Added pull coordinate geometry dihedral (angle)
[gromacs.git] / docs / manual / special.tex
blob14ce7da762a1ce070eb361b0fa1a01bc0e0fc770
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 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
336 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
337 vector when these vectors are projected onto a plane normal to the second vector (the axis vector).
338 As an example, consider a dihedral angle involving four groups: 1, 5, 8 and 9. Here, the {\tt .mdp} option
339 {\tt pull-coord?-groups = 8 1 1 5 5 9} specifies the three vectors that define the dihedral angle:
340 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.
341 The dihedral angle takes values in the interval (-180, 180] deg and has periodic boundaries.
343 \subsubsection{Limitations}
344 There is one theoretical limitation:
345 strictly speaking, constraint forces can only be calculated between
346 groups that are not connected by constraints to the rest of the system.
347 If a group contains part of a molecule of which the bond lengths
348 are constrained, the pull constraint and LINCS or SHAKE bond constraint
349 algorithms should be iterated simultaneously. This is not done in {\gromacs}.
350 This means that for simulations with {\tt constraints = all-bonds}
351 in the {\tt .mdp} file pulling is, strictly speaking,
352 limited to whole molecules or groups of molecules.
353 In some cases this limitation can be avoided by using the free energy code,
354 see \secref{fepmf}.
355 In practice, the errors caused by not iterating the two constraint
356 algorithms can be negligible when the pull group consists of a large
357 amount of atoms and/or the pull force is small.
358 In such cases, the constraint correction displacement of the pull group
359 is small compared to the bond lengths.
363 \section{\normindex{Enforced Rotation}}
364 \index{rotational pulling|see{enforced rotation}}
365 \index{pulling, rotational|see{enforced rotation}}
366 \label{sec:rotation}
368 \mathchardef\mhyphen="2D
369 \newcommand{\rotiso }{^\mathrm{iso}}
370 \newcommand{\rotisopf }{^\mathrm{iso\mhyphen pf}}
371 \newcommand{\rotpm }{^\mathrm{pm}}
372 \newcommand{\rotpmpf }{^\mathrm{pm\mhyphen pf}}
373 \newcommand{\rotrm }{^\mathrm{rm}}
374 \newcommand{\rotrmpf }{^\mathrm{rm\mhyphen pf}}
375 \newcommand{\rotrmtwo }{^\mathrm{rm2}}
376 \newcommand{\rotrmtwopf }{^\mathrm{rm2\mhyphen pf}}
377 \newcommand{\rotflex }{^\mathrm{flex}}
378 \newcommand{\rotflext }{^\mathrm{flex\mhyphen t}}
379 \newcommand{\rotflextwo }{^\mathrm{flex2}}
380 \newcommand{\rotflextwot}{^\mathrm{flex2\mhyphen t}}
382 This module can be used to enforce the rotation of a group of atoms, as {\eg}
383 a protein subunit. There are a variety of rotation potentials, among them
384 complex ones that allow flexible adaptations of both the rotated subunit as
385 well as the local rotation axis during the simulation. An example application
386 can be found in ref. \cite{Kutzner2011}.
388 \begin{figure}
389 \centerline{\includegraphics[width=13cm]{plots/rotation.pdf}}
390 \caption[Fixed and flexible axis rotation]{Comparison of fixed and flexible axis
391 rotation.
392 {\sf A:} Rotating the sketched shape inside the white tubular cavity can create
393 artifacts when a fixed rotation axis (dashed) is used. More realistically, the
394 shape would revolve like a flexible pipe-cleaner (dotted) inside the bearing (gray).
395 {\sf B:} Fixed rotation around an axis \ve{v} with a pivot point
396 specified by the vector \ve{u}.
397 {\sf C:} Subdividing the rotating fragment into slabs with separate rotation
398 axes ($\uparrow$) and pivot points ($\bullet$) for each slab allows for
399 flexibility. The distance between two slabs with indices $n$ and $n+1$ is $\Delta x$.}
400 \label{fig:rotation}
401 \end{figure}
403 \begin{figure}
404 \centerline{\includegraphics[width=13cm]{plots/equipotential.pdf}}
405 \caption{Selection of different rotation potentials and definition of notation.
406 All four potentials $V$ (color coded) are shown for a single atom at position
407 $\ve{x}_j(t)$.
408 {\sf A:} Isotropic potential $V\rotiso$,
409 {\sf B:} radial motion potential $V\rotrm$ and flexible potential
410 $V\rotflex$,
411 {\sf C--D:} radial motion\,2 potential $V\rotrmtwo$ and
412 flexible\,2 potential $V\rotflextwo$ for $\epsilon' = 0$\,nm$^2$ {\sf (C)}
413 and $\epsilon' = 0.01$\,nm$^2$ {\sf (D)}. The rotation axis is perpendicular to
414 the plane and marked by $\otimes$. The light gray contours indicate Boltzmann factors
415 $e^{-V/(k_B T)}$ in the $\ve{x}_j$-plane for $T=300$\,K and
416 $k=200$\,kJ/(mol$\cdot$nm$^2$). The green arrow shows the direction of the
417 force $\ve{F}_{\!j}$ acting on atom $j$; the blue dashed line indicates the
418 motion of the reference position.}
419 \label{fig:equipotential}
420 \end{figure}
422 \subsection{Fixed Axis Rotation}
423 \subsubsection{Stationary Axis with an Isotropic Potential}
424 In the fixed axis approach (see \figref{rotation}B), torque on a group of $N$
425 atoms with positions $\ve{x}_i$ (denoted ``rotation group'') is applied by
426 rotating a reference set of atomic positions -- usually their initial positions
427 $\ve{y}_i^0$ -- at a constant angular velocity $\omega$ around an axis
428 defined by a direction vector $\hat{\ve{v}}$ and a pivot point \ve{u}.
429 To that aim, each atom with position $\ve{x}_i$ is attracted by a
430 ``virtual spring'' potential to its moving reference position
431 $\ve{y}_i = \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{u})$,
432 where $\mathbf{\Omega}(t)$ is a matrix that describes the rotation around the
433 axis. In the simplest case, the ``springs'' are described by a harmonic
434 potential,
435 \beq
436 V\rotiso = \frac{k}{2} \sum_{i=1}^{N} w_i \left[ \mathbf{\Omega}(t)
437 (\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u}) \right]^2 ,
438 \label{eqn:potiso}
439 \eeq
440 with optional mass-weighted prefactors $w_i = N \, m_i/M$ with total mass
441 $M = \sum_{i=1}^N m_i$.
442 The rotation matrix $\mathbf{\Omega}(t)$ is
443 \newcommand{\omcost}{\,\xi\,} % abbreviation
444 \begin{displaymath}
445 \mathbf{\Omega}(t) =
446 \left(
447 \begin{array}{ccc}
448 \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\\
449 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\\
450 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 \\
451 \end{array}
452 \right) ,
453 \end{displaymath}
454 where $v_x$, $v_y$, and $v_z$ are the components of the normalized rotation vector
455 $\hat{\ve{v}}$, and $\omcost := 1-\cos(\omega t)$. As illustrated in
456 \figref{equipotential}A for a single atom $j$, the
457 rotation matrix $\mathbf{\Omega}(t)$ operates on the initial reference positions
458 $\ve{y}_j^0 = \ve{x}_j(t_0)$ of atom $j$ at $t=t_0$. At a later
459 time $t$, the reference position has rotated away from its initial place
460 (along the blue dashed line), resulting in the force
461 \beq
462 \ve{F}_{\!j}\rotiso
463 = -\nabla_{\!j} \, V\rotiso
464 = k \, w_j \left[
465 \mathbf{\Omega}(t) (\ve{y}_j^0 - \ve{u}) - (\ve{x}_j - \ve{u} ) \right] ,
466 \label{eqn:force_fixed}
467 \eeq
468 which is directed towards the reference position.
471 \subsubsection{Pivot-Free Isotropic Potential}
472 Instead of a fixed pivot vector \ve{u} this potential uses the center of
473 mass $\ve{x}_c$ of the rotation group as pivot for the rotation axis,
474 \beq
475 \ve{x}_c = \frac{1}{M} \sum_{i=1}^N m_i \ve{x}_i
476 \label{eqn:com}
477 \mbox{\hspace{4ex}and\hspace{4ex}}
478 \ve{y}_c^0 = \frac{1}{M} \sum_{i=1}^N m_i \ve{y}_i^0 \ ,
479 \eeq
480 which yields the ``pivot-free'' isotropic potential
481 \beq
482 V\rotisopf = \frac{k}{2} \sum_{i=1}^{N} w_i \left[ \mathbf{\Omega}(t)
483 (\ve{y}_i^0 - \ve{y}_c^0) - (\ve{x}_i - \ve{x}_c) \right]^2 ,
484 \label{eqn:potisopf}
485 \eeq
486 with forces
487 \beq
488 \mathbf{F}_{\!j}\rotisopf = k \, w_j
489 \left[
490 \mathbf{\Omega}(t) ( \ve{y}_j^0 - \ve{y}_c^0)
491 - ( \ve{x}_j - \ve{x}_c )
492 \right] .
493 \label{eqn:force_isopf}
494 \eeq
495 Without mass-weighting, the pivot $\ve{x}_c$ is the geometrical center of
496 the group.
497 \label{sec:fixed}
499 \subsubsection{Parallel Motion Potential Variant}
500 The forces generated by the isotropic potentials
501 (\eqnsref{potiso}{potisopf}) also contain components parallel
502 to the rotation axis and thereby restrain motions along the axis of either the
503 whole rotation group (in case of $V\rotiso$) or within
504 the rotation group (in case of $V\rotisopf$). For cases where
505 unrestrained motion along the axis is preferred, we have implemented a
506 ``parallel motion'' variant by eliminating all components parallel to the
507 rotation axis for the potential. This is achieved by projecting the distance
508 vectors between reference and actual positions
509 \beq
510 \ve{r}_i = \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u})
511 \eeq
512 onto the plane perpendicular to the rotation vector,
513 \beq
514 \label{eqn:project}
515 \ve{r}_i^\perp := \ve{r}_i - (\ve{r}_i \cdot \hat{\ve{v}})\hat{\ve{v}} \ ,
516 \eeq
517 yielding
518 \bea
519 \nonumber
520 V\rotpm &=& \frac{k}{2} \sum_{i=1}^{N} w_i ( \ve{r}_i^\perp )^2 \\
521 &=& \frac{k}{2} \sum_{i=1}^{N} w_i
522 \left\lbrace
523 \mathbf{\Omega}(t)
524 (\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u}) \right. \nonumber \\
525 && \left. - \left\lbrace
526 \left[ \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u}) - (\ve{x}_i - \ve{u}) \right] \cdot\hat{\ve{v}}
527 \right\rbrace\hat{\ve{v}} \right\rbrace^2 ,
528 \label{eqn:potpm}
529 \eea
530 and similarly
531 \beq
532 \ve{F}_{\!j}\rotpm = k \, w_j \, \ve{r}_j^\perp .
533 \label{eqn:force_pm}
534 \eeq
536 \subsubsection{Pivot-Free Parallel Motion Potential}
537 Replacing in \eqnref{potpm} the fixed pivot \ve{u} by the center
538 of mass $\ve{x_c}$ yields the pivot-free variant of the parallel motion
539 potential. With
540 \beq
541 \ve{s}_i = \mathbf{\Omega}(t) (\ve{y}_i^0 - \ve{y}_c^0) - (\ve{x}_i - \ve{x}_c)
542 \eeq
543 the respective potential and forces are
544 \bea
545 V\rotpmpf &=& \frac{k}{2} \sum_{i=1}^{N} w_i ( \ve{s}_i^\perp )^2 \ , \\
546 \label{eqn:potpmpf}
547 \ve{F}_{\!j}\rotpmpf &=& k \, w_j \, \ve{s}_j^\perp .
548 \label{eqn:force_pmpf}
549 \eea
551 \subsubsection{Radial Motion Potential}
552 In the above variants, the minimum of the rotation potential is either a single
553 point at the reference position $\ve{y}_i$ (for the isotropic potentials) or a
554 single line through $\ve{y}_i$ parallel to the rotation axis (for the
555 parallel motion potentials). As a result, radial forces restrict radial motions
556 of the atoms. The two subsequent types of rotation potentials, $V\rotrm$
557 and $V\rotrmtwo$, drastically reduce or even eliminate this effect. The first
558 variant, $V\rotrm$ (\figref{equipotential}B), eliminates all force
559 components parallel to the vector connecting the reference atom and the
560 rotation axis,
561 \beq
562 V\rotrm = \frac{k}{2} \sum_{i=1}^{N} w_i \left[
563 \ve{p}_i
564 \cdot(\ve{x}_i - \ve{u}) \right]^2 ,
565 \label{eqn:potrm}
566 \eeq
567 with
568 \beq
569 \ve{p}_i :=
570 \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})\|} \ .
571 \eeq
572 This variant depends only on the distance $\ve{p}_i \cdot (\ve{x}_i -
573 \ve{u})$ of atom $i$ from the plane spanned by $\hat{\ve{v}}$ and
574 $\mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u})$. The resulting force is
575 \beq
576 \mathbf{F}_{\!j}\rotrm =
577 -k \, w_j \left[ \ve{p}_j\cdot(\ve{x}_j - \ve{u}) \right] \,\ve{p}_j \, .
578 \label{eqn:potrm_force}
579 \eeq
581 \subsubsection{Pivot-Free Radial Motion Potential}
582 Proceeding similar to the pivot-free isotropic potential yields a pivot-free
583 version of the above potential. With
584 \beq
585 \ve{q}_i :=
586 \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)\|} \, ,
587 \eeq
588 the potential and force for the pivot-free variant of the radial motion potential read
589 \bea
590 V\rotrmpf & = & \frac{k}{2} \sum_{i=1}^{N} w_i \left[
591 \ve{q}_i
592 \cdot(\ve{x}_i - \ve{x}_c)
593 \right]^2 \, , \\
594 \label{eqn:potrmpf}
595 \mathbf{F}_{\!j}\rotrmpf & = &
596 -k \, w_j \left[ \ve{q}_j\cdot(\ve{x}_j - \ve{x}_c) \right] \,\ve{q}_j
597 + k \frac{m_j}{M} \sum_{i=1}^{N} w_i \left[
598 \ve{q}_i\cdot(\ve{x}_i - \ve{x}_c) \right]\,\ve{q}_i \, .
599 \label{eqn:potrmpf_force}
600 \eea
602 \subsubsection{Radial Motion 2 Alternative Potential}
603 As seen in \figref{equipotential}B, the force resulting from
604 $V\rotrm$ still contains a small, second-order radial component. In most
605 cases, this perturbation is tolerable; if not, the following
606 alternative, $V\rotrmtwo$, fully eliminates the radial contribution to the
607 force, as depicted in \figref{equipotential}C,
608 \beq
609 V\rotrmtwo =
610 \frac{k}{2} \sum_{i=1}^{N} w_i\,
611 \frac{\left[ (\hat{\ve{v}} \times ( \ve{x}_i - \ve{u} ))
612 \cdot \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u}) \right]^2}
613 {\| \hat{\ve{v}} \times (\ve{x}_i - \ve{u}) \|^2 +
614 \epsilon'} \, ,
615 \label{eqn:potrm2}
616 \eeq
617 where a small parameter $\epsilon'$ has been introduced to avoid singularities.
618 For $\epsilon'=0$\,nm$^2$, the equipotential planes are spanned by $\ve{x}_i -
619 \ve{u}$ and $\hat{\ve{v}}$, yielding a force
620 perpendicular to $\ve{x}_i - \ve{u}$, thus not contracting or
621 expanding structural parts that moved away from or toward the rotation axis.
623 Choosing a small positive $\epsilon'$ ({\eg},
624 $\epsilon'=0.01$\,nm$^2$, \figref{equipotential}D) in the denominator of
625 \eqnref{potrm2} yields a well-defined potential and continuous forces also
626 close to the rotation axis, which is not the case for $\epsilon'=0$\,nm$^2$
627 (\figref{equipotential}C). With
628 \bea
629 \ve{r}_i & := & \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{u})\\
630 \ve{s}_i & := & \frac{\hat{\ve{v}} \times (\ve{x}_i -
631 \ve{u} ) }{ \| \hat{\ve{v}} \times (\ve{x}_i - \ve{u})
632 \| } \equiv \; \Psi_{i} \;\; {\hat{\ve{v}} \times
633 (\ve{x}_i-\ve{u} ) }\\
634 \Psi_i^{*} & := & \frac{1}{ \| \hat{\ve{v}} \times
635 (\ve{x}_i-\ve{u}) \|^2 + \epsilon'}
636 \eea
637 the force on atom $j$ reads
638 \beq
639 \ve{F}_{\!j}\rotrmtwo =
640 - k\;
641 \left\lbrace w_j\;
642 (\ve{s}_j\cdot\ve{r}_{\!j})\;
643 \left[ \frac{\Psi_{\!j}^* }{\Psi_{\!j} } \ve{r}_{\!j}
644 - \frac{\Psi_{\!j}^{*2}}{\Psi_{\!j}^3}
645 (\ve{s}_j\cdot\ve{r}_{\!j})\ve{s}_j \right]
646 \right\rbrace \times \hat{\ve{v}} .
647 \label{eqn:potrm2_force}
648 \eeq
650 \subsubsection{Pivot-Free Radial Motion 2 Potential}
651 The pivot-free variant of the above potential is
652 \beq
653 V\rotrmtwopf =
654 \frac{k}{2} \sum_{i=1}^{N} w_i\,
655 \frac{\left[ (\hat{\ve{v}} \times ( \ve{x}_i - \ve{x}_c ))
656 \cdot \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c) \right]^2}
657 {\| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c) \|^2 +
658 \epsilon'} \, .
659 \label{eqn:potrm2pf}
660 \eeq
661 With
662 \bea
663 \ve{r}_i & := & \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c)\\
664 \ve{s}_i & := & \frac{\hat{\ve{v}} \times (\ve{x}_i -
665 \ve{x}_c ) }{ \| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c)
666 \| } \equiv \; \Psi_{i} \;\; {\hat{\ve{v}} \times
667 (\ve{x}_i-\ve{x}_c ) }\\ \Psi_i^{*} & := & \frac{1}{ \| \hat{\ve{v}} \times
668 (\ve{x}_i-\ve{x}_c) \|^2 + \epsilon'}
669 \eea
670 the force on atom $j$ reads
671 \bea
672 \nonumber
673 \ve{F}_{\!j}\rotrmtwopf & = &
674 - k\;
675 \left\lbrace w_j\;
676 (\ve{s}_j\cdot\ve{r}_{\!j})\;
677 \left[ \frac{\Psi_{\!j}^* }{\Psi_{\!j} } \ve{r}_{\!j}
678 - \frac{\Psi_{\!j}^{*2}}{\Psi_{\!j}^3}
679 (\ve{s}_j\cdot\ve{r}_{\!j})\ve{s}_j \right]
680 \right\rbrace \times \hat{\ve{v}}\\
682 + k\;\frac{m_j}{M} \left\lbrace \sum_{i=1}^{N}
683 w_i\;(\ve{s}_i\cdot\ve{r}_i) \;
684 \left[ \frac{\Psi_i^* }{\Psi_i } \ve{r}_i
685 - \frac{\Psi_i^{*2}}{\Psi_i^3} (\ve{s}_i\cdot\ve{r}_i )\;
686 \ve{s}_i \right] \right\rbrace \times \hat{\ve{v}} \, .
687 \label{eqn:potrm2pf_force}
688 \eea
690 \subsection{Flexible Axis Rotation}
691 As sketched in \figref{rotation}A--B, the rigid body behavior of
692 the fixed axis rotation scheme is a drawback for many applications. In
693 particular, deformations of the rotation group are suppressed when the
694 equilibrium atom positions directly depend on the reference positions.
695 To avoid this limitation, \eqnsref{potrmpf}{potrm2pf}
696 will now be generalized towards a ``flexible axis'' as sketched in
697 \figref{rotation}C. This will be achieved by subdividing the
698 rotation group into a set of equidistant slabs perpendicular to
699 the rotation vector, and by applying a separate rotation potential to each
700 of these slabs. \figref{rotation}C shows the midplanes of the slabs
701 as dotted straight lines and the centers as thick black dots.
703 To avoid discontinuities in the potential and in the forces, we define
704 ``soft slabs'' by weighing the contributions of each
705 slab $n$ to the total potential function $V\rotflex$ by a Gaussian
706 function
707 \beq
708 \label{eqn:gaussian}
709 g_n(\ve{x}_i) = \Gamma \ \mbox{exp} \left(
710 -\frac{\beta_n^2(\ve{x}_i)}{2\sigma^2} \right) ,
711 \eeq
712 centered at the midplane of the $n$th slab. Here $\sigma$ is the width
713 of the Gaussian function, $\Delta x$ the distance between adjacent slabs, and
714 \beq
715 \beta_n(\ve{x}_i) := \ve{x}_i \cdot \hat{\ve{v}} - n \, \Delta x \, .
716 \eeq
718 \begin{figure}
719 \centerline{\includegraphics[width=6.5cm]{plots/gaussians.pdf}}
720 \caption{Gaussian functions $g_n$ centered at $n \, \Delta x$ for a slab
721 distance $\Delta x = 1.5$ nm and $n \geq -2$. Gaussian function $g_0$ is
722 highlighted in bold; the dashed line depicts the sum of the shown Gaussian
723 functions.}
724 \label{fig:gaussians}
725 \end{figure}
727 A most convenient choice is $\sigma = 0.7 \Delta x$ and
728 \begin{displaymath}
729 1/\Gamma = \sum_{n \in Z}
730 \mbox{exp}
731 \left(-\frac{(n - \frac{1}{4})^2}{2\cdot 0.7^2}\right)
732 \approx 1.75464 \, ,
733 \end{displaymath}
734 which yields a nearly constant sum, essentially independent of $\ve{x}_i$
735 (dashed line in \figref{gaussians}), {\ie},
736 \beq
737 \sum_{n \in Z} g_n(\ve{x}_i) = 1 + \epsilon(\ve{x}_i) \, ,
738 \label{eqn:normal}
739 \eeq
740 with $ | \epsilon(\ve{x}_i) | < 1.3\cdot 10^{-4}$. This choice also
741 implies that the individual contributions to the force from the slabs add up to
742 unity such that no further normalization is required.
744 To each slab center $\ve{x}_c^n$, all atoms contribute by their
745 Gaussian-weighted (optionally also mass-weighted) position vectors
746 $g_n(\ve{x}_i) \, \ve{x}_i$. The
747 instantaneous slab centers $\ve{x}_c^n$ are calculated from the
748 current positions $\ve{x}_i$,
749 \beq
750 \label{eqn:defx0}
751 \ve{x}_c^n =
752 \frac{\sum_{i=1}^N g_n(\ve{x}_i) \, m_i \, \ve{x}_i}
753 {\sum_{i=1}^N g_n(\ve{x}_i) \, m_i} \, ,\\
754 \eeq
755 while the reference centers $\ve{y}_c^n$ are calculated from the reference
756 positions $\ve{y}_i^0$,
757 \beq
758 \label{eqn:defy0}
759 \ve{y}_c^n =
760 \frac{\sum_{i=1}^N g_n(\ve{y}_i^0) \, m_i \, \ve{y}_i^0}
761 {\sum_{i=1}^N g_n(\ve{y}_i^0) \, m_i} \, .
762 \eeq
763 Due to the rapid decay of $g_n$, each slab
764 will essentially involve contributions from atoms located within $\approx
765 3\Delta x$ from the slab center only.
767 \subsubsection{Flexible Axis Potential}
768 We consider two flexible axis variants. For the first variant,
769 the slab segmentation procedure with Gaussian weighting is applied to the radial
770 motion potential (\eqnref{potrmpf}\,/\,\figref{equipotential}B),
771 yielding as the contribution of slab $n$
772 \begin{displaymath}
773 V^n =
774 \frac{k}{2} \sum_{i=1}^{N} w_i \, g_n(\ve{x}_i)
775 \left[
776 \ve{q}_i^n
777 \cdot
778 (\ve{x}_i - \ve{x}_c^n)
779 \right]^2 ,
780 \label{eqn:flexpot}
781 \end{displaymath}
782 and a total potential function
783 \beq
784 V\rotflex = \sum_n V^n \, .
785 \label{eqn:potflex}
786 \eeq
787 Note that the global center of mass $\ve{x}_c$ used in
788 \eqnref{potrmpf} is now replaced by $\ve{x}_c^n$, the center of mass of
789 the slab. With
790 \bea
791 \ve{q}_i^n & := & \frac{\hat{\ve{v}} \times
792 \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n) }{ \| \hat{\ve{v}}
793 \times \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n) \| } \\
794 b_i^n & := & \ve{q}_i^n \cdot (\ve{x}_i - \ve{x}_c^n) \, ,
795 \eea
796 the resulting force on atom $j$ reads
797 \bea
798 \nonumber\hspace{-15mm}
799 \ve{F}_{\!j}\rotflex &=&
800 - \, k \, w_j \sum_n g_n(\ve{x}_j) \, b_j^n \left\lbrace \ve{q}_j^n -
801 b_j^n \frac{\beta_n(\ve{x}_j)}{2\sigma^2} \hat{\ve{v}} \right\rbrace \\ & &
802 + \, k \, m_j \sum_n \frac{g_n(\ve{x}_j)}{\sum_h g_n(\ve{x}_h)}
803 \sum_{i=1}^{N} w_i \, g_n(\ve{x}_i) \, b_i^n \left\lbrace
804 \ve{q}_i^n -\frac{\beta_n(\ve{x}_j)}{\sigma^2}
805 \left[ \ve{q}_i^n \cdot (\ve{x}_j - \ve{x}_c^n )\right]
806 \hat{\ve{v}} \right\rbrace .
807 \label{eqn:potflex_force}
808 \eea
810 Note that for $V\rotflex$, as defined, the slabs are fixed in space and so
811 are the reference centers $\ve{y}_c^n$. If during the simulation the
812 rotation group moves too far in $\ve{v}$ direction, it may enter a
813 region where -- due to the lack of nearby reference positions -- no reference
814 slab centers are defined, rendering the potential evaluation impossible.
815 We therefore have included a slightly modified version of this potential that
816 avoids this problem by attaching the midplane of slab $n=0$ to the center of mass
817 of the rotation group, yielding slabs that move with the rotation group.
818 This is achieved by subtracting the center of mass $\ve{x}_c$ of the
819 group from the positions,
820 \beq
821 \tilde{\ve{x}}_i = \ve{x}_i - \ve{x}_c \, , \mbox{\ \ \ and \ \ }
822 \tilde{\ve{y}}_i^0 = \ve{y}_i^0 - \ve{y}_c^0 \, ,
823 \label{eqn:trafo}
824 \eeq
825 such that
826 \bea
827 V\rotflext
828 & = & \frac{k}{2} \sum_n \sum_{i=1}^{N} w_i \, g_n(\tilde{\ve{x}}_i)
829 \left[ \frac{\hat{\ve{v}} \times \mathbf{\Omega}(t)(\tilde{\ve{y}}_i^0
830 - \tilde{\ve{y}}_c^n) }{ \| \hat{\ve{v}} \times
831 \mathbf{\Omega}(t)(\tilde{\ve{y}}_i^0 -
832 \tilde{\ve{y}}_c^n) \| }
833 \cdot
834 (\tilde{\ve{x}}_i - \tilde{\ve{x}}_c^n)
835 \right]^2 .
836 \label{eqn:potflext}
837 \eea
838 To simplify the force derivation, and for efficiency reasons, we here assume
839 $\ve{x}_c$ to be constant, and thus $\partial \ve{x}_c / \partial x =
840 \partial \ve{x}_c / \partial y = \partial \ve{x}_c / \partial z = 0$. The
841 resulting force error is small (of order $O(1/N)$ or $O(m_j/M)$ if
842 mass-weighting is applied) and can therefore be tolerated. With this assumption,
843 the forces $\ve{F}\rotflext$ have the same form as
844 \eqnref{potflex_force}.
846 \subsubsection{Flexible Axis 2 Alternative Potential}
847 In this second variant, slab segmentation is applied to $V\rotrmtwo$
848 (\eqnref{potrm2pf}), resulting in a flexible axis potential without radial
849 force contributions (\figref{equipotential}C),
850 \beq
851 V\rotflextwo =
852 \frac{k}{2} \sum_{i=1}^{N} \sum_n w_i\,g_n(\ve{x}_i)
853 \frac{\left[ (\hat{\ve{v}} \times ( \ve{x}_i - \ve{x}_c^n ))
854 \cdot \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n) \right]^2}
855 {\| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c^n) \|^2 +
856 \epsilon'} \, .
857 \label{eqn:potflex2}
858 \eeq
859 With
860 \bea
861 \ve{r}_i^n & := & \mathbf{\Omega}(t)(\ve{y}_i^0 - \ve{y}_c^n)\\
862 \ve{s}_i^n & := & \frac{\hat{\ve{v}} \times (\ve{x}_i -
863 \ve{x}_c^n ) }{ \| \hat{\ve{v}} \times (\ve{x}_i - \ve{x}_c^n)
864 \| } \equiv \; \psi_{i} \;\; {\hat{\ve{v}} \times (\ve{x}_i-\ve{x}_c^n ) }\\
865 \psi_i^{*} & := & \frac{1}{ \| \hat{\ve{v}} \times (\ve{x}_i-\ve{x}_c^n) \|^2 + \epsilon'}\\
866 W_j^n & := & \frac{g_n(\ve{x}_j)\,m_j}{\sum_h g_n(\ve{x}_h)\,m_h}\\
867 \ve{S}^n & := &
868 \sum_{i=1}^{N} w_i\;g_n(\ve{x}_i)
869 \; (\ve{s}_i^n\cdot\ve{r}_i^n)
870 \left[ \frac{\psi_i^* }{\psi_i } \ve{r}_i^n
871 - \frac{\psi_i^{*2}}{\psi_i^3} (\ve{s}_i^n\cdot\ve{r}_i^n )\;
872 \ve{s}_i^n \right] \label{eqn:Sn}
873 \eea
874 the force on atom $j$ reads
875 \bea
876 \nonumber
877 \ve{F}_{\!j}\rotflextwo & = &
878 - k\;
879 \left\lbrace \sum_n w_j\;g_n(\ve{x}_j)\;
880 (\ve{s}_j^n\cdot\ve{r}_{\!j}^n)\;
881 \left[ \frac{\psi_j^* }{\psi_j } \ve{r}_{\!j}^n
882 - \frac{\psi_j^{*2}}{\psi_j^3} (\ve{s}_j^n\cdot\ve{r}_{\!j}^n)\;
883 \ve{s}_{\!j}^n \right] \right\rbrace \times \hat{\ve{v}} \\
884 \nonumber
886 + k \left\lbrace \sum_n W_{\!j}^n \, \ve{S}^n \right\rbrace \times
887 \hat{\ve{v}}
888 - k \left\lbrace \sum_n W_{\!j}^n \; \frac{\beta_n(\ve{x}_j)}{\sigma^2} \frac{1}{\psi_j}\;\;
889 \ve{s}_j^n \cdot
890 \ve{S}^n \right\rbrace \hat{\ve{v}}\\
891 & &
892 + \frac{k}{2} \left\lbrace \sum_n w_j\;g_n(\ve{x}_j)
893 \frac{\beta_n(\ve{x}_j)}{\sigma^2}
894 \frac{\psi_j^*}{\psi_j^2}( \ve{s}_j^n \cdot \ve{r}_{\!j}^n )^2 \right\rbrace
895 \hat{\ve{v}} .
896 \label{eqn:potflex2_force}
897 \eea
899 Applying transformation (\ref{eqn:trafo}) yields a ``translation-tolerant''
900 version of the flexible\,2 potential, $V\rotflextwot$. Again,
901 assuming that $\partial \ve{x}_c / \partial x$, $\partial \ve{x}_c /
902 \partial y$, $\partial \ve{x}_c / \partial z$ are small, the
903 resulting equations for $V\rotflextwot$ and $\ve{F}\rotflextwot$ are
904 similar to those of $V\rotflextwo$ and $\ve{F}\rotflextwo$.
906 \subsection{Usage}
907 To apply enforced rotation, the particles $i$ that are to
908 be subjected to one of the rotation potentials are defined via index groups
909 {\tt rot-group0}, {\tt rot-group1}, etc., in the {\tt .mdp} input file.
910 The reference positions $\ve{y}_i^0$ are
911 read from a special {\tt .trr} file provided to {\tt grompp}. If no such file is found,
912 $\ve{x}_i(t=0)$ are used as reference positions and written to {\tt .trr} such
913 that they can be used for subsequent setups. All parameters of the potentials
914 such as $k$, $\epsilon'$, etc. (\tabref{vars}) are provided as {\tt .mdp}
915 parameters; {\tt rot-type} selects the type of the potential.
916 The option {\tt rot-massw} allows to choose whether or not to use
917 mass-weighted averaging.
918 For the flexible potentials, a cutoff value $g_n^\mathrm{min}$
919 (typically $g_n^\mathrm{min}=0.001$) makes shure that only
920 significant contributions to $V$ and \ve{F} are evaluated, {\ie} terms with
921 $g_n(\ve{x}) < g_n^\mathrm{min}$ are omitted.
922 \tabref{quantities} summarizes observables that are written
923 to additional output files and which are described below.
926 \begin{table}[tbp]
927 \caption{Parameters used by the various rotation potentials.
928 {\sf x}'s indicate which parameter is actually used for a given potential.}
929 %\small
931 \newcommand{\kunit}{$\frac{\mathrm{kJ}}{\mathrm{mol} \cdot \mathrm{nm}^2}$}
932 \newcommand{\smtt}[1]{{\hspace{-0.5ex}\small #1\hspace{-0.5ex}}}
933 \label{tab:vars}
934 \begin{center}
935 \begin{tabular}{l>{$}l<{$}rccccccc}
936 \hline
937 parameter & & & $k$ & $\hat{\ve{v}}$ & $\ve{u}$ & $\omega$ & $\epsilon'$ & $\Delta x$ & $g_n^\mathrm{min}$ \\
938 \multicolumn{3}{l}{{\tt .mdp} input variable name} & \smtt{k} & \smtt{vec} & \smtt{pivot} & \smtt{rate} & \smtt{eps} & \smtt{slab-dist} & \smtt{min-gauss} \\
939 unit & & & \kunit & - & nm & $^\circ$/ps & nm$^2$ & nm & \,-\, \\[1mm]
940 \hline \multicolumn{2}{l}{fixed axis potentials:} & eqn.\\
941 isotropic & V\rotiso & (\ref{eqn:potiso}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - & - \\
942 --- pivot-free & V\rotisopf & (\ref{eqn:potisopf}) & {\sf x} & {\sf x} & - & {\sf x} & - & - & - \\
943 parallel motion & V\rotpm & (\ref{eqn:potpm}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - & - \\
944 --- pivot-free & V\rotpmpf & (\ref{eqn:potpmpf}) & {\sf x} & {\sf x} & - & {\sf x} & - & - & - \\
945 radial motion & V\rotrm & (\ref{eqn:potrm}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - & - \\
946 --- pivot-free & V\rotrmpf & (\ref{eqn:potrmpf}) & {\sf x} & {\sf x} & - & {\sf x} & - & - & - \\
947 radial motion\,2 & V\rotrmtwo & (\ref{eqn:potrm2}) & {\sf x} & {\sf x} & {\sf x} & {\sf x} & {\sf x} & - & - \\
948 --- pivot-free & V\rotrmtwopf & (\ref{eqn:potrm2pf}) & {\sf x} & {\sf x} & - & {\sf x} & {\sf x} & - & - \\ \hline
949 \multicolumn{2}{l}{flexible axis potentials:} & eqn.\\
950 flexible & V\rotflex & (\ref{eqn:potflex}) & {\sf x} & {\sf x} & - & {\sf x} & - & {\sf x} & {\sf x} \\
951 --- transl. tol. & V\rotflext & (\ref{eqn:potflext}) & {\sf x} & {\sf x} & - & {\sf x} & - & {\sf x} & {\sf x} \\
952 flexible\,2 & V\rotflextwo & (\ref{eqn:potflex2}) & {\sf x} & {\sf x} & - & {\sf x} & {\sf x} & {\sf x} & {\sf x} \\
953 --- transl. tol. & V\rotflextwot & - & {\sf x} & {\sf x} & - & {\sf x} & {\sf x} & {\sf x} & {\sf x} \\
954 \hline
955 \end{tabular}
956 \end{center}
957 \end{table}
959 \begin{table}
960 \caption{Quantities recorded in output files during enforced rotation simulations.
961 All slab-wise data is written every {\tt nstsout} steps, other rotation data every {\tt nstrout} steps.}
962 \label{tab:quantities}
963 \begin{center}
964 \begin{tabular}{llllcc}
965 \hline
966 quantity & unit & equation & output file & fixed & flexible\\ \hline
967 $V(t)$ & kJ/mol & see \ref{tab:vars} & {\tt rotation} & {\sf x} & {\sf x} \\
968 $\theta_\mathrm{ref}(t)$ & degrees & $\theta_\mathrm{ref}(t)=\omega t$ & {\tt rotation} & {\sf x} & {\sf x} \\
969 $\theta_\mathrm{av}(t)$ & degrees & (\ref{eqn:avangle}) & {\tt rotation} & {\sf x} & - \\
970 $\theta_\mathrm{fit}(t)$, $\theta_\mathrm{fit}(t,n)$ & degrees & (\ref{eqn:rmsdfit}) & {\tt rotangles} & - & {\sf x} \\
971 $\ve{y}_0(n)$, $\ve{x}_0(t,n)$ & nm & (\ref{eqn:defx0}, \ref{eqn:defy0})& {\tt rotslabs} & - & {\sf x} \\
972 $\tau(t)$ & kJ/mol & (\ref{eqn:torque}) & {\tt rotation} & {\sf x} & - \\
973 $\tau(t,n)$ & kJ/mol & (\ref{eqn:torque}) & {\tt rottorque} & - & {\sf x} \\ \hline
974 \end{tabular}
975 \end{center}
976 \end{table}
979 \subsubsection*{Angle of Rotation Groups: Fixed Axis}
980 For fixed axis rotation, the average angle $\theta_\mathrm{av}(t)$ of the
981 group relative to the reference group is determined via the distance-weighted
982 angular deviation of all rotation group atoms from their reference positions,
983 \beq
984 \theta_\mathrm{av} = \left. \sum_{i=1}^{N} r_i \ \theta_i \right/ \sum_{i=1}^N r_i \ .
985 \label{eqn:avangle}
986 \eeq
987 Here, $r_i$ is the distance of the reference position to the rotation axis, and
988 the difference angles $\theta_i$ are determined from the atomic positions,
989 projected onto a plane perpendicular to the rotation axis through pivot point
990 $\ve{u}$ (see \eqnref{project} for the definition of $\perp$),
991 \beq
992 \cos \theta_i =
993 \frac{(\ve{y}_i-\ve{u})^\perp \cdot (\ve{x}_i-\ve{u})^\perp}
994 { \| (\ve{y}_i-\ve{u})^\perp \cdot (\ve{x}_i-\ve{u})^\perp
995 \| } \ .
996 \eeq
998 The sign of $\theta_\mathrm{av}$ is chosen such that
999 $\theta_\mathrm{av} > 0$ if the actual structure rotates ahead of the reference.
1001 \subsubsection*{Angle of Rotation Groups: Flexible Axis}
1002 For flexible axis rotation, two outputs are provided, the angle of the
1003 entire rotation group, and separate angles for the segments in the slabs.
1004 The angle of the entire rotation group is determined by an RMSD fit
1005 of $\ve{x}_i$
1006 to the reference positions $\ve{y}_i^0$ at $t=0$, yielding $\theta_\mathrm{fit}$
1007 as the angle by which the reference has to be rotated around $\hat{\ve{v}}$
1008 for the optimal fit,
1009 \beq
1010 \mathrm{RMSD} \big( \ve{x}_i,\ \mathbf{\Omega}(\theta_\mathrm{fit})
1011 \ve{y}_i^0 \big) \stackrel{!}{=} \mathrm{min} \, .
1012 \label{eqn:rmsdfit}
1013 \eeq
1014 To determine the local angle for each slab $n$, both reference and actual
1015 positions are weighted with the Gaussian function of slab $n$, and
1016 $\theta_\mathrm{fit}(t,n)$ is calculated as in \eqnref{rmsdfit}) from the
1017 Gaussian-weighted positions.
1019 For all angles, the {\tt .mdp} input option {\tt rot-fit-method} controls
1020 whether a normal RMSD fit is performed or whether for the fit each
1021 position $\ve{x}_i$ is put at the same distance to the rotation axis as its
1022 reference counterpart $\ve{y}_i^0$. In the latter case, the RMSD
1023 measures only angular differences, not radial ones.
1026 \subsubsection*{Angle Determination by Searching the Energy Minimum}
1027 Alternatively, for {\tt rot-fit-method = potential}, the angle of the rotation
1028 group is determined as the angle for which the rotation potential energy is minimal.
1029 Therefore, the used rotation potential is additionally evaluated for a set of angles
1030 around the current reference angle. In this case, the {\tt rotangles.log} output file
1031 contains the values of the rotation potential at the chosen set of angles, while
1032 {\tt rotation.xvg} lists the angle with minimal potential energy.
1035 \subsubsection*{Torque}
1036 \label{torque}
1037 The torque $\ve{\tau}(t)$ exerted by the rotation potential is calculated for fixed
1038 axis rotation via
1039 \beq
1040 \ve{\tau}(t) = \sum_{i=1}^{N} \ve{r}_i(t) \times \ve{f}_{\!i}^\perp(t) ,
1041 \label{eqn:torque}
1042 \eeq
1043 where $\ve{r}_i(t)$ is the distance vector from the rotation axis to
1044 $\ve{x}_i(t)$ and $\ve{f}_{\!i}^\perp(t)$ is the force component
1045 perpendicular to $\ve{r}_i(t)$ and $\hat{\ve{v}}$. For flexible axis
1046 rotation, torques $\ve{\tau}_{\!n}$ are calculated for each slab using the
1047 local rotation axis of the slab and the Gaussian-weighted positions.
1050 \section{\normindex{Computational Electrophysiology}}
1051 \label{sec:compel}
1053 The Computational Electrophysiology (CompEL) protocol \cite{Kutzner2011b} allows the simulation of
1054 ion flux through membrane channels, driven by transmembrane potentials or ion
1055 concentration gradients. Just as in real cells, CompEL establishes transmembrane
1056 potentials by sustaining a small imbalance of charges $\Delta q$ across the membrane,
1057 which gives rise to a potential difference $\Delta U$ according to the membrane capacitance:
1058 \beq
1059 \Delta U = \Delta q / C_{membrane}
1060 \eeq
1061 The transmembrane electric field and concentration gradients are controlled by
1062 {\tt .mdp} options, which allow the user to set reference counts for the ions on either side
1063 of the membrane. If a difference between the actual and the reference numbers persists
1064 over a certain time span, specified by the user, a number of ion/water pairs are
1065 exchanged between the compartments until the reference numbers are restored.
1066 Alongside the calculation of channel conductance and ion selectivity, CompEL simulations also
1067 enable determination of the channel reversal potential, an important
1068 characteristic obtained in electrophysiology experiments.
1070 In a CompEL setup, the simulation system is divided into two compartments {\bf A} and {\bf B}
1071 with independent ion concentrations. This is best achieved by using double bilayer systems with
1072 a copy (or copies) of the channel/pore of interest in each bilayer (\figref{compelsetup} A, B).
1073 If the channel axes point in the same direction, channel flux is observed
1074 simultaneously at positive and negative potentials in this way, which is for instance
1075 important for studying channel rectification.
1077 \begin{figure}
1078 \centerline{\includegraphics[width=13.5cm]{plots/compelsetup.pdf}}
1079 \caption{Typical double-membrane setup for CompEL simulations (A, B).
1080 Ion\,/\,water molecule exchanges will be performed as needed
1081 between the two light blue volumes around the dotted black lines (A).
1082 Plot (C) shows the potential difference $\Delta U$ resulting
1083 from the selected charge imbalance $\Delta q_{ref}$ between the compartments.}
1084 \label{fig:compelsetup}
1085 \end{figure}
1087 The potential difference $\Delta U$ across the membrane is easily calculated with the
1088 {\tt gmx potential} utility. By this, the potential drop along $z$ or the
1089 pore axis is exactly known in each time interval of the simulation (\figref{compelsetup} C).
1090 Type and number of ions $n_i$ of charge $q_i$, traversing the channel in the simulation,
1091 are written to the {\tt swapions.xvg} output file, from which the average channel
1092 conductance $G$ in each interval $\Delta t$ is determined by:
1093 \beq
1094 G = \frac{\sum_{i} n_{i}q_{i}}{\Delta t \, \Delta U} \, .
1095 \eeq
1096 The ion selectivity is calculated as the number flux ratio of different species.
1097 Best results are obtained by averaging these values over several overlapping time intervals.
1099 The calculation of reversal potentials is best achieved using a small set of simulations in which a given
1100 transmembrane concentration gradient is complemented with small ion imbalances of varying magnitude. For
1101 example, if one compartment contains 1\,M salt and the other 0.1\,M, and given charge neutrality otherwise,
1102 a set of simulations with $\Delta q = 0\,e$, $\Delta q = 2\,e$, $\Delta q = 4\,e$ could
1103 be used. Fitting a straight line through the current-voltage relationship of all obtained
1104 $I$-$U$ pairs near zero current will then yield $U_{rev}$.
1106 \subsection{Usage}
1107 The following {\tt .mdp} options control the CompEL protocol:
1108 {\small
1109 \begin{verbatim}
1110 swapcoords = Z ; Swap positions: no, X, Y, Z
1111 swap-frequency = 100 ; Swap attempt frequency
1112 \end{verbatim}}
1113 Choose {\tt Z} if your membrane is in the $xy$-plane (\figref{compelsetup}).
1114 Ions will be exchanged between compartments depending on their $z$-positions alone.
1115 {\tt swap-frequency} determines how often a swap attempt will be made.
1116 This step requires that the positions of the split groups, the ions, and possibly the solvent molecules are
1117 communicated between the parallel processes, so if chosen too small it can decrease the simulation
1118 performance. The {\tt Position swapping} entry in the cycle and time accounting
1119 table at the end of the {\tt md.log} file summarizes the amount of runtime spent
1120 in the swap module.
1121 {\small
1122 \begin{verbatim}
1123 split-group0 = channel0 ; Defines compartment boundary
1124 split-group1 = channel1 ; Defines other compartment boundary
1125 massw-split0 = no ; use mass-weighted center?
1126 massw-split1 = no
1127 \end{verbatim}}
1128 {\tt split-group0} and {\tt split-group1} are two index groups that define the boundaries
1129 between the two compartments, which are usually the centers of the channels.
1130 If {\tt massw-split0} or {\tt massw-split1} are set to {\tt yes}, the center of mass
1131 of each index group is used as boundary, here in $z$-direction. Otherwise, the
1132 geometrical centers will be used ($\times$ in \figref{compelsetup} A). If, such as here, a membrane
1133 channel is selected as split group, the center of the channel will define the dividing
1134 plane between the compartments (dashed horizontal lines). All index groups
1135 must be defined in the index file.
1137 If, to restore the requested ion counts, an ion from one compartment has to be exchanged
1138 with a water molecule from the other compartment, then those molecules are swapped
1139 which have the largest distance to the compartment-defining boundaries
1140 (dashed horizontal lines). Depending on the ion concentration,
1141 this effectively results in exchanges of molecules between the light blue volumes.
1142 If a channel is very asymmetric in $z$-direction and would extend into one of the
1143 swap volumes, one can offset the swap exchange plane with the {\tt bulk-offset}
1144 parameter. A value of 0.0 means no offset $b$, values $-1.0 < b < 0$ move the
1145 swap exchange plane closer to the lower, values $0 < b < 1.0$ closer to the upper
1146 membrane. \figref{compelsetup} A (left) depicts that for the {\bf A} compartment.
1148 {\small
1149 \begin{verbatim}
1150 solvent-group = SOL ; Group containing the solvent molecules
1151 iontypes = 3 ; Number of different ion types to control
1152 iontype0-name = NA ; Group name of the ion type
1153 iontype0-in-A = 51 ; Reference count of ions of type 0 in A
1154 iontype0-in-B = 35 ; Reference count of ions of type 0 in B
1155 iontype1-name = K
1156 iontype1-in-A = 10
1157 iontype1-in-B = 38
1158 iontype2-name = CL
1159 iontype2-in-A = -1
1160 iontype2-in-B = -1
1161 \end{verbatim}}
1162 The group name of solvent molecules acting as exchange partners for the ions
1163 has to be set with {\tt solvent-group}. The number of different ionic species under
1164 control of the CompEL protocol is given by the {\tt iontypes} parameter,
1165 while {\tt iontype0-name} gives the name of the index group containing the
1166 atoms of this ionic species. The reference number of ions of this type
1167 can be set with the {\tt iontype0-in-A} and {\tt iontype0-in-B} options
1168 for compartments {\bf A} and {\bf B}, respectively. Obviously,
1169 the sum of {\tt iontype0-in-A} and {\tt iontype0-in-B} needs to equal the number
1170 of ions in the group defined by {\tt iontype0-name}.
1171 A reference number of {\tt -1} means: use the number of ions as found at the beginning
1172 of the simulation as the reference value.
1174 {\small
1175 \begin{verbatim}
1176 coupl-steps = 10 ; Average over these many swap steps
1177 threshold = 1 ; Do not swap if < threshold
1178 \end{verbatim}}
1179 If {\tt coupl-steps} is set to 1, then the momentary ion distribution determines
1180 whether ions are exchanged. {\tt coupl-steps} \textgreater\ 1 will use the time-average
1181 of ion distributions over the selected number of attempt steps instead. This can be useful, for example,
1182 when ions diffuse near compartment boundaries, which would lead to numerous unproductive
1183 ion exchanges. A {\tt threshold} of 1 means that a swap is performed if the average ion
1184 count in a compartment differs by at least 1 from the requested values. Higher thresholds
1185 will lead to toleration of larger differences. Ions are exchanged until the requested
1186 number $\pm$ the threshold is reached.
1188 {\small
1189 \begin{verbatim}
1190 cyl0-r = 5.0 ; Split cylinder 0 radius (nm)
1191 cyl0-up = 0.75 ; Split cylinder 0 upper extension (nm)
1192 cyl0-down = 0.75 ; Split cylinder 0 lower extension (nm)
1193 cyl1-r = 5.0 ; same for other channel
1194 cyl1-up = 0.75
1195 cyl1-down = 0.75
1196 \end{verbatim}}
1197 The cylinder options are used to define virtual geometric cylinders around the
1198 channel's pore to track how many ions of which type have passed each channel.
1199 Ions will be counted as having traveled through a channel
1200 according to the definition of the channel's cylinder radius, upper and lower extension,
1201 relative to the location of the respective split group. This will not affect the actual
1202 flux or exchange, but will provide you with the ion permeation numbers across
1203 each of the channels. Note that an ion can only be counted as passing through a particular
1204 channel if it is detected \emph{within} the defined split cylinder in a swap step.
1205 If {\tt swap-frequency} is chosen too high, a particular ion may be detected in compartment {\bf A}
1206 in one swap step, and in compartment {\bf B} in the following swap step, so it will be unclear
1207 through which of the channels it has passed.
1209 A double-layered system for CompEL simulations can be easily prepared by
1210 duplicating an existing membrane/channel MD system in the direction of the membrane
1211 normal (typically $z$) with {\tt gmx editconf -translate 0 0 <l_z>}, where {\tt l_z}
1212 is the box length in that direction. If you have already defined index groups for
1213 the channel for the single-layered system, {\tt gmx make_ndx -n index.ndx -twin} will
1214 provide you with the groups for the double-layered system.
1216 To suppress large fluctuations of the membranes along the swap direction,
1217 it may be useful to apply a harmonic potential (acting only in the swap dimension)
1218 between each of the two channel and/or bilayer centers using umbrella pulling
1219 (see section~\ref{sec:pull}).
1221 \subsection*{Multimeric channels}
1222 If a split group consists of more than one molecule, the correct PBC image of all molecules
1223 with respect to each other has to be chosen such that the channel center can be correctly
1224 determined. \gromacs\ assumes that the starting structure in the {\tt .tpr}
1225 file has the correct PBC representation. Set the following environment variable
1226 to check whether that is the case:
1227 \begin{itemize}
1228 \item {\tt GMX_COMPELDUMP}: output the starting structure after it has been made whole to
1229 {\tt .pdb} file.
1230 \end{itemize}
1233 \section{Calculating a PMF using the free-energy code}
1234 \label{sec:fepmf}
1235 \index{potentials of mean force}
1236 \index{free energy calculations}
1237 The free-energy coupling-parameter approach (see~\secref{fecalc})
1238 provides several ways to calculate potentials of mean force.
1239 A potential of mean force between two atoms can be calculated
1240 by connecting them with a harmonic potential or a constraint.
1241 For this purpose there are special potentials that avoid the generation of
1242 extra exclusions, see~\secref{excl}.
1243 When the position of the minimum or the constraint length is 1 nm more
1244 in state B than in state A, the restraint or constraint force is given
1245 by $\partial H/\partial \lambda$.
1246 The distance between the atoms can be changed as a function of $\lambda$
1247 and time by setting {\tt delta-lambda} in the {\tt .mdp} file.
1248 The results should be identical (although not numerically
1249 due to the different implementations) to the results of the pull code
1250 with umbrella sampling and constraint pulling.
1251 Unlike the pull code, the free energy code can also handle atoms that
1252 are connected by constraints.
1254 Potentials of mean force can also be calculated using position restraints.
1255 With position restraints, atoms can be linked to a position in space
1256 with a harmonic potential (see \ssecref{positionrestraint}).
1257 These positions can be made a function of the coupling parameter $\lambda$.
1258 The positions for the A and the B states are supplied to {\tt grompp} with
1259 the {\tt -r} and {\tt -rb} options, respectively.
1260 One could use this approach to do \normindex{targeted MD};
1261 note that we do not encourage the use of targeted MD for proteins.
1262 A protein can be forced from one conformation to another by using
1263 these conformations as position restraint coordinates for state A and B.
1264 One can then slowly change $\lambda$ from 0 to 1.
1265 The main drawback of this approach is that the conformational freedom
1266 of the protein is severely limited by the position restraints,
1267 independent of the change from state A to B.
1268 Also, the protein is forced from state A to B in an almost straight line,
1269 whereas the real pathway might be very different.
1270 An example of a more fruitful application is a solid system or a liquid
1271 confined between walls where one wants to measure the force required
1272 to change the separation between the boundaries or walls.
1273 Because the boundaries (or walls) already need to be fixed,
1274 the position restraints do not limit the system in its sampling.
1276 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1277 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1278 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1279 \newcommand{\amine}{\sf -NH$_2$}
1280 \newcommand{\amines}{\sf -NH-}
1281 \newcommand{\aminep}{\sf -NH$_3^+$}
1282 \section{Removing fastest \swapindex{degrees of}{freedom}}
1283 \label{sec:rmfast}
1284 The maximum time step in MD simulations is limited by the smallest
1285 oscillation period that can be found in the simulated
1286 system. Bond-stretching vibrations are in their quantum-mechanical
1287 ground state and are therefore better represented by a constraint
1288 instead of a harmonic potential.
1290 For the remaining degrees of freedom, the shortest oscillation period
1291 (as measured from a simulation) is 13~fs for bond-angle vibrations
1292 involving hydrogen atoms. Taking as a guideline that with a Verlet
1293 (leap-frog) integration scheme a minimum of 5 numerical integration
1294 steps should be performed per period of a harmonic oscillation in
1295 order to integrate it with reasonable accuracy, the maximum time step
1296 will be about 3~fs. Disregarding these very fast oscillations of
1297 period 13~fs, the next shortest periods are around 20~fs, which will
1298 allow a maximum time step of about 4~fs.
1300 Removing the bond-angle degrees of freedom from hydrogen atoms can
1301 best be done by defining them as \normindex{virtual interaction sites}
1302 instead of normal atoms. Whereas a normal atom is connected to the molecule
1303 with bonds, angles and dihedrals, a virtual site's position is calculated
1304 from the position of three nearby heavy atoms in a predefined manner
1305 (see also \secref{virtual_sites}). For the hydrogens in water and in
1306 hydroxyl, sulfhydryl, or amine groups, no degrees of freedom can be
1307 removed, because rotational freedom should be preserved. The only
1308 other option available to slow down these motions is to increase the
1309 mass of the hydrogen atoms at the expense of the mass of the connected
1310 heavy atom. This will increase the moment of inertia of the water
1311 molecules and the hydroxyl, sulfhydryl, or amine groups, without
1312 affecting the equilibrium properties of the system and without
1313 affecting the dynamical properties too much. These constructions will
1314 shortly be described in \secref{vsitehydro} and have previously
1315 been described in full detail~\cite{feenstra99}.
1317 Using both virtual sites and \swapindex{modified}{mass}es, the next
1318 bottleneck is likely to be formed by the improper dihedrals (which are
1319 used to preserve planarity or chirality of molecular groups) and the
1320 peptide dihedrals. The peptide dihedral cannot be changed without
1321 affecting the physical behavior of the protein. The improper dihedrals
1322 that preserve planarity mostly deal with aromatic residues. Bonds,
1323 angles, and dihedrals in these residues can also be replaced with
1324 somewhat elaborate virtual site constructions.
1326 All modifications described in this section can be performed using the
1327 {\gromacs} topology building tool {\tt \normindex{pdb2gmx}}. Separate
1328 options exist to increase hydrogen masses, virtualize all hydrogen atoms,
1329 or also virtualize all aromatic residues. {\bf Note} that when all hydrogen
1330 atoms are virtualized, those inside the aromatic residues will be
1331 virtualized as well, {\ie} hydrogens in the aromatic residues are treated
1332 differently depending on the treatment of the aromatic residues.
1334 Parameters for the virtual site constructions for the hydrogen atoms are
1335 inferred from the force-field parameters ({\em vis}. bond lengths and
1336 angles) directly by {\tt \normindex{grompp}} while processing the
1337 topology file. The constructions for the aromatic residues are based
1338 on the bond lengths and angles for the geometry as described in the
1339 force fields, but these parameters are hard-coded into {\tt
1340 \normindex{pdb2gmx}} due to the complex nature of the construction
1341 needed for a whole aromatic group.
1343 \subsection{Hydrogen bond-angle vibrations}
1344 \label{sec:vsitehydro}
1345 \subsubsection{Construction of virtual sites} %%%%%%%%%%%%%%%%%%%%%%%%%
1346 \begin{figure}
1347 \centerline{\includegraphics[width=11cm]{plots/dumtypes}}
1348 \caption[Virtual site constructions for hydrogen atoms.]{The different
1349 types of virtual site constructions used for hydrogen atoms. The atoms
1350 used in the construction of the virtual site(s) are depicted as black
1351 circles, virtual sites as gray ones. Hydrogens are smaller than heavy
1352 atoms. {\sf A}: fixed bond angle, note that here the hydrogen is not a
1353 virtual site; {\sf B}: in the plane of three atoms, with fixed distance;
1354 {\sf C}: in the plane of three atoms, with fixed angle and distance;
1355 {\sf D}: construction for amine groups ({\amine} or {\aminep}), see
1356 text for details.}
1357 \label{fig:vsitehydro}
1358 \end{figure}
1360 The goal of defining hydrogen atoms as virtual sites is to remove all
1361 high-frequency degrees of freedom from them. In some cases, not all
1362 degrees of freedom of a hydrogen atom should be removed, {\eg} in the
1363 case of hydroxyl or amine groups the rotational freedom of the
1364 hydrogen atom(s) should be preserved. Care should be taken that no
1365 unwanted correlations are introduced by the construction of virtual
1366 sites, {\eg} bond-angle vibration between the constructing atoms could
1367 translate into hydrogen bond-length vibration. Additionally, since
1368 virtual sites are by definition massless, in order to preserve total
1369 system mass, the mass of each hydrogen atom that is treated as virtual
1370 site should be added to the bonded heavy atom.
1372 Taking into account these considerations, the hydrogen atoms in a
1373 protein naturally fall into several categories, each requiring a
1374 different approach (see also \figref{vsitehydro}).
1376 \begin{itemize}
1378 \item{\em hydroxyl ({\sf -OH}) or sulfhydryl ({\sf -SH})
1379 hydrogen:\/} The only internal degree of freedom in a hydroxyl group
1380 that can be constrained is the bending of the {\sf C-O-H} angle. This
1381 angle is fixed by defining an additional bond of appropriate length,
1382 see \figref{vsitehydro}A. Doing so removes the high-frequency angle bending,
1383 but leaves the dihedral rotational freedom. The same goes for a
1384 sulfhydryl group. {\bf Note} that in these cases the hydrogen is not treated
1385 as a virtual site.
1387 \item{\em single amine or amide ({\amines}) and aromatic hydrogens
1388 ({\sf -CH-}):\/} The position of these hydrogens cannot be constructed
1389 from a linear combination of bond vectors, because of the flexibility
1390 of the angle between the heavy atoms. Instead, the hydrogen atom is
1391 positioned at a fixed distance from the bonded heavy atom on a line
1392 going through the bonded heavy atom and a point on the line through
1393 both second bonded atoms, see \figref{vsitehydro}B.
1395 \item{\em planar amine ({\amine}) hydrogens:\/} The method used for
1396 the single amide hydrogen is not well suited for planar amine groups,
1397 because no suitable two heavy atoms can be found to define the
1398 direction of the hydrogen atoms. Instead, the hydrogen is constructed
1399 at a fixed distance from the nitrogen atom, with a fixed angle to the
1400 carbon atom, in the plane defined by one of the other heavy atoms, see
1401 \figref{vsitehydro}C.
1403 \item{\em amine group (umbrella {\amine} or {\aminep}) hydrogens:\/}
1404 Amine hydrogens with rotational freedom cannot be constructed as virtual
1405 sites from the heavy atoms they are connected to, since this would
1406 result in loss of the rotational freedom of the amine group. To
1407 preserve the rotational freedom while removing the hydrogen bond-angle
1408 degrees of freedom, two ``dummy masses'' are constructed with the same
1409 total mass, moment of inertia (for rotation around the {\sf C-N} bond)
1410 and center of mass as the amine group. These dummy masses have no
1411 interaction with any other atom, except for the fact that they are
1412 connected to the carbon and to each other, resulting in a rigid
1413 triangle. From these three particles, the positions of the nitrogen and
1414 hydrogen atoms are constructed as linear combinations of the two
1415 carbon-mass vectors and their outer product, resulting in an amine
1416 group with rotational freedom intact, but without other internal
1417 degrees of freedom. See \figref{vsitehydro}D.
1419 \end{itemize}
1421 \begin{figure}
1422 \centerline{\includegraphics[width=15cm]{plots/dumaro}}
1423 \caption[Virtual site constructions for aromatic residues.]{The
1424 different types of virtual site constructions used for aromatic
1425 residues. The atoms used in the construction of the virtual site(s) are
1426 depicted as black circles, virtual sites as gray ones. Hydrogens are
1427 smaller than heavy atoms. {\sf A}: phenylalanine; {\sf B}: tyrosine
1428 (note that the hydroxyl hydrogen is {\em not} a virtual site); {\sf C}:
1429 tryptophan; {\sf D}: histidine.}
1430 \label{fig:vistearo}
1431 \end{figure}
1433 \subsection{Out-of-plane vibrations in aromatic groups}
1434 \label{sec:vsitearo}
1435 The planar arrangements in the side chains of the aromatic residues
1436 lends itself perfectly to a virtual-site construction, giving a
1437 perfectly planar group without the inherently unstable constraints
1438 that are necessary to keep normal atoms in a plane. The basic approach
1439 is to define three atoms or dummy masses with constraints between them
1440 to fix the geometry and create the rest of the atoms as simple virtual
1441 sites type (see \secref{virtual_sites}) from these three. Each of
1442 the aromatic residues require a different approach:
1444 \begin{itemize}
1446 \item{\em Phenylalanine:\/} {\sf C}$_\gamma$, {\sf C}$_{{\epsilon}1}$,
1447 and {\sf C}$_{{\epsilon}2}$ are kept as normal atoms, but with each a
1448 mass of one third the total mass of the phenyl group. See
1449 \figref{vsitehydro}A.
1451 \item{\em Tyrosine:\/} The ring is treated identically to the
1452 phenylalanine ring. Additionally, constraints are defined between {\sf
1453 C}$_{{\epsilon}1}$, {\sf C}$_{{\epsilon}2}$, and {\sf O}$_{\eta}$.
1454 The original improper dihedral angles will keep both triangles (one
1455 for the ring and one with {\sf O}$_{\eta}$) in a plane, but due to the
1456 larger moments of inertia this construction will be much more
1457 stable. The bond-angle in the hydroxyl group will be constrained by a
1458 constraint between {\sf C}$_\gamma$ and {\sf H}$_{\eta}$. {\bf Note} that
1459 the hydrogen is not treated as a virtual site. See
1460 \figref{vsitehydro}B.
1462 \item{\em Tryptophan:\/} {\sf C}$_\beta$ is kept as a normal atom
1463 and two dummy masses are created at the center of mass of each of the
1464 rings, each with a mass equal to the total mass of the respective ring
1465 ({\sf C}$_{{\delta}2}$ and {\sf C}$_{{\epsilon}2}$ are each
1466 counted half for each ring). This keeps the overall center of mass and
1467 the moment of inertia almost (but not quite) equal to what it was. See
1468 \figref{vsitehydro}C.
1470 \item{\em Histidine:\/} {\sf C}$_\gamma$, {\sf C}$_{{\epsilon}1}$
1471 and {\sf N}$_{{\epsilon}2}$ are kept as normal atoms, but with masses
1472 redistributed such that the center of mass of the ring is
1473 preserved. See \figref{vsitehydro}D.
1475 \end{itemize}
1477 \section{Viscosity calculation\index{viscosity}}
1479 The shear viscosity is a property of liquids that can be determined easily
1480 by experiment. It is useful for parameterizing a force field
1481 because it is a kinetic property, while most other properties
1482 which are used for parameterization are thermodynamic.
1483 The viscosity is also an important property, since it influences
1484 the rates of conformational changes of molecules solvated in the liquid.
1486 The viscosity can be calculated from an equilibrium simulation using
1487 an Einstein relation:
1488 \beq
1489 \eta = \frac{1}{2}\frac{V}{k_B T} \lim_{t \rightarrow \infty}
1490 \frac{\mbox{d}}{\mbox{d} t} \left\langle
1491 \left( \int_{t_0}^{{t_0}+t} P_{xz}(t') \mbox{d} t' \right)^2
1492 \right\rangle_{t_0}
1493 \eeq
1494 This can be done with {\tt g_energy}.
1495 This method converges very slowly~\cite{Hess2002a}, and as such
1496 a nanosecond simulation might not
1497 be long enough for an accurate determination of the viscosity.
1498 The result is very dependent on the treatment of the electrostatics.
1499 Using a (short) cut-off results in large noise on the off-diagonal
1500 pressure elements, which can increase the calculated viscosity by an order
1501 of magnitude.
1503 {\gromacs} also has a non-equilibrium method for determining
1504 the viscosity~\cite{Hess2002a}.
1505 This makes use of the fact that energy, which is fed into system by
1506 external forces, is dissipated through viscous friction. The generated heat
1507 is removed by coupling to a heat bath. For a Newtonian liquid adding a
1508 small force will result in a velocity gradient according to the following
1509 equation:
1510 \beq
1511 a_x(z) + \frac{\eta}{\rho} \frac{\partial^2 v_x(z)}{\partial z^2} = 0
1512 \eeq
1513 Here we have applied an acceleration $a_x(z)$ in the $x$-direction, which
1514 is a function of the $z$-coordinate.
1515 In {\gromacs} the acceleration profile is:
1516 \beq
1517 a_x(z) = A \cos\left(\frac{2\pi z}{l_z}\right)
1518 \eeq
1519 where $l_z$ is the height of the box. The generated velocity profile is:
1520 \beq
1521 v_x(z) = V \cos\left(\frac{2\pi z}{l_z}\right)
1522 \eeq
1523 \beq
1524 V = A \frac{\rho}{\eta}\left(\frac{l_z}{2\pi}\right)^2
1525 \eeq
1526 The viscosity can be calculated from $A$ and $V$:
1527 \beq
1528 \label{visc}
1529 \eta = \frac{A}{V}\rho \left(\frac{l_z}{2\pi}\right)^2
1530 \eeq
1532 In the simulation $V$ is defined as:
1533 \beq
1534 V = \frac{\displaystyle \sum_{i=1}^N m_i v_{i,x} 2 \cos\left(\frac{2\pi z}{l_z}\right)}
1535 {\displaystyle \sum_{i=1}^N m_i}
1536 \eeq
1537 The generated velocity profile is not coupled to the heat bath. Moreover,
1538 the velocity profile is excluded from the kinetic energy.
1539 One would like $V$ to be as large as possible to get good statistics.
1540 However, the shear rate should not be so high that the system gets too far
1541 from equilibrium. The maximum shear rate occurs where the cosine is zero,
1542 the rate being:
1543 \beq
1544 \mbox{sh}_{\max} = \max_z \left| \frac{\partial v_x(z)}{\partial z} \right|
1545 = A \frac{\rho}{\eta} \frac{l_z}{2\pi}
1546 \eeq
1547 For a simulation with: $\eta=10^{-3}$ [kg\,m$^{-1}$\,s$^{-1}$],
1548 $\rho=10^3$\,[kg\,m$^{-3}$] and $l_z=2\pi$\,[nm],
1549 $\mbox{sh}_{\max}=1$\,[ps\,nm$^{-1}$] $A$.
1550 This shear rate should be smaller than one over the longest
1551 correlation time in the system. For most liquids, this will be the rotation
1552 correlation time, which is around 10 ps. In this case, $A$ should
1553 be smaller than 0.1\,[nm\,ps$^{-2}$].
1554 When the shear rate is too high, the observed viscosity will be too low.
1555 Because $V$ is proportional to the square of the box height,
1556 the optimal box is elongated in the $z$-direction.
1557 In general, a simulation length of 100 ps is enough to obtain an
1558 accurate value for the viscosity.
1560 The heat generated by the viscous friction is removed by coupling to a heat
1561 bath. Because this coupling is not instantaneous the real temperature of the
1562 liquid will be slightly lower than the observed temperature.
1563 Berendsen derived this temperature shift~\cite{Berendsen91}, which can
1564 be written in terms of the shear rate as:
1565 \beq
1566 T_s = \frac{\eta\,\tau}{2 \rho\,C_v} \mbox{sh}_{\max}^2
1567 \eeq
1568 where $\tau$ is the coupling time for the Berendsen thermostat and
1569 $C_v$ is the heat capacity. Using the values of the example above,
1570 $\tau=10^{-13}$ [s] and $C_v=2 \cdot 10^3$\,[J kg$^{-1}$\,K$^{-1}$], we
1571 get: $T_s=25$\,[K\,ps$^{-2}$]\,sh$_{\max}^2$. When we want the shear
1572 rate to be smaller than $1/10$\,[ps$^{-1}$], $T_s$ is smaller than
1573 0.25\,[K], which is negligible.
1575 {\bf Note} that the system has to build up the velocity profile when starting
1576 from an equilibrium state. This build-up time is of the order of the
1577 correlation time of the liquid.
1579 Two quantities are written to the energy file, along with their averages
1580 and fluctuations: $V$ and $1/\eta$, as obtained from (\ref{visc}).
1582 \section{Tabulated interaction functions\index{tabulated interaction functions}}
1583 \subsection{Cubic splines for potentials}
1584 \label{subsec:cubicspline}
1585 In some of the inner loops of {\gromacs}, look-up tables are used
1586 for computation of potential and forces.
1587 The tables are interpolated using a cubic
1588 spline algorithm.
1589 There are separate tables for electrostatic, dispersion, and repulsion
1590 interactions,
1591 but for the sake of caching performance these have been combined
1592 into a single array.
1593 The cubic spline interpolation for $x_i \leq x < x_{i+1}$ looks like this:
1594 \beq
1595 V_s(x) = A_0 + A_1 \,\epsilon + A_2 \,\epsilon^2 + A_3 \,\epsilon^3
1596 \label{eqn:spline}
1597 \eeq
1598 where the table spacing $h$ and fraction $\epsilon$ are given by:
1599 \bea
1600 h &=& x_{i+1} - x_i \\
1601 \epsilon&=& (x - x_i)/h
1602 \eea
1603 so that $0 \le \epsilon < 1$.
1604 From this, we can calculate the derivative in order to determine the forces:
1605 \beq
1606 -V_s'(x) ~=~
1607 -\frac{{\rm d}V_s(x)}{{\rm d}\epsilon}\frac{{\rm d}\epsilon}{{\rm d}x} ~=~
1608 -(A_1 + 2 A_2 \,\epsilon + 3 A_3 \,\epsilon^2)/h
1609 \eeq
1610 The four coefficients are determined from the four conditions
1611 that $V_s$ and $-V_s'$ at both ends of each interval should match
1612 the exact potential $V$ and force $-V'$.
1613 This results in the following errors for each interval:
1614 \bea
1615 |V_s - V |_{max} &=& V'''' \frac{h^4}{384} + O(h^5) \\
1616 |V_s' - V' |_{max} &=& V'''' \frac{h^3}{72\sqrt{3}} + O(h^4) \\
1617 |V_s''- V''|_{max} &=& V'''' \frac{h^2}{12} + O(h^3)
1618 \eea
1619 V and V' are continuous, while V'' is the first discontinuous
1620 derivative.
1621 The number of points per nanometer is 500 and 2000
1622 for mixed- and double-precision versions of {\gromacs}, respectively.
1623 This means that the errors in the potential and force will usually
1624 be smaller than the mixed precision accuracy.
1626 {\gromacs} stores $A_0$, $A_1$, $A_2$ and $A_3$.
1627 The force routines get a table with these four parameters and
1628 a scaling factor $s$ that is equal to the number of points per nm.
1629 ({\bf Note} that $h$ is $s^{-1}$).
1630 The algorithm goes a little something like this:
1631 \begin{enumerate}
1632 \item Calculate distance vector (\ve{r}$_{ij}$) and distance r$_{ij}$
1633 \item Multiply r$_{ij}$ by $s$ and truncate to an integer value $n_0$
1634 to get a table index
1635 \item Calculate fractional component ($\epsilon$ = $s$r$_{ij} - n_0$)
1636 and $\epsilon^2$
1637 \item Do the interpolation to calculate the potential $V$ and the scalar force $f$
1638 \item Calculate the vector force \ve{F} by multiplying $f$ with \ve{r}$_{ij}$
1639 \end{enumerate}
1641 {\bf Note} that table look-up is significantly {\em
1642 slower} than computation of the most simple Lennard-Jones and Coulomb
1643 interaction. However, it is much faster than the shifted Coulomb
1644 function used in conjunction with the PPPM method. Finally, it is much
1645 easier to modify a table for the potential (and get a graphical
1646 representation of it) than to modify the inner loops of the MD
1647 program.
1649 \subsection{User-specified potential functions}
1650 \label{subsec:userpot}
1651 You can also use your own potential functions\index{potential function} without
1652 editing the {\gromacs} code. The potential function should be according to the
1653 following equation
1654 \beq
1655 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})
1656 \eeq
1657 where $f$, $g$, and $h$ are user defined functions. {\bf Note} that if $g(r)$ represents a
1658 normal dispersion interaction, $g(r)$ should be $<$ 0. C$_6$, C$_{12}$
1659 and the charges are read from the topology. Also note that combination
1660 rules are only supported for Lennard-Jones and Buckingham, and that
1661 your tables should match the parameters in the binary topology.
1663 When you add the following lines in your {\tt .mdp} file:
1665 {\small
1666 \begin{verbatim}
1667 rlist = 1.0
1668 coulombtype = User
1669 rcoulomb = 1.0
1670 vdwtype = User
1671 rvdw = 1.0
1672 \end{verbatim}}
1674 {\tt mdrun} will read a single non-bonded table file,
1675 or multiple when {\tt energygrp-table} is set (see below).
1676 The name of the file(s) can be set with the {\tt mdrun} option {\tt -table}.
1677 The table file should contain seven columns of table look-up data in the
1678 order: $x$, $f(x)$, $-f'(x)$, $g(x)$, $-g'(x)$, $h(x)$, $-h'(x)$.
1679 The $x$ should run from 0 to $r_c+1$ (the value of {\tt table_extension} can be
1680 changed in the {\tt .mdp} file).
1681 You can choose the spacing you like; for the standard tables {\gromacs}
1682 uses a spacing of 0.002 and 0.0005 nm when you run in mixed
1683 and double precision, respectively. In this
1684 context, $r_c$ denotes the maximum of the two cut-offs {\tt rvdw} and
1685 {\tt rcoulomb} (see above). These variables need not be the same (and
1686 need not be 1.0 either). Some functions used for potentials contain a
1687 singularity at $x = 0$, but since atoms are normally not closer to each
1688 other than 0.1 nm, the function value at $x = 0$ is not important.
1689 Finally, it is also
1690 possible to combine a standard Coulomb with a modified LJ potential
1691 (or vice versa). One then specifies {\eg} {\tt coulombtype = Cut-off} or
1692 {\tt coulombtype = PME}, combined with {\tt vdwtype = User}. The table file must
1693 always contain the 7 columns however, and meaningful data (i.e. not
1694 zeroes) must be entered in all columns. A number of pre-built table
1695 files can be found in the {\tt GMXLIB} directory for 6-8, 6-9, 6-10, 6-11, and 6-12
1696 Lennard-Jones potentials combined with a normal Coulomb.
1698 If you want to have different functional forms between different
1699 groups of atoms, this can be set through energy groups.
1700 Different tables can be used for non-bonded interactions between
1701 different energy groups pairs through the {\tt .mdp} option {\tt energygrp-table}
1702 (see details in the User Guide).
1703 Atoms that should interact with a different potential should
1704 be put into different energy groups.
1705 Between group pairs which are not listed in {\tt energygrp-table},
1706 the normal user tables will be used. This makes it easy to use
1707 a different functional form between a few types of atoms.
1709 \section{Mixed Quantum-Classical simulation techniques}
1711 In a molecular mechanics (MM) force field, the influence of electrons
1712 is expressed by empirical parameters that are assigned on the basis of
1713 experimental data, or on the basis of results from high-level quantum
1714 chemistry calculations. These are valid for the ground state of a
1715 given covalent structure, and the MM approximation is usually
1716 sufficiently accurate for ground-state processes in which the overall
1717 connectivity between the atoms in the system remains
1718 unchanged. However, for processes in which the connectivity does
1719 change, such as chemical reactions, or processes that involve multiple
1720 electronic states, such as photochemical conversions, electrons can no
1721 longer be ignored, and a quantum mechanical description is required
1722 for at least those parts of the system in which the reaction takes
1723 place.
1725 One approach to the simulation of chemical reactions in solution, or
1726 in enzymes, is to use a combination of quantum mechanics (QM) and
1727 molecular mechanics (MM). The reacting parts of the system are treated
1728 quantum mechanically, with the remainder being modeled using the
1729 force field. The current version of {\gromacs} provides interfaces to
1730 several popular Quantum Chemistry packages (MOPAC~\cite{mopac},
1731 GAMESS-UK~\cite{gamess-uk}, Gaussian~\cite{g03} and CPMD~\cite{Car85a}).
1733 {\gromacs} interactions between the two subsystems are
1734 either handled as described by Field {\em et al.}~\cite{Field90a} or
1735 within the ONIOM approach by Morokuma and coworkers~\cite{Maseras96a,
1736 Svensson96a}.
1738 \subsection{Overview}
1740 Two approaches for describing the interactions between the QM and MM
1741 subsystems are supported in this version:
1743 \begin{enumerate}
1744 \item{\textbf{Electronic Embedding}} The electrostatic interactions
1745 between the electrons of the QM region and the MM atoms and between
1746 the QM nuclei and the MM atoms are included in the Hamiltonian for
1747 the QM subsystem: \beq H^{QM/MM} =
1748 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}},
1749 \eeq where $n$ and $N$ are the number of electrons and nuclei in the
1750 QM region, respectively, and $M$ is the number of charged MM
1751 atoms. The first term on the right hand side is the original
1752 electronic Hamiltonian of an isolated QM system. The first of the
1753 double sums is the total electrostatic interaction between the QM
1754 electrons and the MM atoms. The total electrostatic interaction of the
1755 QM nuclei with the MM atoms is given by the second double sum. Bonded
1756 interactions between QM and MM atoms are described at the MM level by
1757 the appropriate force-field terms. Chemical bonds that connect the two
1758 subsystems are capped by a hydrogen atom to complete the valence of
1759 the QM region. The force on this atom, which is present in the QM
1760 region only, is distributed over the two atoms of the bond. The cap
1761 atom is usually referred to as a link atom.
1763 \item{\textbf{ONIOM}} In the ONIOM approach, the energy and gradients
1764 are first evaluated for the isolated QM subsystem at the desired level
1765 of {\it{ab initio}} theory. Subsequently, the energy and gradients of
1766 the total system, including the QM region, are computed using the
1767 molecular mechanics force field and added to the energy and gradients
1768 calculated for the isolated QM subsystem. Finally, in order to correct
1769 for counting the interactions inside the QM region twice, a molecular
1770 mechanics calculation is performed on the isolated QM subsystem and
1771 the energy and gradients are subtracted. This leads to the following
1772 expression for the total QM/MM energy (and gradients likewise): \beq
1773 E_{tot} = E_{I}^{QM}
1774 +E_{I+II}^{MM}-E_{I}^{MM}, \eeq where the
1775 subscripts I and II refer to the QM and MM subsystems,
1776 respectively. The superscripts indicate at what level of theory the
1777 energies are computed. The ONIOM scheme has the
1778 advantage that it is not restricted to a two-layer QM/MM description,
1779 but can easily handle more than two layers, with each layer described
1780 at a different level of theory.
1781 \end{enumerate}
1783 \subsection{Usage}
1785 To make use of the QM/MM functionality in {\gromacs}, one needs to:
1787 \begin{enumerate}
1788 \item introduce link atoms at the QM/MM boundary, if needed;
1789 \item specify which atoms are to be treated at a QM level;
1790 \item specify the QM level, basis set, type of QM/MM interface and so on.
1791 \end{enumerate}
1793 \subsubsection{Adding link atoms}
1795 At the bond that connects the QM and MM subsystems, a link atoms is
1796 introduced. In {\gromacs} the link atom has special atomtype, called
1797 LA. This atomtype is treated as a hydrogen atom in the QM calculation,
1798 and as a virtual site in the force-field calculation. The link atoms, if
1799 any, are part of the system, but have no interaction with any other
1800 atom, except that the QM force working on it is distributed over the
1801 two atoms of the bond. In the topology, the link atom (LA), therefore,
1802 is defined as a virtual site atom:
1804 {\small
1805 \begin{verbatim}
1806 [ virtual_sites2 ]
1807 LA QMatom MMatom 1 0.65
1808 \end{verbatim}}
1810 See~\secref{vsitetop} for more details on how virtual sites are
1811 treated. The link atom is replaced at every step of the simulation.
1813 In addition, the bond itself is replaced by a constraint:
1815 {\small
1816 \begin{verbatim}
1817 [ constraints ]
1818 QMatom MMatom 2 0.153
1819 \end{verbatim}}
1821 {\bf Note} that, because in our system the QM/MM bond is a carbon-carbon
1822 bond (0.153 nm), we use a constraint length of 0.153 nm, and dummy
1823 position of 0.65. The latter is the ratio between the ideal C-H
1824 bond length and the ideal C-C bond length. With this ratio, the link
1825 atom is always 0.1 nm away from the {\tt QMatom}, consistent with the
1826 carbon-hydrogen bond length. If the QM and MM subsystems are connected
1827 by a different kind of bond, a different constraint and a different
1828 dummy position, appropriate for that bond type, are required.
1830 \subsubsection{Specifying the QM atoms}
1832 Atoms that should be treated at a QM level of theory, including the
1833 link atoms, are added to the index file. In addition, the chemical
1834 bonds between the atoms in the QM region are to be defined as
1835 connect bonds (bond type 5) in the topology file:
1837 {\small
1838 \begin{verbatim}
1839 [ bonds ]
1840 QMatom1 QMatom2 5
1841 QMatom2 QMatom3 5
1842 \end{verbatim}}
1844 \subsubsection{Specifying the QM/MM simulation parameters}
1846 In the {\tt .mdp} file, the following parameters control a QM/MM simulation.
1848 \begin{description}
1850 \item[\tt QMMM = no]\mbox{}\\ If this is set to {\tt yes}, a QM/MM
1851 simulation is requested. Several groups of atoms can be described at
1852 different QM levels separately. These are specified in the QMMM-grps
1853 field separated by spaces. The level of {\it{ab initio}} theory at which the
1854 groups are described is specified by {\tt QMmethod} and {\tt QMbasis}
1855 Fields. Describing the groups at different levels of theory is only
1856 possible with the ONIOM QM/MM scheme, specified by {\tt QMMMscheme}.
1858 \item[\tt QMMM-grps =]\mbox{}\\groups to be described at the QM level
1860 \item[\tt QMMMscheme = normal]\mbox{}\\Options are {\tt normal} and
1861 {\tt ONIOM}. This selects the QM/MM interface. {\tt normal} implies
1862 that the QM subsystem is electronically embedded in the MM
1863 subsystem. There can only be one {\tt QMMM-grps} that is modeled at
1864 the {\tt QMmethod} and {\tt QMbasis} level of {\it{ ab initio}}
1865 theory. The rest of the system is described at the MM level. The QM
1866 and MM subsystems interact as follows: MM point charges are included
1867 in the QM one-electron Hamiltonian and all Lennard-Jones interactions
1868 are described at the MM level. If {\tt ONIOM} is selected, the
1869 interaction between the subsystem is described using the ONIOM method
1870 by Morokuma and co-workers. There can be more than one QMMM-grps each
1871 modeled at a different level of QM theory (QMmethod and QMbasis).
1873 \item[\tt QMmethod = ]\mbox{}\\Method used to compute the energy
1874 and gradients on the QM atoms. Available methods are AM1, PM3, RHF,
1875 UHF, DFT, B3LYP, MP2, CASSCF, MMVB and CPMD. For CASSCF, the number of
1876 electrons and orbitals included in the active space is specified by
1877 {\tt CASelectrons} and {\tt CASorbitals}. For CPMD, the plane-wave
1878 cut-off is specified by the {\tt planewavecutoff} keyword.
1880 \item[\tt QMbasis = ]\mbox{}\\Gaussian basis set used to expand the
1881 electronic wave-function. Only Gaussian basis sets are currently
1882 available, i.e. STO-3G, 3-21G, 3-21G*, 3-21+G*, 6-21G, 6-31G, 6-31G*,
1883 6-31+G*, and 6-311G. For CPMD, which uses plane wave expansion rather
1884 than atom-centered basis functions, the {\tt planewavecutoff} keyword
1885 controls the plane wave expansion.
1887 \item[\tt QMcharge = ]\mbox{}\\The total charge in {\it{e}} of the {\tt
1888 QMMM-grps}. In case there are more than one {\tt QMMM-grps}, the total
1889 charge of each ONIOM layer needs to be specified separately.
1891 \item[\tt QMmult = ]\mbox{}\\The multiplicity of the {\tt
1892 QMMM-grps}. In case there are more than one {\tt QMMM-grps}, the
1893 multiplicity of each ONIOM layer needs to be specified separately.
1895 \item[\tt CASorbitals = ]\mbox{}\\The number of orbitals to be
1896 included in the active space when doing a CASSCF computation.
1898 \item[\tt CASelectrons = ]\mbox{}\\The number of electrons to be
1899 included in the active space when doing a CASSCF computation.
1901 \item[\tt SH = no]\mbox{}\\If this is set to yes, a QM/MM MD
1902 simulation on the excited state-potential energy surface and enforce a
1903 diabatic hop to the ground-state when the system hits the conical
1904 intersection hyperline in the course the simulation. This option only
1905 works in combination with the CASSCF method.
1907 \end{description}
1909 \subsection{Output}
1911 The energies and gradients computed in the QM calculation are added to
1912 those computed by {\gromacs}. In the {\tt .edr} file there is a section
1913 for the total QM energy.
1915 \subsection{Future developments}
1917 Several features are currently under development to increase the
1918 accuracy of the QM/MM interface. One useful feature is the use of
1919 delocalized MM charges in the QM computations. The most important
1920 benefit of using such smeared-out charges is that the Coulombic
1921 potential has a finite value at interatomic distances. In the point
1922 charge representation, the partially-charged MM atoms close to the QM
1923 region tend to ``over-polarize'' the QM system, which leads to artifacts
1924 in the calculation.
1926 What is needed as well is a transition state optimizer.
1928 \section{Using VMD plug-ins for trajectory file I/O}
1929 \index{VMD plug-ins}\index{trajectory file}{\gromacs} tools are able
1930 to use the plug-ins found in an existing installation of
1931 \href{http://www.ks.uiuc.edu/Research/vmd}{VMD} in order to read and
1932 write trajectory files in formats that are not native to
1933 {\gromacs}. You will be able to supply an AMBER DCD-format trajectory
1934 filename directly to {\gromacs} tools, for example.
1936 This requires a VMD installation not older than version 1.8, that your
1937 system provides the dlopen function so that programs can determine at
1938 run time what plug-ins exist, and that you build shared libraries when
1939 building {\gromacs}. CMake will find the vmd executable in your path, and
1940 from it, or the environment variable {\tt VMDDIR} at configuration or
1941 run time, locate the plug-ins. Alternatively, the {\tt VMD_PLUGIN_PATH}
1942 can be used at run time to specify a path where these plug-ins can be
1943 found. Note that these plug-ins are in a binary format, and that format
1944 must match the architecture of the machine attempting to use them.
1947 \section{\normindex{Interactive Molecular Dynamics}}
1948 {\gromacs} supports the interactive molecular dynamics (IMD) protocol as implemented
1949 by \href{http://www.ks.uiuc.edu/Research/vmd}{VMD} to control a running simulation
1950 in NAMD. IMD allows to monitor a running {\gromacs} simulation from a VMD client.
1951 In addition, the user can interact with the simulation by pulling on atoms, residues
1952 or fragments with a mouse or a force-feedback device. Additional information about
1953 the {\gromacs} implementation and an exemplary {\gromacs} IMD system can be found
1954 \href{http://www.mpibpc.mpg.de/grubmueller/interactivemd}{on this homepage}.
1956 \subsection{Simulation input preparation}
1957 The {\gromacs} implementation allows transmission and interaction with a part of the
1958 running simulation only, e.g.\ in cases where no water molecules should be transmitted
1959 or pulled. The group is specified via the {\tt .mdp} option {\tt IMD-group}. When
1960 {\tt IMD-group} is empty, the IMD protocol is disabled and cannot be enabled via the
1961 switches in {\tt mdrun}. To interact with the entire system, {\tt IMD-group} can
1962 be set to {\tt System}. When using {\tt grompp}, a {\tt .gro} file
1963 to be used as VMD input is written out ({\tt -imd} switch of {\tt grompp}).
1965 \subsection{Starting the simulation}
1966 Communication between VMD and {\gromacs} is achieved via TCP sockets and thus enables
1967 controlling an {\tt mdrun} running locally or on a remote cluster. The port for the
1968 connection can be specified with the {\tt -imdport} switch of {\tt mdrun}, 8888 is
1969 the default. If a port number of 0 or smaller is provided, {\gromacs} automatically
1970 assigns a free port to use with IMD.
1972 Every $N$ steps, the {\tt mdrun} client receives the applied forces from VMD and sends the new
1973 positions to the client. VMD permits increasing or decreasing the communication frequency
1974 interactively.
1975 By default, the simulation starts and runs even if no IMD client is connected. This
1976 behavior is changed by the {\tt -imdwait} switch of {\tt mdrun}. After startup and
1977 whenever the client has disconnected, the integration stops until reconnection of
1978 the client.
1979 When the {\tt -imdterm} switch is used, the simulation can be terminated by pressing
1980 the stop button in VMD. This is disabled by default.
1981 Finally, to allow interacting with the simulation (i.e. pulling from VMD) the {\tt -imdpull}
1982 switch has to be used.
1983 Therefore, a simulation can only be monitored but not influenced from the VMD client
1984 when none of {\tt -imdwait}, {\tt -imdterm} or {\tt -imdpull} are set. However, since
1985 the IMD protocol requires no authentication, it is not advisable to run simulations on
1986 a host directly reachable from an insecure environment. Secure shell forwarding of TCP
1987 can be used to connect to running simulations not directly reachable from the interacting host.
1988 Note that the IMD command line switches of {\tt mdrun} are hidden by default and show
1989 up in the help text only with {\tt gmx mdrun -h -hidden}.
1991 \subsection{Connecting from VMD}
1992 In VMD, first the structure corresponding to the IMD group has to be loaded ({\it File
1993 $\rightarrow$ New Molecule}). Then the IMD connection window has to be used
1994 ({\it Extensions $\rightarrow$ Simulation $\rightarrow$ IMD Connect (NAMD)}). In the IMD
1995 connection window, hostname and port have to be specified and followed by pressing
1996 {\it Connect}. {\it Detach Sim} allows disconnecting without terminating the simulation, while
1997 {\it Stop Sim} ends the simulation on the next neighbor searching step (if allowed by
1998 {\tt -imdterm}).
2000 The timestep transfer rate allows adjusting the communication frequency between simulation
2001 and IMD client. Setting the keep rate loads every $N^\mathrm{th}$ frame into VMD instead
2002 of discarding them when a new one is received. The displayed energies are in SI units
2003 in contrast to energies displayed from NAMD simulations.
2006 % LocalWords: PMF pmf kJ mol Jarzynski BT bilayer rup mdp AFM fepmf fecalc rb
2007 % LocalWords: posre grompp fs Verlet dihedrals hydrogens hydroxyl sulfhydryl
2008 % LocalWords: vsitehydro planarity chirality pdb gmx virtualize virtualized xz
2009 % LocalWords: vis massless tryptophan histidine phenyl parameterizing ij PPPM
2010 % LocalWords: parameterization Berendsen rlist coulombtype rcoulomb vdwtype LJ
2011 % LocalWords: rvdw energygrp mdrun pre GMXLIB MOPAC GAMESS CPMD ONIOM
2012 % LocalWords: Morokuma iJ AQ AJ initio atomtype QMatom MMatom QMMM grps et al
2013 % LocalWords: QMmethod QMbasis QMMMscheme RHF DFT LYP CASSCF MMVB CASelectrons
2014 % LocalWords: CASorbitals planewavecutoff STO QMcharge QMmult diabatic edr
2015 % LocalWords: hyperline delocalized Coulombic indices nm ccc th ps
2016 % LocalWords: GTX CPUs GHz md sd bd vv avek tcoupl andersen tc OPLSAA GROMOS
2017 % LocalWords: OBC obc CCMA tol pbc xyz barostat pcoupl acc gpu PLUGIN Cmake GX
2018 % LocalWords: MSVC gcc installpath cmake DGMX DCMAKE functionalities GPGPU GTS
2019 % LocalWords: OpenCL DeviceID gromacs gpus html GeForce Quadro FX Plex CX GF
2020 % LocalWords: VMD DCD GROningen MAchine BIOSON Groningen der Spoel Drunen Comp
2021 % LocalWords: Phys Comm moltype intramol vdw Waals fep multivector pf
2022 % LocalWords: pymbar dhdl xvg LINCS Entropic entropic solutes ref com iso pm
2023 % LocalWords: rm prefactors equipotential potiso potisopf potpm trr
2024 % LocalWords: potrm potrmpf midplanes midplane gaussians potflex vars massw av
2025 % LocalWords: shure observables rccccccc vec eps dist min eqn transl nstsout
2026 % LocalWords: nstrout rotangles rotslabs rottorque RMSD rmsdfit excl NH amine
2027 % LocalWords: positionrestraint es SH phenylalanine solvated sh nanometer QM
2028 % LocalWords: Lennard Buckingham UK Svensson ab vsitetop co UHF MP interatomic
2029 % LocalWords: cg grp coords SPC userpot
2030 % LocalWords: itp sitesn atomtypes ff csg ndx Tesla CHARMM AA gauss
2031 % LocalWords: CMAP nocmap fators Monte performant lib LD DIR llllcc
2032 % LocalWords: CMake homepage DEV overclocking GT dlopen vmd VMDDIR
2033 % LocalWords: versa PME atomperatom graining forcefields hy spc OPLS
2034 % LocalWords: topol multi