Update for FFTW 3.3.5
[gromacs.git] / docs / manual / special.tex
blobfd5b51bbb2d0cc2ea2684d9e2b4b57be48c8d928
2 % This file is part of the GROMACS molecular simulation package.
4 % Copyright (c) 2013,2014,2015,2016, 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{Special Topics}
36 \label{ch:special}
39 \section{Free energy implementation}
40 \label{sec:dgimplement}
41 For free energy calculations, there are two things that must be
42 specified; the end states, and the pathway connecting the end states.
43 The end states can be specified in two ways. The most straightforward
44 is through the specification of end states in the topology file. Most
45 potential forms support both an $A$ state and a $B$ state. Whenever both
46 states are specified, then the $A$ state corresponds to the initial free
47 energy state, and the $B$ state corresponds to the final state.
49 In some cases, the end state can also be defined in some cases without
50 altering the topology, solely through the {\tt .mdp} file, through the use
51 of the {\tt couple-moltype},{\tt couple-lambda0}, {\tt couple-lambda1}, and
52 {\tt couple-intramol} mdp keywords. Any molecule type selected in
53 {\tt couple-moltype} will automatically have a $B$ state implicitly
54 constructed (and the $A$ state redefined) according to the {\tt couple-lambda}
55 keywords. {\tt couple-lambda0} and {\tt couple-lambda1} define the non-bonded
56 parameters that are present in the $A$ state ({\tt couple-lambda0})
57 and the $B$ state ({\tt couple-lambda1}). The choices are 'q','vdw', and
58 'vdw-q'; these indicate the Coulombic, van der Waals, or both parameters
59 that are turned on in the respective state.
61 Once the end states are defined, then the path between the end states
62 has to be defined. This path is defined solely in the .mdp file.
63 Starting in 4.6, $\lambda$ is a vector of components, with Coulombic,
64 van der Waals, bonded, restraint, and mass components all able to be
65 adjusted independently. This makes it possible to turn off the
66 Coulombic term linearly, and then the van der Waals using soft core,
67 all in the same simulation. This is especially useful for replica
68 exchange or expanded ensemble simulations, where it is important to
69 sample all the way from interacting to non-interacting states in the
70 same simulation to improve sampling.
72 {\tt fep-lambdas} is the default array of $\lambda$ values ranging
73 from 0 to 1. All of the other lambda arrays use the values in this
74 array if they are not specified. The previous behavior, where the
75 pathway is controlled by a single $\lambda$ variable, can be preserved
76 by using only {\tt fep-lambdas} to define the pathway.
78 For example, if you wanted to first to change the Coulombic terms,
79 then the van der Waals terms, changing bonded at the same time rate as
80 the van der Waals, but changing the restraints throughout the first
81 two-thirds of the simulation, then you could use this $\lambda$ vector:
83 \begin{verbatim}
84 coul-lambdas = 0.0 0.2 0.5 1.0 1.0 1.0 1.0 1.0 1.0 1.0
85 vdw-lambdas = 0.0 0.0 0.0 0.0 0.4 0.5 0.6 0.7 0.8 1.0
86 bonded-lambdas = 0.0 0.0 0.0 0.0 0.4 0.5 0.6 0.7 0.8 1.0
87 restraint-lambdas = 0.0 0.0 0.1 0.2 0.3 0.5 0.7 1.0 1.0 1.0
88 \end{verbatim}
90 This is also equivalent to:
92 \begin{verbatim}
93 fep-lambdas = 0.0 0.0 0.0 0.0 0.4 0.5 0.6 0.7 0.8 1.0
94 coul-lambdas = 0.0 0.2 0.5 1.0 1.0 1.0 1.0 1.0 1.0 1.0
95 restraint-lambdas = 0.0 0.0 0.1 0.2 0.3 0.5 0.7 1.0 1.0 1.0
96 \end{verbatim}
97 The {\tt fep-lambda array}, in this case, is being used as the default to
98 fill in the bonded and van der Waals $\lambda$ arrays. Usually, it's best to fill
99 in all arrays explicitly, just to make sure things are properly
100 assigned.
102 If you want to turn on only restraints going from $A$ to $B$, then it would be:
103 \begin{verbatim}
104 restraint-lambdas = 0.0 0.1 0.2 0.4 0.6 1.0
105 \end{verbatim}
106 and all of the other components of the $\lambda$ vector would be left in the $A$ state.
108 To compute free energies with a vector $\lambda$ using
109 thermodynamic integration, then the TI equation becomes vector equation:
110 \beq
111 \Delta F = \int \langle \nabla H \rangle \cdot d\vec{\lambda}
112 \eeq
113 or for finite differences:
114 \beq
115 \Delta F \approx \int \sum \langle \nabla H \rangle \cdot \Delta\lambda
116 \eeq
118 The external {\tt pymbar} script downloaded from https://SimTK.org/home/pymbar can
119 compute this integral automatically from the {\gromacs} dhdl.xvg output.
121 \section{Potential of mean force}
123 A potential of mean force (PMF) is a potential that is obtained
124 by integrating the mean force from an ensemble of configurations.
125 In {\gromacs}, there are several different methods to calculate the mean force.
126 Each method has its limitations, which are listed below.
127 \begin{itemize}
128 \item{\bf pull code:} between the centers of mass of molecules or groups of molecules.
129 \item{\bf free-energy code with harmonic bonds or constraints:} between single atoms.
130 \item{\bf free-energy code with position restraints:} changing the conformation of a relatively immobile group of atoms.
131 \item{\bf pull code in limited cases:} between groups of atoms that are
132 part of a larger molecule for which the bonds are constrained with
133 SHAKE or LINCS. If the pull group if relatively large,
134 the pull code can be used.
135 \end{itemize}
136 The pull and free-energy code a described in more detail
137 in the following two sections.
139 \subsubsection{Entropic effects}
140 When a distance between two atoms or the centers of mass of two groups
141 is constrained or restrained, there will be a purely entropic contribution
142 to the PMF due to the rotation of the two groups~\cite{RMNeumann1980a}.
143 For a system of two non-interacting masses the potential of mean force is:
144 \beq
145 V_{pmf}(r) = -(n_c - 1) k_B T \log(r)
146 \eeq
147 where $n_c$ is the number of dimensions in which the constraint works
148 (i.e. $n_c=3$ for a normal constraint and $n_c=1$ when only
149 the $z$-direction is constrained).
150 Whether one needs to correct for this contribution depends on what
151 the PMF should represent. When one wants to pull a substrate
152 into a protein, this entropic term indeed contributes to the work to
153 get the substrate into the protein. But when calculating a PMF
154 between two solutes in a solvent, for the purpose of simulating
155 without solvent, the entropic contribution should be removed.
156 {\bf Note} that this term can be significant; when at 300K the distance is halved,
157 the contribution is 3.5 kJ~mol$^{-1}$.
159 \section{Non-equilibrium pulling}
160 When the distance between two groups is changed continuously,
161 work is applied to the system, which means that the system is no longer
162 in equilibrium. Although in the limit of very slow pulling
163 the system is again in equilibrium, for many systems this limit
164 is not reachable within reasonable computational time.
165 However, one can use the Jarzynski relation~\cite{Jarzynski1997a}
166 to obtain the equilibrium free-energy difference $\Delta G$
167 between two distances from many non-equilibrium simulations:
168 \begin{equation}
169 \Delta G_{AB} = -k_BT \log \left\langle e^{-\beta W_{AB}} \right\rangle_A
170 \label{eq:Jarz}
171 \end{equation}
172 where $W_{AB}$ is the work performed to force the system along one path
173 from state A to B, the angular bracket denotes averaging over
174 a canonical ensemble of the initial state A and $\beta=1/k_B T$.
177 \section{The pull code}
178 \index{center-of-mass pulling}
179 \label{sec:pull}
180 The pull code applies forces or constraints between the centers
181 of mass of one or more pairs of groups of atoms.
182 Each pull reaction coordinate is called a ``coordinate'' and it operates
183 on usually two, but sometimes more, pull groups. A pull group can be part of one or more pull
184 coordinates. Furthermore, a coordinate can also operate on a single group
185 and an absolute reference position in space.
186 The distance between a pair of groups can be determined
187 in 1, 2 or 3 dimensions, or can be along a user-defined vector.
188 The reference distance can be constant or can change linearly with time.
189 Normally all atoms are weighted by their mass, but an additional
190 weighting factor can also be used.
191 \begin{figure}
192 \centerline{\includegraphics[width=6cm,angle=270]{plots/pull}}
193 \caption{Schematic picture of pulling a lipid out of a lipid bilayer
194 with umbrella pulling. $V_{rup}$ is the velocity at which the spring is
195 retracted, $Z_{link}$ is the atom to which the spring is attached and
196 $Z_{spring}$ is the location of the spring.}
197 \label{fig:pull}
198 \end{figure}
200 Several different pull types, i.e. ways to apply the pull force, are supported,
201 and in all cases the reference distance can be constant
202 or linearly changing with time.
203 \begin{enumerate}
204 \item{\textbf{Umbrella pulling}\swapindexquiet{umbrella}{pulling}}
205 A harmonic potential is applied between
206 the centers of mass of two groups.
207 Thus, the force is proportional to the displacement.
208 \item{\textbf{Constraint pulling\swapindexquiet{constraint}{pulling}}}
209 The distance between the centers of mass of two groups is constrained.
210 The constraint force can be written to a file.
211 This method uses the SHAKE algorithm but only needs 1 iteration to be
212 exact if only two groups are constrained.
213 \item{\textbf{Constant force pulling}}
214 A constant force is applied between the centers of mass of two groups.
215 Thus, the potential is linear.
216 In this case there is no reference distance of pull rate.
217 \item{\textbf{Flat bottom pulling}}
218 Like umbrella pulling, but the potential and force are zero for
219 coordinate values below ({\tt pull-coord?-type = flat-bottom}) or above
220 ({\tt pull-coord?-type = flat-bottom-high}) a reference value.
221 This is useful for restraining e.g. the distance
222 between two molecules to a certain region.
223 \end{enumerate}
224 In addition, there are different types of reaction coordinates, so-called pull geometries.
225 These are set with the {\tt .mdp} option {\tt pull-coord?-geometry}.
227 \subsubsection{Definition of the center of mass}
229 In {\gromacs}, there are three ways to define the center of mass of a group.
230 The standard way is a ``plain'' center of mass, possibly with additional
231 weighting factors. With periodic boundary conditions it is no longer
232 possible to uniquely define the center of mass of a group of atoms.
233 Therefore, a reference atom is used. For determining the center of mass,
234 for all other atoms in the group, the closest periodic image to the reference
235 atom is used. This uniquely defines the center of mass.
236 By default, the middle (determined by the order in the topology) atom
237 is used as a reference atom, but the user can also select any other atom
238 if it would be closer to center of the group.
240 For a layered system, for instance a lipid bilayer, it may be of interest
241 to calculate the PMF of a lipid as function of its distance
242 from the whole bilayer. The whole bilayer can be taken as reference
243 group in that case, but it might also be of interest to define the
244 reaction coordinate for the PMF more locally. The {\tt .mdp} option
245 {\tt pull-coord?-geometry = cylinder} does not
246 use all the atoms of the reference group, but instead dynamically only those
247 within a cylinder with radius {\tt pull-cylinder-r} around the pull vector going
248 through the pull group. This only
249 works for distances defined in one dimension, and the cylinder is
250 oriented with its long axis along this one dimension. To avoid jumps in
251 the pull force, contributions of atoms are weighted as a function of distance
252 (in addition to the mass weighting):
253 \bea
254 w(r < r_\mathrm{cyl}) & = &
255 1-2 \left(\frac{r}{r_\mathrm{cyl}}\right)^2 + \left(\frac{r}{r_\mathrm{cyl}}\right)^4 \\
256 w(r \geq r_\mathrm{cyl}) & = & 0
257 \eea
258 Note that the radial dependence on the weight causes a radial force on
259 both cylinder group and the other pull group. This is an undesirable,
260 but unavoidable effect. To minimize this effect, the cylinder radius should
261 be chosen sufficiently large. The effective mass is 0.47 times that of
262 a cylinder with uniform weights and equal to the mass of uniform cylinder
263 of 0.79 times the radius.
265 \begin{figure}
266 \centerline{\includegraphics[width=6cm]{plots/pullref}}
267 \caption{Comparison of a plain center of mass reference group versus a cylinder
268 reference group applied to interface systems. C is the reference group.
269 The circles represent the center of mass of two groups plus the reference group,
270 $d_c$ is the reference distance.}
271 \label{fig:pullref}
272 \end{figure}
274 For a group of molecules in a periodic system, a plain reference group
275 might not be well-defined. An example is a water slab that is connected
276 periodically in $x$ and $y$, but has two liquid-vapor interfaces along $z$.
277 In such a setup, water molecules can evaporate from the liquid and they
278 will move through the vapor, through the periodic boundary, to the other
279 interface. Such a system is inherently periodic and there is no proper way
280 of defining a ``plain'' center of mass along $z$. A proper solution is to using
281 a cosine shaped weighting profile for all atoms in the reference group.
282 The profile is a cosine with a single period in the unit cell. Its phase
283 is optimized to give the maximum sum of weights, including mass weighting.
284 This provides a unique and continuous reference position that is nearly
285 identical to the plain center of mass position in case all atoms are all
286 within a half of the unit-cell length. See ref \cite{Engin2010a} for details.
288 When relative weights $w_i$ are used during the calculations, either
289 by supplying weights in the input or due to cylinder geometry
290 or due to cosine weighting,
291 the weights need to be scaled to conserve momentum:
292 \beq
293 w'_i = w_i
294 \left. \sum_{j=1}^N w_j \, m_j \right/ \sum_{j=1}^N w_j^2 \, m_j
295 \eeq
296 where $m_j$ is the mass of atom $j$ of the group.
297 The mass of the group, required for calculating the constraint force, is:
298 \beq
299 M = \sum_{i=1}^N w'_i \, m_i
300 \eeq
301 The definition of the weighted center of mass is:
302 \beq
303 \ve{r}_{com} = \left. \sum_{i=1}^N w'_i \, m_i \, \ve{r}_i \right/ M
304 \eeq
305 From the centers of mass the AFM, constraint, or umbrella force $\ve{F}_{\!com}$
306 on each group can be calculated.
307 The force on the center of mass of a group is redistributed to the atoms
308 as follows:
309 \beq
310 \ve{F}_{\!i} = \frac{w'_i \, m_i}{M} \, \ve{F}_{\!com}
311 \eeq
313 \subsubsection{Definition of the pull direction}
315 The most common setup is to pull along the direction of the vector containing
316 the two pull groups, this is selected with
317 {\tt pull-coord?-geometry = distance}. You might want to pull along a certain
318 vector instead, which is selected with {\tt pull-coord?-geometry = direction}.
319 But this can cause unwanted torque forces in the system, unless you pull against a reference group with (nearly) fixed orientation, e.g. a membrane protein embedded in a membrane along x/y while pulling along z. If your reference group does not have a fixed orientation, you should probably use
320 {\tt pull-coord?-geometry = direction-relative}, see \figref{pulldirrel}.
321 Since the potential now depends on the coordinates of two additional groups defining the orientation, the torque forces will work on these two groups.
323 \begin{figure}
324 \centerline{\includegraphics[width=5cm]{plots/pulldirrel}}
325 \caption{The pull setup for geometry {\tt direction-relative}. The ``normal'' pull groups are 1 and 2. Groups 3 and 4 define the pull direction and thus the direction of the normal pull forces (red). This leads to reaction forces (blue) on groups 3 and 4, which are perpendicular to the pull direction. Their magnitude is given by the ``normal'' pull force times the ratio of $d_p$ and the distance between groups 3 and 4.}
326 \label{fig:pulldirrel}
327 \end{figure}
329 \subsubsection{Definition of the angle and dihedral pull geometries}
330 Four pull groups are required for {\tt pull-coord?-geometry = angle}. In the same way as for geometries with two groups, each consecutive pair of groups
331 $i$ and $i+1$ define a vector connecting the COMs of groups $i$ and $i+1$. The angle is defined as the angle between the two resulting vectors.
332 E.g., the {\tt .mdp} option {\tt pull-coord?-groups = 1 2 2 4} defines the angle between the vector from the COM of group 1 to the COM of group 2
333 and the vector from the COM of group 2 to the COM of group 4.
334 The angle takes values in the closed interval [0, 180] deg.
335 For {\tt pull-coord?-geometry = angle-axis} the angle is defined with respect to a reference axis given by {\tt pull-coord?-vec} and only two groups need to be given.
336 The dihedral geometry requires six pull groups. These pair up in the same way as described above and so define three vectors. The dihedral angle is defined as the angle between the two
337 planes spanned by the two first and the two last vectors. Equivalently, the dihedral angle can be seen as the angle between the first and the third
338 vector when these vectors are projected onto a plane normal to the second vector (the axis vector).
339 As an example, consider a dihedral angle involving four groups: 1, 5, 8 and 9. Here, the {\tt .mdp} option
340 {\tt pull-coord?-groups = 8 1 1 5 5 9} specifies the three vectors that define the dihedral angle:
341 the first vector is the COM distance vector from group 8 to 1, the second vector is the COM distance vector from group 1 to 5, and the third vector is the COM distance vector from group 5 to 9.
342 The dihedral angle takes values in the interval (-180, 180] deg and has periodic boundaries.
344 \subsubsection{Limitations}
345 There is one theoretical limitation:
346 strictly speaking, constraint forces can only be calculated between
347 groups that are not connected by constraints to the rest of the system.
348 If a group contains part of a molecule of which the bond lengths
349 are constrained, the pull constraint and LINCS or SHAKE bond constraint
350 algorithms should be iterated simultaneously. This is not done in {\gromacs}.
351 This means that for simulations with {\tt constraints = all-bonds}
352 in the {\tt .mdp} file pulling is, strictly speaking,
353 limited to whole molecules or groups of molecules.
354 In some cases this limitation can be avoided by using the free energy code,
355 see \secref{fepmf}.
356 In practice, the errors caused by not iterating the two constraint
357 algorithms can be negligible when the pull group consists of a large
358 amount of atoms and/or the pull force is small.
359 In such cases, the constraint correction displacement of the pull group
360 is small compared to the bond lengths.
364 \section{\normindex{Enforced Rotation}}
365 \index{rotational pulling|see{enforced rotation}}
366 \index{pulling, rotational|see{enforced rotation}}
367 \label{sec:rotation}
369 \mathchardef\mhyphen="2D
370 \newcommand{\rotiso }{^\mathrm{iso}}
371 \newcommand{\rotisopf }{^\mathrm{iso\mhyphen pf}}
372 \newcommand{\rotpm }{^\mathrm{pm}}
373 \newcommand{\rotpmpf }{^\mathrm{pm\mhyphen pf}}
374 \newcommand{\rotrm }{^\mathrm{rm}}
375 \newcommand{\rotrmpf }{^\mathrm{rm\mhyphen pf}}
376 \newcommand{\rotrmtwo }{^\mathrm{rm2}}
377 \newcommand{\rotrmtwopf }{^\mathrm{rm2\mhyphen pf}}
378 \newcommand{\rotflex }{^\mathrm{flex}}
379 \newcommand{\rotflext }{^\mathrm{flex\mhyphen t}}
380 \newcommand{\rotflextwo }{^\mathrm{flex2}}
381 \newcommand{\rotflextwot}{^\mathrm{flex2\mhyphen t}}
383 This module can be used to enforce the rotation of a group of atoms, as {\eg}
384 a protein subunit. There are a variety of rotation potentials, among them
385 complex ones that allow flexible adaptations of both the rotated subunit as
386 well as the local rotation axis during the simulation. An example application
387 can be found in ref. \cite{Kutzner2011}.
389 \begin{figure}
390 \centerline{\includegraphics[width=13cm]{plots/rotation.pdf}}
391 \caption[Fixed and flexible axis rotation]{Comparison of fixed and flexible axis
392 rotation.
393 {\sf A:} Rotating the sketched shape inside the white tubular cavity can create
394 artifacts when a fixed rotation axis (dashed) is used. More realistically, the
395 shape would revolve like a flexible pipe-cleaner (dotted) inside the bearing (gray).
396 {\sf B:} Fixed rotation around an axis \ve{v} with a pivot point
397 specified by the vector \ve{u}.
398 {\sf C:} Subdividing the rotating fragment into slabs with separate rotation
399 axes ($\uparrow$) and pivot points ($\bullet$) for each slab allows for
400 flexibility. The distance between two slabs with indices $n$ and $n+1$ is $\Delta x$.}
401 \label{fig:rotation}
402 \end{figure}
404 \begin{figure}
405 \centerline{\includegraphics[width=13cm]{plots/equipotential.pdf}}
406 \caption{Selection of different rotation potentials and definition of notation.
407 All four potentials $V$ (color coded) are shown for a single atom at position
408 $\ve{x}_j(t)$.
409 {\sf A:} Isotropic potential $V\rotiso$,
410 {\sf B:} radial motion potential $V\rotrm$ and flexible potential
411 $V\rotflex$,
412 {\sf C--D:} radial motion\,2 potential $V\rotrmtwo$ and
413 flexible\,2 potential $V\rotflextwo$ for $\epsilon' = 0$\,nm$^2$ {\sf (C)}
414 and $\epsilon' = 0.01$\,nm$^2$ {\sf (D)}. The rotation axis is perpendicular to
415 the plane and marked by $\otimes$. The light gray contours indicate Boltzmann factors
416 $e^{-V/(k_B T)}$ in the $\ve{x}_j$-plane for $T=300$\,K and
417 $k=200$\,kJ/(mol$\cdot$nm$^2$). The green arrow shows the direction of the
418 force $\ve{F}_{\!j}$ acting on atom $j$; the blue dashed line indicates the
419 motion of the reference position.}
420 \label{fig:equipotential}
421 \end{figure}
423 \subsection{Fixed Axis Rotation}
424 \subsubsection{Stationary Axis with an Isotropic Potential}
425 In the fixed axis approach (see \figref{rotation}B), torque on a group of $N$
426 atoms with positions $\ve{x}_i$ (denoted ``rotation group'') is applied by
427 rotating a reference set of atomic positions -- usually their initial positions
428 $\ve{y}_i^0$ -- at a constant angular velocity $\omega$ around an axis
429 defined by a direction vector $\hat{\ve{v}}$ and a pivot point \ve{u}.
430 To that aim, each atom with position $\ve{x}_i$ is attracted by a
431 ``virtual spring'' potential to its moving reference position
432 $\ve{y}_i = \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{u})$,
433 where $\mathbf{\Omega}(t)$ is a matrix that describes the rotation around the
434 axis. In the simplest case, the ``springs'' are described by a harmonic
435 potential,
436 \beq
437 V\rotiso = \frac{k}{2} \sum_{i=1}^{N} w_i \left[ \mathbf{\Omega}(t)
438 (\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u}) \right]^2 ,
439 \label{eqn:potiso}
440 \eeq
441 with optional mass-weighted prefactors $w_i = N \, m_i/M$ with total mass
442 $M = \sum_{i=1}^N m_i$.
443 The rotation matrix $\mathbf{\Omega}(t)$ is
444 \newcommand{\omcost}{\,\xi\,} % abbreviation
445 \begin{displaymath}
446 \mathbf{\Omega}(t) =
447 \left(
448 \begin{array}{ccc}
449 \cos\omega t + v_x^2\omcost & v_x v_y\omcost - v_z\sin\omega t & v_x v_z\omcost + v_y\sin\omega t\\
450 v_x v_y\omcost + v_z\sin\omega t & \cos\omega t + v_y^2\omcost & v_y v_z\omcost - v_x\sin\omega t\\
451 v_x v_z\omcost - v_y\sin\omega t & v_y v_z\omcost + v_x\sin\omega t & \cos\omega t + v_z^2\omcost \\
452 \end{array}
453 \right) ,
454 \end{displaymath}
455 where $v_x$, $v_y$, and $v_z$ are the components of the normalized rotation vector
456 $\hat{\ve{v}}$, and $\omcost := 1-\cos(\omega t)$. As illustrated in
457 \figref{equipotential}A for a single atom $j$, the
458 rotation matrix $\mathbf{\Omega}(t)$ operates on the initial reference positions
459 $\ve{y}_j^0 = \ve{x}_j(t_0)$ of atom $j$ at $t=t_0$. At a later
460 time $t$, the reference position has rotated away from its initial place
461 (along the blue dashed line), resulting in the force
462 \beq
463 \ve{F}_{\!j}\rotiso
464 = -\nabla_{\!j} \, V\rotiso
465 = k \, w_j \left[
466 \mathbf{\Omega}(t) (\ve{y}_j^0 - \ve{u}) - (\ve{x}_j - \ve{u} ) \right] ,
467 \label{eqn:force_fixed}
468 \eeq
469 which is directed towards the reference position.
472 \subsubsection{Pivot-Free Isotropic Potential}
473 Instead of a fixed pivot vector \ve{u} this potential uses the center of
474 mass $\ve{x}_c$ of the rotation group as pivot for the rotation axis,
475 \beq
476 \ve{x}_c = \frac{1}{M} \sum_{i=1}^N m_i \ve{x}_i
477 \label{eqn:com}
478 \mbox{\hspace{4ex}and\hspace{4ex}}
479 \ve{y}_c^0 = \frac{1}{M} \sum_{i=1}^N m_i \ve{y}_i^0 \ ,
480 \eeq
481 which yields the ``pivot-free'' isotropic potential
482 \beq
483 V\rotisopf = \frac{k}{2} \sum_{i=1}^{N} w_i \left[ \mathbf{\Omega}(t)
484 (\ve{y}_i^0 - \ve{y}_c^0) - (\ve{x}_i - \ve{x}_c) \right]^2 ,
485 \label{eqn:potisopf}
486 \eeq
487 with forces
488 \beq
489 \mathbf{F}_{\!j}\rotisopf = k \, w_j
490 \left[
491 \mathbf{\Omega}(t) ( \ve{y}_j^0 - \ve{y}_c^0)
492 - ( \ve{x}_j - \ve{x}_c )
493 \right] .
494 \label{eqn:force_isopf}
495 \eeq
496 Without mass-weighting, the pivot $\ve{x}_c$ is the geometrical center of
497 the group.
498 \label{sec:fixed}
500 \subsubsection{Parallel Motion Potential Variant}
501 The forces generated by the isotropic potentials
502 (\eqnsref{potiso}{potisopf}) also contain components parallel
503 to the rotation axis and thereby restrain motions along the axis of either the
504 whole rotation group (in case of $V\rotiso$) or within
505 the rotation group (in case of $V\rotisopf$). For cases where
506 unrestrained motion along the axis is preferred, we have implemented a
507 ``parallel motion'' variant by eliminating all components parallel to the
508 rotation axis for the potential. This is achieved by projecting the distance
509 vectors between reference and actual positions
510 \beq
511 \ve{r}_i = \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u})
512 \eeq
513 onto the plane perpendicular to the rotation vector,
514 \beq
515 \label{eqn:project}
516 \ve{r}_i^\perp := \ve{r}_i - (\ve{r}_i \cdot \hat{\ve{v}})\hat{\ve{v}} \ ,
517 \eeq
518 yielding
519 \bea
520 \nonumber
521 V\rotpm &=& \frac{k}{2} \sum_{i=1}^{N} w_i ( \ve{r}_i^\perp )^2 \\
522 &=& \frac{k}{2} \sum_{i=1}^{N} w_i
523 \left\lbrace
524 \mathbf{\Omega}(t)
525 (\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u}) \right. \nonumber \\
526 && \left. - \left\lbrace
527 \left[ \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u}) \right] \cdot\hat{\ve{v}}
528 \right\rbrace\hat{\ve{v}} \right\rbrace^2 ,
529 \label{eqn:potpm}
530 \eea
531 and similarly
532 \beq
533 \ve{F}_{\!j}\rotpm = k \, w_j \, \ve{r}_j^\perp .
534 \label{eqn:force_pm}
535 \eeq
537 \subsubsection{Pivot-Free Parallel Motion Potential}
538 Replacing in \eqnref{potpm} the fixed pivot \ve{u} by the center
539 of mass $\ve{x_c}$ yields the pivot-free variant of the parallel motion
540 potential. With
541 \beq
542 \ve{s}_i = \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{y}_c^0) - (\ve{x}_i - \ve{x}_c)
543 \eeq
544 the respective potential and forces are
545 \bea
546 V\rotpmpf &=& \frac{k}{2} \sum_{i=1}^{N} w_i ( \ve{s}_i^\perp )^2 \ , \\
547 \label{eqn:potpmpf}
548 \ve{F}_{\!j}\rotpmpf &=& k \, w_j \, \ve{s}_j^\perp .
549 \label{eqn:force_pmpf}
550 \eea
552 \subsubsection{Radial Motion Potential}
553 In the above variants, the minimum of the rotation potential is either a single
554 point at the reference position $\ve{y}_i$ (for the isotropic potentials) or a
555 single line through $\ve{y}_i$ parallel to the rotation axis (for the
556 parallel motion potentials). As a result, radial forces restrict radial motions
557 of the atoms. The two subsequent types of rotation potentials, $V\rotrm$
558 and $V\rotrmtwo$, drastically reduce or even eliminate this effect. The first
559 variant, $V\rotrm$ (\figref{equipotential}B), eliminates all force
560 components parallel to the vector connecting the reference atom and the
561 rotation axis,
562 \beq
563 V\rotrm = \frac{k}{2} \sum_{i=1}^{N} w_i \left[
564 \ve{p}_i
565 \cdot(\ve{x}_i - \ve{u}) \right]^2 ,
566 \label{eqn:potrm}
567 \eeq
568 with
569 \beq
570 \ve{p}_i :=
571 \frac{\hat{\ve{v}}\times \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{u})} {\| \hat{\ve{v}}\times \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{u})\|} \ .
572 \eeq
573 This variant depends only on the distance $\ve{p}_i \cdot (\ve{x}_i -
574 \ve{u})$ of atom $i$ from the plane spanned by $\hat{\ve{v}}$ and
575 $\mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u})$. The resulting force is
576 \beq
577 \mathbf{F}_{\!j}\rotrm =
578 -k \, w_j \left[ \ve{p}_j\cdot(\ve{x}_j - \ve{u}) \right] \,\ve{p}_j \, .
579 \label{eqn:potrm_force}
580 \eeq
582 \subsubsection{Pivot-Free Radial Motion Potential}
583 Proceeding similar to the pivot-free isotropic potential yields a pivot-free
584 version of the above potential. With
585 \beq
586 \ve{q}_i :=
587 \frac{\hat{\ve{v}}\times \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{y}_c^0)} {\| \hat{\ve{v}}\times \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{y}_c^0)\|} \, ,
588 \eeq
589 the potential and force for the pivot-free variant of the radial motion potential read
590 \bea
591 V\rotrmpf & = & \frac{k}{2} \sum_{i=1}^{N} w_i \left[
592 \ve{q}_i
593 \cdot(\ve{x}_i - \ve{x}_c)
594 \right]^2 \, , \\
595 \label{eqn:potrmpf}
596 \mathbf{F}_{\!j}\rotrmpf & = &
597 -k \, w_j \left[ \ve{q}_j\cdot(\ve{x}_j - \ve{x}_c) \right] \,\ve{q}_j
598 + k \frac{m_j}{M} \sum_{i=1}^{N} w_i \left[
599 \ve{q}_i\cdot(\ve{x}_i - \ve{x}_c) \right]\,\ve{q}_i \, .
600 \label{eqn:potrmpf_force}
601 \eea
603 \subsubsection{Radial Motion 2 Alternative Potential}
604 As seen in \figref{equipotential}B, the force resulting from
605 $V\rotrm$ still contains a small, second-order radial component. In most
606 cases, this perturbation is tolerable; if not, the following
607 alternative, $V\rotrmtwo$, fully eliminates the radial contribution to the
608 force, as depicted in \figref{equipotential}C,
609 \beq
610 V\rotrmtwo =
611 \frac{k}{2} \sum_{i=1}^{N} w_i\,
612 \frac{\left[ (\hat{\ve{v}} \times ( \ve{x}_i - \ve{u} ))
613 \cdot \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u}) \right]^2}
614 {\| \hat{\ve{v}} \times (\ve{x}_i - \ve{u}) \|^2 +
615 \epsilon'} \, ,
616 \label{eqn:potrm2}
617 \eeq
618 where a small parameter $\epsilon'$ has been introduced to avoid singularities.
619 For $\epsilon'=0$\,nm$^2$, the equipotential planes are spanned by $\ve{x}_i -
620 \ve{u}$ and $\hat{\ve{v}}$, yielding a force
621 perpendicular to $\ve{x}_i - \ve{u}$, thus not contracting or
622 expanding structural parts that moved away from or toward the rotation axis.
624 Choosing a small positive $\epsilon'$ ({\eg},
625 $\epsilon'=0.01$\,nm$^2$, \figref{equipotential}D) in the denominator of
626 \eqnref{potrm2} yields a well-defined potential and continuous forces also
627 close to the rotation axis, which is not the case for $\epsilon'=0$\,nm$^2$
628 (\figref{equipotential}C). With
629 \bea
630 \ve{r}_i & := & \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u})\\
631 \ve{s}_i & := & \frac{\hat{\ve{v}} \times (\ve{x}_i -
632 \ve{u} ) }{ \| \hat{\ve{v}} \times (\ve{x}_i - \ve{u})
633 \| } \equiv \; \Psi_{i} \;\; {\hat{\ve{v}} \times
634 (\ve{x}_i-\ve{u} ) }\\
635 \Psi_i^{*} & := & \frac{1}{ \| \hat{\ve{v}} \times
636 (\ve{x}_i-\ve{u}) \|^2 + \epsilon'}
637 \eea
638 the force on atom $j$ reads
639 \beq
640 \ve{F}_{\!j}\rotrmtwo =
641 - k\;
642 \left\lbrace w_j\;
643 (\ve{s}_j\cdot\ve{r}_{\!j})\;
644 \left[ \frac{\Psi_{\!j}^* }{\Psi_{\!j} } \ve{r}_{\!j}
645 - \frac{\Psi_{\!j}^{*2}}{\Psi_{\!j}^3}
646 (\ve{s}_j\cdot\ve{r}_{\!j})\ve{s}_j \right]
647 \right\rbrace \times \hat{\ve{v}} .
648 \label{eqn:potrm2_force}
649 \eeq
651 \subsubsection{Pivot-Free Radial Motion 2 Potential}
652 The pivot-free variant of the above potential is
653 \beq
654 V\rotrmtwopf =
655 \frac{k}{2} \sum_{i=1}^{N} w_i\,
656 \frac{\left[ (\hat{\ve{v}} \times ( \ve{x}_i - \ve{x}_c ))
657 \cdot \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c) \right]^2}
658 {\| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c) \|^2 +
659 \epsilon'} \, .
660 \label{eqn:potrm2pf}
661 \eeq
662 With
663 \bea
664 \ve{r}_i & := & \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c)\\
665 \ve{s}_i & := & \frac{\hat{\ve{v}} \times (\ve{x}_i -
666 \ve{x}_c ) }{ \| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c)
667 \| } \equiv \; \Psi_{i} \;\; {\hat{\ve{v}} \times
668 (\ve{x}_i-\ve{x}_c ) }\\ \Psi_i^{*} & := & \frac{1}{ \| \hat{\ve{v}} \times
669 (\ve{x}_i-\ve{x}_c) \|^2 + \epsilon'}
670 \eea
671 the force on atom $j$ reads
672 \bea
673 \nonumber
674 \ve{F}_{\!j}\rotrmtwopf & = &
675 - k\;
676 \left\lbrace w_j\;
677 (\ve{s}_j\cdot\ve{r}_{\!j})\;
678 \left[ \frac{\Psi_{\!j}^* }{\Psi_{\!j} } \ve{r}_{\!j}
679 - \frac{\Psi_{\!j}^{*2}}{\Psi_{\!j}^3}
680 (\ve{s}_j\cdot\ve{r}_{\!j})\ve{s}_j \right]
681 \right\rbrace \times \hat{\ve{v}}\\
683 + k\;\frac{m_j}{M} \left\lbrace \sum_{i=1}^{N}
684 w_i\;(\ve{s}_i\cdot\ve{r}_i) \;
685 \left[ \frac{\Psi_i^* }{\Psi_i } \ve{r}_i
686 - \frac{\Psi_i^{*2}}{\Psi_i^3} (\ve{s}_i\cdot\ve{r}_i )\;
687 \ve{s}_i \right] \right\rbrace \times \hat{\ve{v}} \, .
688 \label{eqn:potrm2pf_force}
689 \eea
691 \subsection{Flexible Axis Rotation}
692 As sketched in \figref{rotation}A--B, the rigid body behavior of
693 the fixed axis rotation scheme is a drawback for many applications. In
694 particular, deformations of the rotation group are suppressed when the
695 equilibrium atom positions directly depend on the reference positions.
696 To avoid this limitation, \eqnsref{potrmpf}{potrm2pf}
697 will now be generalized towards a ``flexible axis'' as sketched in
698 \figref{rotation}C. This will be achieved by subdividing the
699 rotation group into a set of equidistant slabs perpendicular to
700 the rotation vector, and by applying a separate rotation potential to each
701 of these slabs. \figref{rotation}C shows the midplanes of the slabs
702 as dotted straight lines and the centers as thick black dots.
704 To avoid discontinuities in the potential and in the forces, we define
705 ``soft slabs'' by weighing the contributions of each
706 slab $n$ to the total potential function $V\rotflex$ by a Gaussian
707 function
708 \beq
709 \label{eqn:gaussian}
710 g_n(\ve{x}_i) = \Gamma \ \mbox{exp} \left(
711 -\frac{\beta_n^2(\ve{x}_i)}{2\sigma^2} \right) ,
712 \eeq
713 centered at the midplane of the $n$th slab. Here $\sigma$ is the width
714 of the Gaussian function, $\Delta x$ the distance between adjacent slabs, and
715 \beq
716 \beta_n(\ve{x}_i) := \ve{x}_i \cdot \hat{\ve{v}} - n \, \Delta x \, .
717 \eeq
719 \begin{figure}
720 \centerline{\includegraphics[width=6.5cm]{plots/gaussians.pdf}}
721 \caption{Gaussian functions $g_n$ centered at $n \, \Delta x$ for a slab
722 distance $\Delta x = 1.5$ nm and $n \geq -2$. Gaussian function $g_0$ is
723 highlighted in bold; the dashed line depicts the sum of the shown Gaussian
724 functions.}
725 \label{fig:gaussians}
726 \end{figure}
728 A most convenient choice is $\sigma = 0.7 \Delta x$ and
729 \begin{displaymath}
730 1/\Gamma = \sum_{n \in Z}
731 \mbox{exp}
732 \left(-\frac{(n - \frac{1}{4})^2}{2\cdot 0.7^2}\right)
733 \approx 1.75464 \, ,
734 \end{displaymath}
735 which yields a nearly constant sum, essentially independent of $\ve{x}_i$
736 (dashed line in \figref{gaussians}), {\ie},
737 \beq
738 \sum_{n \in Z} g_n(\ve{x}_i) = 1 + \epsilon(\ve{x}_i) \, ,
739 \label{eqn:normal}
740 \eeq
741 with $ | \epsilon(\ve{x}_i) | < 1.3\cdot 10^{-4}$. This choice also
742 implies that the individual contributions to the force from the slabs add up to
743 unity such that no further normalization is required.
745 To each slab center $\ve{x}_c^n$, all atoms contribute by their
746 Gaussian-weighted (optionally also mass-weighted) position vectors
747 $g_n(\ve{x}_i) \, \ve{x}_i$. The
748 instantaneous slab centers $\ve{x}_c^n$ are calculated from the
749 current positions $\ve{x}_i$,
750 \beq
751 \label{eqn:defx0}
752 \ve{x}_c^n =
753 \frac{\sum_{i=1}^N g_n(\ve{x}_i) \, m_i \, \ve{x}_i}
754 {\sum_{i=1}^N g_n(\ve{x}_i) \, m_i} \, ,\\
755 \eeq
756 while the reference centers $\ve{y}_c^n$ are calculated from the reference
757 positions $\ve{y}_i^0$,
758 \beq
759 \label{eqn:defy0}
760 \ve{y}_c^n =
761 \frac{\sum_{i=1}^N g_n(\ve{y}_i^0) \, m_i \, \ve{y}_i^0}
762 {\sum_{i=1}^N g_n(\ve{y}_i^0) \, m_i} \, .
763 \eeq
764 Due to the rapid decay of $g_n$, each slab
765 will essentially involve contributions from atoms located within $\approx
766 3\Delta x$ from the slab center only.
768 \subsubsection{Flexible Axis Potential}
769 We consider two flexible axis variants. For the first variant,
770 the slab segmentation procedure with Gaussian weighting is applied to the radial
771 motion potential (\eqnref{potrmpf}\,/\,\figref{equipotential}B),
772 yielding as the contribution of slab $n$
773 \begin{displaymath}
774 V^n =
775 \frac{k}{2} \sum_{i=1}^{N} w_i \, g_n(\ve{x}_i)
776 \left[
777 \ve{q}_i^n
778 \cdot
779 (\ve{x}_i - \ve{x}_c^n)
780 \right]^2 ,
781 \label{eqn:flexpot}
782 \end{displaymath}
783 and a total potential function
784 \beq
785 V\rotflex = \sum_n V^n \, .
786 \label{eqn:potflex}
787 \eeq
788 Note that the global center of mass $\ve{x}_c$ used in
789 \eqnref{potrmpf} is now replaced by $\ve{x}_c^n$, the center of mass of
790 the slab. With
791 \bea
792 \ve{q}_i^n & := & \frac{\hat{\ve{v}} \times
793 \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n) }{ \| \hat{\ve{v}}
794 \times \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n) \| } \\
795 b_i^n & := & \ve{q}_i^n \cdot (\ve{x}_i - \ve{x}_c^n) \, ,
796 \eea
797 the resulting force on atom $j$ reads
798 \bea
799 \nonumber\hspace{-15mm}
800 \ve{F}_{\!j}\rotflex &=&
801 - \, k \, w_j \sum_n g_n(\ve{x}_j) \, b_j^n \left\lbrace \ve{q}_j^n -
802 b_j^n \frac{\beta_n(\ve{x}_j)}{2\sigma^2} \hat{\ve{v}} \right\rbrace \\ & &
803 + \, k \, m_j \sum_n \frac{g_n(\ve{x}_j)}{\sum_h g_n(\ve{x}_h)}
804 \sum_{i=1}^{N} w_i \, g_n(\ve{x}_i) \, b_i^n \left\lbrace
805 \ve{q}_i^n -\frac{\beta_n(\ve{x}_j)}{\sigma^2}
806 \left[ \ve{q}_i^n \cdot (\ve{x}_j - \ve{x}_c^n )\right]
807 \hat{\ve{v}} \right\rbrace .
808 \label{eqn:potflex_force}
809 \eea
811 Note that for $V\rotflex$, as defined, the slabs are fixed in space and so
812 are the reference centers $\ve{y}_c^n$. If during the simulation the
813 rotation group moves too far in $\ve{v}$ direction, it may enter a
814 region where -- due to the lack of nearby reference positions -- no reference
815 slab centers are defined, rendering the potential evaluation impossible.
816 We therefore have included a slightly modified version of this potential that
817 avoids this problem by attaching the midplane of slab $n=0$ to the center of mass
818 of the rotation group, yielding slabs that move with the rotation group.
819 This is achieved by subtracting the center of mass $\ve{x}_c$ of the
820 group from the positions,
821 \beq
822 \tilde{\ve{x}}_i = \ve{x}_i - \ve{x}_c \, , \mbox{\ \ \ and \ \ }
823 \tilde{\ve{y}}_i^0 = \ve{y}_i^0 - \ve{y}_c^0 \, ,
824 \label{eqn:trafo}
825 \eeq
826 such that
827 \bea
828 V\rotflext
829 & = & \frac{k}{2} \sum_n \sum_{i=1}^{N} w_i \, g_n(\tilde{\ve{x}}_i)
830 \left[ \frac{\hat{\ve{v}} \times \mathbf{\Omega}(t)(\tilde{\ve{y}}_i^0
831 - \tilde{\ve{y}}_c^n) }{ \| \hat{\ve{v}} \times
832 \mathbf{\Omega}(t)(\tilde{\ve{y}}_i^0 -
833 \tilde{\ve{y}}_c^n) \| }
834 \cdot
835 (\tilde{\ve{x}}_i - \tilde{\ve{x}}_c^n)
836 \right]^2 .
837 \label{eqn:potflext}
838 \eea
839 To simplify the force derivation, and for efficiency reasons, we here assume
840 $\ve{x}_c$ to be constant, and thus $\partial \ve{x}_c / \partial x =
841 \partial \ve{x}_c / \partial y = \partial \ve{x}_c / \partial z = 0$. The
842 resulting force error is small (of order $O(1/N)$ or $O(m_j/M)$ if
843 mass-weighting is applied) and can therefore be tolerated. With this assumption,
844 the forces $\ve{F}\rotflext$ have the same form as
845 \eqnref{potflex_force}.
847 \subsubsection{Flexible Axis 2 Alternative Potential}
848 In this second variant, slab segmentation is applied to $V\rotrmtwo$
849 (\eqnref{potrm2pf}), resulting in a flexible axis potential without radial
850 force contributions (\figref{equipotential}C),
851 \beq
852 V\rotflextwo =
853 \frac{k}{2} \sum_{i=1}^{N} \sum_n w_i\,g_n(\ve{x}_i)
854 \frac{\left[ (\hat{\ve{v}} \times ( \ve{x}_i - \ve{x}_c^n ))
855 \cdot \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n) \right]^2}
856 {\| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c^n) \|^2 +
857 \epsilon'} \, .
858 \label{eqn:potflex2}
859 \eeq
860 With
861 \bea
862 \ve{r}_i^n & := & \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n)\\
863 \ve{s}_i^n & := & \frac{\hat{\ve{v}} \times (\ve{x}_i -
864 \ve{x}_c^n ) }{ \| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c^n)
865 \| } \equiv \; \psi_{i} \;\; {\hat{\ve{v}} \times (\ve{x}_i-\ve{x}_c^n ) }\\
866 \psi_i^{*} & := & \frac{1}{ \| \hat{\ve{v}} \times (\ve{x}_i-\ve{x}_c^n) \|^2 + \epsilon'}\\
867 W_j^n & := & \frac{g_n(\ve{x}_j)\,m_j}{\sum_h g_n(\ve{x}_h)\,m_h}\\
868 \ve{S}^n & := &
869 \sum_{i=1}^{N} w_i\;g_n(\ve{x}_i)
870 \; (\ve{s}_i^n\cdot\ve{r}_i^n)
871 \left[ \frac{\psi_i^* }{\psi_i } \ve{r}_i^n
872 - \frac{\psi_i^{*2}}{\psi_i^3} (\ve{s}_i^n\cdot\ve{r}_i^n )\;
873 \ve{s}_i^n \right] \label{eqn:Sn}
874 \eea
875 the force on atom $j$ reads
876 \bea
877 \nonumber
878 \ve{F}_{\!j}\rotflextwo & = &
879 - k\;
880 \left\lbrace \sum_n w_j\;g_n(\ve{x}_j)\;
881 (\ve{s}_j^n\cdot\ve{r}_{\!j}^n)\;
882 \left[ \frac{\psi_j^* }{\psi_j } \ve{r}_{\!j}^n
883 - \frac{\psi_j^{*2}}{\psi_j^3} (\ve{s}_j^n\cdot\ve{r}_{\!j}^n)\;
884 \ve{s}_{\!j}^n \right] \right\rbrace \times \hat{\ve{v}} \\
885 \nonumber
887 + k \left\lbrace \sum_n W_{\!j}^n \, \ve{S}^n \right\rbrace \times
888 \hat{\ve{v}}
889 - k \left\lbrace \sum_n W_{\!j}^n \; \frac{\beta_n(\ve{x}_j)}{\sigma^2} \frac{1}{\psi_j}\;\;
890 \ve{s}_j^n \cdot
891 \ve{S}^n \right\rbrace \hat{\ve{v}}\\
892 & &
893 + \frac{k}{2} \left\lbrace \sum_n w_j\;g_n(\ve{x}_j)
894 \frac{\beta_n(\ve{x}_j)}{\sigma^2}
895 \frac{\psi_j^*}{\psi_j^2}( \ve{s}_j^n \cdot \ve{r}_{\!j}^n )^2 \right\rbrace
896 \hat{\ve{v}} .
897 \label{eqn:potflex2_force}
898 \eea
900 Applying transformation (\ref{eqn:trafo}) yields a ``translation-tolerant''
901 version of the flexible\,2 potential, $V\rotflextwot$. Again,
902 assuming that $\partial \ve{x}_c / \partial x$, $\partial \ve{x}_c /
903 \partial y$, $\partial \ve{x}_c / \partial z$ are small, the
904 resulting equations for $V\rotflextwot$ and $\ve{F}\rotflextwot$ are
905 similar to those of $V\rotflextwo$ and $\ve{F}\rotflextwo$.
907 \subsection{Usage}
908 To apply enforced rotation, the particles $i$ that are to
909 be subjected to one of the rotation potentials are defined via index groups
910 {\tt rot-group0}, {\tt rot-group1}, etc., in the {\tt .mdp} input file.
911 The reference positions $\ve{y}_i^0$ are
912 read from a special {\tt .trr} file provided to {\tt grompp}. If no such file is found,
913 $\ve{x}_i(t=0)$ are used as reference positions and written to {\tt .trr} such
914 that they can be used for subsequent setups. All parameters of the potentials
915 such as $k$, $\epsilon'$, etc. (\tabref{vars}) are provided as {\tt .mdp}
916 parameters; {\tt rot-type} selects the type of the potential.
917 The option {\tt rot-massw} allows to choose whether or not to use
918 mass-weighted averaging.
919 For the flexible potentials, a cutoff value $g_n^\mathrm{min}$
920 (typically $g_n^\mathrm{min}=0.001$) makes shure that only
921 significant contributions to $V$ and \ve{F} are evaluated, {\ie} terms with
922 $g_n(\ve{x}) < g_n^\mathrm{min}$ are omitted.
923 \tabref{quantities} summarizes observables that are written
924 to additional output files and which are described below.
927 \begin{table}[tbp]
928 \caption{Parameters used by the various rotation potentials.
929 {\sf x}'s indicate which parameter is actually used for a given potential.}
930 %\small
932 \newcommand{\kunit}{$\frac{\mathrm{kJ}}{\mathrm{mol} \cdot \mathrm{nm}^2}$}
933 \newcommand{\smtt}[1]{{\hspace{-0.5ex}\small #1\hspace{-0.5ex}}}
934 \label{tab:vars}
935 \begin{center}
936 \begin{tabular}{l>{$}l<{$}rccccccc}
937 \hline
938 parameter & & & $k$ & $\hat{\ve{v}}$ & $\ve{u}$ & $\omega$ & $\epsilon'$ & $\Delta x$ & $g_n^\mathrm{min}$ \\
939 \multicolumn{3}{l}{{\tt .mdp} input variable name} & \smtt{k} & \smtt{vec} & \smtt{pivot} & \smtt{rate} & \smtt{eps} & \smtt{slab-dist} & \smtt{min-gauss} \\
940 unit & & & \kunit & - & nm & $^\circ$/ps & nm$^2$ & nm & \,-\, \\[1mm]
941 \hline \multicolumn{2}{l}{fixed axis potentials:} & eqn.\\
942 isotropic & V\rotiso & (\ref{eqn:potiso}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - & - \\
943 --- pivot-free & V\rotisopf & (\ref{eqn:potisopf}) & {\sf x} & {\sf x} & - & {\sf x} & - & - & - \\
944 parallel motion & V\rotpm & (\ref{eqn:potpm}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - & - \\
945 --- pivot-free & V\rotpmpf & (\ref{eqn:potpmpf}) & {\sf x} & {\sf x} & - & {\sf x} & - & - & - \\
946 radial motion & V\rotrm & (\ref{eqn:potrm}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - & - \\
947 --- pivot-free & V\rotrmpf & (\ref{eqn:potrmpf}) & {\sf x} & {\sf x} & - & {\sf x} & - & - & - \\
948 radial motion\,2 & V\rotrmtwo & (\ref{eqn:potrm2}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - \\
949 --- pivot-free & V\rotrmtwopf & (\ref{eqn:potrm2pf}) & {\sf x} & {\sf x} & - & {\sf x} & {\sf x} & - & - \\ \hline
950 \multicolumn{2}{l}{flexible axis potentials:} & eqn.\\
951 flexible & V\rotflex & (\ref{eqn:potflex}) & {\sf x} & {\sf x} & - & {\sf x} & - & {\sf x} & {\sf x} \\
952 --- transl. tol. & V\rotflext & (\ref{eqn:potflext}) & {\sf x} & {\sf x} & - & {\sf x} & - & {\sf x} & {\sf x} \\
953 flexible\,2 & V\rotflextwo & (\ref{eqn:potflex2}) & {\sf x} & {\sf x} & - & {\sf x} & {\sf x} & {\sf x} & {\sf x} \\
954 --- transl. tol. & V\rotflextwot & - & {\sf x} & {\sf x} & - & {\sf x} & {\sf x} & {\sf x} & {\sf x} \\
955 \hline
956 \end{tabular}
957 \end{center}
958 \end{table}
960 \begin{table}
961 \caption{Quantities recorded in output files during enforced rotation simulations.
962 All slab-wise data is written every {\tt nstsout} steps, other rotation data every {\tt nstrout} steps.}
963 \label{tab:quantities}
964 \begin{center}
965 \begin{tabular}{llllcc}
966 \hline
967 quantity & unit & equation & output file & fixed & flexible\\ \hline
968 $V(t)$ & kJ/mol & see \ref{tab:vars} & {\tt rotation} & {\sf x} & {\sf x} \\
969 $\theta_\mathrm{ref}(t)$ & degrees & $\theta_\mathrm{ref}(t)=\omega t$ & {\tt rotation} & {\sf x} & {\sf x} \\
970 $\theta_\mathrm{av}(t)$ & degrees & (\ref{eqn:avangle}) & {\tt rotation} & {\sf x} & - \\
971 $\theta_\mathrm{fit}(t)$, $\theta_\mathrm{fit}(t,n)$ & degrees & (\ref{eqn:rmsdfit}) & {\tt rotangles} & - & {\sf x} \\
972 $\ve{y}_0(n)$, $\ve{x}_0(t,n)$ & nm & (\ref{eqn:defx0}, \ref{eqn:defy0})& {\tt rotslabs} & - & {\sf x} \\
973 $\tau(t)$ & kJ/mol & (\ref{eqn:torque}) & {\tt rotation} & {\sf x} & - \\
974 $\tau(t,n)$ & kJ/mol & (\ref{eqn:torque}) & {\tt rottorque} & - & {\sf x} \\ \hline
975 \end{tabular}
976 \end{center}
977 \end{table}
980 \subsubsection*{Angle of Rotation Groups: Fixed Axis}
981 For fixed axis rotation, the average angle $\theta_\mathrm{av}(t)$ of the
982 group relative to the reference group is determined via the distance-weighted
983 angular deviation of all rotation group atoms from their reference positions,
984 \beq
985 \theta_\mathrm{av} = \left. \sum_{i=1}^{N} r_i \ \theta_i \right/ \sum_{i=1}^N r_i \ .
986 \label{eqn:avangle}
987 \eeq
988 Here, $r_i$ is the distance of the reference position to the rotation axis, and
989 the difference angles $\theta_i$ are determined from the atomic positions,
990 projected onto a plane perpendicular to the rotation axis through pivot point
991 $\ve{u}$ (see \eqnref{project} for the definition of $\perp$),
992 \beq
993 \cos \theta_i =
994 \frac{(\ve{y}_i-\ve{u})^\perp \cdot (\ve{x}_i-\ve{u})^\perp}
995 { \| (\ve{y}_i-\ve{u})^\perp \cdot (\ve{x}_i-\ve{u})^\perp
996 \| } \ .
997 \eeq
999 The sign of $\theta_\mathrm{av}$ is chosen such that
1000 $\theta_\mathrm{av} > 0$ if the actual structure rotates ahead of the reference.
1002 \subsubsection*{Angle of Rotation Groups: Flexible Axis}
1003 For flexible axis rotation, two outputs are provided, the angle of the
1004 entire rotation group, and separate angles for the segments in the slabs.
1005 The angle of the entire rotation group is determined by an RMSD fit
1006 of $\ve{x}_i$
1007 to the reference positions $\ve{y}_i^0$ at $t=0$, yielding $\theta_\mathrm{fit}$
1008 as the angle by which the reference has to be rotated around $\hat{\ve{v}}$
1009 for the optimal fit,
1010 \beq
1011 \mathrm{RMSD} \big( \ve{x}_i,\ \mathbf{\Omega}(\theta_\mathrm{fit})
1012 \ve{y}_i^0 \big) \stackrel{!}{=} \mathrm{min} \, .
1013 \label{eqn:rmsdfit}
1014 \eeq
1015 To determine the local angle for each slab $n$, both reference and actual
1016 positions are weighted with the Gaussian function of slab $n$, and
1017 $\theta_\mathrm{fit}(t,n)$ is calculated as in \eqnref{rmsdfit}) from the
1018 Gaussian-weighted positions.
1020 For all angles, the {\tt .mdp} input option {\tt rot-fit-method} controls
1021 whether a normal RMSD fit is performed or whether for the fit each
1022 position $\ve{x}_i$ is put at the same distance to the rotation axis as its
1023 reference counterpart $\ve{y}_i^0$. In the latter case, the RMSD
1024 measures only angular differences, not radial ones.
1027 \subsubsection*{Angle Determination by Searching the Energy Minimum}
1028 Alternatively, for {\tt rot-fit-method = potential}, the angle of the rotation
1029 group is determined as the angle for which the rotation potential energy is minimal.
1030 Therefore, the used rotation potential is additionally evaluated for a set of angles
1031 around the current reference angle. In this case, the {\tt rotangles.log} output file
1032 contains the values of the rotation potential at the chosen set of angles, while
1033 {\tt rotation.xvg} lists the angle with minimal potential energy.
1036 \subsubsection*{Torque}
1037 \label{torque}
1038 The torque $\ve{\tau}(t)$ exerted by the rotation potential is calculated for fixed
1039 axis rotation via
1040 \beq
1041 \ve{\tau}(t) = \sum_{i=1}^{N} \ve{r}_i(t) \times \ve{f}_{\!i}^\perp(t) ,
1042 \label{eqn:torque}
1043 \eeq
1044 where $\ve{r}_i(t)$ is the distance vector from the rotation axis to
1045 $\ve{x}_i(t)$ and $\ve{f}_{\!i}^\perp(t)$ is the force component
1046 perpendicular to $\ve{r}_i(t)$ and $\hat{\ve{v}}$. For flexible axis
1047 rotation, torques $\ve{\tau}_{\!n}$ are calculated for each slab using the
1048 local rotation axis of the slab and the Gaussian-weighted positions.
1051 \section{\normindex{Computational Electrophysiology}}
1052 \label{sec:compel}
1054 The Computational Electrophysiology (CompEL) protocol \cite{Kutzner2011b} allows the simulation of
1055 ion flux through membrane channels, driven by transmembrane potentials or ion
1056 concentration gradients. Just as in real cells, CompEL establishes transmembrane
1057 potentials by sustaining a small imbalance of charges $\Delta q$ across the membrane,
1058 which gives rise to a potential difference $\Delta U$ according to the membrane capacitance:
1059 \beq
1060 \Delta U = \Delta q / C_{membrane}
1061 \eeq
1062 The transmembrane electric field and concentration gradients are controlled by
1063 {\tt .mdp} options, which allow the user to set reference counts for the ions on either side
1064 of the membrane. If a difference between the actual and the reference numbers persists
1065 over a certain time span, specified by the user, a number of ion/water pairs are
1066 exchanged between the compartments until the reference numbers are restored.
1067 Alongside the calculation of channel conductance and ion selectivity, CompEL simulations also
1068 enable determination of the channel reversal potential, an important
1069 characteristic obtained in electrophysiology experiments.
1071 In a CompEL setup, the simulation system is divided into two compartments {\bf A} and {\bf B}
1072 with independent ion concentrations. This is best achieved by using double bilayer systems with
1073 a copy (or copies) of the channel/pore of interest in each bilayer (\figref{compelsetup} A, B).
1074 If the channel axes point in the same direction, channel flux is observed
1075 simultaneously at positive and negative potentials in this way, which is for instance
1076 important for studying channel rectification.
1078 \begin{figure}
1079 \centerline{\includegraphics[width=13.5cm]{plots/compelsetup.pdf}}
1080 \caption{Typical double-membrane setup for CompEL simulations (A, B).
1081 Ion\,/\,water molecule exchanges will be performed as needed
1082 between the two light blue volumes around the dotted black lines (A).
1083 Plot (C) shows the potential difference $\Delta U$ resulting
1084 from the selected charge imbalance $\Delta q_{ref}$ between the compartments.}
1085 \label{fig:compelsetup}
1086 \end{figure}
1088 The potential difference $\Delta U$ across the membrane is easily calculated with the
1089 {\tt gmx potential} utility. By this, the potential drop along $z$ or the
1090 pore axis is exactly known in each time interval of the simulation (\figref{compelsetup} C).
1091 Type and number of ions $n_i$ of charge $q_i$, traversing the channel in the simulation,
1092 are written to the {\tt swapions.xvg} output file, from which the average channel
1093 conductance $G$ in each interval $\Delta t$ is determined by:
1094 \beq
1095 G = \frac{\sum_{i} n_{i}q_{i}}{\Delta t \, \Delta U} \, .
1096 \eeq
1097 The ion selectivity is calculated as the number flux ratio of different species.
1098 Best results are obtained by averaging these values over several overlapping time intervals.
1100 The calculation of reversal potentials is best achieved using a small set of simulations in which a given
1101 transmembrane concentration gradient is complemented with small ion imbalances of varying magnitude. For
1102 example, if one compartment contains 1\,M salt and the other 0.1\,M, and given charge neutrality otherwise,
1103 a set of simulations with $\Delta q = 0\,e$, $\Delta q = 2\,e$, $\Delta q = 4\,e$ could
1104 be used. Fitting a straight line through the current-voltage relationship of all obtained
1105 $I$-$U$ pairs near zero current will then yield $U_{rev}$.
1107 \subsection{Usage}
1108 The following {\tt .mdp} options control the CompEL protocol:
1109 {\small
1110 \begin{verbatim}
1111 swapcoords = Z ; Swap positions: no, X, Y, Z
1112 swap-frequency = 100 ; Swap attempt frequency
1113 \end{verbatim}}
1114 Choose {\tt Z} if your membrane is in the $xy$-plane (\figref{compelsetup}).
1115 Ions will be exchanged between compartments depending on their $z$-positions alone.
1116 {\tt swap-frequency} determines how often a swap attempt will be made.
1117 This step requires that the positions of the split groups, the ions, and possibly the solvent molecules are
1118 communicated between the parallel processes, so if chosen too small it can decrease the simulation
1119 performance. The {\tt Position swapping} entry in the cycle and time accounting
1120 table at the end of the {\tt md.log} file summarizes the amount of runtime spent
1121 in the swap module.
1122 {\small
1123 \begin{verbatim}
1124 split-group0 = channel0 ; Defines compartment boundary
1125 split-group1 = channel1 ; Defines other compartment boundary
1126 massw-split0 = no ; use mass-weighted center?
1127 massw-split1 = no
1128 \end{verbatim}}
1129 {\tt split-group0} and {\tt split-group1} are two index groups that define the boundaries
1130 between the two compartments, which are usually the centers of the channels.
1131 If {\tt massw-split0} or {\tt massw-split1} are set to {\tt yes}, the center of mass
1132 of each index group is used as boundary, here in $z$-direction. Otherwise, the
1133 geometrical centers will be used ($\times$ in \figref{compelsetup} A). If, such as here, a membrane
1134 channel is selected as split group, the center of the channel will define the dividing
1135 plane between the compartments (dashed horizontal lines). All index groups
1136 must be defined in the index file.
1138 If, to restore the requested ion counts, an ion from one compartment has to be exchanged
1139 with a water molecule from the other compartment, then those molecules are swapped
1140 which have the largest distance to the compartment-defining boundaries
1141 (dashed horizontal lines). Depending on the ion concentration,
1142 this effectively results in exchanges of molecules between the light blue volumes.
1143 If a channel is very asymmetric in $z$-direction and would extend into one of the
1144 swap volumes, one can offset the swap exchange plane with the {\tt bulk-offset}
1145 parameter. A value of 0.0 means no offset $b$, values $-1.0 < b < 0$ move the
1146 swap exchange plane closer to the lower, values $0 < b < 1.0$ closer to the upper
1147 membrane. \figref{compelsetup} A (left) depicts that for the {\bf A} compartment.
1149 {\small
1150 \begin{verbatim}
1151 solvent-group = SOL ; Group containing the solvent molecules
1152 iontypes = 3 ; Number of different ion types to control
1153 iontype0-name = NA ; Group name of the ion type
1154 iontype0-in-A = 51 ; Reference count of ions of type 0 in A
1155 iontype0-in-B = 35 ; Reference count of ions of type 0 in B
1156 iontype1-name = K
1157 iontype1-in-A = 10
1158 iontype1-in-B = 38
1159 iontype2-name = CL
1160 iontype2-in-A = -1
1161 iontype2-in-B = -1
1162 \end{verbatim}}
1163 The group name of solvent molecules acting as exchange partners for the ions
1164 has to be set with {\tt solvent-group}. The number of different ionic species under
1165 control of the CompEL protocol is given by the {\tt iontypes} parameter,
1166 while {\tt iontype0-name} gives the name of the index group containing the
1167 atoms of this ionic species. The reference number of ions of this type
1168 can be set with the {\tt iontype0-in-A} and {\tt iontype0-in-B} options
1169 for compartments {\bf A} and {\bf B}, respectively. Obviously,
1170 the sum of {\tt iontype0-in-A} and {\tt iontype0-in-B} needs to equal the number
1171 of ions in the group defined by {\tt iontype0-name}.
1172 A reference number of {\tt -1} means: use the number of ions as found at the beginning
1173 of the simulation as the reference value.
1175 {\small
1176 \begin{verbatim}
1177 coupl-steps = 10 ; Average over these many swap steps
1178 threshold = 1 ; Do not swap if < threshold
1179 \end{verbatim}}
1180 If {\tt coupl-steps} is set to 1, then the momentary ion distribution determines
1181 whether ions are exchanged. {\tt coupl-steps} \textgreater\ 1 will use the time-average
1182 of ion distributions over the selected number of attempt steps instead. This can be useful, for example,
1183 when ions diffuse near compartment boundaries, which would lead to numerous unproductive
1184 ion exchanges. A {\tt threshold} of 1 means that a swap is performed if the average ion
1185 count in a compartment differs by at least 1 from the requested values. Higher thresholds
1186 will lead to toleration of larger differences. Ions are exchanged until the requested
1187 number $\pm$ the threshold is reached.
1189 {\small
1190 \begin{verbatim}
1191 cyl0-r = 5.0 ; Split cylinder 0 radius (nm)
1192 cyl0-up = 0.75 ; Split cylinder 0 upper extension (nm)
1193 cyl0-down = 0.75 ; Split cylinder 0 lower extension (nm)
1194 cyl1-r = 5.0 ; same for other channel
1195 cyl1-up = 0.75
1196 cyl1-down = 0.75
1197 \end{verbatim}}
1198 The cylinder options are used to define virtual geometric cylinders around the
1199 channel's pore to track how many ions of which type have passed each channel.
1200 Ions will be counted as having traveled through a channel
1201 according to the definition of the channel's cylinder radius, upper and lower extension,
1202 relative to the location of the respective split group. This will not affect the actual
1203 flux or exchange, but will provide you with the ion permeation numbers across
1204 each of the channels. Note that an ion can only be counted as passing through a particular
1205 channel if it is detected \emph{within} the defined split cylinder in a swap step.
1206 If {\tt swap-frequency} is chosen too high, a particular ion may be detected in compartment {\bf A}
1207 in one swap step, and in compartment {\bf B} in the following swap step, so it will be unclear
1208 through which of the channels it has passed.
1210 A double-layered system for CompEL simulations can be easily prepared by
1211 duplicating an existing membrane/channel MD system in the direction of the membrane
1212 normal (typically $z$) with {\tt gmx editconf -translate 0 0 <l_z>}, where {\tt l_z}
1213 is the box length in that direction. If you have already defined index groups for
1214 the channel for the single-layered system, {\tt gmx make_ndx -n index.ndx -twin} will
1215 provide you with the groups for the double-layered system.
1217 To suppress large fluctuations of the membranes along the swap direction,
1218 it may be useful to apply a harmonic potential (acting only in the swap dimension)
1219 between each of the two channel and/or bilayer centers using umbrella pulling
1220 (see section~\ref{sec:pull}).
1222 \subsection*{Multimeric channels}
1223 If a split group consists of more than one molecule, the correct PBC image of all molecules
1224 with respect to each other has to be chosen such that the channel center can be correctly
1225 determined. \gromacs\ assumes that the starting structure in the {\tt .tpr}
1226 file has the correct PBC representation. Set the following environment variable
1227 to check whether that is the case:
1228 \begin{itemize}
1229 \item {\tt GMX_COMPELDUMP}: output the starting structure after it has been made whole to
1230 {\tt .pdb} file.
1231 \end{itemize}
1234 \section{Calculating a PMF using the free-energy code}
1235 \label{sec:fepmf}
1236 \index{potentials of mean force}
1237 \index{free energy calculations}
1238 The free-energy coupling-parameter approach (see~\secref{fecalc})
1239 provides several ways to calculate potentials of mean force.
1240 A potential of mean force between two atoms can be calculated
1241 by connecting them with a harmonic potential or a constraint.
1242 For this purpose there are special potentials that avoid the generation of
1243 extra exclusions, see~\secref{excl}.
1244 When the position of the minimum or the constraint length is 1 nm more
1245 in state B than in state A, the restraint or constraint force is given
1246 by $\partial H/\partial \lambda$.
1247 The distance between the atoms can be changed as a function of $\lambda$
1248 and time by setting {\tt delta-lambda} in the {\tt .mdp} file.
1249 The results should be identical (although not numerically
1250 due to the different implementations) to the results of the pull code
1251 with umbrella sampling and constraint pulling.
1252 Unlike the pull code, the free energy code can also handle atoms that
1253 are connected by constraints.
1255 Potentials of mean force can also be calculated using position restraints.
1256 With position restraints, atoms can be linked to a position in space
1257 with a harmonic potential (see \ssecref{positionrestraint}).
1258 These positions can be made a function of the coupling parameter $\lambda$.
1259 The positions for the A and the B states are supplied to {\tt grompp} with
1260 the {\tt -r} and {\tt -rb} options, respectively.
1261 One could use this approach to do \normindex{targeted MD};
1262 note that we do not encourage the use of targeted MD for proteins.
1263 A protein can be forced from one conformation to another by using
1264 these conformations as position restraint coordinates for state A and B.
1265 One can then slowly change $\lambda$ from 0 to 1.
1266 The main drawback of this approach is that the conformational freedom
1267 of the protein is severely limited by the position restraints,
1268 independent of the change from state A to B.
1269 Also, the protein is forced from state A to B in an almost straight line,
1270 whereas the real pathway might be very different.
1271 An example of a more fruitful application is a solid system or a liquid
1272 confined between walls where one wants to measure the force required
1273 to change the separation between the boundaries or walls.
1274 Because the boundaries (or walls) already need to be fixed,
1275 the position restraints do not limit the system in its sampling.
1277 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1278 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1279 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1280 \newcommand{\amine}{\sf -NH$_2$}
1281 \newcommand{\amines}{\sf -NH-}
1282 \newcommand{\aminep}{\sf -NH$_3^+$}
1283 \section{Removing fastest \swapindex{degrees of}{freedom}}
1284 \label{sec:rmfast}
1285 The maximum time step in MD simulations is limited by the smallest
1286 oscillation period that can be found in the simulated
1287 system. Bond-stretching vibrations are in their quantum-mechanical
1288 ground state and are therefore better represented by a constraint
1289 instead of a harmonic potential.
1291 For the remaining degrees of freedom, the shortest oscillation period
1292 (as measured from a simulation) is 13~fs for bond-angle vibrations
1293 involving hydrogen atoms. Taking as a guideline that with a Verlet
1294 (leap-frog) integration scheme a minimum of 5 numerical integration
1295 steps should be performed per period of a harmonic oscillation in
1296 order to integrate it with reasonable accuracy, the maximum time step
1297 will be about 3~fs. Disregarding these very fast oscillations of
1298 period 13~fs, the next shortest periods are around 20~fs, which will
1299 allow a maximum time step of about 4~fs.
1301 Removing the bond-angle degrees of freedom from hydrogen atoms can
1302 best be done by defining them as \normindex{virtual interaction sites}
1303 instead of normal atoms. Whereas a normal atom is connected to the molecule
1304 with bonds, angles and dihedrals, a virtual site's position is calculated
1305 from the position of three nearby heavy atoms in a predefined manner
1306 (see also \secref{virtual_sites}). For the hydrogens in water and in
1307 hydroxyl, sulfhydryl, or amine groups, no degrees of freedom can be
1308 removed, because rotational freedom should be preserved. The only
1309 other option available to slow down these motions is to increase the
1310 mass of the hydrogen atoms at the expense of the mass of the connected
1311 heavy atom. This will increase the moment of inertia of the water
1312 molecules and the hydroxyl, sulfhydryl, or amine groups, without
1313 affecting the equilibrium properties of the system and without
1314 affecting the dynamical properties too much. These constructions will
1315 shortly be described in \secref{vsitehydro} and have previously
1316 been described in full detail~\cite{feenstra99}.
1318 Using both virtual sites and \swapindex{modified}{mass}es, the next
1319 bottleneck is likely to be formed by the improper dihedrals (which are
1320 used to preserve planarity or chirality of molecular groups) and the
1321 peptide dihedrals. The peptide dihedral cannot be changed without
1322 affecting the physical behavior of the protein. The improper dihedrals
1323 that preserve planarity mostly deal with aromatic residues. Bonds,
1324 angles, and dihedrals in these residues can also be replaced with
1325 somewhat elaborate virtual site constructions.
1327 All modifications described in this section can be performed using the
1328 {\gromacs} topology building tool {\tt \normindex{pdb2gmx}}. Separate
1329 options exist to increase hydrogen masses, virtualize all hydrogen atoms,
1330 or also virtualize all aromatic residues. {\bf Note} that when all hydrogen
1331 atoms are virtualized, those inside the aromatic residues will be
1332 virtualized as well, {\ie} hydrogens in the aromatic residues are treated
1333 differently depending on the treatment of the aromatic residues.
1335 Parameters for the virtual site constructions for the hydrogen atoms are
1336 inferred from the force-field parameters ({\em vis}. bond lengths and
1337 angles) directly by {\tt \normindex{grompp}} while processing the
1338 topology file. The constructions for the aromatic residues are based
1339 on the bond lengths and angles for the geometry as described in the
1340 force fields, but these parameters are hard-coded into {\tt
1341 \normindex{pdb2gmx}} due to the complex nature of the construction
1342 needed for a whole aromatic group.
1344 \subsection{Hydrogen bond-angle vibrations}
1345 \label{sec:vsitehydro}
1346 \subsubsection{Construction of virtual sites} %%%%%%%%%%%%%%%%%%%%%%%%%
1347 \begin{figure}
1348 \centerline{\includegraphics[width=11cm]{plots/dumtypes}}
1349 \caption[Virtual site constructions for hydrogen atoms.]{The different
1350 types of virtual site constructions used for hydrogen atoms. The atoms
1351 used in the construction of the virtual site(s) are depicted as black
1352 circles, virtual sites as gray ones. Hydrogens are smaller than heavy
1353 atoms. {\sf A}: fixed bond angle, note that here the hydrogen is not a
1354 virtual site; {\sf B}: in the plane of three atoms, with fixed distance;
1355 {\sf C}: in the plane of three atoms, with fixed angle and distance;
1356 {\sf D}: construction for amine groups ({\amine} or {\aminep}), see
1357 text for details.}
1358 \label{fig:vsitehydro}
1359 \end{figure}
1361 The goal of defining hydrogen atoms as virtual sites is to remove all
1362 high-frequency degrees of freedom from them. In some cases, not all
1363 degrees of freedom of a hydrogen atom should be removed, {\eg} in the
1364 case of hydroxyl or amine groups the rotational freedom of the
1365 hydrogen atom(s) should be preserved. Care should be taken that no
1366 unwanted correlations are introduced by the construction of virtual
1367 sites, {\eg} bond-angle vibration between the constructing atoms could
1368 translate into hydrogen bond-length vibration. Additionally, since
1369 virtual sites are by definition massless, in order to preserve total
1370 system mass, the mass of each hydrogen atom that is treated as virtual
1371 site should be added to the bonded heavy atom.
1373 Taking into account these considerations, the hydrogen atoms in a
1374 protein naturally fall into several categories, each requiring a
1375 different approach (see also \figref{vsitehydro}).
1377 \begin{itemize}
1379 \item{\em hydroxyl ({\sf -OH}) or sulfhydryl ({\sf -SH})
1380 hydrogen:\/} The only internal degree of freedom in a hydroxyl group
1381 that can be constrained is the bending of the {\sf C-O-H} angle. This
1382 angle is fixed by defining an additional bond of appropriate length,
1383 see \figref{vsitehydro}A. Doing so removes the high-frequency angle bending,
1384 but leaves the dihedral rotational freedom. The same goes for a
1385 sulfhydryl group. {\bf Note} that in these cases the hydrogen is not treated
1386 as a virtual site.
1388 \item{\em single amine or amide ({\amines}) and aromatic hydrogens
1389 ({\sf -CH-}):\/} The position of these hydrogens cannot be constructed
1390 from a linear combination of bond vectors, because of the flexibility
1391 of the angle between the heavy atoms. Instead, the hydrogen atom is
1392 positioned at a fixed distance from the bonded heavy atom on a line
1393 going through the bonded heavy atom and a point on the line through
1394 both second bonded atoms, see \figref{vsitehydro}B.
1396 \item{\em planar amine ({\amine}) hydrogens:\/} The method used for
1397 the single amide hydrogen is not well suited for planar amine groups,
1398 because no suitable two heavy atoms can be found to define the
1399 direction of the hydrogen atoms. Instead, the hydrogen is constructed
1400 at a fixed distance from the nitrogen atom, with a fixed angle to the
1401 carbon atom, in the plane defined by one of the other heavy atoms, see
1402 \figref{vsitehydro}C.
1404 \item{\em amine group (umbrella {\amine} or {\aminep}) hydrogens:\/}
1405 Amine hydrogens with rotational freedom cannot be constructed as virtual
1406 sites from the heavy atoms they are connected to, since this would
1407 result in loss of the rotational freedom of the amine group. To
1408 preserve the rotational freedom while removing the hydrogen bond-angle
1409 degrees of freedom, two ``dummy masses'' are constructed with the same
1410 total mass, moment of inertia (for rotation around the {\sf C-N} bond)
1411 and center of mass as the amine group. These dummy masses have no
1412 interaction with any other atom, except for the fact that they are
1413 connected to the carbon and to each other, resulting in a rigid
1414 triangle. From these three particles, the positions of the nitrogen and
1415 hydrogen atoms are constructed as linear combinations of the two
1416 carbon-mass vectors and their outer product, resulting in an amine
1417 group with rotational freedom intact, but without other internal
1418 degrees of freedom. See \figref{vsitehydro}D.
1420 \end{itemize}
1422 \begin{figure}
1423 \centerline{\includegraphics[width=15cm]{plots/dumaro}}
1424 \caption[Virtual site constructions for aromatic residues.]{The
1425 different types of virtual site constructions used for aromatic
1426 residues. The atoms used in the construction of the virtual site(s) are
1427 depicted as black circles, virtual sites as gray ones. Hydrogens are
1428 smaller than heavy atoms. {\sf A}: phenylalanine; {\sf B}: tyrosine
1429 (note that the hydroxyl hydrogen is {\em not} a virtual site); {\sf C}:
1430 tryptophan; {\sf D}: histidine.}
1431 \label{fig:vistearo}
1432 \end{figure}
1434 \subsection{Out-of-plane vibrations in aromatic groups}
1435 \label{sec:vsitearo}
1436 The planar arrangements in the side chains of the aromatic residues
1437 lends itself perfectly to a virtual-site construction, giving a
1438 perfectly planar group without the inherently unstable constraints
1439 that are necessary to keep normal atoms in a plane. The basic approach
1440 is to define three atoms or dummy masses with constraints between them
1441 to fix the geometry and create the rest of the atoms as simple virtual
1442 sites type (see \secref{virtual_sites}) from these three. Each of
1443 the aromatic residues require a different approach:
1445 \begin{itemize}
1447 \item{\em Phenylalanine:\/} {\sf C}$_\gamma$, {\sf C}$_{{\epsilon}1}$,
1448 and {\sf C}$_{{\epsilon}2}$ are kept as normal atoms, but with each a
1449 mass of one third the total mass of the phenyl group. See
1450 \figref{vsitehydro}A.
1452 \item{\em Tyrosine:\/} The ring is treated identically to the
1453 phenylalanine ring. Additionally, constraints are defined between {\sf
1454 C}$_{{\epsilon}1}$, {\sf C}$_{{\epsilon}2}$, and {\sf O}$_{\eta}$.
1455 The original improper dihedral angles will keep both triangles (one
1456 for the ring and one with {\sf O}$_{\eta}$) in a plane, but due to the
1457 larger moments of inertia this construction will be much more
1458 stable. The bond-angle in the hydroxyl group will be constrained by a
1459 constraint between {\sf C}$_\gamma$ and {\sf H}$_{\eta}$. {\bf Note} that
1460 the hydrogen is not treated as a virtual site. See
1461 \figref{vsitehydro}B.
1463 \item{\em Tryptophan:\/} {\sf C}$_\beta$ is kept as a normal atom
1464 and two dummy masses are created at the center of mass of each of the
1465 rings, each with a mass equal to the total mass of the respective ring
1466 ({\sf C}$_{{\delta}2}$ and {\sf C}$_{{\epsilon}2}$ are each
1467 counted half for each ring). This keeps the overall center of mass and
1468 the moment of inertia almost (but not quite) equal to what it was. See
1469 \figref{vsitehydro}C.
1471 \item{\em Histidine:\/} {\sf C}$_\gamma$, {\sf C}$_{{\epsilon}1}$
1472 and {\sf N}$_{{\epsilon}2}$ are kept as normal atoms, but with masses
1473 redistributed such that the center of mass of the ring is
1474 preserved. See \figref{vsitehydro}D.
1476 \end{itemize}
1478 \section{Viscosity calculation\index{viscosity}}
1480 The shear viscosity is a property of liquids that can be determined easily
1481 by experiment. It is useful for parameterizing a force field
1482 because it is a kinetic property, while most other properties
1483 which are used for parameterization are thermodynamic.
1484 The viscosity is also an important property, since it influences
1485 the rates of conformational changes of molecules solvated in the liquid.
1487 The viscosity can be calculated from an equilibrium simulation using
1488 an Einstein relation:
1489 \beq
1490 \eta = \frac{1}{2}\frac{V}{k_B T} \lim_{t \rightarrow \infty}
1491 \frac{\mbox{d}}{\mbox{d} t} \left\langle
1492 \left( \int_{t_0}^{{t_0}+t} P_{xz}(t') \mbox{d} t' \right)^2
1493 \right\rangle_{t_0}
1494 \eeq
1495 This can be done with {\tt g_energy}.
1496 This method converges very slowly~\cite{Hess2002a}, and as such
1497 a nanosecond simulation might not
1498 be long enough for an accurate determination of the viscosity.
1499 The result is very dependent on the treatment of the electrostatics.
1500 Using a (short) cut-off results in large noise on the off-diagonal
1501 pressure elements, which can increase the calculated viscosity by an order
1502 of magnitude.
1504 {\gromacs} also has a non-equilibrium method for determining
1505 the viscosity~\cite{Hess2002a}.
1506 This makes use of the fact that energy, which is fed into system by
1507 external forces, is dissipated through viscous friction. The generated heat
1508 is removed by coupling to a heat bath. For a Newtonian liquid adding a
1509 small force will result in a velocity gradient according to the following
1510 equation:
1511 \beq
1512 a_x(z) + \frac{\eta}{\rho} \frac{\partial^2 v_x(z)}{\partial z^2} = 0
1513 \eeq
1514 Here we have applied an acceleration $a_x(z)$ in the $x$-direction, which
1515 is a function of the $z$-coordinate.
1516 In {\gromacs} the acceleration profile is:
1517 \beq
1518 a_x(z) = A \cos\left(\frac{2\pi z}{l_z}\right)
1519 \eeq
1520 where $l_z$ is the height of the box. The generated velocity profile is:
1521 \beq
1522 v_x(z) = V \cos\left(\frac{2\pi z}{l_z}\right)
1523 \eeq
1524 \beq
1525 V = A \frac{\rho}{\eta}\left(\frac{l_z}{2\pi}\right)^2
1526 \eeq
1527 The viscosity can be calculated from $A$ and $V$:
1528 \beq
1529 \label{visc}
1530 \eta = \frac{A}{V}\rho \left(\frac{l_z}{2\pi}\right)^2
1531 \eeq
1533 In the simulation $V$ is defined as:
1534 \beq
1535 V = \frac{\displaystyle \sum_{i=1}^N m_i v_{i,x} 2 \cos\left(\frac{2\pi z}{l_z}\right)}
1536 {\displaystyle \sum_{i=1}^N m_i}
1537 \eeq
1538 The generated velocity profile is not coupled to the heat bath. Moreover,
1539 the velocity profile is excluded from the kinetic energy.
1540 One would like $V$ to be as large as possible to get good statistics.
1541 However, the shear rate should not be so high that the system gets too far
1542 from equilibrium. The maximum shear rate occurs where the cosine is zero,
1543 the rate being:
1544 \beq
1545 \mbox{sh}_{\max} = \max_z \left| \frac{\partial v_x(z)}{\partial z} \right|
1546 = A \frac{\rho}{\eta} \frac{l_z}{2\pi}
1547 \eeq
1548 For a simulation with: $\eta=10^{-3}$ [kg\,m$^{-1}$\,s$^{-1}$],
1549 $\rho=10^3$\,[kg\,m$^{-3}$] and $l_z=2\pi$\,[nm],
1550 $\mbox{sh}_{\max}=1$\,[ps\,nm$^{-1}$] $A$.
1551 This shear rate should be smaller than one over the longest
1552 correlation time in the system. For most liquids, this will be the rotation
1553 correlation time, which is around 10 ps. In this case, $A$ should
1554 be smaller than 0.1\,[nm\,ps$^{-2}$].
1555 When the shear rate is too high, the observed viscosity will be too low.
1556 Because $V$ is proportional to the square of the box height,
1557 the optimal box is elongated in the $z$-direction.
1558 In general, a simulation length of 100 ps is enough to obtain an
1559 accurate value for the viscosity.
1561 The heat generated by the viscous friction is removed by coupling to a heat
1562 bath. Because this coupling is not instantaneous the real temperature of the
1563 liquid will be slightly lower than the observed temperature.
1564 Berendsen derived this temperature shift~\cite{Berendsen91}, which can
1565 be written in terms of the shear rate as:
1566 \beq
1567 T_s = \frac{\eta\,\tau}{2 \rho\,C_v} \mbox{sh}_{\max}^2
1568 \eeq
1569 where $\tau$ is the coupling time for the Berendsen thermostat and
1570 $C_v$ is the heat capacity. Using the values of the example above,
1571 $\tau=10^{-13}$ [s] and $C_v=2 \cdot 10^3$\,[J kg$^{-1}$\,K$^{-1}$], we
1572 get: $T_s=25$\,[K\,ps$^{-2}$]\,sh$_{\max}^2$. When we want the shear
1573 rate to be smaller than $1/10$\,[ps$^{-1}$], $T_s$ is smaller than
1574 0.25\,[K], which is negligible.
1576 {\bf Note} that the system has to build up the velocity profile when starting
1577 from an equilibrium state. This build-up time is of the order of the
1578 correlation time of the liquid.
1580 Two quantities are written to the energy file, along with their averages
1581 and fluctuations: $V$ and $1/\eta$, as obtained from (\ref{visc}).
1583 \section{Tabulated interaction functions\index{tabulated interaction functions}}
1584 \subsection{Cubic splines for potentials}
1585 \label{subsec:cubicspline}
1586 In some of the inner loops of {\gromacs}, look-up tables are used
1587 for computation of potential and forces.
1588 The tables are interpolated using a cubic
1589 spline algorithm.
1590 There are separate tables for electrostatic, dispersion, and repulsion
1591 interactions,
1592 but for the sake of caching performance these have been combined
1593 into a single array.
1594 The cubic spline interpolation for $x_i \leq x < x_{i+1}$ looks like this:
1595 \beq
1596 V_s(x) = A_0 + A_1 \,\epsilon + A_2 \,\epsilon^2 + A_3 \,\epsilon^3
1597 \label{eqn:spline}
1598 \eeq
1599 where the table spacing $h$ and fraction $\epsilon$ are given by:
1600 \bea
1601 h &=& x_{i+1} - x_i \\
1602 \epsilon&=& (x - x_i)/h
1603 \eea
1604 so that $0 \le \epsilon < 1$.
1605 From this, we can calculate the derivative in order to determine the forces:
1606 \beq
1607 -V_s'(x) ~=~
1608 -\frac{{\rm d}V_s(x)}{{\rm d}\epsilon}\frac{{\rm d}\epsilon}{{\rm d}x} ~=~
1609 -(A_1 + 2 A_2 \,\epsilon + 3 A_3 \,\epsilon^2)/h
1610 \eeq
1611 The four coefficients are determined from the four conditions
1612 that $V_s$ and $-V_s'$ at both ends of each interval should match
1613 the exact potential $V$ and force $-V'$.
1614 This results in the following errors for each interval:
1615 \bea
1616 |V_s - V |_{max} &=& V'''' \frac{h^4}{384} + O(h^5) \\
1617 |V_s' - V' |_{max} &=& V'''' \frac{h^3}{72\sqrt{3}} + O(h^4) \\
1618 |V_s''- V''|_{max} &=& V'''' \frac{h^2}{12} + O(h^3)
1619 \eea
1620 V and V' are continuous, while V'' is the first discontinuous
1621 derivative.
1622 The number of points per nanometer is 500 and 2000
1623 for mixed- and double-precision versions of {\gromacs}, respectively.
1624 This means that the errors in the potential and force will usually
1625 be smaller than the mixed precision accuracy.
1627 {\gromacs} stores $A_0$, $A_1$, $A_2$ and $A_3$.
1628 The force routines get a table with these four parameters and
1629 a scaling factor $s$ that is equal to the number of points per nm.
1630 ({\bf Note} that $h$ is $s^{-1}$).
1631 The algorithm goes a little something like this:
1632 \begin{enumerate}
1633 \item Calculate distance vector (\ve{r}$_{ij}$) and distance r$_{ij}$
1634 \item Multiply r$_{ij}$ by $s$ and truncate to an integer value $n_0$
1635 to get a table index
1636 \item Calculate fractional component ($\epsilon$ = $s$r$_{ij} - n_0$)
1637 and $\epsilon^2$
1638 \item Do the interpolation to calculate the potential $V$ and the scalar force $f$
1639 \item Calculate the vector force \ve{F} by multiplying $f$ with \ve{r}$_{ij}$
1640 \end{enumerate}
1642 {\bf Note} that table look-up is significantly {\em
1643 slower} than computation of the most simple Lennard-Jones and Coulomb
1644 interaction. However, it is much faster than the shifted Coulomb
1645 function used in conjunction with the PPPM method. Finally, it is much
1646 easier to modify a table for the potential (and get a graphical
1647 representation of it) than to modify the inner loops of the MD
1648 program.
1650 \subsection{User-specified potential functions}
1651 \label{subsec:userpot}
1652 You can also use your own potential functions\index{potential function} without
1653 editing the {\gromacs} code. The potential function should be according to the
1654 following equation
1655 \beq
1656 V(r_{ij}) ~=~ \frac{q_i q_j}{4 \pi\epsilon_0} f(r_{ij}) + C_6 \,g(r_{ij}) + C_{12} \,h(r_{ij})
1657 \eeq
1658 where $f$, $g$, and $h$ are user defined functions. {\bf Note} that if $g(r)$ represents a
1659 normal dispersion interaction, $g(r)$ should be $<$ 0. C$_6$, C$_{12}$
1660 and the charges are read from the topology. Also note that combination
1661 rules are only supported for Lennard-Jones and Buckingham, and that
1662 your tables should match the parameters in the binary topology.
1664 When you add the following lines in your {\tt .mdp} file:
1666 {\small
1667 \begin{verbatim}
1668 rlist = 1.0
1669 coulombtype = User
1670 rcoulomb = 1.0
1671 vdwtype = User
1672 rvdw = 1.0
1673 \end{verbatim}}
1675 {\tt mdrun} will read a single non-bonded table file,
1676 or multiple when {\tt energygrp-table} is set (see below).
1677 The name of the file(s) can be set with the {\tt mdrun} option {\tt -table}.
1678 The table file should contain seven columns of table look-up data in the
1679 order: $x$, $f(x)$, $-f'(x)$, $g(x)$, $-g'(x)$, $h(x)$, $-h'(x)$.
1680 The $x$ should run from 0 to $r_c+1$ (the value of {\tt table_extension} can be
1681 changed in the {\tt .mdp} file).
1682 You can choose the spacing you like; for the standard tables {\gromacs}
1683 uses a spacing of 0.002 and 0.0005 nm when you run in mixed
1684 and double precision, respectively. In this
1685 context, $r_c$ denotes the maximum of the two cut-offs {\tt rvdw} and
1686 {\tt rcoulomb} (see above). These variables need not be the same (and
1687 need not be 1.0 either). Some functions used for potentials contain a
1688 singularity at $x = 0$, but since atoms are normally not closer to each
1689 other than 0.1 nm, the function value at $x = 0$ is not important.
1690 Finally, it is also
1691 possible to combine a standard Coulomb with a modified LJ potential
1692 (or vice versa). One then specifies {\eg} {\tt coulombtype = Cut-off} or
1693 {\tt coulombtype = PME}, combined with {\tt vdwtype = User}. The table file must
1694 always contain the 7 columns however, and meaningful data (i.e. not
1695 zeroes) must be entered in all columns. A number of pre-built table
1696 files can be found in the {\tt GMXLIB} directory for 6-8, 6-9, 6-10, 6-11, and 6-12
1697 Lennard-Jones potentials combined with a normal Coulomb.
1699 If you want to have different functional forms between different
1700 groups of atoms, this can be set through energy groups.
1701 Different tables can be used for non-bonded interactions between
1702 different energy groups pairs through the {\tt .mdp} option {\tt energygrp-table}
1703 (see details in the User Guide).
1704 Atoms that should interact with a different potential should
1705 be put into different energy groups.
1706 Between group pairs which are not listed in {\tt energygrp-table},
1707 the normal user tables will be used. This makes it easy to use
1708 a different functional form between a few types of atoms.
1710 \section{Mixed Quantum-Classical simulation techniques}
1712 In a molecular mechanics (MM) force field, the influence of electrons
1713 is expressed by empirical parameters that are assigned on the basis of
1714 experimental data, or on the basis of results from high-level quantum
1715 chemistry calculations. These are valid for the ground state of a
1716 given covalent structure, and the MM approximation is usually
1717 sufficiently accurate for ground-state processes in which the overall
1718 connectivity between the atoms in the system remains
1719 unchanged. However, for processes in which the connectivity does
1720 change, such as chemical reactions, or processes that involve multiple
1721 electronic states, such as photochemical conversions, electrons can no
1722 longer be ignored, and a quantum mechanical description is required
1723 for at least those parts of the system in which the reaction takes
1724 place.
1726 One approach to the simulation of chemical reactions in solution, or
1727 in enzymes, is to use a combination of quantum mechanics (QM) and
1728 molecular mechanics (MM). The reacting parts of the system are treated
1729 quantum mechanically, with the remainder being modeled using the
1730 force field. The current version of {\gromacs} provides interfaces to
1731 several popular Quantum Chemistry packages (MOPAC~\cite{mopac},
1732 GAMESS-UK~\cite{gamess-uk}, Gaussian~\cite{g03} and CPMD~\cite{Car85a}).
1734 {\gromacs} interactions between the two subsystems are
1735 either handled as described by Field {\em et al.}~\cite{Field90a} or
1736 within the ONIOM approach by Morokuma and coworkers~\cite{Maseras96a,
1737 Svensson96a}.
1739 \subsection{Overview}
1741 Two approaches for describing the interactions between the QM and MM
1742 subsystems are supported in this version:
1744 \begin{enumerate}
1745 \item{\textbf{Electronic Embedding}} The electrostatic interactions
1746 between the electrons of the QM region and the MM atoms and between
1747 the QM nuclei and the MM atoms are included in the Hamiltonian for
1748 the QM subsystem: \beq H^{QM/MM} =
1749 H^{QM}_e-\sum_i^n\sum_J^M\frac{e^2Q_J}{4\pi\epsilon_0r_{iJ}}+\sum_A^N\sum_J^M\frac{e^2Z_AQ_J}{e\pi\epsilon_0R_{AJ}},
1750 \eeq where $n$ and $N$ are the number of electrons and nuclei in the
1751 QM region, respectively, and $M$ is the number of charged MM
1752 atoms. The first term on the right hand side is the original
1753 electronic Hamiltonian of an isolated QM system. The first of the
1754 double sums is the total electrostatic interaction between the QM
1755 electrons and the MM atoms. The total electrostatic interaction of the
1756 QM nuclei with the MM atoms is given by the second double sum. Bonded
1757 interactions between QM and MM atoms are described at the MM level by
1758 the appropriate force-field terms. Chemical bonds that connect the two
1759 subsystems are capped by a hydrogen atom to complete the valence of
1760 the QM region. The force on this atom, which is present in the QM
1761 region only, is distributed over the two atoms of the bond. The cap
1762 atom is usually referred to as a link atom.
1764 \item{\textbf{ONIOM}} In the ONIOM approach, the energy and gradients
1765 are first evaluated for the isolated QM subsystem at the desired level
1766 of {\it{ab initio}} theory. Subsequently, the energy and gradients of
1767 the total system, including the QM region, are computed using the
1768 molecular mechanics force field and added to the energy and gradients
1769 calculated for the isolated QM subsystem. Finally, in order to correct
1770 for counting the interactions inside the QM region twice, a molecular
1771 mechanics calculation is performed on the isolated QM subsystem and
1772 the energy and gradients are subtracted. This leads to the following
1773 expression for the total QM/MM energy (and gradients likewise): \beq
1774 E_{tot} = E_{I}^{QM}
1775 +E_{I+II}^{MM}-E_{I}^{MM}, \eeq where the
1776 subscripts I and II refer to the QM and MM subsystems,
1777 respectively. The superscripts indicate at what level of theory the
1778 energies are computed. The ONIOM scheme has the
1779 advantage that it is not restricted to a two-layer QM/MM description,
1780 but can easily handle more than two layers, with each layer described
1781 at a different level of theory.
1782 \end{enumerate}
1784 \subsection{Usage}
1786 To make use of the QM/MM functionality in {\gromacs}, one needs to:
1788 \begin{enumerate}
1789 \item introduce link atoms at the QM/MM boundary, if needed;
1790 \item specify which atoms are to be treated at a QM level;
1791 \item specify the QM level, basis set, type of QM/MM interface and so on.
1792 \end{enumerate}
1794 \subsubsection{Adding link atoms}
1796 At the bond that connects the QM and MM subsystems, a link atoms is
1797 introduced. In {\gromacs} the link atom has special atomtype, called
1798 LA. This atomtype is treated as a hydrogen atom in the QM calculation,
1799 and as a virtual site in the force-field calculation. The link atoms, if
1800 any, are part of the system, but have no interaction with any other
1801 atom, except that the QM force working on it is distributed over the
1802 two atoms of the bond. In the topology, the link atom (LA), therefore,
1803 is defined as a virtual site atom:
1805 {\small
1806 \begin{verbatim}
1807 [ virtual_sites2 ]
1808 LA QMatom MMatom 1 0.65
1809 \end{verbatim}}
1811 See~\secref{vsitetop} for more details on how virtual sites are
1812 treated. The link atom is replaced at every step of the simulation.
1814 In addition, the bond itself is replaced by a constraint:
1816 {\small
1817 \begin{verbatim}
1818 [ constraints ]
1819 QMatom MMatom 2 0.153
1820 \end{verbatim}}
1822 {\bf Note} that, because in our system the QM/MM bond is a carbon-carbon
1823 bond (0.153 nm), we use a constraint length of 0.153 nm, and dummy
1824 position of 0.65. The latter is the ratio between the ideal C-H
1825 bond length and the ideal C-C bond length. With this ratio, the link
1826 atom is always 0.1 nm away from the {\tt QMatom}, consistent with the
1827 carbon-hydrogen bond length. If the QM and MM subsystems are connected
1828 by a different kind of bond, a different constraint and a different
1829 dummy position, appropriate for that bond type, are required.
1831 \subsubsection{Specifying the QM atoms}
1833 Atoms that should be treated at a QM level of theory, including the
1834 link atoms, are added to the index file. In addition, the chemical
1835 bonds between the atoms in the QM region are to be defined as
1836 connect bonds (bond type 5) in the topology file:
1838 {\small
1839 \begin{verbatim}
1840 [ bonds ]
1841 QMatom1 QMatom2 5
1842 QMatom2 QMatom3 5
1843 \end{verbatim}}
1845 \subsubsection{Specifying the QM/MM simulation parameters}
1847 In the {\tt .mdp} file, the following parameters control a QM/MM simulation.
1849 \begin{description}
1851 \item[\tt QMMM = no]\mbox{}\\ If this is set to {\tt yes}, a QM/MM
1852 simulation is requested. Several groups of atoms can be described at
1853 different QM levels separately. These are specified in the QMMM-grps
1854 field separated by spaces. The level of {\it{ab initio}} theory at which the
1855 groups are described is specified by {\tt QMmethod} and {\tt QMbasis}
1856 Fields. Describing the groups at different levels of theory is only
1857 possible with the ONIOM QM/MM scheme, specified by {\tt QMMMscheme}.
1859 \item[\tt QMMM-grps =]\mbox{}\\groups to be described at the QM level
1861 \item[\tt QMMMscheme = normal]\mbox{}\\Options are {\tt normal} and
1862 {\tt ONIOM}. This selects the QM/MM interface. {\tt normal} implies
1863 that the QM subsystem is electronically embedded in the MM
1864 subsystem. There can only be one {\tt QMMM-grps} that is modeled at
1865 the {\tt QMmethod} and {\tt QMbasis} level of {\it{ ab initio}}
1866 theory. The rest of the system is described at the MM level. The QM
1867 and MM subsystems interact as follows: MM point charges are included
1868 in the QM one-electron Hamiltonian and all Lennard-Jones interactions
1869 are described at the MM level. If {\tt ONIOM} is selected, the
1870 interaction between the subsystem is described using the ONIOM method
1871 by Morokuma and co-workers. There can be more than one QMMM-grps each
1872 modeled at a different level of QM theory (QMmethod and QMbasis).
1874 \item[\tt QMmethod = ]\mbox{}\\Method used to compute the energy
1875 and gradients on the QM atoms. Available methods are AM1, PM3, RHF,
1876 UHF, DFT, B3LYP, MP2, CASSCF, MMVB and CPMD. For CASSCF, the number of
1877 electrons and orbitals included in the active space is specified by
1878 {\tt CASelectrons} and {\tt CASorbitals}. For CPMD, the plane-wave
1879 cut-off is specified by the {\tt planewavecutoff} keyword.
1881 \item[\tt QMbasis = ]\mbox{}\\Gaussian basis set used to expand the
1882 electronic wave-function. Only Gaussian basis sets are currently
1883 available, i.e. STO-3G, 3-21G, 3-21G*, 3-21+G*, 6-21G, 6-31G, 6-31G*,
1884 6-31+G*, and 6-311G. For CPMD, which uses plane wave expansion rather
1885 than atom-centered basis functions, the {\tt planewavecutoff} keyword
1886 controls the plane wave expansion.
1888 \item[\tt QMcharge = ]\mbox{}\\The total charge in {\it{e}} of the {\tt
1889 QMMM-grps}. In case there are more than one {\tt QMMM-grps}, the total
1890 charge of each ONIOM layer needs to be specified separately.
1892 \item[\tt QMmult = ]\mbox{}\\The multiplicity of the {\tt
1893 QMMM-grps}. In case there are more than one {\tt QMMM-grps}, the
1894 multiplicity of each ONIOM layer needs to be specified separately.
1896 \item[\tt CASorbitals = ]\mbox{}\\The number of orbitals to be
1897 included in the active space when doing a CASSCF computation.
1899 \item[\tt CASelectrons = ]\mbox{}\\The number of electrons to be
1900 included in the active space when doing a CASSCF computation.
1902 \item[\tt SH = no]\mbox{}\\If this is set to yes, a QM/MM MD
1903 simulation on the excited state-potential energy surface and enforce a
1904 diabatic hop to the ground-state when the system hits the conical
1905 intersection hyperline in the course the simulation. This option only
1906 works in combination with the CASSCF method.
1908 \end{description}
1910 \subsection{Output}
1912 The energies and gradients computed in the QM calculation are added to
1913 those computed by {\gromacs}. In the {\tt .edr} file there is a section
1914 for the total QM energy.
1916 \subsection{Future developments}
1918 Several features are currently under development to increase the
1919 accuracy of the QM/MM interface. One useful feature is the use of
1920 delocalized MM charges in the QM computations. The most important
1921 benefit of using such smeared-out charges is that the Coulombic
1922 potential has a finite value at interatomic distances. In the point
1923 charge representation, the partially-charged MM atoms close to the QM
1924 region tend to ``over-polarize'' the QM system, which leads to artifacts
1925 in the calculation.
1927 What is needed as well is a transition state optimizer.
1929 \section{Using VMD plug-ins for trajectory file I/O}
1930 \index{VMD plug-ins}\index{trajectory file}{\gromacs} tools are able
1931 to use the plug-ins found in an existing installation of
1932 \href{http://www.ks.uiuc.edu/Research/vmd}{VMD} in order to read and
1933 write trajectory files in formats that are not native to
1934 {\gromacs}. You will be able to supply an AMBER DCD-format trajectory
1935 filename directly to {\gromacs} tools, for example.
1937 This requires a VMD installation not older than version 1.8, that your
1938 system provides the dlopen function so that programs can determine at
1939 run time what plug-ins exist, and that you build shared libraries when
1940 building {\gromacs}. CMake will find the vmd executable in your path, and
1941 from it, or the environment variable {\tt VMDDIR} at configuration or
1942 run time, locate the plug-ins. Alternatively, the {\tt VMD_PLUGIN_PATH}
1943 can be used at run time to specify a path where these plug-ins can be
1944 found. Note that these plug-ins are in a binary format, and that format
1945 must match the architecture of the machine attempting to use them.
1948 \section{\normindex{Interactive Molecular Dynamics}}
1949 {\gromacs} supports the interactive molecular dynamics (IMD) protocol as implemented
1950 by \href{http://www.ks.uiuc.edu/Research/vmd}{VMD} to control a running simulation
1951 in NAMD. IMD allows to monitor a running {\gromacs} simulation from a VMD client.
1952 In addition, the user can interact with the simulation by pulling on atoms, residues
1953 or fragments with a mouse or a force-feedback device. Additional information about
1954 the {\gromacs} implementation and an exemplary {\gromacs} IMD system can be found
1955 \href{http://www.mpibpc.mpg.de/grubmueller/interactivemd}{on this homepage}.
1957 \subsection{Simulation input preparation}
1958 The {\gromacs} implementation allows transmission and interaction with a part of the
1959 running simulation only, e.g.\ in cases where no water molecules should be transmitted
1960 or pulled. The group is specified via the {\tt .mdp} option {\tt IMD-group}. When
1961 {\tt IMD-group} is empty, the IMD protocol is disabled and cannot be enabled via the
1962 switches in {\tt mdrun}. To interact with the entire system, {\tt IMD-group} can
1963 be set to {\tt System}. When using {\tt grompp}, a {\tt .gro} file
1964 to be used as VMD input is written out ({\tt -imd} switch of {\tt grompp}).
1966 \subsection{Starting the simulation}
1967 Communication between VMD and {\gromacs} is achieved via TCP sockets and thus enables
1968 controlling an {\tt mdrun} running locally or on a remote cluster. The port for the
1969 connection can be specified with the {\tt -imdport} switch of {\tt mdrun}, 8888 is
1970 the default. If a port number of 0 or smaller is provided, {\gromacs} automatically
1971 assigns a free port to use with IMD.
1973 Every $N$ steps, the {\tt mdrun} client receives the applied forces from VMD and sends the new
1974 positions to the client. VMD permits increasing or decreasing the communication frequency
1975 interactively.
1976 By default, the simulation starts and runs even if no IMD client is connected. This
1977 behavior is changed by the {\tt -imdwait} switch of {\tt mdrun}. After startup and
1978 whenever the client has disconnected, the integration stops until reconnection of
1979 the client.
1980 When the {\tt -imdterm} switch is used, the simulation can be terminated by pressing
1981 the stop button in VMD. This is disabled by default.
1982 Finally, to allow interacting with the simulation (i.e. pulling from VMD) the {\tt -imdpull}
1983 switch has to be used.
1984 Therefore, a simulation can only be monitored but not influenced from the VMD client
1985 when none of {\tt -imdwait}, {\tt -imdterm} or {\tt -imdpull} are set. However, since
1986 the IMD protocol requires no authentication, it is not advisable to run simulations on
1987 a host directly reachable from an insecure environment. Secure shell forwarding of TCP
1988 can be used to connect to running simulations not directly reachable from the interacting host.
1989 Note that the IMD command line switches of {\tt mdrun} are hidden by default and show
1990 up in the help text only with {\tt gmx mdrun -h -hidden}.
1992 \subsection{Connecting from VMD}
1993 In VMD, first the structure corresponding to the IMD group has to be loaded ({\it File
1994 $\rightarrow$ New Molecule}). Then the IMD connection window has to be used
1995 ({\it Extensions $\rightarrow$ Simulation $\rightarrow$ IMD Connect (NAMD)}). In the IMD
1996 connection window, hostname and port have to be specified and followed by pressing
1997 {\it Connect}. {\it Detach Sim} allows disconnecting without terminating the simulation, while
1998 {\it Stop Sim} ends the simulation on the next neighbor searching step (if allowed by
1999 {\tt -imdterm}).
2001 The timestep transfer rate allows adjusting the communication frequency between simulation
2002 and IMD client. Setting the keep rate loads every $N^\mathrm{th}$ frame into VMD instead
2003 of discarding them when a new one is received. The displayed energies are in SI units
2004 in contrast to energies displayed from NAMD simulations.
2006 \section{\normindex{Embedding proteins into the membranes}}
2007 \label{sec:membed}
2009 GROMACS is capable of inserting the protein into pre-equilibrated
2010 lipid bilayers with minimal perturbation of the lipids using the
2011 method, which was initially described as a ProtSqueeze
2012 technique,\cite{Yesylevskyy2007} and later implemented as g\_membed
2013 tool.\cite{Wolf2010} Currently the functionality of g\_membed is
2014 available in mdrun as described in the user guide.
2016 This method works by first artificially shrinking the protein in the
2017 $xy$-plane, then it removes lipids that overlap with that much smaller
2018 core. Then the protein atoms are gradually resized back to their
2019 initial configuration, using normal dynamics for the rest of the
2020 system, so the lipids adapt to the protein. Further lipids are removed
2021 as required.
2024 % LocalWords: PMF pmf kJ mol Jarzynski BT bilayer rup mdp AFM fepmf fecalc rb
2025 % LocalWords: posre grompp fs Verlet dihedrals hydrogens hydroxyl sulfhydryl
2026 % LocalWords: vsitehydro planarity chirality pdb gmx virtualize virtualized xz
2027 % LocalWords: vis massless tryptophan histidine phenyl parameterizing ij PPPM
2028 % LocalWords: parameterization Berendsen rlist coulombtype rcoulomb vdwtype LJ
2029 % LocalWords: rvdw energygrp mdrun pre GMXLIB MOPAC GAMESS CPMD ONIOM
2030 % LocalWords: Morokuma iJ AQ AJ initio atomtype QMatom MMatom QMMM grps et al
2031 % LocalWords: QMmethod QMbasis QMMMscheme RHF DFT LYP CASSCF MMVB CASelectrons
2032 % LocalWords: CASorbitals planewavecutoff STO QMcharge QMmult diabatic edr
2033 % LocalWords: hyperline delocalized Coulombic indices nm ccc th ps
2034 % LocalWords: GTX CPUs GHz md sd bd vv avek tcoupl andersen tc OPLSAA GROMOS
2035 % LocalWords: OBC obc CCMA tol pbc xyz barostat pcoupl acc gpu PLUGIN Cmake GX
2036 % LocalWords: MSVC gcc installpath cmake DGMX DCMAKE functionalities GPGPU GTS
2037 % LocalWords: OpenCL DeviceID gromacs gpus html GeForce Quadro FX Plex CX GF
2038 % LocalWords: VMD DCD GROningen MAchine BIOSON Groningen der Spoel Drunen Comp
2039 % LocalWords: Phys Comm moltype intramol vdw Waals fep multivector pf
2040 % LocalWords: pymbar dhdl xvg LINCS Entropic entropic solutes ref com iso pm
2041 % LocalWords: rm prefactors equipotential potiso potisopf potpm trr
2042 % LocalWords: potrm potrmpf midplanes midplane gaussians potflex vars massw av
2043 % LocalWords: shure observables rccccccc vec eps dist min eqn transl nstsout
2044 % LocalWords: nstrout rotangles rotslabs rottorque RMSD rmsdfit excl NH amine
2045 % LocalWords: positionrestraint es SH phenylalanine solvated sh nanometer QM
2046 % LocalWords: Lennard Buckingham UK Svensson ab vsitetop co UHF MP interatomic
2047 % LocalWords: cg grp coords SPC userpot
2048 % LocalWords: itp sitesn atomtypes ff csg ndx Tesla CHARMM AA gauss
2049 % LocalWords: CMAP nocmap fators Monte performant lib LD DIR llllcc
2050 % LocalWords: CMake homepage DEV overclocking GT dlopen vmd VMDDIR
2051 % LocalWords: versa PME atomperatom graining forcefields hy spc OPLS
2052 % LocalWords: topol multi