Remove cycle suppression
[gromacs.git] / docs / manual / algorithms.tex
blobbe04de79c9b588edb69ed3bff2b62cf7bd185420
2 % This file is part of the GROMACS molecular simulation package.
4 % Copyright (c) 2013,2014,2015,2016,2017,2018, 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 \newcommand{\nproc}{\mbox{$M$}}
36 \newcommand{\natom}{\mbox{$N$}}
37 \newcommand{\nx}{\mbox{$n_x$}}
38 \newcommand{\ny}{\mbox{$n_y$}}
39 \newcommand{\nz}{\mbox{$n_z$}}
40 \newcommand{\nsgrid}{NS grid}
41 \newcommand{\fftgrid}{FFT grid}
42 \newcommand{\dgrid}{\mbox{$\delta_{grid}$}}
43 \newcommand{\bfv}[1]{{\mbox{\boldmath{$#1$}}}}
44 % non-italicized boldface for math (e.g. matrices)
45 \newcommand{\bfm}[1]{{\bf #1}}
46 \newcommand{\dt}{\Delta t}
47 \newcommand{\rv}{\bfv{r}}
48 \newcommand{\vv}{\bfv{v}}
49 \newcommand{\F}{\bfv{F}}
50 \newcommand{\pb}{\bfv{p}}
51 \newcommand{\veps}{v_{\epsilon}}
52 \newcommand{\peps}{p_{\epsilon}}
53 \newcommand{\sinhx}[1]{\frac{\sinh{\left( #1\right)}}{#1}}
54 \chapter{Algorithms}
55 \label{ch:algorithms}
56 \section{Introduction}
57 In this chapter we first give describe some general concepts used in
58 {\gromacs}: {\em periodic boundary conditions} (\secref{pbc})
59 and the {\em group concept} (\secref{groupconcept}). The MD algorithm is
60 described in \secref{MD}: first a global form of the algorithm is
61 given, which is refined in subsequent subsections. The (simple) EM
62 (Energy Minimization) algorithm is described in \secref{EM}. Some
63 other algorithms for special purpose dynamics are described after
64 this.
66 A few issues are of general interest. In all cases the {\em system}
67 must be defined, consisting of molecules. Molecules again consist of
68 particles with defined interaction functions. The detailed
69 description of the {\em topology} of the molecules and of the {\em force
70 field} and the calculation of forces is given in
71 \chref{ff}. In the present chapter we describe
72 other aspects of the algorithm, such as pair list generation, update of
73 velocities and positions, coupling to external temperature and
74 pressure, conservation of constraints.
75 The {\em analysis} of the data generated by an MD simulation is treated in \chref{analysis}.
77 \section{Periodic boundary conditions\index{periodic boundary conditions}}
78 \label{sec:pbc}
79 \begin{figure}
80 \centerline{\includegraphics[width=9cm]{plots/pbctric}}
81 \caption {Periodic boundary conditions in two dimensions.}
82 \label{fig:pbc}
83 \end{figure}
84 The classical way to minimize edge effects in a finite system is to
85 apply {\em periodic boundary conditions}. The atoms of the system to
86 be simulated are put into a space-filling box, which is surrounded by
87 translated copies of itself (\figref{pbc}). Thus there are no
88 boundaries of the system; the artifact caused by unwanted boundaries
89 in an isolated cluster is now replaced by the artifact of periodic
90 conditions. If the system is crystalline, such boundary conditions are
91 desired (although motions are naturally restricted to periodic motions
92 with wavelengths fitting into the box). If one wishes to simulate
93 non-periodic systems, such as liquids or solutions, the periodicity by
94 itself causes errors. The errors can be evaluated by comparing various
95 system sizes; they are expected to be less severe than the errors
96 resulting from an unnatural boundary with vacuum.
98 There are several possible shapes for space-filling unit cells. Some,
99 like the {\em \normindex{rhombic dodecahedron}} and the
100 {\em \normindex{truncated octahedron}}~\cite{Adams79} are closer to being a sphere
101 than a cube is, and are therefore better suited to the
102 study of an approximately spherical macromolecule in solution, since
103 fewer solvent molecules are required to fill the box given a minimum
104 distance between macromolecular images. At the same time, rhombic
105 dodecahedra and truncated octahedra are special cases of {\em triclinic}
106 unit cells\index{triclinic unit cell}; the most general space-filling unit cells
107 that comprise all possible space-filling shapes~\cite{Bekker95}.
108 For this reason, {\gromacs} is based on the triclinic unit cell.
110 {\gromacs} uses periodic boundary conditions, combined with the {\em
111 \normindex{minimum image convention}}: only one -- the nearest -- image of each
112 particle is considered for short-range non-bonded interaction terms.
113 For long-range electrostatic interactions this is not always accurate
114 enough, and {\gromacs} therefore also incorporates lattice sum methods
115 such as Ewald Sum, PME and PPPM.
117 {\gromacs} supports triclinic boxes of any shape.
118 The simulation box (unit cell) is defined by the 3 box vectors
119 ${\bf a}$,${\bf b}$ and ${\bf c}$.
120 The box vectors must satisfy the following conditions:
121 \beq
122 \label{eqn:box_rot}
123 a_y = a_z = b_z = 0
124 \eeq
125 \beq
126 \label{eqn:box_shift1}
127 a_x>0,~~~~b_y>0,~~~~c_z>0
128 \eeq
129 \beq
130 \label{eqn:box_shift2}
131 |b_x| \leq \frac{1}{2} \, a_x,~~~~
132 |c_x| \leq \frac{1}{2} \, a_x,~~~~
133 |c_y| \leq \frac{1}{2} \, b_y
134 \eeq
135 Equations \ref{eqn:box_rot} can always be satisfied by rotating the box.
136 Inequalities (\ref{eqn:box_shift1}) and (\ref{eqn:box_shift2}) can always be
137 satisfied by adding and subtracting box vectors.
139 Even when simulating using a triclinic box, {\gromacs} always keeps the
140 particles in a brick-shaped volume for efficiency,
141 as illustrated in \figref{pbc} for a 2-dimensional system.
142 Therefore, from the output trajectory it might seem that the simulation was
143 done in a rectangular box. The program {\tt trjconv} can be used to convert
144 the trajectory to a different unit-cell representation.
146 It is also possible to simulate without periodic boundary conditions,
147 but it is usually more efficient to simulate an isolated cluster of molecules
148 in a large periodic box, since fast grid searching can only be used
149 in a periodic system.
151 \begin{figure}
152 \centerline{
153 \includegraphics[width=5cm]{plots/rhododec}
154 ~~~~\includegraphics[width=5cm]{plots/truncoct}
156 \caption {A rhombic dodecahedron and truncated octahedron
157 (arbitrary orientations).}
158 \label{fig:boxshapes}
159 \end{figure}
161 \subsection{Some useful box types}
162 \begin{table}
163 \centerline{
164 \begin{tabular}{|c|c|c|ccc|ccc|}
165 \dline
166 box type & image & box & \multicolumn{3}{c|}{box vectors} & \multicolumn{3}{c|}{box vector angles} \\
167 & distance & volume & ~{\bf a}~ & {\bf b} & {\bf c} &
168 $\angle{\bf bc}$ & $\angle{\bf ac}$ & $\angle{\bf ab}$ \\
169 \dline
170 & & & $d$ & 0 & 0 & & & \\
171 cubic & $d$ & $d^3$ & 0 & $d$ & 0 & $90^\circ$ & $90^\circ$ & $90^\circ$ \\
172 & & & 0 & 0 & $d$ & & & \\
173 \hline
174 rhombic & & & $d$ & 0 & $\frac{1}{2}\,d$ & & & \\
175 dodecahedron & $d$ & $\frac{1}{2}\sqrt{2}\,d^3$ & 0 & $d$ & $\frac{1}{2}\,d$ & $60^\circ$ & $60^\circ$ & $90^\circ$ \\
176 (xy-square) & & $0.707\,d^3$ & 0 & 0 & $\frac{1}{2}\sqrt{2}\,d$ & & & \\
177 \hline
178 rhombic & & & $d$ & $\frac{1}{2}\,d$ & $\frac{1}{2}\,d$ & & & \\
179 dodecahedron & $d$ & $\frac{1}{2}\sqrt{2}\,d^3$ & 0 & $\frac{1}{2}\sqrt{3}\,d$ & $\frac{1}{6}\sqrt{3}\,d$ & $60^\circ$ & $60^\circ$ & $60^\circ$ \\
180 (xy-hexagon) & & $0.707\,d^3$ & 0 & 0 & $\frac{1}{3}\sqrt{6}\,d$ & & & \\
181 \hline
182 truncated & & & $d$ & $\frac{1}{3}\,d$ & $-\frac{1}{3}\,d$ & & &\\
183 octahedron & $d$ & $\frac{4}{9}\sqrt{3}\,d^3$ & 0 & $\frac{2}{3}\sqrt{2}\,d$ & $\frac{1}{3}\sqrt{2}\,d$ & $71.53^\circ$ & $109.47^\circ$ & $71.53^\circ$ \\
184 & & $0.770\,d^3$ & 0 & 0 & $\frac{1}{3}\sqrt{6}\,d$ & & & \\
185 \dline
186 \end{tabular}
188 \caption{The cubic box, the rhombic \normindex{dodecahedron} and the truncated
189 \normindex{octahedron}.}
190 \label{tab:boxtypes}
191 \end{table}
192 The three most useful box types for simulations of solvated systems
193 are described in \tabref{boxtypes}. The rhombic dodecahedron
194 (\figref{boxshapes}) is the smallest and most regular space-filling
195 unit cell. Each of the 12 image cells is at the same distance. The
196 volume is 71\% of the volume of a cube having the same image
197 distance. This saves about 29\% of CPU-time when simulating a
198 spherical or flexible molecule in solvent. There are two different
199 orientations of a rhombic dodecahedron that satisfy equations
200 \ref{eqn:box_rot}, \ref{eqn:box_shift1} and \ref{eqn:box_shift2}.
201 The program {\tt editconf} produces the orientation
202 which has a square intersection with the xy-plane. This orientation
203 was chosen because the first two box vectors coincide with the x and
204 y-axis, which is easier to comprehend. The other orientation can be
205 useful for simulations of membrane proteins. In this case the
206 cross-section with the xy-plane is a hexagon, which has an area which
207 is 14\% smaller than the area of a square with the same image
208 distance. The height of the box ($c_z$) should be changed to obtain
209 an optimal spacing. This box shape not only saves CPU time, it
210 also results in a more uniform arrangement of the proteins.
212 \subsection{Cut-off restrictions}
213 The \normindex{minimum image convention} implies that the cut-off radius used to
214 truncate non-bonded interactions may not exceed half the shortest box
215 vector:
216 \beq
217 \label{eqn:physicalrc}
218 R_c < \half \min(\|{\bf a}\|,\|{\bf b}\|,\|{\bf c}\|),
219 \eeq
220 because otherwise more than one image would be within the cut-off distance
221 of the force. When a macromolecule, such as a protein, is studied in
222 solution, this restriction alone is not sufficient: in principle, a single
223 solvent molecule should not be able
224 to `see' both sides of the macromolecule. This means that the length of
225 each box vector must exceed the length of the macromolecule in the
226 direction of that edge {\em plus} two times the cut-off radius $R_c$.
227 It is, however, common to compromise in this respect, and make the solvent
228 layer somewhat smaller in order to reduce the computational cost.
229 For efficiency reasons the cut-off with triclinic boxes is more restricted.
230 For grid search the extra restriction is weak:
231 \beq
232 \label{eqn:gridrc}
233 R_c < \min(a_x,b_y,c_z)
234 \eeq
235 For simple search the extra restriction is stronger:
236 \beq
237 \label{eqn:simplerc}
238 R_c < \half \min(a_x,b_y,c_z)
239 \eeq
241 Each unit cell (cubic, rectangular or triclinic)
242 is surrounded by 26 translated images. A
243 particular image can therefore always be identified by an index pointing to one
244 of 27 {\em translation vectors} and constructed by applying a
245 translation with the indexed vector (see \ssecref{forces}).
246 Restriction (\ref{eqn:gridrc}) ensures that only 26 images need to be
247 considered.
249 \section{The group concept}
250 \label{sec:groupconcept}\index{group}
251 The {\gromacs} MD and analysis programs use user-defined {\em groups} of
252 atoms to perform certain actions on. The maximum number of groups is
253 256, but each atom can only belong to six different groups, one
254 each of the following:
255 \begin{description}
256 \item[\swapindex{temperature-coupling}{group}]
257 The \normindex{temperature coupling} parameters (reference
258 temperature, time constant, number of degrees of freedom, see
259 \ssecref{update}) can be defined for each T-coupling group
260 separately. For example, in a solvated macromolecule the solvent (that
261 tends to generate more heating by force and integration errors) can be
262 coupled with a shorter time constant to a bath than is a macromolecule,
263 or a surface can be kept cooler than an adsorbing molecule. Many
264 different T-coupling groups may be defined. See also center of mass
265 groups below.
267 \item[\swapindex{freeze}{group}\index{frozen atoms}]
268 Atoms that belong to a freeze group are kept stationary in the
269 dynamics. This is useful during equilibration, {\eg} to avoid badly
270 placed solvent molecules giving unreasonable kicks to protein atoms,
271 although the same effect can also be obtained by putting a restraining
272 potential on the atoms that must be protected. The freeze option can
273 be used, if desired, on just one or two coordinates of an atom,
274 thereby freezing the atoms in a plane or on a line. When an atom is
275 partially frozen, constraints will still be able to move it, even in a
276 frozen direction. A fully frozen atom can not be moved by constraints.
277 Many freeze groups can be defined. Frozen coordinates are unaffected
278 by pressure scaling; in some cases this can produce unwanted results,
279 particularly when constraints are also used (in this case you will
280 get very large pressures). Accordingly, it is recommended to avoid
281 combining freeze groups with constraints and pressure coupling. For the
282 sake of equilibration it could suffice to start with freezing in a
283 constant volume simulation, and afterward use position restraints in
284 conjunction with constant pressure.
286 \item[\swapindex{accelerate}{group}]
287 On each atom in an ``accelerate group'' an acceleration
288 $\ve{a}^g$ is imposed. This is equivalent to an external
289 force. This feature makes it possible to drive the system into a
290 non-equilibrium state and enables the performance of
291 \swapindex{non-equilibrium}{MD} and hence to obtain transport properties.
293 \item[\swapindex{energy-monitor}{group}]
294 Mutual interactions between all energy-monitor groups are compiled
295 during the simulation. This is done separately for Lennard-Jones and
296 Coulomb terms. In principle up to 256 groups could be defined, but
297 that would lead to 256$\times$256 items! Better use this concept
298 sparingly.
300 All non-bonded interactions between pairs of energy-monitor groups can
301 be excluded\index{exclusions} (see details in the User Guide).
302 Pairs of particles from excluded pairs of energy-monitor groups
303 are not put into the pair list.
304 This can result in a significant speedup
305 for simulations where interactions within or between parts of the system
306 are not required.
308 \item[\swapindex{center of mass}{group}\index{removing COM motion}]
309 In \gromacs\ the center of mass (COM) motion can be removed, for
310 either the complete system or for groups of atoms. The latter is
311 useful, {\eg} for systems where there is limited friction ({\eg} gas
312 systems) to prevent center of mass motion to occur. It makes sense to
313 use the same groups for temperature coupling and center of mass motion
314 removal.
316 \item[\swapindex{Compressed position output}{group}]
318 In order to further reduce the size of the compressed trajectory file
319 ({\tt .xtc{\index{XTC}}} or {\tt .tng{\index{TNG}}}), it is possible
320 to store only a subset of all particles. All x-compression groups that
321 are specified are saved, the rest are not. If no such groups are
322 specified, than all atoms are saved to the compressed trajectory file.
324 \end{description}
325 The use of groups in {\gromacs} tools is described in
326 \secref{usinggroups}.
328 \section{Molecular Dynamics}
329 \label{sec:MD}
330 \begin{figure}
331 \begin{center}
332 \addtolength{\fboxsep}{0.5cm}
333 \begin{shadowenv}[12cm]
334 {\large \bf THE GLOBAL MD ALGORITHM}
335 \rule{\textwidth}{2pt} \\
336 {\bf 1. Input initial conditions}\\[2ex]
337 Potential interaction $V$ as a function of atom positions\\
338 Positions $\ve{r}$ of all atoms in the system\\
339 Velocities $\ve{v}$ of all atoms in the system \\
340 $\Downarrow$\\
341 \rule{\textwidth}{1pt}\\
342 {\bf repeat 2,3,4} for the required number of steps:\\
343 \rule{\textwidth}{1pt}\\
344 {\bf 2. Compute forces} \\[1ex]
345 The force on any atom \\[1ex]
346 $\ve{F}_i = - \displaystyle\frac{\partial V}{\partial \ve{r}_i}$ \\[1ex]
347 is computed by calculating the force between non-bonded atom pairs: \\
348 $\ve{F}_i = \sum_j \ve{F}_{ij}$ \\
349 plus the forces due to bonded interactions (which may depend on 1, 2,
350 3, or 4 atoms), plus restraining and/or external forces. \\
351 The potential and kinetic energies and the pressure tensor may be computed. \\
352 $\Downarrow$\\
353 {\bf 3. Update configuration} \\[1ex]
354 The movement of the atoms is simulated by numerically solving Newton's
355 equations of motion \\[1ex]
356 $\displaystyle
357 \frac {\de^2\ve{r}_i}{\de t^2} = \frac{\ve{F}_i}{m_i} $ \\
358 or \\
359 $\displaystyle
360 \frac{\de\ve{r}_i}{\de t} = \ve{v}_i ; \;\;
361 \frac{\de\ve{v}_i}{\de t} = \frac{\ve{F}_i}{m_i} $ \\[1ex]
362 $\Downarrow$ \\
363 {\bf 4.} if required: {\bf Output step} \\
364 write positions, velocities, energies, temperature, pressure, etc. \\
365 \end{shadowenv}
366 \caption{The global MD algorithm}
367 \label{fig:global}
368 \end{center}
369 \end{figure}
370 A global flow scheme for MD is given in \figref{global}. Each
371 MD or EM run requires as input a set of initial coordinates and --
372 optionally -- initial velocities of all particles involved. This
373 chapter does not describe how these are obtained; for the setup of an
374 actual MD run check the online manual at {\wwwpage}.
376 \subsection{Initial conditions}
377 \subsubsection{Topology and force field}
378 The system topology, including a description of the force field, must
379 be read in.
380 Force fields and topologies are described in \chref{ff}
381 and \ref{ch:top}, respectively.
382 All this information is static; it is never modified during the run.
384 \subsubsection{Coordinates and velocities}
385 \begin{figure}
386 \centerline{\includegraphics[width=8cm]{plots/maxwell}}
387 \caption{A Maxwell-Boltzmann velocity distribution, generated from
388 random numbers.}
389 \label{fig:maxwell}
390 \end{figure}
392 Then, before a run starts, the box size and the coordinates and
393 velocities of all particles are required. The box size and shape is
394 determined by three vectors (nine numbers) $\ve{b}_1, \ve{b}_2, \ve{b}_3$,
395 which represent the three basis vectors of the periodic box.
397 If the run starts at $t=t_0$, the coordinates at $t=t_0$ must be
398 known. The {\em leap-frog algorithm}, the default algorithm used to
399 update the time step with $\Dt$ (see \ssecref{update}), also requires
400 that the velocities at $t=t_0 - \hDt$ are known. If velocities are not
401 available, the program can generate initial atomic velocities
402 $v_i, i=1\ldots 3N$ with a \index{Maxwell-Boltzmann distribution}
403 (\figref{maxwell}) at a given absolute temperature $T$:
404 \beq
405 p(v_i) = \sqrt{\frac{m_i}{2 \pi kT}}\exp\left(-\frac{m_i v_i^2}{2kT}\right)
406 \eeq
407 where $k$ is Boltzmann's constant (see \chref{defunits}).
408 To accomplish this, normally distributed random numbers are generated
409 by adding twelve random numbers $R_k$ in the range $0 \le R_k < 1$ and
410 subtracting 6.0 from their sum. The result is then multiplied by the
411 standard deviation of the velocity distribution $\sqrt{kT/m_i}$. Since
412 the resulting total energy will not correspond exactly to the required
413 temperature $T$, a correction is made: first the center-of-mass motion
414 is removed and then all velocities are scaled so that the total
415 energy corresponds exactly to $T$ (see \eqnref{E-T}).
416 % Why so complicated? What's wrong with Box-Mueller transforms?
418 \subsubsection{Center-of-mass motion\index{removing COM motion}}
419 The \swapindex{center-of-mass}{velocity} is normally set to zero at
420 every step; there is (usually) no net external force acting on the
421 system and the center-of-mass velocity should remain constant. In
422 practice, however, the update algorithm introduces a very slow change in
423 the center-of-mass velocity, and therefore in the total kinetic energy of
424 the system -- especially when temperature coupling is used. If such
425 changes are not quenched, an appreciable center-of-mass motion
426 can develop in long runs, and the temperature will be
427 significantly misinterpreted. Something similar may happen due to overall
428 rotational motion, but only when an isolated cluster is simulated. In
429 periodic systems with filled boxes, the overall rotational motion is
430 coupled to other degrees of freedom and does not cause such problems.
433 \subsection{Neighbor searching\swapindexquiet{neighbor}{searching}}
434 \label{subsec:ns}
435 As mentioned in \chref{ff}, internal forces are
436 either generated from fixed (static) lists, or from dynamic lists.
437 The latter consist of non-bonded interactions between any pair of particles.
438 When calculating the non-bonded forces, it is convenient to have all
439 particles in a rectangular box.
440 As shown in \figref{pbc}, it is possible to transform a
441 triclinic box into a rectangular box.
442 The output coordinates are always in a rectangular box, even when a
443 dodecahedron or triclinic box was used for the simulation.
444 Equation \ref{eqn:box_rot} ensures that we can reset particles
445 in a rectangular box by first shifting them with
446 box vector ${\bf c}$, then with ${\bf b}$ and finally with ${\bf a}$.
447 Equations \ref{eqn:box_shift2}, \ref{eqn:physicalrc} and \ref{eqn:gridrc}
448 ensure that we can find the 14 nearest triclinic images within
449 a linear combination that does not involve multiples of box vectors.
451 \subsubsection{Pair lists generation}
452 The non-bonded pair forces need to be calculated only for those pairs
453 $i,j$ for which the distance $r_{ij}$ between $i$ and the
454 \swapindex{nearest}{image}
455 of $j$ is less than a given cut-off radius $R_c$. Some of the particle
456 pairs that fulfill this criterion are excluded, when their interaction
457 is already fully accounted for by bonded interactions. {\gromacs}
458 employs a {\em pair list} that contains those particle pairs for which
459 non-bonded forces must be calculated. The pair list contains particles
460 $i$, a displacement vector for particle $i$, and all particles $j$ that
461 are within \verb'rlist' of this particular image of particle $i$. The
462 list is updated every \verb'nstlist' steps.
464 To make the \normindex{neighbor list}, all particles that are close
465 ({\ie} within the neighbor list cut-off) to a given particle must be found.
466 This searching, usually called neighbor search (NS) or pair search,
467 involves periodic boundary conditions and determining the {\em image}
468 (see \secref{pbc}). The search algorithm is $O(N)$, although a simpler
469 $O(N^2)$ algorithm is still available under some conditions.
471 \subsubsection{\normindex{Cut-off schemes}: group versus Verlet}
472 From version 4.6, {\gromacs} supports two different cut-off scheme
473 setups: the original one based on particle groups and one using a Verlet
474 buffer. There are some important differences that affect results,
475 performance and feature support. The group scheme can be made to work
476 (almost) like the Verlet scheme, but this will lead to a decrease in
477 performance. The group scheme is especially fast for water molecules,
478 which are abundant in many simulations, but on the most recent x86
479 processors, this advantage is negated by the better instruction-level
480 parallelism available in the Verlet-scheme implementation. The group
481 scheme is deprecated in version 5.0, and will be removed in a future
482 version. For practical details of choosing and setting up
483 cut-off schemes, please see the User Guide.
485 In the group scheme, a neighbor list is generated consisting of pairs
486 of groups of at least one particle. These groups were originally
487 \swapindex{charge}{group}s (see
488 \secref{chargegroup}), but with a proper treatment of long-range
489 electrostatics, performance in unbuffered simulations is their only advantage. A pair of groups
490 is put into the neighbor list when their center of geometry is within
491 the cut-off distance. Interactions between all particle pairs (one from
492 each charge group) are calculated for a certain number of MD steps,
493 until the neighbor list is updated. This setup is efficient, as the
494 neighbor search only checks distance between charge-group pair, not
495 particle pairs (saves a factor of $3 \times 3 = 9$ with a three-particle water
496 model) and the non-bonded force kernels can be optimized for, say, a
497 water molecule ``group''. Without explicit buffering, this setup leads
498 to energy drift as some particle pairs which are within the cut-off don't
499 interact and some outside the cut-off do interact. This can be caused
501 \begin{itemize}
502 \item particles moving across the cut-off between neighbor search steps, and/or
503 \item for charge groups consisting of more than one particle, particle pairs
504 moving in/out of the cut-off when their charge group center of
505 geometry distance is outside/inside of the cut-off.
506 \end{itemize}
507 Explicitly adding a buffer to the neighbor list will remove such
508 artifacts, but this comes at a high computational cost. How severe the
509 artifacts are depends on the system, the properties in which you are
510 interested, and the cut-off setup.
512 The Verlet cut-off scheme uses a buffered pair list by default. It
513 also uses clusters of particles, but these are not static as in the group
514 scheme. Rather, the clusters are defined spatially and consist of 4 or
515 8 particles, which is convenient for stream computing, using e.g. SSE, AVX
516 or CUDA on GPUs. At neighbor search steps, a pair list is created
517 with a Verlet buffer, ie. the pair-list cut-off is larger than the
518 interaction cut-off. In the non-bonded kernels, interactions are only
519 computed when a particle pair is within the cut-off distance at that
520 particular time step. This ensures that as particles move between pair
521 search steps, forces between nearly all particles within the cut-off
522 distance are calculated. We say {\em nearly} all particles, because
523 {\gromacs} uses a fixed pair list update frequency for
524 efficiency. A particle-pair, whose distance was outside the cut-off,
525 could possibly move enough during this fixed number of
526 steps that its distance is now within the cut-off. This
527 small chance results in a small energy drift, and the size of the
528 chance depends on the temperature. When temperature
529 coupling is used, the buffer size can be determined automatically,
530 given a certain tolerance on the energy drift.
532 The Verlet cut-off scheme is implemented in a very efficient fashion
533 based on clusters of particles. The simplest example is a cluster size
534 of 4 particles. The pair list is then constructed based on cluster
535 pairs. The cluster-pair search is much faster searching based on
536 particle pairs, because $4 \times 4 = 16$ particle pairs are put in
537 the list at once. The non-bonded force calculation kernel can then
538 calculate many particle-pair interactions at once, which maps nicely
539 to SIMD or SIMT units on modern hardware, which can perform multiple
540 floating operations at once. These non-bonded kernels
541 are much faster than the kernels used in the group scheme for most
542 types of systems, particularly on newer hardware.
544 Additionally, when the list buffer is determined automatically as
545 described below, we also apply dynamic pair list pruning. The pair list
546 can be constructed infrequently, but that can lead to a lot of pairs
547 in the list that are outside the cut-off range for all or most of
548 the life time of this pair list. Such pairs can be pruned out by
549 applying a cluster-pair kernel that only determines which clusters
550 are in range. Because of the way the non-bonded data is regularized
551 in {\gromacs}, this kernel is an order of magnitude faster than
552 the search and the interaction kernel. On the GPU this pruning is
553 overlapped with the integration on the CPU, so it is free in most
554 cases. Therefore we can prune every 4-10 integration steps with
555 little overhead and significantly reduce the number of cluster pairs
556 in the interaction kernel. This procedure is applied automatically,
557 unless the user set the pair-list buffer size manually.
559 \subsubsection{Energy drift and pair-list buffering}
560 For a canonical (NVT) ensemble, the average energy error caused by
561 diffusion of $j$ particles from outside the pair-list cut-off
562 $r_\ell$ to inside the interaction cut-off $r_c$ over the lifetime
563 of the list can be determined from the atomic
564 displacements and the shape of the potential at the cut-off.
565 %Since we are interested in the small drift regime, we will assume
566 %#that atoms will only move within the cut-off distance in the last step,
567 %$n_\mathrm{ps}-1$, of the pair list update interval $n_\mathrm{ps}$.
568 %Over this number of steps the displacment of an atom with mass $m$
569 The displacement distribution along one dimension for a freely moving
570 particle with mass $m$ over time $t$ at temperature $T$ is
571 a Gaussian $G(x)$
572 of zero mean and variance $\sigma^2 = t^2 k_B T/m$. For the distance
573 between two particles, the variance changes to $\sigma^2 = \sigma_{12}^2 =
574 t^2 k_B T(1/m_1+1/m_2)$. Note that in practice particles usually
575 interact with (bump into) other particles over time $t$ and therefore the real
576 displacement distribution is much narrower. Given a non-bonded
577 interaction cut-off distance of $r_c$ and a pair-list cut-off
578 $r_\ell=r_c+r_b$ for $r_b$ the Verlet buffer size, we can then
579 write the average energy error after time $t$ for all missing pair
580 interactions between a single $i$ particle of type 1 surrounded
581 by all $j$ particles that are of type 2 with number density $\rho_2$,
582 when the inter-particle distance changes from $r_0$ to $r_t$, as:
583 \beq
584 \langle \Delta V \rangle =
585 \int_{0}^{r_c} \int_{r_\ell}^\infty 4 \pi r_0^2 \rho_2 V(r_t) G\!\left(\frac{r_t-r_0}{\sigma}\right) d r_0\, d r_t
586 \eeq
587 To evaluate this analytically, we need to make some approximations. First we replace $V(r_t)$ by a Taylor expansion around $r_c$, then we can move the lower bound of the integral over $r_0$ to $-\infty$ which will simplify the result:
588 \begin{eqnarray}
589 \langle \Delta V \rangle &\approx&
590 \int_{-\infty}^{r_c} \int_{r_\ell}^\infty 4 \pi r_0^2 \rho_2 \Big[ V'(r_c) (r_t - r_c) +
591 \nonumber\\
593 \phantom{\int_{-\infty}^{r_c} \int_{r_\ell}^\infty 4 \pi r_0^2 \rho_2 \Big[}
594 V''(r_c)\frac{1}{2}(r_t - r_c)^2 +
595 \nonumber\\
597 \phantom{\int_{-\infty}^{r_c} \int_{r_\ell}^\infty 4 \pi r_0^2 \rho_2 \Big[}
598 V'''(r_c)\frac{1}{6}(r_t - r_c)^3 +
599 \nonumber\\
601 \phantom{\int_{-\infty}^{r_c} \int_{r_\ell}^\infty 4 \pi r_0^2 \rho_2 \Big[}
602 O \! \left((r_t - r_c)^4 \right)\Big] G\!\left(\frac{r_t-r_0}{\sigma}\right) d r_0 \, d r_t
603 \end{eqnarray}
604 Replacing the factor $r_0^2$ by $(r_\ell + \sigma)^2$, which results in a slight overestimate, allows us to calculate the integrals analytically:
605 \begin{eqnarray}
606 \langle \Delta V \rangle \!
607 &\approx&
608 4 \pi (r_\ell+\sigma)^2 \rho_2
609 \int_{-\infty}^{r_c} \int_{r_\ell}^\infty \Big[ V'(r_c) (r_t - r_c) +
610 \nonumber\\
612 \phantom{4 \pi (r_\ell+\sigma)^2 \rho_2 \int_{-\infty}^{r_c} \int_{r_\ell}^\infty \Big[}
613 V''(r_c)\frac{1}{2}(r_t - r_c)^2 +
614 \nonumber\\
616 \phantom{4 \pi (r_\ell+\sigma)^2 \rho_2 \int_{-\infty}^{r_c} \int_{r_\ell}^\infty \Big[}
617 V'''(r_c)\frac{1}{6}(r_t - r_c)^3 \Big] G\!\left(\frac{r_t-r_0}{\sigma}\right)
618 d r_0 \, d r_t\\
620 4 \pi (r_\ell+\sigma)^2 \rho_2 \bigg\{
621 \frac{1}{2}V'(r_c)\left[r_b \sigma G\!\left(\frac{r_b}{\sigma}\right) - (r_b^2+\sigma^2)E\!\left(\frac{r_b}{\sigma}\right) \right] +
622 \nonumber\\
624 \phantom{4 \pi (r_\ell+\sigma)^2 \rho_2 \bigg\{ }
625 \frac{1}{6}V''(r_c)\left[ \sigma(r_b^2+2\sigma^2) G\!\left(\frac{r_b}{\sigma}\right) - r_b(r_b^2+3\sigma^2 ) E\!\left(\frac{r_b}{\sigma}\right) \right] +
626 \nonumber\\
628 \phantom{4 \pi (r_\ell+\sigma)^2 \rho_2 \bigg\{ }
629 \frac{1}{24}V'''(r_c)\bigg[ r_b\sigma(r_b^2+5\sigma^2) G\!\left(\frac{r_b}{\sigma}\right)
630 \nonumber\\
632 \phantom{4 \pi (r_\ell+\sigma)^2 \rho_2 \bigg\{ \frac{1}{24}V'''(r_c)\bigg[ }
633 - (r_b^4+6r_b^2\sigma^2+3\sigma^4 ) E\!\left(\frac{r_b}{\sigma}\right) \bigg]
634 \bigg\}
635 \end{eqnarray}
637 where $G(x)$ is a Gaussian distribution with 0 mean and unit variance and
638 $E(x)=\frac{1}{2}\mathrm{erfc}(x/\sqrt{2})$. We always want to achieve
639 small energy error, so $\sigma$ will be small compared to both $r_c$
640 and $r_\ell$, thus the approximations in the equations above are good,
641 since the Gaussian distribution decays rapidly. The energy error needs
642 to be averaged over all particle pair types and weighted with the
643 particle counts. In {\gromacs} we don't allow cancellation of error
644 between pair types, so we average the absolute values. To obtain the
645 average energy error per unit time, it needs to be divided by the
646 neighbor-list life time $t = ({\tt nstlist} - 1)\times{\tt dt}$. The
647 function can not be inverted analytically, so we use bisection to
648 obtain the buffer size $r_b$ for a target drift. Again we note that
649 in practice the error we usually be much smaller than this estimate,
650 as in the condensed phase particle displacements will be much smaller
651 than for freely moving particles, which is the assumption used here.
653 When (bond) constraints are present, some particles will have fewer
654 degrees of freedom. This will reduce the energy errors. For simplicity,
655 we only consider one constraint per particle, the heaviest particle
656 in case a particle is involved in multiple constraints.
657 This simplification overestimates the displacement. The motion of
658 a constrained particle is a superposition of the 3D motion of the
659 center of mass of both particles and a 2D rotation around the center of mass.
660 The displacement in an arbitrary direction of a particle with 2 degrees
661 of freedom is not Gaussian, but rather follows the complementary error
662 function:
663 \beq
664 \frac{\sqrt{\pi}}{2\sqrt{2}\sigma}\,\mathrm{erfc}\left(\frac{|r|}{\sqrt{2}\,\sigma}\right)
665 \label{eqn:2D_disp}
666 \eeq
667 where $\sigma^2$ is again $t^2 k_B T/m$. This distribution can no
668 longer be integrated analytically to obtain the energy error. But we
669 can generate a tight upper bound using a scaled and shifted Gaussian
670 distribution (not shown). This Gaussian distribution can then be used
671 to calculate the energy error as described above. The rotation displacement
672 around the center of mass can not be more than the length of the arm.
673 To take this into account, we scale $\sigma$ in \eqnref{2D_disp} (details
674 not presented here) to obtain an overestimate of the real displacement.
675 This latter effect significantly reduces the buffer size for longer
676 neighborlist lifetimes in e.g. water, as constrained hydrogens are by far
677 the fastest particles, but they can not move further than 0.1 nm
678 from the heavy atom they are connected to.
681 There is one important implementation detail that reduces the energy
682 errors caused by the finite Verlet buffer list size. The derivation
683 above assumes a particle pair-list. However, the {\gromacs}
684 implementation uses a cluster pair-list for efficiency. The pair list
685 consists of pairs of clusters of 4 particles in most cases, also
686 called a $4 \times 4$ list, but the list can also be $4 \times 8$ (GPU
687 CUDA kernels and AVX 256-bit single precision kernels) or $4 \times 2$
688 (SSE double-precision kernels). This means that the pair-list is
689 effectively much larger than the corresponding $1 \times 1$ list. Thus
690 slightly beyond the pair-list cut-off there will still be a large
691 fraction of particle pairs present in the list. This fraction can be
692 determined in a simulation and accurately estimated under some
693 reasonable assumptions. The fraction decreases with increasing
694 pair-list range, meaning that a smaller buffer can be used. For
695 typical all-atom simulations with a cut-off of 0.9 nm this fraction is
696 around 0.9, which gives a reduction in the energy errors of a factor of
697 10. This reduction is taken into account during the automatic Verlet
698 buffer calculation and results in a smaller buffer size.
700 \begin{figure}
701 \centerline{\includegraphics[width=9cm]{plots/verlet-drift}}
702 \caption {Energy drift per atom for an SPC/E water system at 300K with
703 a time step of 2 fs and a pair-list update period of 10 steps
704 (pair-list life time: 18 fs). PME was used with {\tt ewald-rtol} set
705 to 10$^{-5}$; this parameter affects the shape of the potential at
706 the cut-off. Error estimates due to finite Verlet buffer size are
707 shown for a $1 \times 1$ atom pair list and $4 \times 4$ atom pair
708 list without and with (dashed line) cancellation of positive and
709 negative errors. Real energy drift is shown for simulations using
710 double- and mixed-precision settings. Rounding errors in the SETTLE
711 constraint algorithm from the use of single precision causes
712 the drift to become negative
713 at large buffer size. Note that at zero buffer size, the real drift
714 is small because positive (H-H) and negative (O-H) energy errors
715 cancel.}
716 \label{fig:verletdrift}
717 \end{figure}
719 In \figref{verletdrift} one can see that for small buffer sizes the drift
720 of the total energy is much smaller than the pair energy error tolerance,
721 due to cancellation of errors. For larger buffer size, the error estimate
722 is a factor of 6 higher than drift of the total energy, or alternatively
723 the buffer estimate is 0.024 nm too large. This is because the protons
724 don't move freely over 18 fs, but rather vibrate.
725 %At a buffer size of zero there is cancellation of
726 %drift due to repulsive (H-H) and attractive (O-H) interactions.
728 \subsubsection{Cut-off artifacts and switched interactions}
729 With the Verlet scheme, the pair potentials are shifted to be zero at
730 the cut-off, which makes the potential the integral of the force.
731 This is only possible in the group scheme if the shape of the potential
732 is such that its value is zero at the cut-off distance.
733 However, there can still be energy drift when the
734 forces are non-zero at the cut-off. This effect is extremely small and
735 often not noticeable, as other integration errors (e.g. from constraints)
736 may dominate. To
737 completely avoid cut-off artifacts, the non-bonded forces can be
738 switched exactly to zero at some distance smaller than the neighbor
739 list cut-off (there are several ways to do this in {\gromacs}, see
740 \secref{mod_nb_int}). One then has a buffer with the size equal to the
741 neighbor list cut-off less the longest interaction cut-off.
744 \subsubsection{Simple search\swapindexquiet{simple}{search}}
745 Due to \eqnsref{box_rot}{simplerc}, the vector $\rvij$
746 connecting images within the cut-off $R_c$ can be found by constructing:
747 \bea
748 \ve{r}''' & = & \ve{r}_j-\ve{r}_i \\
749 \ve{r}'' & = & \ve{r}''' - {\bf c}*\verb'round'(r'''_z/c_z) \\
750 \ve{r}' & = & \ve{r}'' - {\bf b}*\verb'round'(r''_y/b_y) \\
751 \ve{r}_{ij} & = & \ve{r}' - {\bf a}*\verb'round'(r'_x/a_x)
752 \eea
753 When distances between two particles in a triclinic box are needed
754 that do not obey \eqnref{box_rot},
755 many shifts of combinations of box vectors need to be considered to find
756 the nearest image.
759 \begin{figure}
760 \centerline{\includegraphics[width=8cm]{plots/nstric}}
761 \caption {Grid search in two dimensions. The arrows are the box vectors.}
762 \label{fig:grid}
763 \end{figure}
765 \subsubsection{Grid search\swapindexquiet{grid}{search}}
766 \label{sec:nsgrid}
767 The grid search is schematically depicted in \figref{grid}. All
768 particles are put on the {\nsgrid}, with the smallest spacing $\ge$
769 $R_c/2$ in each of the directions. In the direction of each box
770 vector, a particle $i$ has three images. For each direction the image
771 may be -1,0 or 1, corresponding to a translation over -1, 0 or +1 box
772 vector. We do not search the surrounding {\nsgrid} cells for neighbors
773 of $i$ and then calculate the image, but rather construct the images
774 first and then search neighbors corresponding to that image of $i$.
775 As \figref{grid} shows, some grid cells may be searched more than once
776 for different images of $i$. This is not a problem, since, due to the
777 minimum image convention, at most one image will ``see'' the
778 $j$-particle. For every particle, fewer than 125 (5$^3$) neighboring
779 cells are searched. Therefore, the algorithm scales linearly with the
780 number of particles. Although the prefactor is large, the scaling
781 behavior makes the algorithm far superior over the standard $O(N^2)$
782 algorithm when there are more than a few hundred particles. The
783 grid search is equally fast for rectangular and triclinic boxes. Thus
784 for most protein and peptide simulations the rhombic dodecahedron will
785 be the preferred box shape.
787 \subsubsection{Charge groups}
788 \label{sec:chargegroup}\swapindexquiet{charge}{group}%
789 Charge groups were originally introduced to reduce cut-off artifacts
790 of Coulomb interactions. When a plain cut-off is used, significant
791 jumps in the potential and forces arise when atoms with (partial) charges
792 move in and out of the cut-off radius. When all chemical moieties have
793 a net charge of zero, these jumps can be reduced by moving groups
794 of atoms with net charge zero, called charge groups, in and
795 out of the neighbor list. This reduces the cut-off effects from
796 the charge-charge level to the dipole-dipole level, which decay
797 much faster. With the advent of full range electrostatics methods,
798 such as particle-mesh Ewald (\secref{pme}), the use of charge groups is
799 no longer required for accuracy. It might even have a slight negative effect
800 on the accuracy or efficiency, depending on how the neighbor list is made
801 and the interactions are calculated.
803 But there is still an important reason for using ``charge groups'': efficiency with the group cut-off scheme.
804 Where applicable, neighbor searching is carried out on the basis of
805 charge groups which are defined in the molecular topology.
806 If the nearest image distance between the {\em
807 geometrical centers} of the atoms of two charge groups is less than
808 the cut-off radius, all atom pairs between the charge groups are
809 included in the pair list.
810 The neighbor searching for a water system, for instance,
811 is $3^2=9$ times faster when each molecule is treated as a charge group.
812 Also the highly optimized water force loops (see \secref{waterloops})
813 only work when all atoms in a water molecule form a single charge group.
814 Currently the name {\em neighbor-search group} would be more appropriate,
815 but the name charge group is retained for historical reasons.
816 When developing a new force field, the advice is to use charge groups
817 of 3 to 4 atoms for optimal performance. For all-atom force fields
818 this is relatively easy, as one can simply put hydrogen atoms, and in some
819 case oxygen atoms, in the same charge group as the heavy atom they
820 are connected to; for example: CH$_3$, CH$_2$, CH, NH$_2$, NH, OH, CO$_2$, CO.
822 With the Verlet cut-off scheme, charge groups are ignored.
825 \subsection{Compute forces}
826 \label{subsec:forces}
828 \subsubsection{Potential energy}
829 When forces are computed, the \swapindex{potential}{energy} of each
830 interaction term is computed as well. The total potential energy is
831 summed for various contributions, such as Lennard-Jones, Coulomb, and
832 bonded terms. It is also possible to compute these contributions for
833 {\em energy-monitor groups} of atoms that are separately defined (see
834 \secref{groupconcept}).
836 \subsubsection{Kinetic energy and temperature}
837 The \normindex{temperature} is given by the total
838 \swapindex{kinetic}{energy} of the $N$-particle system:
839 \beq
840 E_{kin} = \half \sum_{i=1}^N m_i v_i^2
841 \eeq
842 From this the absolute temperature $T$ can be computed using:
843 \beq
844 \half N_{\mathrm{df}} kT = E_{\mathrm{kin}}
845 \label{eqn:E-T}
846 \eeq
847 where $k$ is Boltzmann's constant and $N_{df}$ is the number of
848 degrees of freedom which can be computed from:
849 \beq
850 N_{\mathrm{df}} ~=~ 3 N - N_c - N_{\mathrm{com}}
851 \eeq
852 Here $N_c$ is the number of {\em \normindex{constraints}} imposed on the system.
853 When performing molecular dynamics $N_{\mathrm{com}}=3$ additional degrees of
854 freedom must be removed, because the three
855 center-of-mass velocities are constants of the motion, which are usually
856 set to zero. When simulating in vacuo, the rotation around the center of mass
857 can also be removed, in this case $N_{\mathrm{com}}=6$.
858 When more than one temperature-coupling group\index{temperature-coupling group} is used, the number of degrees
859 of freedom for group $i$ is:
860 \beq
861 N^i_{\mathrm{df}} ~=~ (3 N^i - N^i_c) \frac{3 N - N_c - N_{\mathrm{com}}}{3 N - N_c}
862 \eeq
864 The kinetic energy can also be written as a tensor, which is necessary
865 for pressure calculation in a triclinic system, or systems where shear
866 forces are imposed:
867 \beq
868 {\bf E}_{\mathrm{kin}} = \half \sum_i^N m_i \vvi \otimes \vvi
869 \eeq
871 \subsubsection{Pressure and virial}
872 The \normindex{pressure}
873 tensor {\bf P} is calculated from the difference between
874 kinetic energy $E_{\mathrm{kin}}$ and the \normindex{virial} ${\bf \Xi}$:
875 \beq
876 {\bf P} = \frac{2}{V} ({\bf E}_{\mathrm{kin}}-{\bf \Xi})
877 \label{eqn:P}
878 \eeq
879 where $V$ is the volume of the computational box.
880 The scalar pressure $P$, which can be used for pressure coupling in the case
881 of isotropic systems, is computed as:
882 \beq
883 P = {\rm trace}({\bf P})/3
884 \eeq
886 The virial ${\bf \Xi}$ tensor is defined as:
887 \beq
888 {\bf \Xi} = -\half \sum_{i<j} \rvij \otimes \Fvij
889 \label{eqn:Xi}
890 \eeq
892 The {\gromacs} implementation of the virial computation is described
893 in \secref{virial}.
895 \subsection{The \swapindex{leap-frog}{integrator}}
896 \label{subsec:update}
897 \begin{figure}
898 \centerline{\includegraphics[width=8cm]{plots/leapfrog}}
899 \caption[The Leap-Frog integration method.]{The Leap-Frog integration method. The algorithm is called Leap-Frog because $\ve{r}$ and $\ve{v}$ are leaping
900 like frogs over each other's backs.}
901 \label{fig:leapfrog}
902 \end{figure}
904 The default MD integrator in {\gromacs} is the so-called {\em leap-frog}
905 algorithm~\cite{Hockney74} for the integration of the equations of
906 motion. When extremely accurate integration with temperature
907 and/or pressure coupling is required, the velocity Verlet integrators
908 are also present and may be preferable (see \ssecref{vverlet}). The leap-frog
909 algorithm uses positions $\ve{r}$ at time $t$ and
910 velocities $\ve{v}$ at time $t-\hDt$; it updates positions and
911 velocities using the forces
912 $\ve{F}(t)$ determined by the positions at time $t$ using these relations:
913 \bea
914 \label{eqn:leapfrogv}
915 \ve{v}(t+\hDt) &~=~& \ve{v}(t-\hDt)+\frac{\Dt}{m}\ve{F}(t) \\
916 \ve{r}(t+\Dt) &~=~& \ve{r}(t)+\Dt\ve{v}(t+\hDt)
917 \eea
918 The algorithm is visualized in \figref{leapfrog}.
919 It produces trajectories that are identical to the Verlet~\cite{Verlet67} algorithm,
920 whose position-update relation is
921 \beq
922 \ve{r}(t+\Dt)~=~2\ve{r}(t) - \ve{r}(t-\Dt) + \frac{1}{m}\ve{F}(t)\Dt^2+O(\Dt^4)
923 \eeq
924 The algorithm is of third order in $\ve{r}$ and is time-reversible.
925 See ref.~\cite{Berendsen86b} for the merits of this algorithm and comparison
926 with other time integration algorithms.
928 The \swapindex{equations of}{motion} are modified for temperature
929 coupling and pressure coupling, and extended to include the
930 conservation of constraints, all of which are described below.
932 \subsection{The \swapindex{velocity Verlet}{integrator}}
933 \label{subsec:vverlet}
934 The velocity Verlet algorithm~\cite{Swope82} is also implemented in
935 {\gromacs}, though it is not yet fully integrated with all sets of
936 options. In velocity Verlet, positions $\ve{r}$ and velocities
937 $\ve{v}$ at time $t$ are used to integrate the equations of motion;
938 velocities at the previous half step are not required. \bea
939 \label{eqn:velocityverlet1}
940 \ve{v}(t+\hDt) &~=~& \ve{v}(t)+\frac{\Dt}{2m}\ve{F}(t) \\
941 \ve{r}(t+\Dt) &~=~& \ve{r}(t)+\Dt\,\ve{v}(t+\hDt) \\
942 \ve{v}(t+\Dt) &~=~& \ve{v}(t+\hDt)+\frac{\Dt}{2m}\ve{F}(t+\Dt)
943 \eea
944 or, equivalently,
945 \bea
946 \label{eqn:velocityverlet2}
947 \ve{r}(t+\Dt) &~=~& \ve{r}(t)+ \Dt\,\ve{v} + \frac{\Dt^2}{2m}\ve{F}(t) \\
948 \ve{v}(t+\Dt) &~=~& \ve{v}(t)+ \frac{\Dt}{2m}\left[\ve{F}(t) + \ve{F}(t+\Dt)\right]
949 \eea
950 With no temperature or pressure coupling, and with {\em corresponding}
951 starting points, leap-frog and velocity Verlet will generate identical
952 trajectories, as can easily be verified by hand from the equations
953 above. Given a single starting file with the {\em same} starting
954 point $\ve{x}(0)$ and $\ve{v}(0)$, leap-frog and velocity Verlet will
955 {\em not} give identical trajectories, as leap-frog will interpret the
956 velocities as corresponding to $t=-\hDt$, while velocity Verlet will
957 interpret them as corresponding to the timepoint $t=0$.
959 \subsection{Understanding reversible integrators: The Trotter decomposition}
960 To further understand the relationship between velocity Verlet and
961 leap-frog integration, we introduce the reversible Trotter formulation
962 of dynamics, which is also useful to understanding implementations of
963 thermostats and barostats in {\gromacs}.
965 A system of coupled, first-order differential equations can be evolved
966 from time $t = 0$ to time $t$ by applying the evolution operator
967 \bea
968 \Gamma(t) &=& \exp(iLt) \Gamma(0) \nonumber \\
969 iL &=& \dot{\Gamma}\cdot \nabla_{\Gamma},
970 \eea
971 where $L$ is the Liouville operator, and $\Gamma$ is the
972 multidimensional vector of independent variables (positions and
973 velocities).
974 A short-time approximation to the true operator, accurate at time $\Dt
975 = t/P$, is applied $P$ times in succession to evolve the system as
976 \beq
977 \Gamma(t) = \prod_{i=1}^P \exp(iL\Dt) \Gamma(0)
978 \eeq
979 For NVE dynamics, the Liouville operator is
980 \bea
981 iL = \sum_{i=1}^{N} \vv_i \cdot \nabla_{\rv_i} + \sum_{i=1}^N \frac{1}{m_i}\F(r_i) \cdot \nabla_{\vv_i}.
982 \eea
983 This can be split into two additive operators
984 \bea
985 iL_1 &=& \sum_{i=1}^N \frac{1}{m_i}\F(r_i) \cdot \nabla_{\vv_i} \nonumber \\
986 iL_2 &=& \sum_{i=1}^{N} \vv_i \cdot \nabla_{\rv_i}
987 \eea
988 Then a short-time, symmetric, and thus reversible approximation of the true dynamics will be
989 \bea
990 \exp(iL\Dt) = \exp(iL_2\hDt) \exp(iL_1\Dt) \exp(iL_2\hDt) + \mathcal{O}(\Dt^3).
991 \label{eq:NVE_Trotter}
992 \eea
993 This corresponds to velocity Verlet integration. The first
994 exponential term over $\hDt$ corresponds to a velocity half-step, the
995 second exponential term over $\Dt$ corresponds to a full velocity
996 step, and the last exponential term over $\hDt$ is the final velocity
997 half step. For future times $t = n\Dt$, this becomes
998 \bea
999 \exp(iLn\Dt) &\approx& \left(\exp(iL_2\hDt) \exp(iL_1\Dt) \exp(iL_2\hDt)\right)^n \nonumber \\
1000 &\approx& \exp(iL_2\hDt) \bigg(\exp(iL_1\Dt) \exp(iL_2\Dt)\bigg)^{n-1} \nonumber \\
1001 & & \;\;\;\; \exp(iL_1\Dt) \exp(iL_2\hDt)
1002 \eea
1003 This formalism allows us to easily see the difference between the
1004 different flavors of Verlet integrators. The leap-frog integrator can
1005 be seen as starting with Eq.~\ref{eq:NVE_Trotter} with the
1006 $\exp\left(iL_1 \dt\right)$ term, instead of the half-step velocity
1007 term, yielding
1008 \bea
1009 \exp(iLn\dt) &=& \exp\left(iL_1 \dt\right) \exp\left(iL_2 \Dt \right) + \mathcal{O}(\Dt^3).
1010 \eea
1011 Here, the full step in velocity is between $t-\hDt$ and $t+\hDt$,
1012 since it is a combination of the velocity half steps in velocity
1013 Verlet. For future times $t = n\Dt$, this becomes
1014 \bea
1015 \exp(iLn\dt) &\approx& \bigg(\exp\left(iL_1 \dt\right) \exp\left(iL_2 \Dt \right) \bigg)^{n}.
1016 \eea
1017 Although at first this does not appear symmetric, as long as the full velocity
1018 step is between $t-\hDt$ and $t+\hDt$, then this is simply a way of
1019 starting velocity Verlet at a different place in the cycle.
1021 Even though the trajectory and thus potential energies are identical
1022 between leap-frog and velocity Verlet, the kinetic energy and
1023 temperature will not necessarily be the same. Standard velocity
1024 Verlet uses the velocities at the $t$ to calculate the kinetic energy
1025 and thus the temperature only at time $t$; the kinetic energy is then a sum over all particles
1026 \bea
1027 KE_{\mathrm{full}}(t) &=& \sum_i \left(\frac{1}{2m_i}\ve{v}_i(t)\right)^2 \nonumber\\
1028 &=& \sum_i \frac{1}{2m_i}\left(\frac{1}{2}\ve{v}_i(t-\hDt)+\frac{1}{2}\ve{v}_i(t+\hDt)\right)^2,
1029 \eea
1030 with the square on the {\em outside} of the average. Standard
1031 leap-frog calculates the kinetic energy at time $t$ based on the
1032 average kinetic energies at the timesteps $t+\hDt$ and $t-\hDt$, or
1033 the sum over all particles
1034 \bea
1035 KE_{\mathrm{average}}(t) &=& \sum_i \frac{1}{2m_i}\left(\frac{1}{2}\ve{v}_i(t-\hDt)^2+\frac{1}{2}\ve{v}_i(t+\hDt)^2\right),
1036 \eea
1037 where the square is {\em inside} the average.
1039 A non-standard variant of velocity Verlet which averages the kinetic
1040 energies $KE(t+\hDt)$ and $KE(t-\hDt)$, exactly like leap-frog, is also
1041 now implemented in {\gromacs} (as {\tt .mdp} file option {\tt md-vv-avek}). Without
1042 temperature and pressure coupling, velocity Verlet with
1043 half-step-averaged kinetic energies and leap-frog will be identical up
1044 to numerical precision. For temperature- and pressure-control schemes,
1045 however, velocity Verlet with half-step-averaged kinetic energies and
1046 leap-frog will be different, as will be discussed in the section in
1047 thermostats and barostats.
1049 The half-step-averaged kinetic energy and temperature are slightly more
1050 accurate for a given step size; the difference in average kinetic
1051 energies using the half-step-averaged kinetic energies ({\em md} and
1052 {\em md-vv-avek}) will be closer to the kinetic energy obtained in the
1053 limit of small step size than will the full-step kinetic energy (using
1054 {\em md-vv}). For NVE simulations, this difference is usually not
1055 significant, since the positions and velocities of the particles are
1056 still identical; it makes a difference in the way the the temperature
1057 of the simulations are {\em interpreted}, but {\em not} in the
1058 trajectories that are produced. Although the kinetic energy is more
1059 accurate with the half-step-averaged method, meaning that it changes
1060 less as the timestep gets large, it is also more noisy. The RMS deviation
1061 of the total energy of the system (sum of kinetic plus
1062 potential) in the half-step-averaged kinetic energy case will be
1063 higher (about twice as high in most cases) than the full-step kinetic
1064 energy. The drift will still be the same, however, as again, the
1065 trajectories are identical.
1067 For NVT simulations, however, there {\em will} be a difference, as
1068 discussed in the section on temperature control, since the velocities
1069 of the particles are adjusted such that kinetic energies of the
1070 simulations, which can be calculated either way, reach the
1071 distribution corresponding to the set temperature. In this case, the
1072 three methods will not give identical results.
1074 Because the velocity and position are both defined at the same time
1075 $t$ the velocity Verlet integrator can be used for some methods,
1076 especially rigorously correct pressure control methods, that are not
1077 actually possible with leap-frog. The integration itself takes
1078 negligibly more time than leap-frog, but twice as many communication
1079 calls are currently required. In most cases, and especially for large
1080 systems where communication speed is important for parallelization and
1081 differences between thermodynamic ensembles vanish in the $1/N$ limit,
1082 and when only NVT ensembles are required, leap-frog will likely be the
1083 preferred integrator. For pressure control simulations where the fine
1084 details of the thermodynamics are important, only velocity Verlet
1085 allows the true ensemble to be calculated. In either case, simulation
1086 with double precision may be required to get fine details of
1087 thermodynamics correct.
1089 \subsection{Multiple time stepping}
1090 Several other simulation packages uses multiple time stepping for
1091 bonds and/or the PME mesh forces. In {\gromacs} we have not implemented
1092 this (yet), since we use a different philosophy. Bonds can be constrained
1093 (which is also a more sound approximation of a physical quantum
1094 oscillator), which allows the smallest time step to be increased
1095 to the larger one. This not only halves the number of force calculations,
1096 but also the update calculations. For even larger time steps, angle vibrations
1097 involving hydrogen atoms can be removed using virtual interaction
1098 sites (see \secref{rmfast}),
1099 which brings the shortest time step up to
1100 PME mesh update frequency of a multiple time stepping scheme.
1102 \subsection{Temperature coupling\index{temperature coupling}}
1103 While direct use of molecular dynamics gives rise to the NVE (constant
1104 number, constant volume, constant energy ensemble), most quantities
1105 that we wish to calculate are actually from a constant temperature
1106 (NVT) ensemble, also called the canonical ensemble. {\gromacs} can use
1107 the {\em weak-coupling} scheme of Berendsen~\cite{Berendsen84},
1108 stochastic randomization through the Andersen
1109 thermostat~\cite{Andersen80}, the extended ensemble Nos{\'e}-Hoover
1110 scheme~\cite{Nose84,Hoover85}, or a velocity-rescaling
1111 scheme~\cite{Bussi2007a} to simulate constant temperature, with
1112 advantages of each of the schemes laid out below.
1114 There are several other reasons why it might be necessary to control
1115 the temperature of the system (drift during equilibration, drift as a
1116 result of force truncation and integration errors, heating due to
1117 external or frictional forces), but this is not entirely correct to do
1118 from a thermodynamic standpoint, and in some cases only masks the
1119 symptoms (increase in temperature of the system) rather than the
1120 underlying problem (deviations from correct physics in the dynamics).
1121 For larger systems, errors in ensemble averages and structural
1122 properties incurred by using temperature control to remove slow drifts
1123 in temperature appear to be negligible, but no completely
1124 comprehensive comparisons have been carried out, and some caution must
1125 be taking in interpreting the results.
1127 When using temperature and/or pressure coupling the total energy is
1128 no longer conserved. Instead there is a \normindex{conserved energy quantity}
1129 the formula of which will depend on the combination or temperature and
1130 pressure coupling algorithm used. For all coupling algorithms, except
1131 for Andersen temperature coupling and Parrinello-Rahman pressure coupling
1132 combined with shear stress, the conserved energy quantity is computed
1133 and stored in the energy and log file. Note that this quantity will not
1134 be conserved when external forces are applied to the system, such as
1135 pulling on group with a changing distance or an electric field.
1136 Furthermore, how well the energy is conserved depends on the accuracy
1137 of all algorithms involved in the simulation. Usually the algorithms that
1138 cause most drift are constraints and the pair-list buffer, depending
1139 on the parameters used.
1141 \subsubsection{Berendsen temperature coupling\pawsindexquiet{Berendsen}{temperature coupling}\index{weak coupling}}
1142 The Berendsen algorithm mimics weak coupling with first-order
1143 kinetics to an external heat bath with given temperature $T_0$.
1144 See ref.~\cite{Berendsen91} for a comparison with the
1145 Nos{\'e}-Hoover scheme. The effect of this algorithm is
1146 that a deviation of the system temperature from $T_0$ is slowly
1147 corrected according to:
1148 \beq
1149 \frac{\de T}{\de t} = \frac{T_0-T}{\tau}
1150 \label{eqn:Tcoupling}
1151 \eeq
1152 which means that a temperature deviation decays exponentially with a
1153 time constant $\tau$.
1154 This method of coupling has the advantage that the strength of the
1155 coupling can be varied and adapted to the user requirement: for
1156 equilibration purposes the coupling time can be taken quite short
1157 ({\eg} 0.01 ps), but for reliable equilibrium runs it can be taken much
1158 longer ({\eg} 0.5 ps) in which case it hardly influences the
1159 conservative dynamics.
1161 The Berendsen thermostat suppresses the fluctuations of the kinetic
1162 energy. This means that one does not generate a proper canonical
1163 ensemble, so rigorously, the sampling will be incorrect. This
1164 error scales with $1/N$, so for very large systems most ensemble
1165 averages will not be affected significantly, except for the
1166 distribution of the kinetic energy itself. However, fluctuation
1167 properties, such as the heat capacity, will be affected. A similar
1168 thermostat which does produce a correct ensemble is the velocity
1169 rescaling thermostat~\cite{Bussi2007a} described below.
1171 The heat flow into or out of the system is affected by scaling the
1172 velocities of each particle every step, or every $n_\mathrm{TC}$ steps,
1173 with a time-dependent factor $\lambda$, given by:
1174 \beq
1175 \lambda = \left[ 1 + \frac{n_\mathrm{TC} \Delta t}{\tau_T}
1176 \left\{\frac{T_0}{T(t - \hDt)} - 1 \right\} \right]^{1/2}
1177 \label{eqn:lambda}
1178 \eeq
1179 The parameter $\tau_T$ is close, but not exactly equal, to the time constant
1180 $\tau$ of the temperature coupling (\eqnref{Tcoupling}):
1181 \beq
1182 \tau = 2 C_V \tau_T / N_{df} k
1183 \eeq
1184 where $C_V$ is the total heat capacity of the system, $k$ is Boltzmann's
1185 constant, and $N_{df}$ is the total number of degrees of freedom. The
1186 reason that $\tau \neq \tau_T$ is that the kinetic energy change
1187 caused by scaling the velocities is partly redistributed between
1188 kinetic and potential energy and hence the change in temperature is
1189 less than the scaling energy. In practice, the ratio $\tau / \tau_T$
1190 ranges from 1 (gas) to 2 (harmonic solid) to 3 (water). When we use
1191 the term ``temperature coupling time constant,'' we mean the parameter
1192 \normindex{$\tau_T$}.
1193 {\bf Note} that in practice the scaling factor $\lambda$ is limited to
1194 the range of 0.8 $<= \lambda <=$ 1.25, to avoid scaling by very large
1195 numbers which may crash the simulation. In normal use,
1196 $\lambda$ will always be much closer to 1.0.
1198 The thermostat modifies the kinetic energy at each scaling step by:
1199 \beq
1200 \Delta E_k = (\lambda - 1)^2 E_k
1201 \eeq
1202 The sum of these changes over the run needs to subtracted from the total energy
1203 to obtain the conserved energy quantity.
1205 \subsubsection{Velocity-rescaling temperature coupling\pawsindexquiet{velocity-rescaling}{temperature coupling}}
1206 The velocity-rescaling thermostat~\cite{Bussi2007a} is essentially a Berendsen
1207 thermostat (see above) with an additional stochastic term that ensures
1208 a correct kinetic energy distribution by modifying it according to
1209 \beq
1210 \de K = (K_0 - K) \frac{\de t}{\tau_T} + 2 \sqrt{\frac{K K_0}{N_f}} \frac{\de W}{\sqrt{\tau_T}},
1211 \label{eqn:vrescale}
1212 \eeq
1213 where $K$ is the kinetic energy, $N_f$ the number of degrees of freedom and $\de W$ a Wiener process.
1214 There are no additional parameters, except for a random seed.
1215 This thermostat produces a correct canonical ensemble and still has
1216 the advantage of the Berendsen thermostat: first order decay of
1217 temperature deviations and no oscillations.
1219 \subsubsection{\normindex{Andersen thermostat}}
1220 One simple way to maintain a thermostatted ensemble is to take an
1221 $NVE$ integrator and periodically re-select the velocities of the
1222 particles from a Maxwell-Boltzmann distribution.~\cite{Andersen80}
1223 This can either be done by randomizing all the velocities
1224 simultaneously (massive collision) every $\tau_T/\Dt$ steps ({\tt andersen-massive}), or by
1225 randomizing every particle with some small probability every timestep ({\tt andersen}),
1226 equal to $\Dt/\tau$, where in both cases $\Dt$ is the timestep and
1227 $\tau_T$ is a characteristic coupling time scale.
1228 Because of the way constraints operate, all particles in the same
1229 constraint group must be randomized simultaneously. Because of
1230 parallelization issues, the {\tt andersen} version cannot currently (5.0) be
1231 used in systems with constraints. {\tt andersen-massive} can be used regardless of constraints.
1232 This thermostat is also currently only possible with velocity Verlet algorithms,
1233 because it operates directly on the velocities at each timestep.
1235 This algorithm completely avoids some of the ergodicity issues of other thermostatting
1236 algorithms, as energy cannot flow back and forth between energetically
1237 decoupled components of the system as in velocity scaling motions.
1238 However, it can slow down the kinetics of system by randomizing
1239 correlated motions of the system, including slowing sampling when
1240 $\tau_T$ is at moderate levels (less than 10 ps). This algorithm
1241 should therefore generally not be used when examining kinetics or
1242 transport properties of the system.~\cite{Basconi2013}
1244 \subsubsection{Nos{\'e}-Hoover temperature coupling\index{Nose-Hoover temperature coupling@Nos{\'e}-Hoover temperature coupling|see{temperature coupling, Nos{\'e}-Hoover}}{\index{temperature coupling Nose-Hoover@temperature coupling Nos{\'e}-Hoover}}\index{extended ensemble}}
1246 The Berendsen weak-coupling algorithm is
1247 extremely efficient for relaxing a system to the target temperature,
1248 but once the system has reached equilibrium it might be more
1249 important to probe a correct canonical ensemble. This is unfortunately
1250 not the case for the weak-coupling scheme.
1252 To enable canonical ensemble simulations, {\gromacs} also supports the
1253 extended-ensemble approach first proposed by Nos{\'e}~\cite{Nose84}
1254 and later modified by Hoover~\cite{Hoover85}. The system Hamiltonian is
1255 extended by introducing a thermal reservoir and a friction term in the
1256 equations of motion. The friction force is proportional to the
1257 product of each particle's velocity and a friction parameter, $\xi$.
1258 This friction parameter (or ``heat bath'' variable) is a fully
1259 dynamic quantity with its own momentum ($p_{\xi}$) and equation of
1260 motion; the time derivative is calculated from the difference between
1261 the current kinetic energy and the reference temperature.
1263 In this formulation, the particles' equations of motion in
1264 \figref{global} are replaced by:
1265 \beq
1266 \frac {\de^2\ve{r}_i}{\de t^2} = \frac{\ve{F}_i}{m_i} -
1267 \frac{p_{\xi}}{Q}\frac{\de \ve{r}_i}{\de t} ,
1268 \label{eqn:NH-eqn-of-motion}
1269 \eeq where the equation of motion for the heat bath parameter $\xi$ is:
1270 \beq \frac {\de p_{\xi}}{\de t} = \left( T - T_0 \right). \eeq The
1271 reference temperature is denoted $T_0$, while $T$ is the current
1272 instantaneous temperature of the system. The strength of the coupling
1273 is determined by the constant $Q$ (usually called the ``mass parameter''
1274 of the reservoir) in combination with the reference
1275 temperature.~\footnote{Note that some derivations, an alternative
1276 notation $\xi_{\mathrm{alt}} = v_{\xi} = p_{\xi}/Q$ is used.}
1278 The conserved quantity for the Nos{\'e}-Hoover equations of motion is not
1279 the total energy, but rather
1280 \bea
1281 H = \sum_{i=1}^{N} \frac{\pb_i}{2m_i} + U\left(\rv_1,\rv_2,\ldots,\rv_N\right) +\frac{p_{\xi}^2}{2Q} + N_fkT\xi,
1282 \eea
1283 where $N_f$ is the total number of degrees of freedom.
1285 In our opinion, the mass parameter is a somewhat awkward way of
1286 describing coupling strength, especially due to its dependence on
1287 reference temperature (and some implementations even include the
1288 number of degrees of freedom in your system when defining $Q$). To
1289 maintain the coupling strength, one would have to change $Q$ in
1290 proportion to the change in reference temperature. For this reason, we
1291 prefer to let the {\gromacs} user work instead with the period
1292 $\tau_T$ of the oscillations of kinetic energy between the system and
1293 the reservoir instead. It is directly related to $Q$ and $T_0$ via:
1294 \beq
1295 Q = \frac {\tau_T^2 T_0}{4 \pi^2}.
1296 \eeq
1297 This provides a much more intuitive way of selecting the
1298 Nos{\'e}-Hoover coupling strength (similar to the weak-coupling
1299 relaxation), and in addition $\tau_T$ is independent of system size
1300 and reference temperature.
1302 It is however important to keep the difference between the
1303 weak-coupling scheme and the Nos{\'e}-Hoover algorithm in mind:
1304 Using weak coupling you get a
1305 strongly damped {\em exponential relaxation},
1306 while the Nos{\'e}-Hoover approach
1307 produces an {\em oscillatory relaxation}.
1308 The actual time it takes to relax with Nos{\'e}-Hoover coupling is
1309 several times larger than the period of the
1310 oscillations that you select. These oscillations (in contrast
1311 to exponential relaxation) also means that
1312 the time constant normally should be 4--5 times larger
1313 than the relaxation time used with weak coupling, but your
1314 mileage may vary.
1316 Nos{\'e}-Hoover dynamics in simple systems such as collections of
1317 harmonic oscillators, can be {\em nonergodic}, meaning that only a
1318 subsection of phase space is ever sampled, even if the simulations
1319 were to run for infinitely long. For this reason, the Nos{\'e}-Hoover
1320 chain approach was developed, where each of the Nos{\'e}-Hoover
1321 thermostats has its own Nos{\'e}-Hoover thermostat controlling its
1322 temperature. In the limit of an infinite chain of thermostats, the
1323 dynamics are guaranteed to be ergodic. Using just a few chains can
1324 greatly improve the ergodicity, but recent research has shown that the
1325 system will still be nonergodic, and it is still not entirely clear
1326 what the practical effect of this~\cite{Cooke2008}. Currently, the
1327 default number of chains is 10, but this can be controlled by the
1328 user. In the case of chains, the equations are modified in the
1329 following way to include a chain of thermostatting
1330 particles~\cite{Martyna1992}:
1332 \bea
1333 \frac {\de^2\ve{r}_i}{\de t^2} &~=~& \frac{\ve{F}_i}{m_i} - \frac{p_{{\xi}_1}}{Q_1} \frac{\de \ve{r}_i}{\de t} \nonumber \\
1334 \frac {\de p_{{\xi}_1}}{\de t} &~=~& \left( T - T_0 \right) - p_{{\xi}_1} \frac{p_{{\xi}_2}}{Q_2} \nonumber \\
1335 \frac {\de p_{{\xi}_{i=2\ldots N}}}{\de t} &~=~& \left(\frac{p_{\xi_{i-1}}^2}{Q_{i-1}} -kT\right) - p_{\xi_i} \frac{p_{\xi_{i+1}}}{Q_{i+1}} \nonumber \\
1336 \frac {\de p_{\xi_N}}{\de t} &~=~& \left(\frac{p_{\xi_{N-1}}^2}{Q_{N-1}}-kT\right)
1337 \label{eqn:NH-chain-eqn-of-motion}
1338 \eea
1339 The conserved quantity for Nos{\'e}-Hoover chains is
1340 \bea
1341 H = \sum_{i=1}^{N} \frac{\pb_i}{2m_i} + U\left(\rv_1,\rv_2,\ldots,\rv_N\right) +\sum_{k=1}^M\frac{p^2_{\xi_k}}{2Q^{\prime}_k} + N_fkT\xi_1 + kT\sum_{k=2}^M \xi_k
1342 \eea
1343 The values and velocities of the Nos{\'e}-Hoover thermostat variables
1344 are generally not included in the output, as they take up a fair
1345 amount of space and are generally not important for analysis of
1346 simulations, but by setting an mdp option the values of all
1347 the positions and velocities of all Nos{\'e}-Hoover particles in the
1348 chain are written to the {\tt .edr} file. Leap-frog simulations
1349 currently can only have Nos{\'e}-Hoover chain lengths of 1,
1350 but this will likely be updated in later version.
1352 As described in the integrator section, for temperature coupling, the
1353 temperature that the algorithm attempts to match to the reference
1354 temperature is calculated differently in velocity Verlet and leap-frog
1355 dynamics. Velocity Verlet ({\em md-vv}) uses the full-step kinetic
1356 energy, while leap-frog and {\em md-vv-avek} use the half-step-averaged
1357 kinetic energy.
1359 We can examine the Trotter decomposition again to better understand
1360 the differences between these constant-temperature integrators. In
1361 the case of Nos{\'e}-Hoover dynamics (for simplicity, using a chain
1362 with $N=1$, with more details in Ref.~\cite{Martyna1996}), we split
1363 the Liouville operator as
1364 \beq
1365 iL = iL_1 + iL_2 + iL_{\mathrm{NHC}},
1366 \eeq
1367 where
1368 \bea
1369 iL_1 &=& \sum_{i=1}^N \left[\frac{\pb_i}{m_i}\right]\cdot \frac{\partial}{\partial \rv_i} \nonumber \\
1370 iL_2 &=& \sum_{i=1}^N \F_i\cdot \frac{\partial}{\partial \pb_i} \nonumber \\
1371 iL_{\mathrm{NHC}} &=& \sum_{i=1}^N-\frac{p_{\xi}}{Q}\vv_i\cdot \nabla_{\vv_i} +\frac{p_{\xi}}{Q}\frac{\partial }{\partial \xi} + \left( T - T_0 \right)\frac{\partial }{\partial p_{\xi}}
1372 \eea
1373 For standard velocity Verlet with Nos{\'e}-Hoover temperature control, this becomes
1374 \bea
1375 \exp(iL\dt) &=& \exp\left(iL_{\mathrm{NHC}}\dt/2\right) \exp\left(iL_2 \dt/2\right) \nonumber \\
1376 &&\exp\left(iL_1 \dt\right) \exp\left(iL_2 \dt/2\right) \exp\left(iL_{\mathrm{NHC}}\dt/2\right) + \mathcal{O}(\Dt^3).
1377 \eea
1378 For half-step-averaged temperature control using {\em md-vv-avek},
1379 this decomposition will not work, since we do not have the full step
1380 temperature until after the second velocity step. However, we can
1381 construct an alternate decomposition that is still reversible, by
1382 switching the place of the NHC and velocity portions of the
1383 decomposition:
1384 \bea
1385 \exp(iL\dt) &=& \exp\left(iL_2 \dt/2\right) \exp\left(iL_{\mathrm{NHC}}\dt/2\right)\exp\left(iL_1 \dt\right)\nonumber \\
1386 &&\exp\left(iL_{\mathrm{NHC}}\dt/2\right) \exp\left(iL_2 \dt/2\right)+ \mathcal{O}(\Dt^3)
1387 \label{eq:half_step_NHC_integrator}
1388 \eea
1389 This formalism allows us to easily see the difference between the
1390 different flavors of velocity Verlet integrator. The leap-frog
1391 integrator can be seen as starting with
1392 Eq.~\ref{eq:half_step_NHC_integrator} just before the $\exp\left(iL_1
1393 \dt\right)$ term, yielding:
1394 \bea
1395 \exp(iL\dt) &=& \exp\left(iL_1 \dt\right) \exp\left(iL_{\mathrm{NHC}}\dt/2\right) \nonumber \\
1396 &&\exp\left(iL_2 \dt\right) \exp\left(iL_{\mathrm{NHC}}\dt/2\right) + \mathcal{O}(\Dt^3)
1397 \eea
1398 and then using some algebra tricks to solve for some quantities are
1399 required before they are actually calculated~\cite{Holian95}.
1402 \subsubsection{Group temperature coupling}\index{temperature-coupling group}%
1403 In {\gromacs} temperature coupling can be performed on groups of
1404 atoms, typically a protein and solvent. The reason such algorithms
1405 were introduced is that energy exchange between different components
1406 is not perfect, due to different effects including cut-offs etc. If
1407 now the whole system is coupled to one heat bath, water (which
1408 experiences the largest cut-off noise) will tend to heat up and the
1409 protein will cool down. Typically 100 K differences can be obtained.
1410 With the use of proper electrostatic methods (PME) these difference
1411 are much smaller but still not negligible. The parameters for
1412 temperature coupling in groups are given in the {\tt mdp} file.
1413 Recent investigation has shown that small temperature differences
1414 between protein and water may actually be an artifact of the way
1415 temperature is calculated when there are finite timesteps, and very
1416 large differences in temperature are likely a sign of something else
1417 seriously going wrong with the system, and should be investigated
1418 carefully~\cite{Eastwood2010}.
1420 One special case should be mentioned: it is possible to temperature-couple only
1421 part of the system, leaving other parts without temperature
1422 coupling. This is done by specifying ${-1}$ for the time constant
1423 $\tau_T$ for the group that should not be thermostatted. If only
1424 part of the system is thermostatted, the system will still eventually
1425 converge to an NVT system. In fact, one suggestion for minimizing
1426 errors in the temperature caused by discretized timesteps is that if
1427 constraints on the water are used, then only the water degrees of
1428 freedom should be thermostatted, not protein degrees of freedom, as
1429 the higher frequency modes in the protein can cause larger deviations
1430 from the ``true'' temperature, the temperature obtained with small
1431 timesteps~\cite{Eastwood2010}.
1433 \subsection{Pressure coupling\index{pressure coupling}}
1434 In the same spirit as the temperature coupling, the system can also be
1435 coupled to a ``pressure bath.'' {\gromacs} supports both the Berendsen
1436 algorithm~\cite{Berendsen84} that scales coordinates and box vectors
1437 every step, the extended-ensemble Parrinello-Rahman approach~\cite{Parrinello81,Nose83}, and for
1438 the velocity Verlet variants, the Martyna-Tuckerman-Tobias-Klein
1439 (MTTK) implementation of pressure
1440 control~\cite{Martyna1996}. Parrinello-Rahman and Berendsen can be
1441 combined with any of the temperature coupling methods above. MTTK can
1442 only be used with Nos{\'e}-Hoover temperature control. From 5.1 afterwards,
1443 it can only used when the system does not have constraints.
1445 \subsubsection{Berendsen pressure coupling\pawsindexquiet{Berendsen}{pressure coupling}\index{weak coupling}}
1446 \label{sec:berendsen_pressure_coupling}
1447 The Berendsen algorithm rescales the
1448 coordinates and box vectors every step, or every $n_\mathrm{PC}$ steps,
1449 with a matrix {\boldmath $\mu$},
1450 which has the effect of a first-order kinetic relaxation of the pressure
1451 towards a given reference pressure ${\bf P}_0$ according to
1452 \beq
1453 \frac{\de {\bf P}}{\de t} = \frac{{\bf P}_0-{\bf P}}{\tau_p}.
1454 \eeq
1455 The scaling matrix {\boldmath $\mu$} is given by
1456 \beq
1457 \mu_{ij}
1458 = \delta_{ij} - \frac{n_\mathrm{PC}\Delta t}{3\, \tau_p} \beta_{ij} \{P_{0ij} - P_{ij}(t) \}.
1459 \label{eqn:mu}
1460 \eeq
1461 \index{isothermal compressibility}
1462 \index{compressibility}
1463 Here, {\boldmath $\beta$} is the isothermal compressibility of the system.
1464 In most cases this will be a diagonal matrix, with equal elements on the
1465 diagonal, the value of which is generally not known.
1466 It suffices to take a rough estimate because the value of {\boldmath $\beta$}
1467 only influences the non-critical time constant of the
1468 pressure relaxation without affecting the average pressure itself.
1469 For water at 1 atm and 300 K
1470 $\beta = 4.6 \times 10^{-10}$ Pa$^{-1} = 4.6 \times 10^{-5}$ bar$^{-1}$,
1471 which is $7.6 \times 10^{-4}$ MD units (see \chref{defunits}).
1472 Most other liquids have similar values.
1473 When scaling completely anisotropically, the system has to be rotated in
1474 order to obey \eqnref{box_rot}.
1475 This rotation is approximated in first order in the scaling, which is usually
1476 less than $10^{-4}$. The actual scaling matrix {\boldmath $\mu'$} is
1477 \beq
1478 \mbox{\boldmath $\mu'$} =
1479 \left(\begin{array}{ccc}
1480 \mu_{xx} & \mu_{xy} + \mu_{yx} & \mu_{xz} + \mu_{zx} \\
1481 0 & \mu_{yy} & \mu_{yz} + \mu_{zy} \\
1482 0 & 0 & \mu_{zz}
1483 \end{array}\right).
1484 \eeq
1485 The velocities are neither scaled nor rotated.
1486 Since the equations of motion are modified by pressure coupling, the conserved
1487 energy quantity also needs to be modified. For first order pressure coupling,
1488 the work the barostat applies to the system every step needs to
1489 be subtracted from the total energy to obtain the conserved energy quantity:
1490 \beq
1491 - \sum_{i,j} (\mu_{ij} -\delta_{ij}) P_{ij} V =
1492 \sum_{i,j} 2(\mu_{ij} -\delta_{ij}) \Xi_{ij}
1493 \eeq
1494 where $\delta_{ij}$ is the Kronecker delta and ${\bf \Xi}$ is the virial.
1495 Note that the factor 2 originates from the factor $\frac{1}{2}$
1496 in the virial definition (\eqnref{Xi}).
1499 In {\gromacs}, the Berendsen scaling can also be done isotropically,
1500 which means that instead of $\ve{P}$ a diagonal matrix with elements of size
1501 trace$(\ve{P})/3$ is used. For systems with interfaces, semi-isotropic
1502 scaling can be useful.
1503 In this case, the $x/y$-directions are scaled isotropically and the $z$
1504 direction is scaled independently. The compressibility in the $x/y$ or
1505 $z$-direction can be set to zero, to scale only in the other direction(s).
1507 If you allow full anisotropic deformations and use constraints you
1508 might have to scale more slowly or decrease your timestep to avoid
1509 errors from the constraint algorithms. It is important to note that
1510 although the Berendsen pressure control algorithm yields a simulation
1511 with the correct average pressure, it does not yield the exact NPT
1512 ensemble, and it is not yet clear exactly what errors this approximation
1513 may yield.
1515 \subsubsection{Parrinello-Rahman pressure coupling\pawsindexquiet{Parrinello-Rahman}{pressure coupling}}
1517 In cases where the fluctuations in pressure or volume are important
1518 {\em per se} ({\eg} to calculate thermodynamic properties), especially
1519 for small systems, it may be a problem that the exact ensemble is not
1520 well defined for the weak-coupling scheme, and that it does not
1521 simulate the true NPT ensemble.
1523 {\gromacs} also supports constant-pressure simulations using the
1524 Parrinello-Rahman approach~\cite{Parrinello81,Nose83}, which is similar
1525 to the Nos{\'e}-Hoover temperature coupling, and in theory gives the
1526 true NPT ensemble. With the Parrinello-Rahman barostat, the box
1527 vectors as represented by the matrix \ve{b} obey the matrix equation
1528 of motion\footnote{The box matrix representation \ve{b} in {\gromacs}
1529 corresponds to the transpose of the box matrix representation \ve{h}
1530 in the paper by Nos{\'e} and Klein. Because of this, some of our
1531 equations will look slightly different.}
1532 \beq
1533 \frac{\de \ve{b}^2}{\de t^2}= V \ve{W}^{-1} \ve{b}'^{-1} \left( \ve{P} - \ve{P}_{ref}\right).
1534 \eeq
1536 The volume of the box is denoted $V$, and $\ve{W}$ is a matrix parameter that determines
1537 the strength of the coupling. The matrices \ve{P} and \ve{P}$_{ref}$ are the
1538 current and reference pressures, respectively.
1540 The equations of motion for the particles are also changed, just as
1541 for the Nos{\'e}-Hoover coupling. In most cases you would combine the
1542 Parrinello-Rahman barostat with the Nos{\'e}-Hoover
1543 thermostat, but to keep it simple we only show the Parrinello-Rahman
1544 modification here. The modified Hamiltonian, which will be conserved, is:
1545 \beq
1546 E_\mathrm{pot} + E_\mathrm{kin} + \sum_i P_{ii} V +
1547 \sum_{i,j} \frac{1}{2} W_{ij} \left( \frac{\de b_{ij}}{\de t} \right)^2
1548 \eeq
1549 The equations of motion for the atoms, obtained from the Hamiltonian are:
1550 \bea \frac {\de^2\ve{r}_i}{\de t^2} & = & \frac{\ve{F}_i}{m_i} -
1551 \ve{M} \frac{\de \ve{r}_i}{\de t} , \\ \ve{M} & = & \ve{b}^{-1} \left[
1552 \ve{b} \frac{\de \ve{b}'}{\de t} + \frac{\de \ve{b}}{\de t} \ve{b}'
1553 \right] \ve{b}'^{-1}.
1554 \eea
1555 This extra term has the appearance of a friction, but it should be
1556 noted that it is ficticious, and rather an effect of the
1557 Parrinello-Rahman equations of motion being defined with all
1558 particle coordinates represented relative to the box vectors, while
1559 {\gromacs] uses normal Cartesian coordinates for positions,
1560 velocities and forces. It is worth noting that the kinetic energy too
1561 should formally be calculated based on velocities relative to the
1562 box vectors. This can have an effect e.g. for external constant stress,
1563 but for now we only support coupling to constant external
1564 pressures, and for any normal simulation the velocities of box
1565 vectors should be extremely small compared to particle velocities.
1566 Gang Liu has done some work on deriving this for Cartesian
1567 coordinates\cite{Liu2015} that we will try to implement at
1568 some point in the future together with support for external stress.
1570 The (inverse) mass parameter matrix
1571 $\ve{W}^{-1}$ determines the strength of the coupling, and how the box
1572 can be deformed. The box restriction (\ref{eqn:box_rot}) will be
1573 fulfilled automatically if the corresponding elements of $\ve{W}^{-1}$
1574 are zero. Since the coupling strength also depends on the size of your
1575 box, we prefer to calculate it automatically in {\gromacs}. You only
1576 have to provide the approximate isothermal compressibilities
1577 {\boldmath $\beta$} and the pressure time constant $\tau_p$ in the
1578 input file ($L$ is the largest box matrix element): \beq \left(
1579 \ve{W}^{-1} \right)_{ij} = \frac{4 \pi^2 \beta_{ij}}{3 \tau_p^2 L}.
1580 \eeq Just as for the Nos{\'e}-Hoover thermostat, you should realize
1581 that the Parrinello-Rahman time constant is {\em not} equivalent to
1582 the relaxation time used in the Berendsen pressure coupling algorithm.
1583 In most cases you will need to use a 4--5 times larger time constant
1584 with Parrinello-Rahman coupling. If your pressure is very far from
1585 equilibrium, the Parrinello-Rahman coupling may result in very large
1586 box oscillations that could even crash your run. In that case you
1587 would have to increase the time constant, or (better) use the weak-coupling
1588 scheme to reach the target pressure, and then switch to
1589 Parrinello-Rahman coupling once the system is in equilibrium.
1590 Additionally, using the leap-frog algorithm, the pressure at time $t$
1591 is not available until after the time step has completed, and so the
1592 pressure from the previous step must be used, which makes the algorithm
1593 not directly reversible, and may not be appropriate for high precision
1594 thermodynamic calculations.
1596 \subsubsection{Surface-tension coupling\pawsindexquiet{surface-tension}{pressure coupling}}
1597 When a periodic system consists of more than one phase, separated by
1598 surfaces which are parallel to the $xy$-plane,
1599 the surface tension and the $z$-component of the pressure can be coupled
1600 to a pressure bath. Presently, this only works with the Berendsen
1601 pressure coupling algorithm in {\gromacs}.
1602 The average surface tension $\gamma(t)$ can be calculated from
1603 the difference between the normal and the lateral pressure
1604 \bea
1605 \gamma(t) & = &
1606 \frac{1}{n} \int_0^{L_z}
1607 \left\{ P_{zz}(z,t) - \frac{P_{xx}(z,t) + P_{yy}(z,t)}{2} \right\} \mbox{d}z \\
1608 & = &
1609 \frac{L_z}{n} \left\{ P_{zz}(t) - \frac{P_{xx}(t) + P_{yy}(t)}{2} \right\},
1610 \eea
1611 where $L_z$ is the height of the box and $n$ is the number of surfaces.
1612 The pressure in the z-direction is corrected by scaling the height of
1613 the box with $\mu_{zz}$
1614 \beq
1615 \Delta P_{zz} = \frac{\Delta t}{\tau_p} \{ P_{0zz} - P_{zz}(t) \}
1616 \eeq
1617 \beq
1618 \mu_{zz} = 1 + \beta_{zz} \Delta P_{zz}
1619 \eeq
1620 This is similar to normal pressure coupling, except that the factor
1621 of $1/3$ is missing.
1622 The pressure correction in the $z$-direction is then used to get the
1623 correct convergence for the surface tension to the reference value $\gamma_0$.
1624 The correction factor for the box length in the $x$/$y$-direction is
1625 \beq
1626 \mu_{x/y} = 1 + \frac{\Delta t}{2\,\tau_p} \beta_{x/y}
1627 \left( \frac{n \gamma_0}{\mu_{zz} L_z}
1628 - \left\{ P_{zz}(t)+\Delta P_{zz} - \frac{P_{xx}(t) + P_{yy}(t)}{2} \right\}
1629 \right)
1630 \eeq
1631 The value of $\beta_{zz}$ is more critical than with normal pressure
1632 coupling. Normally an incorrect compressibility will just scale $\tau_p$,
1633 but with surface tension coupling it affects the convergence of the surface
1634 tension.
1635 When $\beta_{zz}$ is set to zero (constant box height), $\Delta P_{zz}$ is also set
1636 to zero, which is necessary for obtaining the correct surface tension.
1638 \subsubsection{MTTK pressure control algorithms}
1640 As mentioned in the previous section, one weakness of leap-frog
1641 integration is in constant pressure simulations, since the pressure
1642 requires a calculation of both the virial and the kinetic energy at
1643 the full time step; for leap-frog, this information is not available
1644 until {\em after} the full timestep. Velocity Verlet does allow the
1645 calculation, at the cost of an extra round of global communication,
1646 and can compute, mod any integration errors, the true NPT ensemble.
1648 The full equations, combining both pressure coupling and temperature
1649 coupling, are taken from Martyna {\em et al.}~\cite{Martyna1996} and
1650 Tuckerman~\cite{Tuckerman2006} and are referred to here as MTTK
1651 equations (Martyna-Tuckerman-Tobias-Klein). We introduce for
1652 convenience $\epsilon = (1/3)\ln (V/V_0)$, where $V_0$ is a reference
1653 volume. The momentum of $\epsilon$ is $\veps = p_{\epsilon}/W =
1654 \dot{\epsilon} = \dot{V}/3V$, and define $\alpha = 1 + 3/N_{dof}$ (see
1655 Ref~\cite{Tuckerman2006})
1657 The isobaric equations are
1658 \bea
1659 \dot{\rv}_i &=& \frac{\pb_i}{m_i} + \frac{\peps}{W} \rv_i \nonumber \\
1660 \frac{\dot{\pb}_i}{m_i} &=& \frac{1}{m_i}\F_i - \alpha\frac{\peps}{W} \frac{\pb_i}{m_i} \nonumber \\
1661 \dot{\epsilon} &=& \frac{\peps}{W} \nonumber \\
1662 \frac{\dot{\peps}}{W} &=& \frac{3V}{W}(P_{\mathrm{int}} - P) + (\alpha-1)\left(\sum_{n=1}^N\frac{\pb_i^2}{m_i}\right),\\
1663 \eea
1664 where
1665 \bea
1666 P_{\mathrm{int}} &=& P_{\mathrm{kin}} -P_{\mathrm{vir}} = \frac{1}{3V}\left[\sum_{i=1}^N \left(\frac{\pb_i^2}{2m_i} - \rv_i \cdot \F_i\
1667 \right)\right].
1668 \eea
1669 The terms including $\alpha$ are required to make phase space
1670 incompressible~\cite{Tuckerman2006}. The $\epsilon$ acceleration term
1671 can be rewritten as
1672 \bea
1673 \frac{\dot{\peps}}{W} &=& \frac{3V}{W}\left(\alpha P_{\mathrm{kin}} - P_{\mathrm{vir}} - P\right)
1674 \eea
1675 In terms of velocities, these equations become
1676 \bea
1677 \dot{\rv}_i &=& \vv_i + \veps \rv_i \nonumber \\
1678 \dot{\vv}_i &=& \frac{1}{m_i}\F_i - \alpha\veps \vv_i \nonumber \\
1679 \dot{\epsilon} &=& \veps \nonumber \\
1680 \dot{\veps} &=& \frac{3V}{W}(P_{\mathrm{int}} - P) + (\alpha-1)\left( \sum_{n=1}^N \frac{1}{2} m_i \vv_i^2\right)\nonumber \\
1681 P_{\mathrm{int}} &=& P_{\mathrm{kin}} - P_{\mathrm{vir}} = \frac{1}{3V}\left[\sum_{i=1}^N \left(\frac{1}{2} m_i\vv_i^2 - \rv_i \cdot \F_i\right)\right]
1682 \eea
1683 For these equations, the conserved quantity is
1684 \bea
1685 H = \sum_{i=1}^{N} \frac{\pb_i^2}{2m_i} + U\left(\rv_1,\rv_2,\ldots,\rv_N\right) + \frac{p_\epsilon}{2W} + PV
1686 \eea
1687 The next step is to add temperature control. Adding Nos{\'e}-Hoover
1688 chains, including to the barostat degree of freedom, where we use
1689 $\eta$ for the barostat Nos{\'e}-Hoover variables, and $Q^{\prime}$
1690 for the coupling constants of the thermostats of the barostats, we get
1691 \bea
1692 \dot{\rv}_i &=& \frac{\pb_i}{m_i} + \frac{\peps}{W} \rv_i \nonumber \\
1693 \frac{\dot{\pb}_i}{m_i} &=& \frac{1}{m_i}\F_i - \alpha\frac{\peps}{W} \frac{\pb_i}{m_i} - \frac{p_{\xi_1}}{Q_1}\frac{\pb_i}{m_i}\nonumber \\
1694 \dot{\epsilon} &=& \frac{\peps}{W} \nonumber \\
1695 \frac{\dot{\peps}}{W} &=& \frac{3V}{W}(\alpha P_{\mathrm{kin}} - P_{\mathrm{vir}} - P) -\frac{p_{\eta_1}}{Q^{\prime}_1}\peps \nonumber \\
1696 \dot{\xi}_k &=& \frac{p_{\xi_k}}{Q_k} \nonumber \\
1697 \dot{\eta}_k &=& \frac{p_{\eta_k}}{Q^{\prime}_k} \nonumber \\
1698 \dot{p}_{\xi_k} &=& G_k - \frac{p_{\xi_{k+1}}}{Q_{k+1}} \;\;\;\; k=1,\ldots, M-1 \nonumber \\
1699 \dot{p}_{\eta_k} &=& G^\prime_k - \frac{p_{\eta_{k+1}}}{Q^\prime_{k+1}} \;\;\;\; k=1,\ldots, M-1 \nonumber \\
1700 \dot{p}_{\xi_M} &=& G_M \nonumber \\
1701 \dot{p}_{\eta_M} &=& G^\prime_M, \nonumber \\
1702 \eea
1703 where
1704 \bea
1705 P_{\mathrm{int}} &=& P_{\mathrm{kin}} - P_{\mathrm{vir}} = \frac{1}{3V}\left[\sum_{i=1}^N \left(\frac{\pb_i^2}{2m_i} - \rv_i \cdot \F_i\right)\right] \nonumber \\
1706 G_1 &=& \sum_{i=1}^N \frac{\pb^2_i}{m_i} - N_f kT \nonumber \\
1707 G_k &=& \frac{p^2_{\xi_{k-1}}}{2Q_{k-1}} - kT \;\; k = 2,\ldots,M \nonumber \\
1708 G^\prime_1 &=& \frac{\peps^2}{2W} - kT \nonumber \\
1709 G^\prime_k &=& \frac{p^2_{\eta_{k-1}}}{2Q^\prime_{k-1}} - kT \;\; k = 2,\ldots,M
1710 \eea
1711 The conserved quantity is now
1712 \bea
1713 H = \sum_{i=1}^{N} \frac{\pb_i}{2m_i} + U\left(\rv_1,\rv_2,\ldots,\rv_N\right) + \frac{p^2_\epsilon}{2W} + PV + \nonumber \\
1714 \sum_{k=1}^M\frac{p^2_{\xi_k}}{2Q_k} +\sum_{k=1}^M\frac{p^2_{\eta_k}}{2Q^{\prime}_k} + N_fkT\xi_1 + kT\sum_{i=2}^M \xi_k + kT\sum_{k=1}^M \eta_k
1715 \eea
1716 Returning to the Trotter decomposition formalism, for pressure control and temperature control~\cite{Martyna1996} we get:
1717 \bea
1718 iL = iL_1 + iL_2 + iL_{\epsilon,1} + iL_{\epsilon,2} + iL_{\mathrm{NHC-baro}} + iL_{\mathrm{NHC}}
1719 \eea
1720 where ``NHC-baro'' corresponds to the Nos{\`e}-Hoover chain of the barostat,
1721 and NHC corresponds to the NHC of the particles,
1722 \bea
1723 iL_1 &=& \sum_{i=1}^N \left[\frac{\pb_i}{m_i} + \frac{\peps}{W}\rv_i\right]\cdot \frac{\partial}{\partial \rv_i} \\
1724 iL_2 &=& \sum_{i=1}^N \F_i - \alpha \frac{\peps}{W}\pb_i \cdot \frac{\partial}{\partial \pb_i} \\
1725 iL_{\epsilon,1} &=& \frac{p_\epsilon}{W} \frac{\partial}{\partial \epsilon}\\
1726 iL_{\epsilon,2} &=& G_{\epsilon} \frac{\partial}{\partial p_\epsilon}
1727 \eea
1728 and where
1729 \bea
1730 G_{\epsilon} = 3V\left(\alpha P_{\mathrm{kin}} - P_{\mathrm{vir}} - P\right)
1731 \eea
1732 Using the Trotter decomposition, we get
1733 \bea
1734 \exp(iL\dt) &=& \exp\left(iL_{\mathrm{NHC-baro}}\dt/2\right)\exp\left(iL_{\mathrm{NHC}}\dt/2\right) \nonumber \nonumber \\
1735 &&\exp\left(iL_{\epsilon,2}\dt/2\right) \exp\left(iL_2 \dt/2\right) \nonumber \nonumber \\
1736 &&\exp\left(iL_{\epsilon,1}\dt\right) \exp\left(iL_1 \dt\right) \nonumber \nonumber \\
1737 &&\exp\left(iL_2 \dt/2\right) \exp\left(iL_{\epsilon,2}\dt/2\right) \nonumber \nonumber \\
1738 &&\exp\left(iL_{\mathrm{NHC}}\dt/2\right)\exp\left(iL_{\mathrm{NHC-baro}}\dt/2\right) + \mathcal{O}(\dt^3)
1739 \eea
1740 The action of $\exp\left(iL_1 \dt\right)$ comes from the solution of
1741 the the differential equation
1742 $\dot{\rv}_i = \vv_i + \veps \rv_i$
1743 with $\vv_i = \pb_i/m_i$ and $\veps$ constant with initial condition
1744 $\rv_i(0)$, evaluate at $t=\Delta t$. This yields the evolution
1745 \beq
1746 \rv_i(\dt) = \rv_i(0)e^{\veps \dt} + \Delta t \vv_i(0) e^{\veps \dt/2} \sinhx{\veps \dt/2}.
1747 \eeq
1748 The action of $\exp\left(iL_2 \dt/2\right)$ comes from the solution
1749 of the differential equation $\dot{\vv}_i = \frac{\F_i}{m_i} -
1750 \alpha\veps\vv_i$, yielding
1751 \beq
1752 \vv_i(\dt/2) = \vv_i(0)e^{-\alpha\veps \dt/2} + \frac{\Delta t}{2m_i}\F_i(0) e^{-\alpha\veps \dt/4}\sinhx{\alpha\veps \dt/4}.
1753 \eeq
1754 {\em md-vv-avek} uses the full step kinetic energies for determining the pressure with the pressure control,
1755 but the half-step-averaged kinetic energy for the temperatures, which can be written as a Trotter decomposition as
1756 \bea
1757 \exp(iL\dt) &=& \exp\left(iL_{\mathrm{NHC-baro}}\dt/2\right)\nonumber \exp\left(iL_{\epsilon,2}\dt/2\right) \exp\left(iL_2 \dt/2\right) \nonumber \\
1758 &&\exp\left(iL_{\mathrm{NHC}}\dt/2\right) \exp\left(iL_{\epsilon,1}\dt\right) \exp\left(iL_1 \dt\right) \exp\left(iL_{\mathrm{NHC}}\dt/2\right) \nonumber \\
1759 &&\exp\left(iL_2 \dt/2\right) \exp\left(iL_{\epsilon,2}\dt/2\right) \exp\left(iL_{\mathrm{NHC-baro}}\dt/2\right) + \mathcal{O}(\dt^3)
1760 \eea
1762 With constraints, the equations become significantly more complicated,
1763 in that each of these equations need to be solved iteratively for the
1764 constraint forces. Before {\gromacs} 5.1, these iterative
1765 constraints were solved as described in~\cite{Yu2010}. From {\gromacs}
1766 5.1 onward, MTTK with constraints has been removed because of
1767 numerical stability issues with the iterations.
1769 \subsubsection{Infrequent evaluation of temperature and pressure coupling}
1771 Temperature and pressure control require global communication to
1772 compute the kinetic energy and virial, which can become costly if
1773 performed every step for large systems. We can rearrange the Trotter
1774 decomposition to give alternate symplectic, reversible integrator with
1775 the coupling steps every $n$ steps instead of every steps. These new
1776 integrators will diverge if the coupling time step is too large, as
1777 the auxiliary variable integrations will not converge. However, in
1778 most cases, long coupling times are more appropriate, as they disturb
1779 the dynamics less~\cite{Martyna1996}.
1781 Standard velocity Verlet with Nos{\'e}-Hoover temperature control has a Trotter expansion
1782 \bea
1783 \exp(iL\dt) &\approx& \exp\left(iL_{\mathrm{NHC}}\dt/2\right) \exp\left(iL_2 \dt/2\right) \nonumber \\
1784 &&\exp\left(iL_1 \dt\right) \exp\left(iL_2 \dt/2\right) \exp\left(iL_{\mathrm{NHC}}\dt/2\right).
1785 \eea
1786 If the Nos{\'e}-Hoover chain is sufficiently slow with respect to the motions of the system, we can
1787 write an alternate integrator over $n$ steps for velocity Verlet as
1788 \bea
1789 \exp(iL\dt) &\approx& (\exp\left(iL_{\mathrm{NHC}}(n\dt/2)\right)\left[\exp\left(iL_2 \dt/2\right)\right. \nonumber \\
1790 &&\left.\exp\left(iL_1 \dt\right) \exp\left(iL_2 \dt/2\right)\right]^n \exp\left(iL_{\mathrm{NHC}}(n\dt/2)\right).
1791 \eea
1792 For pressure control, this becomes
1793 \bea
1794 \exp(iL\dt) &\approx& \exp\left(iL_{\mathrm{NHC-baro}}(n\dt/2)\right)\exp\left(iL_{\mathrm{NHC}}(n\dt/2)\right) \nonumber \nonumber \\
1795 &&\exp\left(iL_{\epsilon,2}(n\dt/2)\right) \left[\exp\left(iL_2 \dt/2\right)\right. \nonumber \nonumber \\
1796 &&\exp\left(iL_{\epsilon,1}\dt\right) \exp\left(iL_1 \dt\right) \nonumber \nonumber \\
1797 &&\left.\exp\left(iL_2 \dt/2\right)\right]^n \exp\left(iL_{\epsilon,2}(n\dt/2)\right) \nonumber \nonumber \\
1798 &&\exp\left(iL_{\mathrm{NHC}}(n\dt/2)\right)\exp\left(iL_{\mathrm{NHC-baro}}(n\dt/2)\right),
1799 \eea
1800 where the box volume integration occurs every step, but the auxiliary variable
1801 integrations happen every $n$ steps.
1805 \subsection{The complete update algorithm}
1806 \begin{figure}
1807 \begin{center}
1808 \addtolength{\fboxsep}{0.5cm}
1809 \begin{shadowenv}[12cm]
1810 {\large \bf THE UPDATE ALGORITHM}
1811 \rule{\textwidth}{2pt} \\
1812 Given:\\
1813 Positions $\ve{r}$ of all atoms at time $t$ \\
1814 Velocities $\ve{v}$ of all atoms at time $t-\hDt$ \\
1815 Accelerations $\ve{F}/m$ on all atoms at time $t$.\\
1816 (Forces are computed disregarding any constraints)\\
1817 Total kinetic energy and virial at $t-\Dt$\\
1818 $\Downarrow$ \\
1819 {\bf 1.} Compute the scaling factors $\lambda$ and $\mu$\\
1820 according to \eqnsref{lambda}{mu}\\
1821 $\Downarrow$ \\
1822 {\bf 2.} Update and scale velocities: $\ve{v}' = \lambda (\ve{v} +
1823 \ve{a} \Delta t)$ \\
1824 $\Downarrow$ \\
1825 {\bf 3.} Compute new unconstrained coordinates: $\ve{r}' = \ve{r} + \ve{v}'
1826 \Delta t$ \\
1827 $\Downarrow$ \\
1828 {\bf 4.} Apply constraint algorithm to coordinates: constrain($\ve{r}^{'} \rightarrow \ve{r}'';
1829 \, \ve{r}$) \\
1830 $\Downarrow$ \\
1831 {\bf 5.} Correct velocities for constraints: $\ve{v} = (\ve{r}'' -
1832 \ve{r}) / \Delta t$ \\
1833 $\Downarrow$ \\
1834 {\bf 6.} Scale coordinates and box: $\ve{r} = \mu \ve{r}''; \ve{b} =
1835 \mu \ve{b}$ \\
1836 \end{shadowenv}
1837 \caption{The MD update algorithm with the leap-frog integrator}
1838 \label{fig:complete-update}
1839 \end{center}
1840 \end{figure}
1841 The complete algorithm for the update of velocities and coordinates is
1842 given using leap-frog in \figref{complete-update}. The SHAKE algorithm of step
1843 4 is explained below.
1845 {\gromacs} has a provision to ``freeze'' (prevent motion of) selected
1846 particles\index{frozen atoms}, which must be defined as a ``\swapindex{freeze}{group}.'' This is implemented
1847 using a {\em freeze factor $\ve{f}_g$}, which is a vector, and differs for each
1848 freeze group (see \secref{groupconcept}). This vector contains only
1849 zero (freeze) or one (don't freeze).
1850 When we take this freeze factor and the external acceleration $\ve{a}_h$ into
1851 account the update algorithm for the velocities becomes
1852 \beq
1853 \ve{v}(t+\hdt)~=~\ve{f}_g * \lambda * \left[ \ve{v}(t-\hdt) +\frac{\ve{F}(t)}{m}\Delta t + \ve{a}_h \Delta t \right],
1854 \eeq
1855 where $g$ and $h$ are group indices which differ per atom.
1857 \subsection{Output step}
1858 The most important output of the MD run is the {\em
1859 \swapindex{trajectory}{file}}, which contains particle coordinates
1860 and (optionally) velocities at regular intervals.
1861 The trajectory file contains frames that could include positions,
1862 velocities and/or forces, as well as information about the dimensions
1863 of the simulation volume, integration step, integration time, etc. The
1864 interpretation of the time varies with the integrator chosen, as
1865 described above. For Velocity Verlet integrators, velocities labeled
1866 at time $t$ are for that time. For other integrators (e.g. leap-frog,
1867 stochastic dynamics), the velocities labeled at time $t$ are for time
1868 $t - \hDt$.
1870 Since the trajectory
1871 files are lengthy, one should not save every step! To retain all
1872 information it suffices to write a frame every 15 steps, since at
1873 least 30 steps are made per period of the highest frequency in the
1874 system, and Shannon's \normindex{sampling} theorem states that two samples per
1875 period of the highest frequency in a band-limited signal contain all
1876 available information. But that still gives very long files! So, if
1877 the highest frequencies are not of interest, 10 or 20 samples per ps
1878 may suffice. Be aware of the distortion of high-frequency motions by
1879 the {\em stroboscopic effect}, called {\em aliasing}: higher frequencies
1880 are mirrored with respect to the sampling frequency and appear as
1881 lower frequencies.
1883 {\gromacs} can also write reduced-precision coordinates for a subset of
1884 the simulation system to a special compressed trajectory file
1885 format. All the other tools can read and write this format. See
1886 the User Guide for details on how to set up your {\tt .mdp} file
1887 to have {\tt mdrun} use this feature.
1889 \section{Shell molecular dynamics}
1890 {\gromacs} can simulate \normindex{polarizability} using the
1891 \normindex{shell model} of Dick and Overhauser~\cite{Dick58}. In such models
1892 a shell particle representing the electronic degrees of freedom is
1893 attached to a nucleus by a spring. The potential energy is minimized with
1894 respect to the shell position at every step of the simulation (see below).
1895 Successful applications of shell models in {\gromacs} have been published
1896 for $N_2$~\cite{Jordan95} and water~\cite{Maaren2001a}.
1898 \subsection{Optimization of the shell positions}
1899 The force \ve{F}$_S$ on a shell particle $S$ can be decomposed into two
1900 components
1901 \begin{equation}
1902 \ve{F}_S ~=~ \ve{F}_{bond} + \ve{F}_{nb}
1903 \end{equation}
1904 where \ve{F}$_{bond}$ denotes the component representing the
1905 polarization energy, usually represented by a harmonic potential and
1906 \ve{F}$_{nb}$ is the sum of Coulomb and van der Waals interactions. If we
1907 assume that \ve{F}$_{nb}$ is almost constant we can analytically derive the
1908 optimal position of the shell, i.e. where \ve{F}$_S$ = 0. If we have the
1909 shell S connected to atom A we have
1910 \begin{equation}
1911 \ve{F}_{bond} ~=~ k_b \left( \ve{x}_S - \ve{x}_A\right).
1912 \end{equation}
1913 In an iterative solver, we have positions \ve{x}$_S(n)$ where $n$ is
1914 the iteration count. We now have at iteration $n$
1915 \begin{equation}
1916 \ve{F}_{nb} ~=~ \ve{F}_S - k_b \left( \ve{x}_S(n) - \ve{x}_A\right)
1917 \end{equation}
1918 and the optimal position for the shells $x_S(n+1)$ thus follows from
1919 \begin{equation}
1920 \ve{F}_S - k_b \left( \ve{x}_S(n) - \ve{x}_A\right) + k_b \left( \ve{x}_S(n+1) - \ve{x}_A\right) = 0
1921 \end{equation}
1922 if we write
1923 \begin{equation}
1924 \Delta \ve{x}_S = \ve{x}_S(n+1) - \ve{x}_S(n)
1925 \end{equation}
1926 we finally obtain
1927 \begin{equation}
1928 \Delta \ve{x}_S = \ve{F}_S/k_b
1929 \end{equation}
1930 which then yields the algorithm to compute the next trial in the optimization
1931 of shell positions
1932 \begin{equation}
1933 \ve{x}_S(n+1) ~=~ \ve{x}_S(n) + \ve{F}_S/k_b.
1934 \end{equation}
1936 \section{Constraint algorithms\index{constraint algorithms}}
1937 Constraints can be imposed in {\gromacs} using LINCS (default) or
1938 the traditional SHAKE method.
1940 \subsection{\normindex{SHAKE}}
1941 \label{subsec:SHAKE}
1942 The SHAKE~\cite{Ryckaert77} algorithm changes a set of unconstrained
1943 coordinates $\ve{r}^{'}$ to a set of coordinates $\ve{r}''$ that
1944 fulfill a list of distance constraints, using a set $\ve{r}$
1945 reference, as
1946 \beq
1947 {\rm SHAKE}(\ve{r}^{'} \rightarrow \ve{r}'';\, \ve{r})
1948 \eeq
1949 This action is consistent with solving a set of Lagrange multipliers
1950 in the constrained equations of motion. SHAKE needs a {\em relative tolerance};
1951 it will continue until all constraints are satisfied within
1952 that relative tolerance. An error message is
1953 given if SHAKE cannot reset the coordinates because the deviation is
1954 too large, or if a given number of iterations is surpassed.
1956 Assume the equations of motion must fulfill $K$ holonomic constraints,
1957 expressed as
1958 \beq
1959 \sigma_k(\ve{r}_1 \ldots \ve{r}_N) = 0; \;\; k=1 \ldots K.
1960 \eeq
1961 For example, $(\ve{r}_1 - \ve{r}_2)^2 - b^2 = 0$.
1962 Then the forces are defined as
1963 \beq
1964 - \frac{\partial}{\partial \ve{r}_i} \left( V + \sum_{k=1}^K \lambda_k
1965 \sigma_k \right),
1966 \eeq
1967 where $\lambda_k$ are Lagrange multipliers which must be solved to
1968 fulfill the constraint equations. The second part of this sum
1969 determines the {\em constraint forces} $\ve{G}_i$, defined by
1970 \beq
1971 \ve{G}_i = -\sum_{k=1}^K \lambda_k \frac{\partial \sigma_k}{\partial
1972 \ve{r}_i}
1973 \eeq
1974 The displacement due to the constraint forces in the leap-frog or
1975 Verlet algorithm is equal to $(\ve{G}_i/m_i)(\Dt)^2$. Solving the
1976 Lagrange multipliers (and hence the displacements) requires the
1977 solution of a set of coupled equations of the second degree. These are
1978 solved iteratively by SHAKE.
1979 \label{subsec:SETTLE}
1980 For the special case of rigid water molecules, that often make up more
1981 than 80\% of the simulation system we have implemented the
1982 \normindex{SETTLE}
1983 algorithm~\cite{Miyamoto92} (\secref{constraints}).
1985 For velocity Verlet, an additional round of constraining must be
1986 done, to constrain the velocities of the second velocity half step,
1987 removing any component of the velocity parallel to the bond vector.
1988 This step is called RATTLE, and is covered in more detail in the
1989 original Andersen paper~\cite{Andersen1983a}.
1995 \newcommand{\fs}[1]{\begin{equation} \label{eqn:#1}}
1996 \newcommand{\fe}{\end{equation}}
1997 \newcommand{\p}{\partial}
1998 \newcommand{\Bm}{\ve{B}}
1999 \newcommand{\M}{\ve{M}}
2000 \newcommand{\iM}{\M^{-1}}
2001 \newcommand{\Tm}{\ve{T}}
2002 \newcommand{\Sm}{\ve{S}}
2003 \newcommand{\fo}{\ve{f}}
2004 \newcommand{\con}{\ve{g}}
2005 \newcommand{\lenc}{\ve{d}}
2007 \subsection{\normindex{LINCS}}
2008 \label{subsec:lincs}
2010 \subsubsection{The LINCS algorithm}
2011 LINCS is an algorithm that resets bonds to their correct lengths
2012 after an unconstrained update~\cite{Hess97}.
2013 The method is non-iterative, as it always uses two steps.
2014 Although LINCS is based on matrices, no matrix-matrix multiplications are
2015 needed. The method is more stable and faster than SHAKE,
2016 but it can only be used with bond constraints and
2017 isolated angle constraints, such as the proton angle in OH.
2018 Because of its stability, LINCS is especially useful for Brownian dynamics.
2019 LINCS has two parameters, which are explained in the subsection parameters.
2020 The parallel version of LINCS, P-LINCS, is described
2021 in subsection \ssecref{plincs}.
2023 \subsubsection{The LINCS formulas}
2024 We consider a system of $N$ particles, with positions given by a
2025 $3N$ vector $\ve{r}(t)$.
2026 For molecular dynamics the equations of motion are given by Newton's Law
2027 \fs{c1}
2028 {\de^2 \ve{r} \over \de t^2} = \iM \ve{F},
2030 where $\ve{F}$ is the $3N$ force vector
2031 and $\M$ is a $3N \times 3N$ diagonal matrix,
2032 containing the masses of the particles.
2033 The system is constrained by $K$ time-independent constraint equations
2034 \fs{c2}
2035 g_i(\ve{r}) = | \ve{r}_{i_1}-\ve{r}_{i_2} | - d_i = 0 ~~~~~~i=1,\ldots,K.
2038 In a numerical integration scheme, LINCS is applied after an
2039 unconstrained update, just like SHAKE. The algorithm works in two
2040 steps (see figure \figref{lincs}). In the first step, the projections
2041 of the new bonds on the old bonds are set to zero. In the second step,
2042 a correction is applied for the lengthening of the bonds due to
2043 rotation. The numerics for the first step and the second step are very
2044 similar. A complete derivation of the algorithm can be found in
2045 \cite{Hess97}. Only a short description of the first step is given
2046 here.
2048 \begin{figure}
2049 \centerline{\includegraphics[height=50mm]{plots/lincs}}
2050 \caption[The three position updates needed for one time step.]{The
2051 three position updates needed for one time step. The dashed line is
2052 the old bond of length $d$, the solid lines are the new bonds. $l=d
2053 \cos \theta$ and $p=(2 d^2 - l^2)^{1 \over 2}$.}
2054 \label{fig:lincs}
2055 \end{figure}
2057 A new notation is introduced for the gradient matrix of the constraint
2058 equations which appears on the right hand side of this equation:
2059 \fs{c3}
2060 B_{hi} = {\p g_h \over \p r_i}
2062 Notice that $\Bm$ is a $K \times 3N$ matrix, it contains the directions
2063 of the constraints.
2064 The following equation shows how the new constrained coordinates
2065 $\ve{r}_{n+1}$ are related to the unconstrained coordinates
2066 $\ve{r}_{n+1}^{unc}$ by
2067 \fs{m0}
2068 \begin{array}{c}
2069 \ve{r}_{n+1}=(\ve{I}-\Tm_n \ve{B}_n) \ve{r}_{n+1}^{unc} + \Tm_n \lenc=
2070 \\[2mm]
2071 \ve{r}_{n+1}^{unc} -
2072 \iM \Bm_n (\Bm_n \iM \Bm_n^T)^{-1} (\Bm_n \ve{r}_{n+1}^{unc} - \lenc)
2073 \end{array}
2075 where $\Tm = \iM \Bm^T (\Bm \iM \Bm^T)^{-1}$.
2076 The derivation of this equation from \eqnsref{c1}{c2} can be found
2077 in \cite{Hess97}.
2079 This first step does not set the real bond lengths to the prescribed lengths,
2080 but the projection of the new bonds onto the old directions of the bonds.
2081 To correct for the rotation of bond $i$, the projection of the
2082 bond, $p_i$, on the old direction is set to
2083 \fs{m1a}
2084 p_i=\sqrt{2 d_i^2 - l_i^2},
2086 where $l_i$ is the bond length after the first projection.
2087 The corrected positions are
2088 \fs{m1b}
2089 \ve{r}_{n+1}^*=(\ve{I}-\Tm_n \Bm_n)\ve{r}_{n+1} + \Tm_n \ve{p}.
2091 This correction for rotational effects is actually an iterative process,
2092 but during MD only one iteration is applied.
2093 The relative constraint deviation after this procedure will be less than
2094 0.0001 for every constraint.
2095 In energy minimization, this might not be accurate enough, so the number
2096 of iterations is equal to the order of the expansion (see below).
2098 Half of the CPU time goes to inverting the constraint coupling
2099 matrix $\Bm_n \iM \Bm_n^T$, which has to be done every time step.
2100 This $K \times K$ matrix
2101 has $1/m_{i_1} + 1/m_{i_2}$ on the diagonal.
2102 The off-diagonal elements are only non-zero when two bonds are connected,
2103 then the element is
2104 $\cos \phi /m_c$, where $m_c$ is
2105 the mass of the atom connecting the
2106 two bonds and $\phi$ is the angle between the bonds.
2108 The matrix $\Tm$ is inverted through a power expansion.
2109 A $K \times K$ matrix $\ve{S}$ is
2110 introduced which is the inverse square root of
2111 the diagonal of $\Bm_n \iM \Bm_n^T$.
2112 This matrix is used to convert the diagonal elements
2113 of the coupling matrix to one:
2114 \fs{m2}
2115 \begin{array}{c}
2116 (\Bm_n \iM \Bm_n^T)^{-1}
2117 = \Sm \Sm^{-1} (\Bm_n \iM \Bm_n^T)^{-1} \Sm^{-1} \Sm \\[2mm]
2118 = \Sm (\Sm \Bm_n \iM \Bm_n^T \Sm)^{-1} \Sm =
2119 \Sm (\ve{I} - \ve{A}_n)^{-1} \Sm
2120 \end{array}
2122 The matrix $\ve{A}_n$ is symmetric and sparse and has zeros on the diagonal.
2123 Thus a simple trick can be used to calculate the inverse:
2124 \fs{m3}
2125 (\ve{I}-\ve{A}_n)^{-1}=
2126 \ve{I} + \ve{A}_n + \ve{A}_n^2 + \ve{A}_n^3 + \ldots
2129 This inversion method is only valid if the absolute values of all the
2130 eigenvalues of $\ve{A}_n$ are smaller than one.
2131 In molecules with only bond constraints, the connectivity is so low
2132 that this will always be true, even if ring structures are present.
2133 Problems can arise in angle-constrained molecules.
2134 By constraining angles with additional distance constraints,
2135 multiple small ring structures are introduced.
2136 This gives a high connectivity, leading to large eigenvalues.
2137 Therefore LINCS should NOT be used with coupled angle-constraints.
2139 For molecules with all bonds constrained the eigenvalues of $A$
2140 are around 0.4. This means that with each additional order
2141 in the expansion \eqnref{m3} the deviations decrease by a factor 0.4.
2142 But for relatively isolated triangles of constraints the largest
2143 eigenvalue is around 0.7.
2144 Such triangles can occur when removing hydrogen angle vibrations
2145 with an additional angle constraint in alcohol groups
2146 or when constraining water molecules with LINCS, for instance
2147 with flexible constraints.
2148 The constraints in such triangles converge twice as slow as
2149 the other constraints. Therefore, starting with {\gromacs} 4,
2150 additional terms are added to the expansion for such triangles
2151 \fs{m3_ang}
2152 (\ve{I}-\ve{A}_n)^{-1} \approx
2153 \ve{I} + \ve{A}_n + \ldots + \ve{A}_n^{N_i} +
2154 \left(\ve{A}^*_n + \ldots + {\ve{A}_n^*}^{N_i} \right) \ve{A}_n^{N_i}
2156 where $N_i$ is the normal order of the expansion and
2157 $\ve{A}^*$ only contains the elements of $\ve{A}$ that couple
2158 constraints within rigid triangles, all other elements are zero.
2159 In this manner, the accuracy of angle constraints comes close
2160 to that of the other constraints, while the series of matrix vector
2161 multiplications required for determining the expansion
2162 only needs to be extended for a few constraint couplings.
2163 This procedure is described in the P-LINCS paper\cite{Hess2008a}.
2165 \subsubsection{The LINCS Parameters}
2166 The accuracy of LINCS depends on the number of matrices used
2167 in the expansion \eqnref{m3}. For MD calculations a fourth order
2168 expansion is enough. For Brownian dynamics with
2169 large time steps an eighth order expansion may be necessary.
2170 The order is a parameter in the {\tt *.mdp} file.
2171 The implementation of LINCS is done in such a way that the
2172 algorithm will never crash. Even when it is impossible to
2173 to reset the constraints LINCS will generate a conformation
2174 which fulfills the constraints as well as possible.
2175 However, LINCS will generate a warning when in one step a bond
2176 rotates over more than a predefined angle.
2177 This angle is set by the user in the {\tt *.mdp} file.
2181 \section{Simulated Annealing}
2182 \label{sec:SA}
2183 The well known \swapindex{simulated}{annealing}
2184 (SA) protocol is supported in {\gromacs}, and you can even couple multiple
2185 groups of atoms separately with an arbitrary number of reference temperatures
2186 that change during the simulation. The annealing is implemented by simply
2187 changing the current reference temperature for each group in the temperature
2188 coupling, so the actual relaxation and coupling properties depends on the
2189 type of thermostat you use and how hard you are coupling it. Since we are
2190 changing the reference temperature it is important to remember that the system
2191 will NOT instantaneously reach this value - you need to allow for the inherent
2192 relaxation time in the coupling algorithm too. If you are changing the
2193 annealing reference temperature faster than the temperature relaxation you
2194 will probably end up with a crash when the difference becomes too large.
2196 The annealing protocol is specified as a series of corresponding times and
2197 reference temperatures for each group, and you can also choose whether you only
2198 want a single sequence (after which the temperature will be coupled to the
2199 last reference value), or if the annealing should be periodic and restart at
2200 the first reference point once the sequence is completed. You can mix and
2201 match both types of annealing and non-annealed groups in your simulation.
2203 \newcommand{\vrond}{\stackrel{\circ}{\ve{r}}}
2204 \newcommand{\rond}{\stackrel{\circ}{r}}
2205 \newcommand{\ruis}{\ve{r}^G}
2207 \section{Stochastic Dynamics\swapindexquiet{stochastic}{dynamics}}
2208 \label{sec:SD}
2209 Stochastic or velocity \swapindex{Langevin}{dynamics} adds a friction
2210 and a noise term to Newton's equations of motion, as
2211 \beq
2212 \label{SDeq}
2213 m_i {\de^2 \ve{r}_i \over \de t^2} =
2214 - m_i \gamma_i {\de \ve{r}_i \over \de t} + \ve{F}_i(\ve{r}) + \vrond_i,
2215 \eeq
2216 where $\gamma_i$ is the friction constant $[1/\mbox{ps}]$ and
2217 $\vrond_i\!\!(t)$ is a noise process with
2218 $\langle \rond_i\!\!(t) \rond_j\!\!(t+s) \rangle =
2219 2 m_i \gamma_i k_B T \delta(s) \delta_{ij}$.
2220 When $1/\gamma_i$ is large compared to the time scales present in the system,
2221 one could see stochastic dynamics as molecular dynamics with stochastic
2222 temperature-coupling. But any processes that take longer than $1/\gamma_i$,
2223 e.g. hydrodynamics, will be dampened. Since each degree of freedom is coupled
2224 independently to a heat bath, equilibration of fast modes occurs rapidly.
2225 For simulating a system in vacuum there is the additional advantage that there is no
2226 accumulation of errors for the overall translational and rotational
2227 degrees of freedom.
2228 When $1/\gamma_i$ is small compared to the time scales present in the system,
2229 the dynamics will be completely different from MD, but the sampling is
2230 still correct.
2232 In {\gromacs} there is one simple and efficient implementation. Its
2233 accuracy is equivalent to the normal MD leap-frog and
2234 Velocity Verlet integrator. It is nearly identical to the common way of discretizing the Langevin equation, but the friction and velocity term are applied in an impulse fashion~\cite{Goga2012}.
2235 It can be described as:
2236 \bea
2237 \label{eqn:sd_int1}
2238 \ve{v}' &~=~& \ve{v}(t-\hDt) + \frac{1}{m}\ve{F}(t)\Dt \\
2239 \Delta\ve{v} &~=~& -\alpha \, \ve{v}'(t+\hDt) + \sqrt{\frac{k_B T}{m}(1 - \alpha^2)} \, \ruis_i \\
2240 \ve{r}(t+\Dt) &~=~& \ve{r}(t)+\left(\ve{v}' +\frac{1}{2}\Delta \ve{v}\right)\Dt \label{eqn:sd1_x_upd}\\
2241 \ve{v}(t+\hDt) &~=~& \ve{v}' + \Delta \ve{v} \\
2242 \alpha &~=~& 1 - e^{-\gamma \Dt}
2243 \eea
2244 where $\ruis_i$ is Gaussian distributed noise with $\mu = 0$, $\sigma = 1$.
2245 The velocity is first updated a full time step without friction and noise to get $\ve{v}'$, identical to the normal update in leap-frog. The friction and noise are then applied as an impulse at step $t+\Dt$. The advantage of this scheme is that the velocity-dependent terms act at the full time step, which makes the correct integration of forces that depend on both coordinates and velocities, such as constraints and dissipative particle dynamics (DPD, not implented yet), straightforward. With constraints, the coordinate update \eqnref{sd1_x_upd} is split into a normal leap-frog update and a $\Delta \ve{v}$. After both of these updates the constraints are applied to coordinates and velocities.
2247 When using SD as a thermostat, an appropriate value for $\gamma$ is e.g. 0.5 ps$^{-1}$,
2248 since this results in a friction that is lower than the internal friction
2249 of water, while it still provides efficient thermostatting.
2252 \section{Brownian Dynamics\swapindexquiet{Brownian}{dynamics}}
2253 \label{sec:BD}
2254 In the limit of high friction, stochastic dynamics reduces to
2255 Brownian dynamics, also called position Langevin dynamics.
2256 This applies to over-damped systems,
2257 {\ie} systems in which the inertia effects are negligible.
2258 The equation is
2259 \beq
2260 {\de \ve{r}_i \over \de t} = \frac{1}{\gamma_i} \ve{F}_i(\ve{r}) + \vrond_i
2261 \eeq
2262 where $\gamma_i$ is the friction coefficient $[\mbox{amu/ps}]$ and
2263 $\vrond_i\!\!(t)$ is a noise process with
2264 $\langle \rond_i\!\!(t) \rond_j\!\!(t+s) \rangle =
2265 2 \delta(s) \delta_{ij} k_B T / \gamma_i$.
2266 In {\gromacs} the equations are integrated with a simple, explicit scheme
2267 \beq
2268 \ve{r}_i(t+\Delta t) = \ve{r}_i(t) +
2269 {\Delta t \over \gamma_i} \ve{F}_i(\ve{r}(t))
2270 + \sqrt{2 k_B T {\Delta t \over \gamma_i}}\, \ruis_i,
2271 \eeq
2272 where $\ruis_i$ is Gaussian distributed noise with $\mu = 0$, $\sigma = 1$.
2273 The friction coefficients $\gamma_i$ can be chosen the same for all
2274 particles or as $\gamma_i = m_i\,\gamma_i$, where the friction constants
2275 $\gamma_i$ can be different for different groups of atoms.
2276 Because the system is assumed to be over-damped, large timesteps
2277 can be used. LINCS should be used for the constraints since SHAKE
2278 will not converge for large atomic displacements.
2279 BD is an option of the {\tt mdrun} program.
2281 \section{Energy Minimization}
2282 \label{sec:EM}\index{energy minimization}%
2283 Energy minimization in {\gromacs} can be done using steepest descent,
2284 conjugate gradients, or l-bfgs (limited-memory
2285 Broyden-Fletcher-Goldfarb-Shanno quasi-Newtonian minimizer...we
2286 prefer the abbreviation). EM is just an option of the {\tt mdrun}
2287 program.
2289 \subsection{Steepest Descent\index{steepest descent}}
2290 Although steepest descent is certainly not the most efficient
2291 algorithm for searching, it is robust and easy to implement.
2293 We define the vector $\ve{r}$ as the vector of all $3N$ coordinates.
2294 Initially a maximum displacement $h_0$ ({\eg} 0.01 nm) must be given.
2296 First the forces $\ve{F}$ and potential energy are calculated.
2297 New positions are calculated by
2298 \beq
2299 \ve{r}_{n+1} = \ve{r}_n + \frac{\ve{F}_n}{\max (|\ve{F}_n|)} h_n,
2300 \eeq
2301 where $h_n$ is the maximum displacement and $\ve{F}_n$ is the force,
2302 or the negative gradient of the potential $V$. The notation $\max
2303 (|\ve{F}_n|)$ means the largest scalar force on any atom.
2304 The forces and energy are again computed for the new positions \\
2305 If ($V_{n+1} < V_n$) the new positions are accepted and $h_{n+1} = 1.2
2306 h_n$. \\
2307 If ($V_{n+1} \geq V_n$) the new positions are rejected and $h_n = 0.2 h_n$.
2309 The algorithm stops when either a user-specified number of force
2310 evaluations has been performed ({\eg} 100), or when the maximum of the absolute
2311 values of the force (gradient) components is smaller than a specified
2312 value $\epsilon$.
2313 Since force truncation produces some noise in the
2314 energy evaluation, the stopping criterion should not be made too tight
2315 to avoid endless iterations. A reasonable value for $\epsilon$ can be
2316 estimated from the root mean square force $f$ a harmonic oscillator would exhibit at a
2317 temperature $T$. This value is
2318 \beq
2319 f = 2 \pi \nu \sqrt{ 2mkT},
2320 \eeq
2321 where $\nu$ is the oscillator frequency, $m$ the (reduced) mass, and
2322 $k$ Boltzmann's constant. For a weak oscillator with a wave number of
2323 100 cm$^{-1}$ and a mass of 10 atomic units, at a temperature of 1 K,
2324 $f=7.7$ kJ~mol$^{-1}$~nm$^{-1}$. A value for $\epsilon$ between 1 and
2325 10 is acceptable.
2327 \subsection{Conjugate Gradient\index{conjugate gradient}}
2328 Conjugate gradient is slower than steepest descent in the early stages
2329 of the minimization, but becomes more efficient closer to the energy
2330 minimum. The parameters and stop criterion are the same as for
2331 steepest descent. In {\gromacs} conjugate gradient can not be used
2332 with constraints, including the SETTLE algorithm for
2333 water~\cite{Miyamoto92}, as this has not been implemented. If water is
2334 present it must be of a flexible model, which can be specified in the
2335 {\tt *.mdp} file by {\tt define = -DFLEXIBLE}.
2337 This is not really a restriction, since the accuracy of conjugate
2338 gradient is only required for minimization prior to a normal-mode
2339 analysis, which cannot be performed with constraints. For most other
2340 purposes steepest descent is efficient enough.
2342 \subsection{\normindex{L-BFGS}}
2343 The original BFGS algorithm works by successively creating better
2344 approximations of the inverse Hessian matrix, and moving the system to
2345 the currently estimated minimum. The memory requirements for this are
2346 proportional to the square of the number of particles, so it is not
2347 practical for large systems like biomolecules. Instead, we use the
2348 L-BFGS algorithm of Nocedal~\cite{Byrd95a,Zhu97a}, which approximates
2349 the inverse Hessian by a fixed number of corrections from previous
2350 steps. This sliding-window technique is almost as efficient as the
2351 original method, but the memory requirements are much lower -
2352 proportional to the number of particles multiplied with the correction
2353 steps. In practice we have found it to converge faster than conjugate
2354 gradients, but due to the correction steps it is not yet parallelized.
2355 It is also noteworthy that switched or shifted interactions usually
2356 improve the convergence, since sharp cut-offs mean the potential
2357 function at the current coordinates is slightly different from the
2358 previous steps used to build the inverse Hessian approximation.
2360 \section{Normal-Mode Analysis\index{normal-mode analysis}\index{NMA}}
2361 Normal-mode analysis~\cite{Levitt83,Go83,BBrooks83b}
2362 can be performed using {\gromacs}, by diagonalization of the mass-weighted
2363 \normindex{Hessian} $H$:
2364 \bea
2365 R^T M^{-1/2} H M^{-1/2} R &=& \mbox{diag}(\lambda_1,\ldots,\lambda_{3N})
2367 \lambda_i &=& (2 \pi \omega_i)^2
2368 \eea
2369 where $M$ contains the atomic masses, $R$ is a matrix that contains
2370 the eigenvectors as columns, $\lambda_i$ are the eigenvalues
2371 and $\omega_i$ are the corresponding frequencies.
2373 First the Hessian matrix, which is a $3N \times 3N$ matrix where $N$
2374 is the number of atoms, needs to be calculated:
2375 \bea
2376 H_{ij} &=& \frac{\partial^2 V}{\partial x_i \partial x_j}
2377 \eea
2378 where $x_i$ and $x_j$ denote the atomic x, y or z coordinates.
2379 In practice, this equation is not used, but the Hessian is
2380 calculated numerically from the force as:
2381 \bea
2382 H_{ij} &=& -
2383 \frac{f_i({\bf x}+h{\bf e}_j) - f_i({\bf x}-h{\bf e}_j)}{2h}
2385 f_i &=& - \frac{\partial V}{\partial x_i}
2386 \eea
2387 where ${\bf e}_j$ is the unit vector in direction $j$.
2388 It should be noted that
2389 for a usual normal-mode calculation, it is necessary to completely minimize
2390 the energy prior to computation of the Hessian.
2391 The tolerance required depends on the type of system,
2392 but a rough indication is 0.001 kJ mol$^{-1}$.
2393 Minimization should be done with conjugate gradients or L-BFGS in double precision.
2395 A number of {\gromacs} programs are involved in these
2396 calculations. First, the energy should be minimized using {\tt mdrun}.
2397 Then, {\tt mdrun} computes the Hessian. {\bf Note} that for generating
2398 the run input file, one should use the minimized conformation from
2399 the full precision trajectory file, as the structure file is not
2400 accurate enough.
2401 {\tt \normindex{gmx nmeig}} does the diagonalization and
2402 the sorting of the normal modes according to their frequencies.
2403 Both {\tt mdrun} and {\tt gmx nmeig} should be run in double precision.
2404 The normal modes can be analyzed with the program {\tt gmx anaeig}.
2405 Ensembles of structures at any temperature and for any subset of
2406 normal modes can be generated with {\tt \normindex{gmx nmens}}.
2407 An overview of normal-mode analysis and the related principal component
2408 analysis (see \secref{covanal}) can be found in~\cite{Hayward95b}.
2411 \section{Free energy calculations\index{free energy calculations}}
2412 \label{sec:fecalc}
2413 \subsection{Slow-growth methods\index{slow-growth methods}}
2414 Free energy calculations can be performed
2415 in {\gromacs} using a number of methods, including ``slow-growth.'' An example problem
2416 might be calculating the difference in free energy of binding of an inhibitor {\bf I}
2417 to an enzyme {\bf E} and to a mutated enzyme {\bf E$^{\prime}$}. It
2418 is not feasible with computer simulations to perform a docking
2419 calculation for such a large complex, or even releasing the inhibitor from
2420 the enzyme in a reasonable amount of computer time with reasonable accuracy.
2421 However, if we consider the free energy cycle in~\figref{free}A
2422 we can write:
2423 \beq
2424 \Delta G_1 - \Delta G_2 = \Delta G_3 - \Delta G_4
2425 \label{eqn:ddg}
2426 \eeq
2427 If we are interested in the left-hand term we can equally well compute
2428 the right-hand term.
2429 \begin{figure}
2430 \centerline{\includegraphics[width=6cm,angle=270]{plots/free1}\hspace{2cm}\includegraphics[width=6cm,angle=270]{plots/free2}}
2431 \caption[Free energy cycles.]{Free energy cycles. {\bf A:} to
2432 calculate $\Delta G_{12}$, the free energy difference between the
2433 binding of inhibitor {\bf I} to enzymes {\bf E} respectively {\bf
2434 E$^{\prime}$}. {\bf B:} to calculate $\Delta G_{12}$, the free energy
2435 difference for binding of inhibitors {\bf I} respectively {\bf I$^{\prime}$} to
2436 enzyme {\bf E}.}
2437 \label{fig:free}
2438 \end{figure}
2440 If we want to compute the difference in free energy of binding of two
2441 inhibitors {\bf I} and {\bf I$^{\prime}$} to an enzyme {\bf E} (\figref{free}B)
2442 we can again use \eqnref{ddg} to compute the desired property.
2444 \newcommand{\sA}{^{\mathrm{A}}}
2445 \newcommand{\sB}{^{\mathrm{B}}}
2446 Free energy differences between two molecular species can
2447 be calculated in {\gromacs} using the ``slow-growth'' method.
2448 Such free energy differences between different molecular species are
2449 physically meaningless, but they can be used to obtain meaningful
2450 quantities employing a thermodynamic cycle.
2451 The method requires a simulation during which the Hamiltonian of the
2452 system changes slowly from that describing one system (A) to that
2453 describing the other system (B). The change must be so slow that the
2454 system remains in equilibrium during the process; if that requirement
2455 is fulfilled, the change is reversible and a slow-growth simulation from B to A
2456 will yield the same results (but with a different sign) as a slow-growth
2457 simulation from A to B. This is a useful check, but the user should be
2458 aware of the danger that equality of forward and backward growth results does
2459 not guarantee correctness of the results.
2461 The required modification of the Hamiltonian $H$ is realized by making
2462 $H$ a function of a \textit{coupling parameter} $\lambda:
2463 H=H(p,q;\lambda)$ in such a way that $\lambda=0$ describes system A
2464 and $\lambda=1$ describes system B:
2465 \beq
2466 H(p,q;0)=H\sA (p,q);~~~~ H(p,q;1)=H\sB (p,q).
2467 \eeq
2468 In {\gromacs}, the functional form of the $\lambda$-dependence is
2469 different for the various force-field contributions and is described
2470 in section \secref{feia}.
2472 The Helmholtz free energy $A$ is related to the
2473 partition function $Q$ of an $N,V,T$ ensemble, which is assumed to be
2474 the equilibrium ensemble generated by a MD simulation at constant
2475 volume and temperature. The generally more useful Gibbs free energy
2476 $G$ is related to the partition function $\Delta$ of an $N,p,T$
2477 ensemble, which is assumed to be the equilibrium ensemble generated by
2478 a MD simulation at constant pressure and temperature:
2479 \bea
2480 A(\lambda) &=& -k_BT \ln Q \\
2481 Q &=& c \int\!\!\int \exp[-\beta H(p,q;\lambda)]\,dp\,dq \\
2482 G(\lambda) &=& -k_BT \ln \Delta \\
2483 \Delta &=& c \int\!\!\int\!\!\int \exp[-\beta H(p,q;\lambda) -\beta
2484 pV]\,dp\,dq\,dV \\
2485 G &=& A + pV,
2486 \eea
2487 where $\beta = 1/(k_BT)$ and $c = (N! h^{3N})^{-1}$.
2488 These integrals over phase space cannot be evaluated from a
2489 simulation, but it is possible to evaluate the derivative with
2490 respect to $\lambda$ as an ensemble average:
2491 \beq
2492 \frac{dA}{d\lambda} = \frac{\int\!\!\int (\partial H/ \partial
2493 \lambda) \exp[-\beta H(p,q;\lambda)]\,dp\,dq}{\int\!\!\int \exp[-\beta
2494 H(p,q;\lambda)]\,dp\,dq} =
2495 \left\langle \frac{\partial H}{\partial \lambda} \right\rangle_{NVT;\lambda},
2496 \eeq
2497 with a similar relation for $dG/d\lambda$ in the $N,p,T$
2498 ensemble. The difference in free energy between A and B can be found
2499 by integrating the derivative over $\lambda$:
2500 \bea
2501 A\sB(V,T)-A\sA(V,T) &=& \int_0^1 \left\langle \frac{\partial
2502 H}{\partial \lambda} \right\rangle_{NVT;\lambda} \,d\lambda
2503 \label{eq:delA} \\
2504 G\sB(p,T)-G\sA(p,T) &=& \int_0^1 \left\langle \frac{\partial
2505 H}{\partial \lambda} \right\rangle_{NpT;\lambda} \,d\lambda.
2506 \label{eq:delG}
2507 \eea
2508 If one wishes to evaluate $G\sB(p,T)-G\sA(p,T)$,
2509 the natural choice is a constant-pressure simulation. However, this
2510 quantity can also be obtained from a slow-growth simulation at
2511 constant volume, starting with system A at pressure $p$ and volume $V$
2512 and ending with system B at pressure $p_B$, by applying the following
2513 small (but, in principle, exact) correction:
2514 \beq
2515 G\sB(p)-G\sA(p) =
2516 A\sB(V)-A\sA(V) - \int_p^{p\sB}[V\sB(p')-V]\,dp'
2517 \eeq
2518 Here we omitted the constant $T$ from the notation. This correction is
2519 roughly equal to $-\frac{1}{2} (p\sB-p)\Delta V=(\Delta V)^2/(2
2520 \kappa V)$, where $\Delta V$ is the volume change at $p$ and $\kappa$
2521 is the isothermal compressibility. This is usually
2522 small; for example, the growth of a water molecule from nothing
2523 in a bath of 1000 water molecules at constant volume would produce an
2524 additional pressure of as much as 22 bar, but a correction to the
2525 Helmholtz free energy of just -1 kJ mol$^{-1}$. %-20 J/mol.
2527 In Cartesian coordinates, the kinetic energy term in the Hamiltonian
2528 depends only on the momenta, and can be separately integrated and, in
2529 fact, removed from the equations. When masses do not change, there is
2530 no contribution from the kinetic energy at all; otherwise the
2531 integrated contribution to the free energy is $-\frac{3}{2} k_BT \ln
2532 (m\sB/m\sA)$. {\bf Note} that this is only true in the absence of constraints.
2534 \subsection{Thermodynamic integration\index{thermodynamic integration}\index{BAR}\index{Bennett's acceptance ratio}}
2535 {\gromacs} offers the possibility to integrate eq.~\ref{eq:delA} or
2536 eq. \ref{eq:delG} in one simulation over the full range from A to
2537 B. However, if the change is large and insufficient sampling can be
2538 expected, the user may prefer to determine the value of $\langle
2539 dG/d\lambda \rangle$ accurately at a number of well-chosen
2540 intermediate values of $\lambda$. This can easily be done by setting
2541 the stepsize {\tt delta_lambda} to zero. Each simulation can be
2542 equilibrated first, and a proper error estimate can be made for each
2543 value of $dG/d\lambda$ from the fluctuation of $\partial H/\partial
2544 \lambda$. The total free energy change is then determined afterward
2545 by an appropriate numerical integration procedure.
2547 {\gromacs} now also supports the use of Bennett's Acceptance Ratio~\cite{Bennett1976}
2548 for calculating values of $\Delta$G for transformations from state A to state B using
2549 the program {\tt \normindex{gmx bar}}. The same data can also be used to calculate free
2550 energies using MBAR~\cite{Shirts2008}, though the analysis currently requires external tools from
2551 the external {\tt pymbar} package, at https://SimTK.org/home/pymbar.
2553 The $\lambda$-dependence for the force-field contributions is
2554 described in detail in section \secref{feia}.
2556 \section{Replica exchange\index{replica exchange}}
2557 Replica exchange molecular dynamics (\normindex{REMD})
2558 is a method that can be used to speed up
2559 the sampling of any type of simulation, especially if
2560 conformations are separated by relatively high energy barriers.
2561 It involves simulating multiple replicas of the same system
2562 at different temperatures and randomly exchanging the complete state
2563 of two replicas at regular intervals with the probability:
2564 \beq
2565 P(1 \leftrightarrow 2)=\min\left(1,\exp\left[
2566 \left(\frac{1}{k_B T_1} - \frac{1}{k_B T_2}\right)(U_1 - U_2)
2567 \right] \right)
2568 \eeq
2569 where $T_1$ and $T_2$ are the reference temperatures and $U_1$ and $U_2$
2570 are the instantaneous potential energies of replicas 1 and 2 respectively.
2571 After exchange the velocities are scaled by $(T_1/T_2)^{\pm0.5}$
2572 and a neighbor search is performed the next step.
2573 This combines the fast sampling and frequent barrier-crossing
2574 of the highest temperature with correct Boltzmann sampling at
2575 all the different temperatures~\cite{Hukushima96a,Sugita99}.
2576 We only attempt exchanges for neighboring temperatures as the probability
2577 decreases very rapidly with the temperature difference.
2578 One should not attempt exchanges for all possible pairs in one step.
2579 If, for instance, replicas 1 and 2 would exchange, the chance of
2580 exchange for replicas 2 and 3 not only depends on the energies of
2581 replicas 2 and 3, but also on the energy of replica 1.
2582 In {\gromacs} this is solved by attempting exchange for all ``odd''
2583 pairs on ``odd'' attempts and for all ``even'' pairs on ``even'' attempts.
2584 If we have four replicas: 0, 1, 2 and 3, ordered in temperature
2585 and we attempt exchange every 1000 steps, pairs 0-1 and 2-3
2586 will be tried at steps 1000, 3000 etc. and pair 1-2 at steps 2000, 4000 etc.
2588 How should one choose the temperatures?
2589 The energy difference can be written as:
2590 \beq
2591 U_1 - U_2 = N_{df} \frac{c}{2} k_B (T_1 - T_2)
2592 \eeq
2593 where $N_{df}$ is the total number of degrees of freedom of one replica
2594 and $c$ is 1 for harmonic potentials and around 2 for protein/water systems.
2595 If $T_2 = (1+\epsilon) T_1$ the probability becomes:
2596 \beq
2597 P(1 \leftrightarrow 2)
2598 = \exp\left( -\frac{\epsilon^2 c\,N_{df}}{2 (1+\epsilon)} \right)
2599 \approx \exp\left(-\epsilon^2 \frac{c}{2} N_{df} \right)
2600 \eeq
2601 Thus for a probability of $e^{-2}\approx 0.135$
2602 one obtains $\epsilon \approx 2/\sqrt{c\,N_{df}}$.
2603 With all bonds constrained one has $N_{df} \approx 2\, N_{atoms}$
2604 and thus for $c$ = 2 one should choose $\epsilon$ as $1/\sqrt{N_{atoms}}$.
2605 However there is one problem when using pressure coupling. The density at
2606 higher temperatures will decrease, leading to higher energy~\cite{Seibert2005a},
2607 which should be taken into account. The {\gromacs} website features a
2608 so-called ``REMD calculator,'' that lets you type in the temperature range and
2609 the number of atoms, and based on that proposes a set of temperatures.
2611 An extension to the REMD for the isobaric-isothermal ensemble was
2612 proposed by Okabe {\em et al.}~\cite{Okabe2001a}. In this work the
2613 exchange probability is modified to:
2614 \beq
2615 P(1 \leftrightarrow 2)=\min\left(1,\exp\left[
2616 \left(\frac{1}{k_B T_1} - \frac{1}{k_B T_2}\right)(U_1 - U_2) +
2617 \left(\frac{P_1}{k_B T_1} - \frac{P_2}{k_B T_2}\right)\left(V_1-V_2\right)
2618 \right] \right)
2619 \eeq
2620 where $P_1$ and $P_2$ are the respective reference pressures and $V_1$ and
2621 $V_2$ are the respective instantaneous volumes in the simulations.
2622 In most cases the differences in volume are so small that the second
2623 term is negligible. It only plays a role when the difference between
2624 $P_1$ and $P_2$ is large or in phase transitions.
2626 Hamiltonian replica exchange is also supported in {\gromacs}. In
2627 Hamiltonian replica exchange, each replica has a different
2628 Hamiltonian, defined by the free energy pathway specified for the simulation. The
2629 exchange probability to maintain the correct ensemble probabilities is:
2630 \beq P(1 \leftrightarrow 2)=\min\left(1,\exp\left[
2631 \left(\frac{1}{k_B T} - \frac{1}{k_B T}\right)((U_1(x_2) - U_1(x_1)) + (U_2(x_1) - U_2(x_2)))
2632 \right]
2633 \right)
2634 \eeq
2635 The separate Hamiltonians are defined by the free energy functionality
2636 of {\gromacs}, with swaps made between the different values of
2637 $\lambda$ defined in the mdp file.
2639 Hamiltonian and temperature replica exchange can also be performed
2640 simultaneously, using the acceptance criteria:
2641 \beq
2642 P(1 \leftrightarrow 2)=\min\left(1,\exp\left[
2643 \left(\frac{1}{k_B T} - \right)(\frac{U_1(x_2) - U_1(x_1)}{k_B T_1} + \frac{U_2(x_1) - U_2(x_2)}{k_B T_2})
2644 \right] \right)
2645 \eeq
2647 Gibbs sampling replica exchange has also been implemented in
2648 {\gromacs}~\cite{Chodera2011}. In Gibbs sampling replica exchange, all
2649 possible pairs are tested for exchange, allowing swaps between
2650 replicas that are not neighbors.
2652 Gibbs sampling replica exchange requires no additional potential
2653 energy calculations. However there is an additional communication
2654 cost in Gibbs sampling replica exchange, as for some permutations,
2655 more than one round of swaps must take place. In some cases, this
2656 extra communication cost might affect the efficiency.
2658 All replica exchange variants are options of the {\tt mdrun}
2659 program. It will only work when MPI is installed, due to the inherent
2660 parallelism in the algorithm. For efficiency each replica can run on a
2661 separate rank. See the manual page of {\tt mdrun} on how to use these
2662 multinode features.
2665 \section{Essential Dynamics sampling\index{essential dynamics}\index{principal component analysis}\seeindexquiet{PCA}{covariance analysis}}
2666 The results from Essential Dynamics (see \secref{covanal})
2667 of a protein can be used to guide MD simulations. The idea is that
2668 from an initial MD simulation (or from other sources) a definition of
2669 the collective fluctuations with largest amplitude is obtained. The
2670 position along one or more of these collective modes can be
2671 constrained in a (second) MD simulation in a number of ways for
2672 several purposes. For example, the position along a certain mode may
2673 be kept fixed to monitor the average force (free-energy gradient) on
2674 that coordinate in that position. Another application is to enhance
2675 sampling efficiency with respect to usual MD
2676 \cite{Degroot96a,Degroot96b}. In this case, the system is encouraged
2677 to sample its available configuration space more systematically than
2678 in a diffusion-like path that proteins usually take.
2680 Another possibility to enhance sampling is \normindex{flooding}.
2681 Here a flooding potential is added to certain
2682 (collective) degrees of freedom to expel the system out
2683 of a region of phase space \cite{Lange2006a}.
2685 The procedure for essential dynamics sampling or flooding is as follows.
2686 First, the eigenvectors and eigenvalues need to be determined
2687 using covariance analysis ({\tt gmx covar})
2688 or normal-mode analysis ({\tt gmx nmeig}).
2689 Then, this information is fed into {\tt make_edi},
2690 which has many options for selecting vectors and setting parameters,
2691 see {\tt gmx make_edi -h}.
2692 The generated {\tt edi} input file is then passed to {\tt mdrun}.
2695 \section{\normindex{Expanded Ensemble}}
2697 In an expanded ensemble simulation~\cite{Lyubartsev1992}, both the coordinates and the
2698 thermodynamic ensemble are treated as configuration variables that can
2699 be sampled over. The probability of any given state can be written as:
2700 \beq
2701 P(\vec{x},k) \propto \exp\left(-\beta_k U_k + g_k\right),
2702 \eeq
2703 where $\beta_k = \frac{1}{k_B T_k}$ is the $\beta$ corresponding to the $k$th
2704 thermodynamic state, and $g_k$ is a user-specified weight factor corresponding
2705 to the $k$th state. This space is therefore a {\em mixed}, {\em generalized}, or {\em
2706 expanded} ensemble which samples from multiple thermodynamic
2707 ensembles simultaneously. $g_k$ is chosen to give a specific weighting
2708 of each subensemble in the expanded ensemble, and can either be fixed,
2709 or determined by an iterative procedure. The set of $g_k$ is
2710 frequently chosen to give each thermodynamic ensemble equal
2711 probability, in which case $g_k$ is equal to the free energy in
2712 non-dimensional units, but they can be set to arbitrary values as
2713 desired. Several different algorithms can be used to equilibrate
2714 these weights, described in the mdp option listings.
2716 In {\gromacs}, this space is sampled by alternating sampling in the $k$
2717 and $\vec{x}$ directions. Sampling in the $\vec{x}$ direction is done
2718 by standard molecular dynamics sampling; sampling between the
2719 different thermodynamics states is done by Monte Carlo, with several
2720 different Monte Carlo moves supported. The $k$ states can be defined
2721 by different temperatures, or choices of the free energy $\lambda$
2722 variable, or both. Expanded ensemble simulations thus represent a
2723 serialization of the replica exchange formalism, allowing a single
2724 simulation to explore many thermodynamic states.
2728 \section{Parallelization\index{parallelization}}
2729 The CPU time required for a simulation can be reduced by running the simulation
2730 in parallel over more than one core.
2731 Ideally, one would want to have linear scaling: running on $N$ cores
2732 makes the simulation $N$ times faster. In practice this can only be
2733 achieved for a small number of cores. The scaling will depend
2734 a lot on the algorithms used. Also, different algorithms can have different
2735 restrictions on the interaction ranges between atoms.
2737 \section{Domain decomposition\index{domain decomposition}}
2738 Since most interactions in molecular simulations are local,
2739 domain decomposition is a natural way to decompose the system.
2740 In domain decomposition, a spatial domain is assigned to each rank,
2741 which will then integrate the equations of motion for the particles
2742 that currently reside in its local domain. With domain decomposition,
2743 there are two choices that have to be made: the division of the unit cell
2744 into domains and the assignment of the forces to domains.
2745 Most molecular simulation packages use the half-shell method for assigning
2746 the forces. But there are two methods that always require less communication:
2747 the eighth shell~\cite{Liem1991} and the midpoint~\cite{Shaw2006} method.
2748 {\gromacs} currently uses the eighth shell method, but for certain systems
2749 or hardware architectures it might be advantageous to use the midpoint
2750 method. Therefore, we might implement the midpoint method in the future.
2751 Most of the details of the domain decomposition can be found
2752 in the {\gromacs} 4 paper~\cite{Hess2008b}.
2754 \subsection{Coordinate and force communication}
2755 In the most general case of a triclinic unit cell,
2756 the space in divided with a 1-, 2-, or 3-D grid in parallelepipeds
2757 that we call domain decomposition cells.
2758 Each cell is assigned to a particle-particle rank.
2759 The system is partitioned over the ranks at the beginning
2760 of each MD step in which neighbor searching is performed.
2761 Since the neighbor searching is based on charge groups, charge groups
2762 are also the units for the domain decomposition.
2763 Charge groups are assigned to the cell where their center of geometry resides.
2764 Before the forces can be calculated, the coordinates from some
2765 neighboring cells need to be communicated,
2766 and after the forces are calculated, the forces need to be communicated
2767 in the other direction.
2768 The communication and force assignment is based on zones that
2769 can cover one or multiple cells.
2770 An example of a zone setup is shown in \figref{ddcells}.
2772 \begin{figure}
2773 \centerline{\includegraphics[width=6cm]{plots/dd-cells}}
2774 \caption{
2775 A non-staggered domain decomposition grid of 3$\times$2$\times$2 cells.
2776 Coordinates in zones 1 to 7 are communicated to the corner cell
2777 that has its home particles in zone 0.
2778 $r_c$ is the cut-off radius.
2779 \label{fig:ddcells}
2781 \end{figure}
2783 The coordinates are communicated by moving data along the ``negative''
2784 direction in $x$, $y$ or $z$ to the next neighbor. This can be done in one
2785 or multiple pulses. In \figref{ddcells} two pulses in $x$ are required,
2786 then one in $y$ and then one in $z$. The forces are communicated by
2787 reversing this procedure. See the {\gromacs} 4 paper~\cite{Hess2008b}
2788 for details on determining which non-bonded and bonded forces
2789 should be calculated on which rank.
2791 \subsection{Dynamic load balancing\swapindexquiet{dynamic}{load balancing}}
2792 When different ranks have a different computational load
2793 (load imbalance), all ranks will have to wait for the one
2794 that takes the most time. One would like to avoid such a situation.
2795 Load imbalance can occur due to four reasons:
2796 \begin{itemize}
2797 \item inhomogeneous particle distribution
2798 \item inhomogeneous interaction cost distribution (charged/uncharged,
2799 water/non-water due to {\gromacs} water innerloops)
2800 \item statistical fluctuation (only with small particle numbers)
2801 \item differences in communication time, due to network topology and/or other jobs on the machine interfering with our communication
2802 \end{itemize}
2803 So we need a dynamic load balancing algorithm
2804 where the volume of each domain decomposition cell
2805 can be adjusted {\em independently}.
2806 To achieve this, the 2- or 3-D domain decomposition grids need to be
2807 staggered. \figref{ddtric} shows the most general case in 2-D.
2808 Due to the staggering, one might require two distance checks
2809 for deciding if a charge group needs to be communicated:
2810 a non-bonded distance and a bonded distance check.
2812 \begin{figure}
2813 \centerline{\includegraphics[width=7cm]{plots/dd-tric}}
2814 \caption{
2815 The zones to communicate to the rank of zone 0,
2816 see the text for details. $r_c$ and $r_b$ are the non-bonded
2817 and bonded cut-off radii respectively, $d$ is an example
2818 of a distance between following, staggered boundaries of cells.
2819 \label{fig:ddtric}
2821 \end{figure}
2823 By default, {\tt mdrun} automatically turns on the dynamic load
2824 balancing during a simulation when the total performance loss
2825 due to the force calculation imbalance is 2\% or more.
2826 {\bf Note} that the reported force load imbalance numbers might be higher,
2827 since the force calculation is only part of work that needs to be done
2828 during an integration step.
2829 The load imbalance is reported in the log file at log output steps
2830 and when the {\tt -v} option is used also on screen.
2831 The average load imbalance and the total performance loss
2832 due to load imbalance are reported at the end of the log file.
2834 There is one important parameter for the dynamic load balancing,
2835 which is the minimum allowed scaling. By default, each dimension
2836 of the domain decomposition cell can scale down by at least
2837 a factor of 0.8. For 3-D domain decomposition this allows cells
2838 to change their volume by about a factor of 0.5, which should allow
2839 for compensation of a load imbalance of 100\%.
2840 The minimum allowed scaling can be changed with the {\tt -dds}
2841 option of {\tt mdrun}.
2843 The load imbalance is measured by timing a single region of the MD step
2844 on each MPI rank. This region can not include MPI communication, as
2845 timing of MPI calls does not allow separating wait due to imbalance
2846 from actual communication.
2847 The domain volumes are then scaled, with under-relaxation, inversely
2848 proportional with the measured time. This procedure will decrease the
2849 load imbalance when the change in load in the measured region correlates
2850 with the change in domain volume and the load outside
2851 the measured region does not depend strongly on the domain volume.
2852 In CPU-only simulations, the load is measured between the coordinate
2853 and the force communication. In simulations with non-bonded
2854 work on GPUs, we overlap communication and
2855 work on the CPU with calculation on the GPU. Therefore we
2856 measure from the last communication before the force calculation to
2857 when the CPU or GPU is finished, whichever is last.
2858 When not using PME ranks, we subtract the time in PME from the CPU time,
2859 as this includes MPI calls and the PME load is independent of domain size.
2860 This generally works well, unless the non-bonded load is low and there is
2861 imbalance in the bonded interactions. Then two issues can arise.
2862 Dynamic load balancing can increase the imbalance in update and constraints
2863 and with PME the coordinate and force redistribution time can go up
2864 significantly. Although dynamic load balancing
2865 can significantly improve performance in cases where there is imbalance in
2866 the bonded interactions on the CPU, there are many situations in which
2867 some domains continue decreasing in size and the load imbalance increases
2868 and/or PME coordinate and force redistribution cost increases significantly.
2869 As of version 2016.1, {\tt mdrun} disables the dynamic load balancing when
2870 measurement indicates that it deteriorates performance. This means that in most
2871 cases the user will get good performance with the default, automated
2872 dynamic load balancing setting.
2874 \subsection{Constraints in parallel\index{constraints}}
2875 \label{subsec:plincs}
2876 Since with domain decomposition parts of molecules can reside
2877 on different ranks, bond constraints can cross cell boundaries.
2878 Therefore a parallel constraint algorithm is required.
2879 {\gromacs} uses the \normindex{P-LINCS} algorithm~\cite{Hess2008a},
2880 which is the parallel version of the \normindex{LINCS} algorithm~\cite{Hess97}
2881 (see \ssecref{lincs}).
2882 The P-LINCS procedure is illustrated in \figref{plincs}.
2883 When molecules cross the cell boundaries, atoms in such molecules
2884 up to ({\tt lincs_order + 1}) bonds away are communicated over the cell boundaries.
2885 Then, the normal LINCS algorithm can be applied to the local bonds
2886 plus the communicated ones. After this procedure, the local bonds
2887 are correctly constrained, even though the extra communicated ones are not.
2888 One coordinate communication step is required for the initial LINCS step
2889 and one for each iteration. Forces do not need to be communicated.
2891 \begin{figure}
2892 \centerline{\includegraphics[width=6cm]{plots/par-lincs2}}
2893 \caption{
2894 Example of the parallel setup of P-LINCS with one molecule
2895 split over three domain decomposition cells, using a matrix
2896 expansion order of 3.
2897 The top part shows which atom coordinates need to be communicated
2898 to which cells. The bottom parts show the local constraints (solid)
2899 and the non-local constraints (dashed) for each of the three cells.
2900 \label{fig:plincs}
2902 \end{figure}
2904 \subsection{Interaction ranges}
2905 Domain decomposition takes advantage of the locality of interactions.
2906 This means that there will be limitations on the range of interactions.
2907 By default, {\tt mdrun} tries to find the optimal balance between
2908 interaction range and efficiency. But it can happen that a simulation
2909 stops with an error message about missing interactions,
2910 or that a simulation might run slightly faster with shorter
2911 interaction ranges. A list of interaction ranges
2912 and their default values is given in \tabref{dd_ranges}.
2914 \begin{table}
2915 \centerline{
2916 \begin{tabular}{|c|c|ll|}
2917 \dline
2918 interaction & range & option & default \\
2919 \dline
2920 non-bonded & $r_c$ = max($r_{\mathrm{list}}$,$r_{\mathrm{VdW}}$,$r_{\mathrm{Coul}}$) & {\tt mdp} file & \\
2921 two-body bonded & max($r_{\mathrm{mb}}$,$r_c$) & {\tt mdrun -rdd} & starting conf. + 10\% \\
2922 multi-body bonded & $r_{\mathrm{mb}}$ & {\tt mdrun -rdd} & starting conf. + 10\% \\
2923 constraints & $r_{\mathrm{con}}$ & {\tt mdrun -rcon} & est. from bond lengths \\
2924 virtual sites & $r_{\mathrm{con}}$ & {\tt mdrun -rcon} & 0 \\
2925 \dline
2926 \end{tabular}
2928 \caption{The interaction ranges with domain decomposition.}
2929 \label{tab:dd_ranges}
2930 \end{table}
2932 In most cases the defaults of {\tt mdrun} should not cause the simulation
2933 to stop with an error message of missing interactions.
2934 The range for the bonded interactions is determined from the distance
2935 between bonded charge-groups in the starting configuration, with 10\% added
2936 for headroom. For the constraints, the value of $r_{\mathrm{con}}$ is determined by
2937 taking the maximum distance that ({\tt lincs_order + 1}) bonds can cover
2938 when they all connect at angles of 120 degrees.
2939 The actual constraint communication is not limited by $r_{\mathrm{con}}$,
2940 but by the minimum cell size $L_C$, which has the following lower limit:
2941 \beq
2942 L_C \geq \max(r_{\mathrm{mb}},r_{\mathrm{con}})
2943 \eeq
2944 Without dynamic load balancing the system is actually allowed to scale
2945 beyond this limit when pressure scaling is used.
2946 {\bf Note} that for triclinic boxes, $L_C$ is not simply the box diagonal
2947 component divided by the number of cells in that direction,
2948 rather it is the shortest distance between the triclinic cells borders.
2949 For rhombic dodecahedra this is a factor of $\sqrt{3/2}$ shorter
2950 along $x$ and $y$.
2952 When $r_{\mathrm{mb}} > r_c$, {\tt mdrun} employs a smart algorithm to reduce
2953 the communication. Simply communicating all charge groups within
2954 $r_{\mathrm{mb}}$ would increase the amount of communication enormously.
2955 Therefore only charge-groups that are connected by bonded interactions
2956 to charge groups which are not locally present are communicated.
2957 This leads to little extra communication, but also to a slightly
2958 increased cost for the domain decomposition setup.
2959 In some cases, {\eg} coarse-grained simulations with a very short cut-off,
2960 one might want to set $r_{\mathrm{mb}}$ by hand to reduce this cost.
2962 \subsection{Multiple-Program, Multiple-Data PME parallelization\index{PME}}
2963 \label{subsec:mpmd_pme}
2964 Electrostatics interactions are long-range, therefore special
2965 algorithms are used to avoid summation over many atom pairs.
2966 In {\gromacs} this is usually
2967 PME (\secref{pme}).
2968 Since with PME all particles interact with each other, global communication
2969 is required. This will usually be the limiting factor for
2970 scaling with domain decomposition.
2971 To reduce the effect of this problem, we have come up with
2972 a Multiple-Program, Multiple-Data approach~\cite{Hess2008b}.
2973 Here, some ranks are selected to do only the PME mesh calculation,
2974 while the other ranks, called particle-particle (PP) ranks,
2975 do all the rest of the work.
2976 For rectangular boxes the optimal PP to PME rank ratio is usually 3:1,
2977 for rhombic dodecahedra usually 2:1.
2978 When the number of PME ranks is reduced by a factor of 4, the number
2979 of communication calls is reduced by about a factor of 16.
2980 Or put differently, we can now scale to 4 times more ranks.
2981 In addition, for modern 4 or 8 core machines in a network,
2982 the effective network bandwidth for PME is quadrupled,
2983 since only a quarter of the cores will be using the network connection
2984 on each machine during the PME calculations.
2986 \begin{figure}
2987 \centerline{\includegraphics[width=12cm]{plots/mpmd-pme}}
2988 \caption{
2989 Example of 8 ranks without (left) and with (right) MPMD.
2990 The PME communication (red arrows) is much higher on the left
2991 than on the right. For MPMD additional PP - PME coordinate
2992 and force communication (blue arrows) is required,
2993 but the total communication complexity is lower.
2994 \label{fig:mpmd_pme}
2996 \end{figure}
2998 {\tt mdrun} will by default interleave the PP and PME ranks.
2999 If the ranks are not number consecutively inside the machines,
3000 one might want to use {\tt mdrun -ddorder pp_pme}.
3001 For machines with a real 3-D torus and proper communication software
3002 that assigns the ranks accordingly one should use
3003 {\tt mdrun -ddorder cartesian}.
3005 To optimize the performance one should usually set up the cut-offs
3006 and the PME grid such that the PME load is 25 to 33\% of the total
3007 calculation load. {\tt grompp} will print an estimate for this load
3008 at the end and also {\tt mdrun} calculates the same estimate
3009 to determine the optimal number of PME ranks to use.
3010 For high parallelization it might be worthwhile to optimize
3011 the PME load with the {\tt mdp} settings and/or the number
3012 of PME ranks with the {\tt -npme} option of {\tt mdrun}.
3013 For changing the electrostatics settings it is useful to know
3014 the accuracy of the electrostatics remains nearly constant
3015 when the Coulomb cut-off and the PME grid spacing are scaled
3016 by the same factor.
3017 {\bf Note} that it is usually better to overestimate than to underestimate
3018 the number of PME ranks, since the number of PME ranks is smaller
3019 than the number of PP ranks, which leads to less total waiting time.
3021 The PME domain decomposition can be 1-D or 2-D along the $x$ and/or
3022 $y$ axis. 2-D decomposition is also known as \normindex{pencil decomposition} because of
3023 the shape of the domains at high parallelization.
3024 1-D decomposition along the $y$ axis can only be used when
3025 the PP decomposition has only 1 domain along $x$. 2-D PME decomposition
3026 has to have the number of domains along $x$ equal to the number of
3027 the PP decomposition. {\tt mdrun} automatically chooses 1-D or 2-D
3028 PME decomposition (when possible with the total given number of ranks),
3029 based on the minimum amount of communication for the coordinate redistribution
3030 in PME plus the communication for the grid overlap and transposes.
3031 To avoid superfluous communication of coordinates and forces
3032 between the PP and PME ranks, the number of DD cells in the $x$
3033 direction should ideally be the same or a multiple of the number
3034 of PME ranks. By default, {\tt mdrun} takes care of this issue.
3036 \subsection{Domain decomposition flow chart}
3037 In \figref{dd_flow} a flow chart is shown for domain decomposition
3038 with all possible communication for different algorithms.
3039 For simpler simulations, the same flow chart applies,
3040 without the algorithms and communication for
3041 the algorithms that are not used.
3043 \begin{figure}
3044 \centerline{\includegraphics[width=12cm]{plots/flowchart}}
3045 \caption{
3046 Flow chart showing the algorithms and communication (arrows)
3047 for a standard MD simulation with virtual sites, constraints
3048 and separate PME-mesh ranks.
3049 \label{fig:dd_flow}
3051 \end{figure}
3054 % LocalWords: GROningen MAchine BIOSON Groningen GROMACS Berendsen der Spoel
3055 % LocalWords: Drunen Comp Phys Comm ROck NS FFT pbc EM ifthenelse gmxlite ff
3056 % LocalWords: octahedra triclinic Ewald PME PPPM trjconv xy solvated
3057 % LocalWords: boxtypes boxshapes editconf Lennard COM XTC TNG kT defunits
3058 % LocalWords: Boltzmann's Mueller nb int mdrun chargegroup simplerc prefactor
3059 % LocalWords: pme waterloops CH NH CO df com virial integrator Verlet vverlet
3060 % LocalWords: integrators ref timepoint timestep timesteps mdp md vv avek NVE
3061 % LocalWords: NVT off's leapfrogv lll LR rmfast SPC fs Nos physicality ps GMX
3062 % LocalWords: Tcoupling nonergodic thermostatting NOSEHOOVER algorithmes ij yx
3063 % LocalWords: Parrinello Rahman rescales atm anisotropically ccc xz zx yy yz
3064 % LocalWords: zy zz se barostat compressibilities MTTK NPT Martyna al isobaric
3065 % LocalWords: Tuckerman vir PV fkT iLt iL Liouville NHC Eq baro mu trj mol bc
3066 % LocalWords: freezegroup Shannon's polarizability Overhauser barostats iLn KE
3067 % LocalWords: negligibly thermostatted Tobias rhombic maxwell et xtc tng TC rlist
3068 % LocalWords: waals LINCS holonomic plincs lincs unc ang SA Langevin SD amu BD
3069 % LocalWords: bfgs Broyden Goldfarb Shanno mkT kJ DFLEXIBLE Nocedal diag nmeig
3070 % LocalWords: diagonalization anaeig nmens covanal ddg feia BT dp dq pV dV dA
3071 % LocalWords: NpT eq stepsize REMD constrainted website Okabe MPI covar edi dd
3072 % LocalWords: progman NMR ddcells innerloops ddtric tric dds rdd conf rcon est
3073 % LocalWords: mb PP MPMD ddorder pp cartesian grompp npme parallelizable edr
3074 % LocalWords: macromolecule nstlist vacuo parallelization dof indices MBAR AVX
3075 % LocalWords: TOL numerics parallelized eigenvectors dG parallelepipeds VdW np
3076 % LocalWords: Coul multi solvation HCT OBC solv cav vdw Schaefer symplectic dt
3077 % LocalWords: pymbar multinode subensemble Monte solute subst groupconcept GPU
3078 % LocalWords: dodecahedron octahedron dodecahedra equilibration usinggroups nm
3079 % LocalWords: topologies rlistlong CUDA GPUs rcoulomb SIMD BlueGene FPUs erfc
3080 % LocalWords: cutoffschemesupport unbuffered bondeds OpenMP ewald rtol
3081 % LocalWords: verletdrift peptide RMS rescaling ergodicity ergodic discretized
3082 % LocalWords: isothermal compressibility isotropically anisotropic iteratively
3083 % LocalWords: incompressible integrations translational biomolecules NMA PCA
3084 % LocalWords: Bennett's equilibrated Hamiltonians covariance equilibrate
3085 % LocalWords: inhomogeneous conformational online other's th