fi
2111 \eeq
2112 The first term is the normal force.
2113 The second term is the force on particle $i$ due to the virtual site, which
2114 can be written in tensor notation:
2115 \newcommand{\partd}[2]{\displaystyle\frac{\partial #1}{\partial #2_i}}
2116 \beq
2117 \Fi = \left[\begin{array}{ccc}
2118 \partd{x_s}{x} & \partd{y_s}{x} & \partd{z_s}{x} \\[1ex]
2119 \partd{x_s}{y} & \partd{y_s}{y} & \partd{z_s}{y} \\[1ex]
2120 \partd{x_s}{z} & \partd{y_s}{z} & \partd{z_s}{z}
2121 \end{array}\right]\Fvis
2122 \label{eqn:fvsite}
2123 \eeq
2124 where $\Fvis$ is the force on the virtual site and $x_s$, $y_s$ and
2125 $z_s$ are the coordinates of the virtual site. In this way, the total
2126 force and the total torque are conserved~\cite{Berendsen84b}.
2128 The computation of the \normindex{virial}
2129 (\eqnref{Xi}) is non-trivial when virtual sites are used. Since the
2130 virial involves a summation over all the atoms (rather than virtual
2131 sites), the forces must be redistributed from the virtual sites to the
2132 atoms (using ~\eqnref{fvsite}) {\em before} computation of the
2133 virial. In some special cases where the forces on the atoms can be
2134 written as a linear combination of the forces on the virtual sites (types 2
2135 and 3 below) there is no difference between computing the virial
2136 before and after the redistribution of forces. However, in the
2137 general case redistribution should be done first.
2139 \begin{figure}
2140 \centerline{\includegraphics[width=15cm]{plots/dummies}}
2141 \caption[Virtual site construction.]{The six different types of virtual
2142 site construction in \protect{\gromacs}. The constructing atoms are
2143 shown as black circles, the virtual sites in gray.}
2144 \label{fig:vsites}
2145 \end{figure}
2147 There are six ways to construct virtual sites from surrounding atoms in
2148 {\gromacs}, which we classify by the number of constructing
2149 atoms. {\bf Note} that all site types mentioned can be constructed from
2150 types 3fd (normalized, in-plane) and 3out (non-normalized, out of
2151 plane). However, the amount of computation involved increases sharply
2152 along this list, so we strongly recommended using the first adequate
2153 virtual site type that will be sufficient for a certain purpose.
2154 \figref{vsites} depicts 6 of the available virtual site constructions.
2155 The conceptually simplest construction types are linear combinations:
2156 \beq
2157 \vvis = \sum_{i=1}^N w_i \, \ve{r}_i
2158 \eeq
2159 The force is then redistributed using the same weights:
2160 \beq
2161 \Fi = w_i \, \Fvis
2162 \eeq
2164 The types of virtual sites supported in {\gromacs} are given in the list below.
2165 Constructing atoms in virtual sites can be virtual sites themselves, but
2166 only if they are higher in the list, i.e. virtual sites can be
2167 constructed from ``particles'' that are simpler virtual sites.
2168 \begin{itemize}
2169 \item[{\bf\sf 2.}]\label{subsec:vsite2}As a linear combination of two atoms
2170 (\figref{vsites} 2):
2171 \beq
2172 w_i = 1 - a ~,~~ w_j = a
2173 \eeq
2174 In this case the virtual site is on the line through atoms $i$ and
2175 $j$.
2177 \item[{\bf\sf 3.}]\label{subsec:vsite3}As a linear combination of three atoms
2178 (\figref{vsites} 3):
2179 \beq
2180 w_i = 1 - a - b ~,~~ w_j = a ~,~~ w_k = b
2181 \eeq
2182 In this case the virtual site is in the plane of the other three
2183 particles.
2185 \item[{\bf\sf 3fd.}]\label{subsec:vsite3fd}In the plane of three atoms, with a fixed distance
2186 (\figref{vsites} 3fd):
2187 \beq
2188 \vvis ~=~ \ve{r}_i + b \frac{ \rvij + a \rvjk }
2189 {| \rvij + a \rvjk |}
2190 \eeq
2191 In this case the virtual site is in the plane of the other three
2192 particles at a distance of $|b|$ from $i$.
2193 The force on particles $i$, $j$ and $k$ due to the force on the virtual
2194 site can be computed as:
2195 \beq
2196 \begin{array}{lcr}
2197 \Fi &=& \displaystyle \Fvis - \gamma ( \Fvis - \ve{p} ) \\[1ex]
2198 \Fj &=& \displaystyle (1-a)\gamma (\Fvis - \ve{p}) \\[1ex]
2199 \Fk &=& \displaystyle a \gamma (\Fvis - \ve{p}) \\
2200 \end{array}
2201 ~\mbox{~ where~ }~
2202 \begin{array}{c}
2203 \displaystyle \gamma = \frac{b}{| \rvij + a \rvjk |} \\[2ex]
2204 \displaystyle \ve{p} = \frac{ \rvis \cdot \Fvis }
2205 { \rvis \cdot \rvis } \rvis
2206 \end{array}
2207 \eeq
2209 \item[{\bf\sf 3fad.}]\label{subsec:vsite3fad}In the plane of three atoms, with a fixed angle and
2210 distance (\figref{vsites} 3fad):
2211 \beq
2212 \label{eqn:vsite2fad-F}
2213 \vvis ~=~ \ve{r}_i +
2214 d \cos \theta \frac{\rvij}{|\rvij|} +
2215 d \sin \theta \frac{\ve{r}_\perp}{|\ve{r}_\perp|}
2216 ~\mbox{~ where~ }~
2217 \ve{r}_\perp ~=~ \rvjk -
2218 \frac{ \rvij \cdot \rvjk }
2219 { \rvij \cdot \rvij }
2220 \rvij
2221 \eeq
2222 In this case the virtual site is in the plane of the other three
2223 particles at a distance of $|d|$ from $i$ at an angle of
2224 $\alpha$ with $\rvij$. Atom $k$ defines the plane and the
2225 direction of the angle. {\bf Note} that in this case $b$ and
2226 $\alpha$ must be specified, instead of $a$ and $b$ (see also
2227 \secref{vsitetop}). The force on particles $i$, $j$ and $k$
2228 due to the force on the virtual site can be computed as (with
2229 $\ve{r}_\perp$ as defined in \eqnref{vsite2fad-F}):
2230 \newcommand{\dfrac}{\displaystyle\frac}
2231 \beq
2232 \begin{array}{c}
2233 \begin{array}{lclllll}
2234 \Fi &=& \Fvis &-&
2235 \dfrac{d \cos \theta}{|\rvij|} \ve{F}_1 &+&
2236 \dfrac{d \sin \theta}{|\ve{r}_\perp|} \left(
2237 \dfrac{ \rvij \cdot \rvjk }
2238 { \rvij \cdot \rvij } \ve{F}_2 +
2239 \ve{F}_3 \right) \\[3ex]
2240 \Fj &=& &&
2241 \dfrac{d \cos \theta}{|\rvij|} \ve{F}_1 &-&
2242 \dfrac{d \sin \theta}{|\ve{r}_\perp|} \left(
2243 \ve{F}_2 +
2244 \dfrac{ \rvij \cdot \rvjk }
2245 { \rvij \cdot \rvij } \ve{F}_2 +
2246 \ve{F}_3 \right) \\[3ex]
2247 \Fk &=& && &&
2248 \dfrac{d \sin \theta}{|\ve{r}_\perp|} \ve{F}_2 \\[3ex]
2249 \end{array} \\[5ex]
2250 \mbox{where ~}
2251 \ve{F}_1 = \Fvis -
2252 \dfrac{ \rvij \cdot \Fvis }
2253 { \rvij \cdot \rvij } \rvij
2254 \mbox{\,, ~}
2255 \ve{F}_2 = \ve{F}_1 -
2256 \dfrac{ \ve{r}_\perp \cdot \Fvis }
2257 { \ve{r}_\perp \cdot \ve{r}_\perp } \ve{r}_\perp
2258 \mbox{~and ~}
2259 \ve{F}_3 = \dfrac{ \rvij \cdot \Fvis }
2260 { \rvij \cdot \rvij } \ve{r}_\perp
2261 \end{array}
2262 \eeq
2264 \item[{\bf\sf 3out.}]\label{subsec:vsite3out}As a non-linear combination of three atoms, out of plane
2265 (\figref{vsites} 3out):
2266 \beq
2267 \vvis ~=~ \ve{r}_i + a \rvij + b \rvik +
2268 c (\rvij \times \rvik)
2269 \eeq
2270 This enables the construction of virtual sites out of the plane of the
2271 other atoms.
2272 The force on particles $i,j$ and $k$ due to the force on the virtual
2273 site can be computed as:
2274 \beq
2275 \begin{array}{lcl}
2276 \vspace{4mm}
2277 \Fj &=& \left[\begin{array}{ccc}
2278 a & -c\,z_{ik} & c\,y_{ik} \\[0.5ex]
2279 c\,z_{ik} & a & -c\,x_{ik} \\[0.5ex]
2280 -c\,y_{ik} & c\,x_{ik} & a
2281 \end{array}\right]\Fvis \\
2282 \vspace{4mm}
2283 \Fk &=& \left[\begin{array}{ccc}
2284 b & c\,z_{ij} & -c\,y_{ij} \\[0.5ex]
2285 -c\,z_{ij} & b & c\,x_{ij} \\[0.5ex]
2286 c\,y_{ij} & -c\,x_{ij} & b
2287 \end{array}\right]\Fvis \\
2288 \Fi &=& \Fvis - \Fj - \Fk
2289 \end{array}
2290 \eeq
2292 \item[{\bf\sf 4fdn.}]\label{subsec:vsite4fdn}From four atoms, with a fixed distance, see separate Fig.\ \ref{fig:vsite-4fdn}.
2293 This construction is a bit
2294 complex, in particular since the previous type (4fd) could be unstable which forced us
2295 to introduce a more elaborate construction:
2297 \begin{figure}
2298 \centerline{\includegraphics[width=5cm]{plots/vsite-4fdn}}
2299 \caption {The new 4fdn virtual site construction, which is stable even when all constructing
2300 atoms are in the same plane.}
2301 \label{fig:vsite-4fdn}
2302 \end{figure}
2304 \begin{eqnarray}
2305 \mathbf{r}_{ja} &=& a\, \mathbf{r}_{ik} - \mathbf{r}_{ij} = a\, (\mathbf{x}_k - \mathbf{x}_i) - (\mathbf{x}_j - \mathbf{x}_i) \nonumber \\
2306 \mathbf{r}_{jb} &=& b\, \mathbf{r}_{il} - \mathbf{r}_{ij} = b\, (\mathbf{x}_l - \mathbf{x}_i) - (\mathbf{x}_j - \mathbf{x}_i) \nonumber \\
2307 \mathbf{r}_m &=& \mathbf{r}_{ja} \times \mathbf{r}_{jb} \nonumber \\
2308 \mathbf{x}_s &=& \mathbf{x}_i + c \frac{\mathbf{r}_m}{|\mathbf{r}_m|}
2309 \label{eq:vsite}
2310 \end{eqnarray}
2312 In this case the virtual site is at a distance of $|c|$ from $i$, while $a$ and $b$ are
2313 parameters. {\bf Note} that the vectors $\mathbf{r}_{ik}$ and $\mathbf{r}_{ij}$ are not normalized
2314 to save floating-point operations.
2315 The force on particles $i$, $j$, $k$ and $l$ due to the force
2316 on the virtual site are computed through chain rule derivatives
2317 of the construction expression. This is exact and conserves energy,
2318 but it does lead to relatively lengthy expressions that we do not
2319 include here (over 200 floating-point operations). The interested reader can
2320 look at the source code in \verb+vsite.c+. Fortunately, this vsite type is normally
2321 only used for chiral centers such as $C_{\alpha}$ atoms in proteins.
2323 The new 4fdn construct is identified with a `type' value of 2 in the topology. The earlier 4fd
2324 type is still supported internally (`type' value 1), but it should not be used for
2325 new simulations. All current {\gromacs} tools will automatically generate type 4fdn instead.
2328 \item[{\bf\sf N.}]\label{subsec:vsiteN} A linear combination of $N$ atoms with relative
2329 weights $a_i$. The weight for atom $i$ is:
2330 \beq
2331 w_i = a_i \left(\sum_{j=1}^N a_j \right)^{-1}
2332 \eeq
2333 There are three options for setting the weights:
2334 \begin{itemize}
2335 \item[COG] center of geometry: equal weights
2336 \item[COM] center of mass: $a_i$ is the mass of atom $i$;
2337 when in free-energy simulations the mass of the atom is changed,
2338 only the mass of the A-state is used for the weight
2339 \item[COW] center of weights: $a_i$ is defined by the user
2340 \end{itemize}
2342 \end{itemize}
2344 \newcommand{\dr}{{\rm d}r}
2345 \newcommand{\avcsix}{\left< C_6 \right>}
2347 \section{Long Range Electrostatics}
2348 \label{sec:lr_elstat}
2349 \subsection{Ewald summation\index{Ewald sum}}
2350 \label{sec:ewald}
2351 The total electrostatic energy of $N$ particles and their periodic
2352 images\index{periodic boundary conditions} is given by
2353 \begin{equation}
2354 V=\frac{f}{2}\sum_{n_x}\sum_{n_y}
2355 \sum_{n_{z}*} \sum_{i}^{N} \sum_{j}^{N}
2356 \frac{q_i q_j}{{\bf r}_{ij,{\bf n}}}.
2357 \label{eqn:totalcoulomb}
2358 \end{equation}
2359 $(n_x,n_y,n_z)={\bf n}$ is the box index vector, and the star indicates that
2360 terms with $i=j$ should be omitted when $(n_x,n_y,n_z)=(0,0,0)$. The
2361 distance ${\bf r}_{ij,{\bf n}}$ is the real distance between the charges and
2362 not the minimum-image. This sum is conditionally convergent, but
2363 very slow.
2365 Ewald summation was first introduced as a method to calculate
2366 long-range interactions of the periodic images in
2367 crystals~\cite{Ewald21}. The idea is to convert the single
2368 slowly-converging sum \eqnref{totalcoulomb} into two
2369 quickly-converging terms and a constant term:
2370 \begin{eqnarray}
2371 V &=& V_{\mathrm{dir}} + V_{\mathrm{rec}} + V_{0} \\[0.5ex]
2372 V_{\mathrm{dir}} &=& \frac{f}{2} \sum_{i,j}^{N}
2373 \sum_{n_x}\sum_{n_y}
2374 \sum_{n_{z}*} q_i q_j \frac{\mbox{erfc}(\beta {r}_{ij,{\bf n}} )}{{r}_{ij,{\bf n}}} \\[0.5ex]
2375 V_{\mathrm{rec}} &=& \frac{f}{2 \pi V} \sum_{i,j}^{N} q_i q_j
2376 \sum_{m_x}\sum_{m_y}
2377 \sum_{m_{z}*} \frac{\exp{\left( -(\pi {\bf m}/\beta)^2 + 2 \pi i
2378 {\bf m} \cdot ({\bf r}_i - {\bf r}_j)\right)}}{{\bf m}^2} \\[0.5ex]
2379 V_{0} &=& -\frac{f \beta}{\sqrt{\pi}}\sum_{i}^{N} q_i^2,
2380 \end{eqnarray}
2381 where $\beta$ is a parameter that determines the relative weight of the
2382 direct and reciprocal sums and ${\bf m}=(m_x,m_y,m_z)$.
2383 In this way we can use a short cut-off (of the order of $1$~nm) in the direct space sum and a
2384 short cut-off in the reciprocal space sum ({\eg} 10 wave vectors in each
2385 direction). Unfortunately, the computational cost of the reciprocal
2386 part of the sum increases as $N^2$
2387 (or $N^{3/2}$ with a slightly better algorithm) and it is therefore not
2388 realistic for use in large systems.
2390 \subsubsection{Using Ewald}
2391 Don't use Ewald unless you are absolutely sure this is what you want -
2392 for almost all cases the PME method below will perform much better.
2393 If you still want to employ classical Ewald summation enter this in
2394 your {\tt .mdp} file, if the side of your box is about $3$~nm:
2396 \begin{verbatim}
2397 coulombtype = Ewald
2398 rvdw = 0.9
2399 rlist = 0.9
2400 rcoulomb = 0.9
2401 fourierspacing = 0.6
2402 ewald-rtol = 1e-5
2403 \end{verbatim}
2405 The ratio of the box dimensions and the {\tt fourierspacing} parameter determines
2406 the highest magnitude of wave vectors $m_x,m_y,m_z$ to use in each
2407 direction. With a 3-nm cubic box this example would use $11$ wave vectors
2408 (from $-5$ to $5$) in each direction. The {\tt ewald-rtol} parameter
2409 is the relative strength of the electrostatic interaction at the
2410 cut-off. Decreasing this gives you a more accurate direct sum, but a
2411 less accurate reciprocal sum.
2413 \subsection{\normindex{PME}}
2414 \label{sec:pme}
2415 Particle-mesh Ewald is a method proposed by Tom
2416 Darden~\cite{Darden93} to improve the performance of the
2417 reciprocal sum. Instead of directly summing wave vectors, the charges
2418 are assigned to a grid using interpolation. The implementation in
2419 {\gromacs} uses cardinal B-spline interpolation~\cite{Essmann95},
2420 which is referred to as smooth PME (SPME).
2421 The grid is then Fourier transformed with a 3D FFT algorithm and the
2422 reciprocal energy term obtained by a single sum over the grid in
2423 k-space.
2425 The potential at the grid points is calculated by inverse
2426 transformation, and by using the interpolation factors we get the
2427 forces on each atom.
2429 The PME algorithm scales as $N \log(N)$, and is substantially faster
2430 than ordinary Ewald summation on medium to large systems. On very
2431 small systems it might still be better to use Ewald to avoid the
2432 overhead in setting up grids and transforms.
2433 For the parallelization of PME see the section on MPMD PME (\ssecref{mpmd_pme}).
2435 With the Verlet cut-off scheme, the PME direct space potential is
2436 shifted by a constant such that the potential is zero at the
2437 cut-off. This shift is small and since the net system charge is close
2438 to zero, the total shift is very small, unlike in the case of the
2439 Lennard-Jones potential where all shifts add up. We apply the shift
2440 anyhow, such that the potential is the exact integral of the force.
2442 \subsubsection{Using PME}
2443 As an example for using Particle-mesh Ewald summation in {\gromacs}, specify the
2444 following lines in your {\tt .mdp} file:
2446 \begin{verbatim}
2447 coulombtype = PME
2448 rvdw = 0.9
2449 rlist = 0.9
2450 rcoulomb = 0.9
2451 fourierspacing = 0.12
2452 pme-order = 4
2453 ewald-rtol = 1e-5
2454 \end{verbatim}
2456 In this case the {\tt fourierspacing} parameter determines the maximum
2457 spacing for the FFT grid (i.e. minimum number of grid points),
2458 and {\tt pme-order} controls the
2459 interpolation order. Using fourth-order (cubic) interpolation and this
2460 spacing should give electrostatic energies accurate to about
2461 $5\cdot10^{-3}$. Since the Lennard-Jones energies are not this
2462 accurate it might even be possible to increase this spacing slightly.
2464 Pressure scaling works with PME, but be aware of the fact that
2465 anisotropic scaling can introduce artificial ordering in some systems.
2467 \subsection{\normindex{P3M-AD}}
2468 \label{sec:pppm}
2469 The \seeindex{Particle-Particle Particle-Mesh}{P3M} methods of
2470 Hockney \& Eastwood can also be applied in {\gromacs} for the
2471 treatment of long range electrostatic interactions~\cite{Hockney81}.
2472 Although the P3M method was the first efficient long-range electrostatics
2473 method for molecular simulation, the smooth PME (SPME) method has largely
2474 replaced P3M as the method of choice in atomistic simulations. One performance
2475 disadvantage of the original P3M method was that it required 3 3D-FFT
2476 back transforms to obtain the forces on the particles. But this is not
2477 required for P3M and the forces can be derived through analytical differentiation
2478 of the potential, as done in PME. The resulting method is termed P3M-AD.
2479 The only remaining difference between P3M-AD and PME is the optimization
2480 of the lattice Green influence function for error minimization that P3M uses.
2481 However, in 2012 it has been shown that the SPME influence function can be
2482 modified to obtain P3M~\cite{Ballenegger2012}.
2483 This means that the advantage of error minimization in P3M-AD can be used
2484 at the same computational cost and with the same code as PME,
2485 just by adding a few lines to modify the influence function.
2486 However, at optimal parameter setting the effect of error minimization
2487 in P3M-AD is less than 10\%. P3M-AD does show large accuracy gains with
2488 interlaced (also known as staggered) grids, but that is not supported
2489 in {\gromacs} (yet).
2491 P3M is used in {\gromacs} with exactly the same options as used with PME
2492 by selecting the electrostatics type:
2493 \begin{verbatim}
2494 coulombtype = P3M-AD
2495 \end{verbatim}
2497 \subsection{Optimizing Fourier transforms and PME calculations}
2498 It is recommended to optimize the parameters for calculation of
2499 electrostatic interaction such as PME grid dimensions and cut-off radii.
2500 This is particularly relevant to do before launching long production runs.
2502 {\tt gmx mdrun} will automatically do a lot of PME optimization, and
2503 {\gromacs} also includes a special tool, {\tt gmx tune_pme}, which
2504 automates the process of selecting the optimal number of PME-only ranks.
2506 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2507 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2508 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2510 \section{Long Range Van der Waals interactions}
2511 \subsection{Dispersion correction\index{dispersion correction}}
2512 In this section, we derive long-range corrections due to the use of a
2513 cut-off for Lennard-Jones or Buckingham interactions.
2514 We assume that the cut-off is
2515 so long that the repulsion term can safely be neglected, and therefore
2516 only the dispersion term is taken into account. Due to the nature of
2517 the dispersion interaction (we are truncating a potential proportional
2518 to $-r^{-6}$), energy and pressure corrections are both negative. While
2519 the energy correction is usually small, it may be important for free
2520 energy calculations where differences between two different Hamiltonians
2521 are considered. In contrast, the pressure correction is very large and
2522 can not be neglected under any circumstances where a correct pressure is
2523 required, especially for any NPT simulations. Although it is, in
2524 principle, possible to parameterize a force field such that the pressure
2525 is close to the desired experimental value without correction, such a
2526 method makes the parameterization dependent on the cut-off and is therefore
2527 undesirable.
2529 \subsubsection{Energy}
2530 \label{sec:ecorr}
2531 The long-range contribution of the dispersion interaction to the
2532 virial can be derived analytically, if we assume a homogeneous
2533 system beyond the cut-off distance $r_c$. The dispersion energy
2534 between two particles is written as:
2535 \beq
2536 V(\rij) ~=~- C_6\,\rij^{-6}
2537 \eeq
2538 and the corresponding force is:
2539 \beq
2540 \Fvij ~=~- 6\,C_6\,\rij^{-8}\rvij
2541 \eeq
2542 In a periodic system it is not easy to calculate the full potentials,
2543 so usually a cut-off is applied, which can be abrupt or smooth.
2544 We will call the potential and force with cut-off $V_c$ and $\ve{F}_c$.
2545 The long-range contribution to the dispersion energy
2546 in a system with $N$ particles and particle density $\rho$ = $N/V$ is:
2547 \beq
2548 \label{eqn:enercorr}
2549 V_{lr} ~=~ \half N \rho\int_0^{\infty} 4\pi r^2 g(r) \left( V(r) -V_c(r) \right) {\dr}
2550 \eeq
2551 We will integrate this for the shift function, which is the most general
2552 form of van der Waals interaction available in {\gromacs}.
2553 The shift function has a constant difference $S$ from 0 to $r_1$
2554 and is 0 beyond the cut-off distance $r_c$.
2555 We can integrate \eqnref{enercorr}, assuming that the density in the sphere
2556 within $r_1$ is equal to the global density and
2557 the radial distribution function $g(r)$ is 1 beyond $r_1$:
2558 \bea
2559 \nonumber
2560 V_{lr} &=& \half N \left(
2561 \rho\int_0^{r_1} 4\pi r^2 g(r) \, C_6 \,S\,{\dr}
2562 + \rho\int_{r_1}^{r_c} 4\pi r^2 \left( V(r) -V_c(r) \right) {\dr}
2563 + \rho\int_{r_c}^{\infty} 4\pi r^2 V(r) \, {\dr}
2564 \right) \\
2565 & = & \half N \left(\left(\frac{4}{3}\pi \rho r_1^{3} - 1\right) C_6 \,S
2566 + \rho\int_{r_1}^{r_c} 4\pi r^2 \left( V(r) -V_c(r) \right) {\dr}
2567 -\frac{4}{3} \pi N \rho\, C_6\,r_c^{-3}
2568 \right)
2569 \eea
2570 where the term $-1$ corrects for the self-interaction.
2571 For a plain cut-off we only need to assume that $g(r)$ is 1 beyond $r_c$
2572 and the correction reduces to~\cite{Allen87}:
2573 \bea
2574 V_{lr} & = & -\frac{2}{3} \pi N \rho\, C_6\,r_c^{-3}
2575 \eea
2576 If we consider, for example, a box of pure water, simulated with a cut-off
2577 of 0.9 nm and a density of 1 g cm$^{-3}$ this correction is
2578 $-0.75$ kJ mol$^{-1}$ per molecule.
2580 For a homogeneous mixture we need to define
2581 an {\em average dispersion constant}:
2582 \beq
2583 \label{eqn:avcsix}
2584 \avcsix = \frac{2}{N(N-1)}\sum_i^N\sum_{j>i}^N C_6(i,j)\\
2585 \eeq
2586 In {\gromacs}, excluded pairs of atoms do not contribute to the average.
2588 In the case of inhomogeneous simulation systems, {\eg} a system with a
2589 lipid interface, the energy correction can be applied if
2590 $\avcsix$ for both components is comparable.
2592 \subsubsection{Virial and pressure}
2593 The scalar virial of the system due to the dispersion interaction between
2594 two particles $i$ and $j$ is given by:
2595 \beq
2596 \Xi~=~-\half \rvij \cdot \Fvij ~=~ 3\,C_6\,\rij^{-6}
2597 \eeq
2598 The pressure is given by:
2599 \beq
2600 P~=~\frac{2}{3\,V}\left(E_{kin} - \Xi\right)
2601 \eeq
2602 The long-range correction to the virial is given by:
2603 \beq
2604 \Xi_{lr} ~=~ \half N \rho \int_0^{\infty} 4\pi r^2 g(r) (\Xi -\Xi_c) \,\dr
2605 \eeq
2606 We can again integrate the long-range contribution to the
2607 virial assuming $g(r)$ is 1 beyond $r_1$:
2608 \bea
2609 \Xi_{lr}&=& \half N \rho \left(
2610 \int_{r_1}^{r_c} 4 \pi r^2 (\Xi -\Xi_c) \,\dr
2611 + \int_{r_c}^{\infty} 4 \pi r^2 3\,C_6\,\rij^{-6}\, \dr
2612 \right) \nonumber\\
2613 &=& \half N \rho \left(
2614 \int_{r_1}^{r_c} 4 \pi r^2 (\Xi -\Xi_c) \, \dr
2615 + 4 \pi C_6 \, r_c^{-3} \right)
2616 \eea
2617 For a plain cut-off the correction to the pressure is~\cite{Allen87}:
2618 \beq
2619 P_{lr}~=~-\frac{4}{3} \pi C_6\, \rho^2 r_c^{-3}
2620 \eeq
2621 Using the same example of a water box, the correction to the virial is
2622 0.75 kJ mol$^{-1}$ per molecule,
2623 the corresponding correction to the pressure for
2624 SPC water is approximately $-280$ bar.
2626 For homogeneous mixtures, we can again use the average dispersion constant
2627 $\avcsix$ (\eqnref{avcsix}):
2628 \beq
2629 P_{lr}~=~-\frac{4}{3} \pi \avcsix \rho^2 r_c^{-3}
2630 \label{eqn:pcorr}
2631 \eeq
2632 For inhomogeneous systems, \eqnref{pcorr} can be applied under the same
2633 restriction as holds for the energy (see \secref{ecorr}).
2635 \subsection{Lennard-Jones PME\index{LJ-PME}}
2637 In order to treat systems, using Lennard-Jones potentials, that are
2638 non-homogeneous outside of the cut-off distance, we can instead use
2639 the Particle-mesh Ewald method as discussed for electrostatics above.
2640 In this case the modified Ewald equations become
2641 \begin{eqnarray}
2642 V &=& V_{\mathrm{dir}} + V_{\mathrm{rec}} + V_{0} \\[0.5ex]
2643 V_{\mathrm{dir}} &=& -\frac{1}{2} \sum_{i,j}^{N}
2644 \sum_{n_x}\sum_{n_y}
2645 \sum_{n_{z}*} \frac{C^{ij}_6 g(\beta {r}_{ij,{\bf n}})}{{r_{ij,{\bf n}}}^6}
2646 \label{eqn:ljpmerealspace}\\[0.5ex]
2647 V_{\mathrm{rec}} &=& \frac{{\pi}^{\frac{3}{2}} \beta^{3}}{2V} \sum_{m_x}\sum_{m_y}\sum_{m_{z}*}
2648 f(\pi |{\mathbf m}|/\beta) \times \sum_{i,j}^{N} C^{ij}_6 {\mathrm{exp}}\left[-2\pi i {\bf m}\cdot({\bf r_i}-{\bf r_j})\right] \\[0.5ex]
2649 V_{0} &=& -\frac{\beta^{6}}{12}\sum_{i}^{N} C^{ii}_6
2650 \end{eqnarray}
2652 where ${\bf m}=(m_x,m_y,m_z)$, $\beta$ is the parameter determining the weight between
2653 direct and reciprocal space, and ${C^{ij}_6}$ is the combined dispersion
2654 parameter for particle $i$ and $j$. The star indicates that terms
2655 with $i = j$ should be omitted when $((n_x,n_y,n_z)=(0,0,0))$, and
2656 ${\bf r}_{ij,{\bf n}}$ is the real distance between the particles.
2657 Following the derivation by Essmann~\cite{Essmann95}, the functions $f$ and $g$ introduced above are defined as
2658 \begin{eqnarray}
2659 f(x)&=&1/3\left[(1-2x^2){\mathrm{exp}}(-x^2) + 2{x^3}\sqrt{\pi}\,{\mathrm{erfc}}(x) \right] \\
2660 g(x)&=&{\mathrm{exp}}(-x^2)(1+x^2+\frac{x^4}{2}).
2661 \end{eqnarray}
2663 The above methodology works fine as long as the dispersion parameters can be combined geometrically (\eqnref{comb}) in the same
2664 way as the charges for electrostatics
2665 \begin{equation}
2666 C^{ij}_{6,\mathrm{geom}} = \left(C^{ii}_6 \, C^{jj}_6\right)^{1/2}
2667 \end{equation}
2668 For Lorentz-Berthelot combination rules (\eqnref{lorentzberthelot}), the reciprocal part of this sum has to be calculated
2669 seven times due to the splitting of the dispersion parameter according to
2670 \begin{equation}
2671 C^{ij}_{6,\mathrm{L-B}} = (\sigma_i+\sigma_j)^6=\sum_{n=0}^{6} P_{n}\sigma_{i}^{n}\sigma_{j}^{(6-n)},
2672 \end{equation}
2673 for $P_{n}$ the Pascal triangle coefficients. This introduces a
2674 non-negligible cost to the reciprocal part, requiring seven separate
2675 FFTs, and therefore this has been the limiting factor in previous
2676 attempts to implement LJ-PME. A solution to this problem is to use
2677 geometrical combination rules in order to calculate an approximate
2678 interaction parameter for the reciprocal part of the potential,
2679 yielding a total interaction of
2680 \begin{eqnarray}
2681 V(r<r_c) & = & \underbrace{C^{\mathrm{dir}}_6 g(\beta r) r^{-6}}_{\mathrm{Direct \ space}} + \underbrace{C^\mathrm{recip}_{6,\mathrm{geom}} [1 - g(\beta r)] r^{-6}}_{\mathrm{Reciprocal \ space}} \nonumber \\
2682 &=& C^\mathrm{recip}_{6,\mathrm{geom}}r^{-6} + \left(C^{\mathrm{dir}}_6-C^\mathrm{recip}_{6,\mathrm{geom}}\right)g(\beta r)r^{-6} \\
2683 V(r>r_c) & = & \underbrace{C^\mathrm{recip}_{6,\mathrm{geom}} [1 - g(\beta r)] r^{-6}}_{\mathrm{Reciprocal \ space}}.
2684 \end{eqnarray}
2685 This will preserve a well-defined Hamiltonian and significantly increase
2686 the performance of the simulations. The approximation does introduce
2687 some errors, but since the difference is located in the interactions
2688 calculated in reciprocal space, the effect will be very small compared
2689 to the total interaction energy. In a simulation of a lipid bilayer,
2690 using a cut-off of 1.0 nm, the relative error in total dispersion
2691 energy was below 0.5\%. A more thorough discussion of this can be
2692 found in \cite{Wennberg13}.
2694 In {\gromacs} we now perform the proper calculation of this interaction
2695 by subtracting, from the direct-space interactions, the contribution
2696 made by the approximate potential that is used in the reciprocal part
2697 \begin{equation}
2698 V_\mathrm{dir} = C^{\mathrm{dir}}_6 r^{-6} - C^\mathrm{recip}_6 [1 - g(\beta r)] r^{-6}.
2699 \label{eqn:ljpmedirectspace}
2700 \end{equation}
2701 This potential will reduce to the expression in \eqnref{ljpmerealspace} when $C^{\mathrm{dir}}_6 = C^\mathrm{recip}_6$,
2702 and the total interaction is given by
2703 \begin{eqnarray}
2704 \nonumber V(r<r_c) &=& \underbrace{C^{\mathrm{dir}}_6 r^{-6} - C^\mathrm{recip}_6 [1 - g(\beta r)] r^{-6}}_{\mathrm{Direct \ space}} + \underbrace{C^\mathrm{recip}_6 [1 - g(\beta r)] r^{-6}}_{\mathrm{Reciprocal \ space}} \\
2705 &=&C^{\mathrm{dir}}_6 r^{-6}
2706 \label {eqn:ljpmecorr2} \\
2707 V(r>r_c) &=& C^\mathrm{recip}_6 [1 - g(\beta r)] r^{-6}.
2708 \end{eqnarray}
2709 For the case when $C^{\mathrm{dir}}_6 \neq C^\mathrm{recip}_6$ this
2710 will retain an unmodified LJ force up to the cut-off, and the error
2711 is an order of magnitude smaller than in simulations where the
2712 direct-space interactions do not account for the approximation used in
2713 reciprocal space. When using a VdW interaction modifier of
2714 potential-shift, the constant
2715 \begin{equation}
2716 \left(-C^{\mathrm{dir}}_6 + C^\mathrm{recip}_6 [1 - g(\beta r_c)]\right) r_c^{-6}
2717 \end{equation}
2718 is added to \eqnref{ljpmecorr2} in order to ensure that the potential
2719 is continuous at the cutoff. Note that, in the same way as \eqnref{ljpmedirectspace}, this degenerates into the
2720 expected $-C_6g(\beta r_c)r^{-6}_c$ when $C^{\mathrm{dir}}_6 =
2721 C^\mathrm{recip}_6$. In addition to this, a long-range dispersion
2722 correction can be applied to correct for the approximation using a
2723 combination rule in reciprocal space. This correction assumes, as for
2724 the cut-off LJ potential, a uniform particle distribution. But since
2725 the error of the combination rule approximation is very small this
2726 long-range correction is not necessary in most cases. Also note that
2727 this homogenous correction does not correct the surface tension, which
2728 is an inhomogeneous property.
2730 \subsubsection{Using LJ-PME}
2731 As an example for using Particle-mesh Ewald summation for Lennard-Jones interactions in {\gromacs}, specify the
2732 following lines in your {\tt .mdp} file:
2733 \begin{verbatim}
2734 vdwtype = PME
2735 rvdw = 0.9
2736 vdw-modifier = Potential-Shift
2737 rlist = 0.9
2738 rcoulomb = 0.9
2739 fourierspacing = 0.12
2740 pme-order = 4
2741 ewald-rtol-lj = 0.001
2742 lj-pme-comb-rule = geometric
2743 \end{verbatim}
2745 The same Fourier grid and interpolation order are used if both
2746 LJ-PME and electrostatic PME are active, so the settings for
2747 {\tt fourierspacing} and {\tt pme-order} are common to both.
2748 {\tt ewald-rtol-lj} controls the
2749 splitting between direct and reciprocal space in the same way as
2750 {\tt ewald-rtol}. In addition to this, the combination rule to be used
2751 in reciprocal space is determined by {\tt lj-pme-comb-rule}. If the
2752 current force field uses Lorentz-Berthelot combination rules, it is
2753 possible to set {\tt lj-pme-comb-rule = geometric} in order to gain a
2754 significant increase in performance for a small loss in accuracy. The
2755 details of this approximation can be found in the section above.
2757 Note that the use of a complete long-range dispersion correction means
2758 that as with Coulomb PME, {\tt rvdw} is now a free parameter in the
2759 method, rather than being necessarily restricted by the force-field
2760 parameterization scheme. Thus it is now possible to optimize the
2761 cutoff, spacing, order and tolerance terms for accuracy and best
2762 performance.
2764 Naturally, the use of LJ-PME rather than LJ cut-off adds computation
2765 and communication done for the reciprocal-space part, so for best
2766 performance in balancing the load of parallel simulations using
2767 PME-only ranks, more such ranks should be used. It may be possible to
2768 improve upon the automatic load-balancing used by {\tt mdrun}.
2771 \section{Force field\index{force field}}
2772 \label{sec:ff}
2773 A force field is built up from two distinct components:
2774 \begin{itemize}
2775 \item The set of equations (called the {\em
2776 potential functions}\index{potential function}) used to generate the potential
2777 energies and their derivatives, the forces. These are described in
2778 detail in the previous chapter.
2779 \item The parameters used in this set of equations. These are not
2780 given in this manual, but in the data files corresponding to your
2781 {\gromacs} distribution.
2782 \end{itemize}
2783 Within one set of equations various sets of parameters can be
2784 used. Care must be taken that the combination of equations and
2785 parameters form a consistent set. It is in general dangerous to make
2786 {\em ad hoc} changes in a subset of parameters, because the various
2787 contributions to the total force are usually interdependent. This
2788 means in principle that every change should be documented, verified by
2789 comparison to experimental data and published in a peer-reviewed
2790 journal before it can be used.
2792 {\gromacs} {\gmxver} includes several force fields, and additional
2793 ones are available on the website. If you do not know which one to
2794 select we recommend \gromosv{96} for united-atom setups and OPLS-AA/L for
2795 all-atom parameters. That said, we describe the available options in
2796 some detail.
2798 \subsubsection{All-hydrogen force field}
2799 The \gromosv{87}-based all-hydrogen force field is almost identical to the
2800 normal \gromosv{87} force field, since the extra hydrogens have no
2801 Lennard-Jones interaction and zero charge. The only differences are in
2802 the bond angle and improper dihedral angle terms. This force field is
2803 only useful when you need the exact hydrogen positions, for instance
2804 for distance restraints derived from NMR measurements. When citing
2805 this force field please read the previous paragraph.
2807 \subsection{\gromosv{96}\index{GROMOS96 force field}}
2808 {\gromacs} supports the \gromosv{96} force fields~\cite{gromos96}.
2809 All parameters for the 43A1, 43A2 (development, improved alkane
2810 dihedrals), 45A3, 53A5, and 53A6 parameter sets are included. All standard
2811 building blocks are included and topologies can be built automatically
2812 by {\tt pdb2gmx}.
2814 The \gromosv{96} force field is a further development of the \gromosv{87} force field.
2815 It has improvements over the \gromosv{87} force field for proteins and small molecules.
2816 {\bf Note} that the sugar parameters present in 53A6 do correspond to those published in
2817 2004\cite{Oostenbrink2004}, which are different from those present in 45A4, which
2818 is not included in {\gromacs} at this time. The 45A4 parameter set corresponds to a later
2819 revision of these parameters.
2820 The \gromosv{96} force field is not, however, recommended for use with long alkanes and
2821 lipids. The \gromosv{96} force field differs from the \gromosv{87}
2822 force field in a few respects:
2823 \begin{itemize}
2824 \item the force field parameters
2825 \item the parameters for the bonded interactions are not linked to atom types
2826 \item a fourth power bond stretching potential (\ssecref{G96bond})
2827 \item an angle potential based on the cosine of the angle (\ssecref{G96angle})
2828 \end{itemize}
2829 There are two differences in implementation between {\gromacs} and \gromosv{96}
2830 which can lead to slightly different results when simulating the same system
2831 with both packages:
2832 \begin{itemize}
2833 \item in \gromosv{96} neighbor searching for solvents is performed on the
2834 first atom of the solvent molecule. This is not implemented in {\gromacs},
2835 but the difference with searching by centers of charge groups is very small
2836 \item the virial in \gromosv{96} is molecule-based. This is not implemented in
2837 {\gromacs}, which uses atomic virials
2838 \end{itemize}
2839 The \gromosv{96} force field was parameterized with a Lennard-Jones cut-off
2840 of 1.4 nm, so be sure to use a Lennard-Jones cut-off ({\tt rvdw}) of at least 1.4.
2841 A larger cut-off is possible because the Lennard-Jones potential and forces
2842 are almost zero beyond 1.4 nm.
2844 \subsubsection{\gromosv{96} files\swapindexquiet{GROMOS96}{files}}
2845 {\gromacs} can read and write \gromosv{96} coordinate and trajectory files.
2846 These files should have the extension {\tt .g96}.
2847 Such a file can be a \gromosv{96} initial/final
2848 configuration file, a coordinate trajectory file, or a combination of both.
2849 The file is fixed format; all floats are written as 15.9, and as such, files can get huge.
2850 {\gromacs} supports the following data blocks in the given order:
2851 \begin{itemize}
2852 \item Header block:
2853 \begin{verbatim}
2854 TITLE (mandatory)
2855 \end{verbatim}
2857 \item Frame blocks:
2858 \begin{verbatim}
2859 TIMESTEP (optional)
2860 POSITION/POSITIONRED (mandatory)
2861 VELOCITY/VELOCITYRED (optional)
2862 BOX (optional)
2863 \end{verbatim}
2865 \end{itemize}
2866 See the \gromosv{96} manual~\cite{gromos96} for a complete description
2867 of the blocks. {\bf Note} that all {\gromacs} programs can read compressed
2868 (.Z) or gzipped (.gz) files.
2870 \subsection{OPLS/AA\index{OPLS/AA force field}}
2872 \subsection{AMBER\index{AMBER force field}}
2874 {\gromacs} provides native support for the following AMBER force fields:
2876 \begin{itemize}
2877 \item AMBER94~\cite{Cornell1995}
2878 \item AMBER96~\cite{Kollman1996}
2879 \item AMBER99~\cite{Wang2000}
2880 \item AMBER99SB~\cite{Hornak2006}
2881 \item AMBER99SB-ILDN~\cite{Lindorff2010}
2882 \item AMBER03~\cite{Duan2003}
2883 \item AMBERGS~\cite{Garcia2002}
2884 \end{itemize}
2886 \subsection{CHARMM\index{CHARMM force field}}
2887 \label{subsec:charmmff}
2889 {\gromacs} supports the CHARMM force field for proteins~\cite{mackerell04, mackerell98}, lipids~\cite{feller00} and nucleic acids~\cite{foloppe00,Mac2000}. The protein parameters (and to some extent the lipid and nucleic acid parameters) were thoroughly tested -- both by comparing potential energies between the port and the standard parameter set in the CHARMM molecular simulation package, as well by how the protein force field behaves together with {\gromacs}-specific techniques such as virtual sites (enabling long time steps) and a fast implicit solvent recently implemented~\cite{Larsson10} -- and the details and results are presented in the paper by Bjelkmar et al.~\cite{Bjelkmar10}. The nucleic acid parameters, as well as the ones for HEME, were converted and tested by Michel Cuendet.
2891 When selecting the CHARMM force field in {\tt \normindex{pdb2gmx}} the default option is to use \normindex{CMAP} (for torsional correction map). To exclude CMAP, use {\tt -nocmap}. The basic form of the CMAP term implemented in {\gromacs} is a function of the $\phi$ and $\psi$ backbone torsion angles. This term is defined in the {\tt .rtp} file by a {\tt [ cmap ]} statement at the end of each residue supporting CMAP. The following five atom names define the two torsional angles. Atoms 1-4 define $\phi$, and atoms 2-5 define $\psi$. The corresponding atom types are then matched to the correct CMAP type in the {\tt cmap.itp} file that contains the correction maps.
2893 A port of the CHARMM36 force field for use with GROMACS is also available at \url{http://mackerell.umaryland.edu/charmm_ff.shtml#gromacs}.
2895 For branched polymers or other topologies not supported by {\tt \normindex{pdb2gmx}}, it is possible to use TopoTools~\cite{kohlmeyer2016} to generate a {\gromacs} top file.
2897 \subsection{Coarse-grained force fields}
2898 \index{force-field, coarse-grained}
2899 \label{sec:cg-forcefields}
2900 Coarse-graining is a systematic way of reducing the number of degrees of freedom representing a system of interest. To achieve this, typically whole groups of atoms are represented by single beads and the coarse-grained force fields describes their effective interactions. Depending on the choice of parameterization, the functional form of such an interaction can be complicated and often tabulated potentials are used.
2902 Coarse-grained models are designed to reproduce certain properties of a reference system. This can be either a full atomistic model or even experimental data. Depending on the properties to reproduce there are different methods to derive such force fields. An incomplete list of methods is given below:
2903 \begin{itemize}
2904 \item Conserving free energies
2905 \begin{itemize}
2906 \item Simplex method
2907 \item MARTINI force field (see next section)
2908 \end{itemize}
2909 \item Conserving distributions (like the radial distribution function), so-called structure-based coarse-graining
2910 \begin{itemize}
2911 \item (iterative) Boltzmann inversion
2912 \item Inverse Monte Carlo
2913 \end{itemize}
2914 \item Conversing forces
2915 \begin{itemize}
2916 \item Force matching
2917 \end{itemize}
2918 \end{itemize}
2920 Note that coarse-grained potentials are state dependent (e.g. temperature, density,...) and should be re-parametrized depending on the system of interest and the simulation conditions. This can for example be done using the \normindex{Versatile Object-oriented Toolkit for Coarse-Graining Applications (VOTCA)}~\cite{ruehle2009}. The package was designed to assists in systematic coarse-graining, provides implementations for most of the algorithms mentioned above and has a well tested interface to {\gromacs}. It is available as open source and further information can be found at \href{http://www.votca.org}{www.votca.org}.
2922 \subsection{MARTINI\index{Martini force field}}
2924 The MARTINI force field is a coarse-grain parameter set that allows for the construction
2925 of many systems, including proteins and membranes.
2927 \subsection{PLUM\index{PLUM force field}}
2929 The \normindex{PLUM force field}~\cite{bereau12} is an example of a solvent-free
2930 protein-membrane model for which the membrane was derived from structure-based
2931 coarse-graining~\cite{wang_jpcb10}. A {\gromacs} implementation can be found at
2932 \href{http://code.google.com/p/plumx/}{code.google.com/p/plumx}.
2934 % LocalWords: dihedrals centro ij dV dr LJ lj rcl jj Bertelot OPLS bh bham rf
2935 % LocalWords: coul defunits grompp crf vcrf fcrf Tironi Debye kgrf cgrf krf dx
2936 % LocalWords: PPPM der Waals erfc lr elstat chirality bstretch bondpot kT kJ
2937 % LocalWords: anharmonic morse mol betaij expminx SPC timestep fs FENE ijk kj
2938 % LocalWords: anglepot CHARMm UB ik rr substituents ijkl Ryckaert Bellemans rb
2939 % LocalWords: alkanes pdb gmx IUPAC IUB jkl cis rbdih crb kcal cubicspline xvg
2940 % LocalWords: topfile mdrun posres ar dihr lcllll NMR nmr lcllllll NOEs lclll
2941 % LocalWords: rav preprocessor ccccccccc ai aj fac disre mdp multi topol tpr
2942 % LocalWords: fc ravdisre nstdisreout dipolar lll ccc orientational MSD const
2943 % LocalWords: orire fitgrp nstorireout Drude intra Noskov et al fecalc coulrf
2944 % LocalWords: polarizabilities parameterized sigeps Ek sc softcore GROMOS NBF
2945 % LocalWords: hydrogens alkane vdwtype coulombtype rlist rcoulomb rvdw
2946 % LocalWords: nstlist virial funcparm VdW jk jl fvsite fd vsites lcr vsitetop
2947 % LocalWords: vsite lclllll lcl parameterize parameterization enercorr avcsix
2948 % LocalWords: pcorr ecorr totalcoulomb dir fourierspacing ewald rtol Darden gz
2949 % LocalWords: FFT parallelization MPMD mpmd pme fft hoc Gromos gromos oxygens
2950 % LocalWords: virials POSITIONRED VELOCITYRED gzipped Charmm Larsson Bjelkmar
2951 % LocalWords: Cuendet CMAP nocmap dihedral Lennard covalent Verlet
2952 % LocalWords: Berthelot nm flexwat ferguson itp harmonicangle versa
2953 % LocalWords: harmonicbond atomtypes dihedraltypes equilibrated fdn
2954 % LocalWords: distancerestraint LINCS Coulombic ja jb il SPME ILDN
2955 % LocalWords: Hamiltonians atomtype AMBERGS rtp cmap graining VOTCA