498 |
|
{\tt minimizer} & string & Chooses a minimizer & Possible minimizers |
499 |
|
are SD and CG. Either {\tt ensemble} or {\tt minimizer} must be specified. \\ |
500 |
|
{\tt ensemble} & string & Sets the ensemble. & Possible ensembles are |
501 |
< |
NVE, NVT, NPTi, NPAT, NPTf, and NPTxyz. Either {\tt ensemble} |
501 |
> |
NVE, NVT, NPTi, NPAT, NPTf, NPTxyz, and LD. Either {\tt ensemble} |
502 |
|
or {\tt minimizer} must be specified. \\ |
503 |
|
{\tt dt} & fs & Sets the time step. & Selection of {\tt dt} should be |
504 |
|
small enough to sample the fastest motion of the simulation. ({\tt |
1905 |
|
& (with changes to box shape) & \\ |
1906 |
|
NPTxyz & approximate isobaric-isothermal & {\tt ensemble = NPTxyz;} \\ |
1907 |
|
& (with separate barostats on each box dimension) & \\ |
1908 |
+ |
LD & Langevin Dynamics & {\tt ensemble = LD;} \\ |
1909 |
+ |
& (approximates the effects of an implicit solvent) & \\ |
1910 |
|
\end{tabular} |
1911 |
|
\end{center} |
1912 |
|
|
2371 |
|
for use with liquid simulations, or in cases where there is |
2372 |
|
orientational anisotropy in the system (i.e. in lipid bilayer |
2373 |
|
simulations). |
2372 |
– |
|
2373 |
– |
|
2374 |
– |
\section{\label{sec:constraints}Constraint Methods} |
2374 |
|
|
2375 |
< |
\subsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond |
2377 |
< |
Constraints} |
2378 |
< |
|
2379 |
< |
In order to satisfy the constraints of fixed bond lengths within {\sc |
2380 |
< |
oopse}, we have implemented the {\sc rattle} algorithm of |
2381 |
< |
Andersen.\cite{andersen83} {\sc rattle} is a velocity-Verlet |
2382 |
< |
formulation of the {\sc shake} method\cite{ryckaert77} for iteratively |
2383 |
< |
solving the Lagrange multipliers which maintain the holonomic |
2384 |
< |
constraints. Both methods are covered in depth in the |
2385 |
< |
literature,\cite{leach01:mm,Allen87} and a detailed description of |
2386 |
< |
this method would be redundant. |
2387 |
< |
|
2388 |
< |
\subsection{\label{oopseSec:zcons}The Z-Constraint Method} |
2389 |
< |
|
2390 |
< |
A force auto-correlation method based on the fluctuation-dissipation |
2391 |
< |
theorem was developed by Roux and Karplus to investigate the dynamics |
2392 |
< |
of ions inside ion channels.\cite{Roux91} The time-dependent friction |
2393 |
< |
coefficient can be calculated from the deviation of the instantaneous |
2394 |
< |
force from its mean value: |
2395 |
< |
\begin{equation} |
2396 |
< |
\xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T, |
2397 |
< |
\end{equation} |
2398 |
< |
where% |
2399 |
< |
\begin{equation} |
2400 |
< |
\delta F(z,t)=F(z,t)-\langle F(z,t)\rangle. |
2401 |
< |
\end{equation} |
2375 |
> |
\section{Langevin Dynamics (LD)\label{LDRB}} |
2376 |
|
|
2377 |
< |
If the time-dependent friction decays rapidly, the static friction |
2378 |
< |
coefficient can be approximated by |
2379 |
< |
\begin{equation} |
2380 |
< |
\xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt. |
2381 |
< |
\end{equation} |
2382 |
< |
|
2409 |
< |
This allows the diffusion constant to then be calculated through the |
2410 |
< |
Einstein relation:\cite{Marrink94} |
2411 |
< |
\begin{equation} |
2412 |
< |
D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty |
2413 |
< |
}\langle\delta F(z,t)\delta F(z,0)\rangle dt}.% |
2414 |
< |
\end{equation} |
2415 |
< |
|
2416 |
< |
The Z-Constraint method, which fixes the $z$ coordinates of a few |
2417 |
< |
``tagged'' molecules with respect to the center of the mass of the |
2418 |
< |
system is a technique that was proposed to obtain the forces required |
2419 |
< |
for the force auto-correlation calculation.\cite{Marrink94} However, |
2420 |
< |
simply resetting the coordinate will move the center of the mass of |
2421 |
< |
the whole system. To avoid this problem, we have developed a new |
2422 |
< |
method that is utilized in {\sc oopse}. Instead of resetting the |
2423 |
< |
coordinates, we reset the forces of $z$-constrained molecules and |
2424 |
< |
subtract the total constraint forces from the rest of the system after |
2425 |
< |
the force calculation at each time step. |
2426 |
< |
|
2427 |
< |
After the force calculation, the total force on molecule $\alpha$ is: |
2428 |
< |
\begin{equation} |
2429 |
< |
G_{\alpha} = \sum_i F_{\alpha i}, |
2430 |
< |
\label{oopseEq:zc1} |
2431 |
< |
\end{equation} |
2432 |
< |
where $F_{\alpha i}$ is the force in the $z$ direction on atom $i$ in |
2433 |
< |
$z$-constrained molecule $\alpha$. The forces on the atoms in the |
2434 |
< |
$z$-constrained molecule are then adjusted to remove the total force |
2435 |
< |
on molecule $\alpha$: |
2436 |
< |
\begin{equation} |
2437 |
< |
F_{\alpha i} = F_{\alpha i} - |
2438 |
< |
\frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}. |
2439 |
< |
\end{equation} |
2440 |
< |
Here, $m_{\alpha i}$ is the mass of atom $i$ in the $z$-constrained |
2441 |
< |
molecule. After the forces have been adjusted, the velocities must |
2442 |
< |
also be modified to subtract out molecule $\alpha$'s center-of-mass |
2443 |
< |
velocity in the $z$ direction. |
2444 |
< |
\begin{equation} |
2445 |
< |
v_{\alpha i} = v_{\alpha i} - |
2446 |
< |
\frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}, |
2447 |
< |
\end{equation} |
2448 |
< |
where $v_{\alpha i}$ is the velocity of atom $i$ in the $z$ direction. |
2449 |
< |
Lastly, all of the accumulated constraint forces must be subtracted |
2450 |
< |
from the rest of the unconstrained system to keep the system center of |
2451 |
< |
mass of the entire system from drifting. |
2452 |
< |
\begin{equation} |
2453 |
< |
F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha} G_{\alpha}} |
2454 |
< |
{\sum_{\beta}\sum_i m_{\beta i}}, |
2455 |
< |
\end{equation} |
2456 |
< |
where $\beta$ denotes all {\it unconstrained} molecules in the |
2457 |
< |
system. Similarly, the velocities of the unconstrained molecules must |
2458 |
< |
also be scaled: |
2459 |
< |
\begin{equation} |
2460 |
< |
v_{\beta i} = v_{\beta i} + \sum_{\alpha} \frac{\sum_i m_{\alpha i} |
2461 |
< |
v_{\alpha i}}{\sum_i m_{\alpha i}}. |
2462 |
< |
\end{equation} |
2463 |
< |
|
2464 |
< |
This method will pin down the centers-of-mass of all of the |
2465 |
< |
$z$-constrained molecules, and will also keep the entire system fixed |
2466 |
< |
at the original system center-of-mass location. |
2377 |
> |
{\sc oopse} implements a Langevin integrator in order to perform |
2378 |
> |
molecular dynamics simulations in implicit solvent environments. This |
2379 |
> |
can result in substantial performance gains when the detailed dynamics |
2380 |
> |
of the solvent is not important. Since {\sc oopse} also handles rigid |
2381 |
> |
bodies of arbitrary composition and shape, the Langevin integrator is |
2382 |
> |
by necessity somewhat more complex than in other simulation packages. |
2383 |
|
|
2384 |
< |
At the very beginning of the simulation, the molecules may not be at |
2469 |
< |
their desired positions. To steer a $z$-constrained molecule to its |
2470 |
< |
specified position, a simple harmonic potential is used: |
2384 |
> |
Consider the Langevin equations of motion in generalized coordinates |
2385 |
|
\begin{equation} |
2472 |
– |
U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},% |
2473 |
– |
\end{equation} |
2474 |
– |
where $k_{\text{Harmonic}}$ is an harmonic force constant, $z(t)$ is |
2475 |
– |
the current $z$ coordinate of the center of mass of the constrained |
2476 |
– |
molecule, and $z_{\text{cons}}$ is the desired constraint |
2477 |
– |
position. The harmonic force operating on the $z$-constrained molecule |
2478 |
– |
at time $t$ can be calculated by |
2479 |
– |
\begin{equation} |
2480 |
– |
F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}= |
2481 |
– |
-k_{\text{Harmonic}}(z(t)-z_{\text{cons}}). |
2482 |
– |
\end{equation} |
2483 |
– |
|
2484 |
– |
The user may also specify the use of a constant velocity method |
2485 |
– |
(steered molecular dynamics) to move the molecules to their desired |
2486 |
– |
initial positions. Based on concepts from atomic force microscopy, |
2487 |
– |
{\sc smd} has been used to study many processes which occur via rare |
2488 |
– |
events on the time scale of a few hundreds of picoseconds. For |
2489 |
– |
example,{\sc smd} has been used to observe the dissociation of |
2490 |
– |
Streptavidin-biotin Complex.\cite{smd} |
2491 |
– |
|
2492 |
– |
To use of the $z$-constraint method in an {\sc oopse} simulation, the |
2493 |
– |
molecules must be specified using the {\tt nZconstraints} keyword in |
2494 |
– |
the meta-data file. The other parameters for modifying the behavior |
2495 |
– |
of the $z$-constraint method are listed in table~\ref{table:zconParams}. |
2496 |
– |
|
2497 |
– |
\begin{longtable}[c]{ABCD} |
2498 |
– |
\caption{Meta-data Keywords: Z-Constraint Parameters} |
2499 |
– |
\\ |
2500 |
– |
{\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline |
2501 |
– |
\endhead |
2502 |
– |
\hline |
2503 |
– |
\endfoot |
2504 |
– |
{\tt zconsTime} & fs & Sets the frequency at which the {\tt .fz} file |
2505 |
– |
is written & \\ |
2506 |
– |
{\tt zconsForcePolicy} & string & The strategy for subtracting |
2507 |
– |
the $z$-constraint force from the {\it unconstrained} molecules & Possible |
2508 |
– |
strategies are {\tt BYMASS} and {\tt BYNUMBER}. The default |
2509 |
– |
strategy is {\tt BYMASS}\\ |
2510 |
– |
{\tt zconsGap} & $\mbox{\AA}$ & Sets the distance between two adjacent |
2511 |
– |
constraint positions&Used mainly to move molecules through a |
2512 |
– |
simulation to estimate potentials of mean force. \\ |
2513 |
– |
{\tt zconsFixtime} & fs & Sets the length of time the $z$-constraint |
2514 |
– |
molecule is held fixed & {\tt zconsFixtime} must be set if {\tt |
2515 |
– |
zconsGap} is set\\ |
2516 |
– |
{\tt zconsUsingSMD} & logical & Flag for using Steered Molecular |
2517 |
– |
Dynamics to move the molecules to the correct constrained positions & |
2518 |
– |
Harmonic Forces are used by default |
2519 |
– |
\label{table:zconParams} |
2520 |
– |
\end{longtable} |
2521 |
– |
|
2522 |
– |
\section{Langevin Dynamics for Rigid Particles of Arbitrary Shape (LANGEVINDYNAMICS)\label{LDRB}} |
2523 |
– |
|
2524 |
– |
{\sc oopse} implements langevin dynamics integrator to perform the |
2525 |
– |
molecular dynamics simulations at implicit solvents environment to |
2526 |
– |
speed up the simulation when the properties of solvents are not |
2527 |
– |
important. Consider the Langevin equations of motion in generalized |
2528 |
– |
coordinates |
2529 |
– |
\begin{equation} |
2386 |
|
{\bf M} \dot{{\bf V}}(t) = {\bf F}_{s}(t) + |
2387 |
|
{\bf F}_{f}(t) + {\bf F}_{r}(t) |
2388 |
|
\label{LDGeneralizedForm} |
2436 |
|
2 k_B T \Xi_R \delta(t - t'). \label{eq:randomForce} |
2437 |
|
\end{equation} |
2438 |
|
$\Xi_R$ is the $6\times6$ resistance tensor at the center of |
2439 |
< |
resistance. Once this tensor is known for a given rigid body (as |
2440 |
< |
described in the previous section) obtaining a stochastic vector that |
2441 |
< |
has the properties in Eq. (\ref{eq:randomForce}) can be done |
2442 |
< |
efficiently by carrying out a one-time Cholesky decomposition to |
2443 |
< |
obtain the square root matrix of the resistance tensor, |
2439 |
> |
resistance. |
2440 |
> |
|
2441 |
> |
For atoms and ellipsoids, there are good approximations for this |
2442 |
> |
tensor that are based on Stokes' law. For arbitrary rigid bodies, the |
2443 |
> |
resistance tensor must be pre-computed before Langevin dynamics can be |
2444 |
> |
used. The {\sc oopse} distribution contains a utitilty program called |
2445 |
> |
Hydro that performs this computation. |
2446 |
> |
|
2447 |
> |
Once this tensor is known for a given {\tt integrableObject}, |
2448 |
> |
obtaining a stochastic vector that has the properties in |
2449 |
> |
Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a |
2450 |
> |
one-time Cholesky decomposition to obtain the square root matrix of |
2451 |
> |
the resistance tensor, |
2452 |
|
\begin{equation} |
2453 |
|
\Xi_R = {\bf S} {\bf S}^{T}, |
2454 |
|
\label{eq:Cholesky} |
2486 |
|
\begin{equation} |
2487 |
|
\frac{\partial}{\partial t}{\bf j}(t) = \tau^{~b}(t) |
2488 |
|
\end{equation} |
2489 |
< |
Embedding the friction and random forces into the the total force and |
2490 |
< |
torque, one can integrate the Langevin equations of motion for a rigid |
2491 |
< |
body of arbitrary shape in a velocity-Verlet style 2-part algorithm, |
2492 |
< |
where $h = \delta t$: |
2489 |
> |
By embedding the friction and random forces into the the total force |
2490 |
> |
and torque, {\sc oopse} integrates the Langevin equations of motion |
2491 |
> |
for a rigid body of arbitrary shape in a velocity-Verlet style 2-part |
2492 |
> |
algorithm, where $h = \delta t$: |
2493 |
|
|
2494 |
|
{\tt move A:} |
2495 |
|
\begin{align*} |
2594 |
|
+ \frac{h}{2} {\bf \tau}^{~b}(t + h) . |
2595 |
|
\end{align*} |
2596 |
|
|
2597 |
< |
The viscosity of the implicit of solvents must be specified using {\tt |
2598 |
< |
viscosity} keywords in the meta-data file to use langevin dynamics |
2599 |
< |
integrator. For simple shaped particles (spheres and ellipsoids), no |
2600 |
< |
further parameters are necessary. However, there are no analytical |
2601 |
< |
solutions for composite shaped particles, the approximate methods have |
2602 |
< |
to be applied to get the resistance tensor. The file which contains |
2603 |
< |
the information about hydro properties is indicated by {\tt |
2604 |
< |
HydroPropFile} keyword in meta-data file. The {\tt HydroPropFile} is |
2605 |
< |
precalculated by {\tt Hydro}. |
2597 |
> |
The viscosity of the implicit solvent must be specified using the {\tt |
2598 |
> |
viscosity} keyword in the meta-data file if the Langevin integrator is |
2599 |
> |
selected. For simple particles (spheres and ellipsoids), no further |
2600 |
> |
parameters are necessary. Since there are no analytic solutions for |
2601 |
> |
the resistance tensors for composite rigid bodies, the approximate |
2602 |
> |
tensors for these objects must also be specified in order to use |
2603 |
> |
Langevin dynamics. The meta-data file must therefore point to another |
2604 |
> |
file which contains the information about the hydrodynamic properties |
2605 |
> |
of all complex rigid bodies being used during the simulation. The |
2606 |
> |
{\tt HydroPropFile} keyword is used to specify the name of this |
2607 |
> |
file. A {\tt HydroPropFile} should be precalculated using the Hydro |
2608 |
> |
program that is included in the {\sc oopse} distribution. |
2609 |
|
|
2610 |
+ |
\begin{longtable}[c]{ABG} |
2611 |
+ |
\caption{Meta-data Keywords: Required parameters for the Langevin integrator} |
2612 |
+ |
\\ |
2613 |
+ |
{\bf keyword} & {\bf units} & {\bf use} \\ \hline |
2614 |
+ |
\endhead |
2615 |
+ |
\hline |
2616 |
+ |
\endfoot |
2617 |
+ |
{\tt viscosity} & centipoise & Sets the value of viscosity of the implicit |
2618 |
+ |
solvent \\ |
2619 |
+ |
{\tt targetTemp} & K & Sets the target temperature of the system. |
2620 |
+ |
This parameter must be specified to use Langevin dynamics. \\ |
2621 |
+ |
{\tt HydroPropFile} & string & Specifies the name of the resistance |
2622 |
+ |
tensor (usually a {\tt .diff} file) which is precalculated by {\tt |
2623 |
+ |
Hydro}. This keyworkd is not necessary if the simulation contains only |
2624 |
+ |
simple bodies (spheres and ellipsoids). \\ |
2625 |
+ |
{\tt beadSize} & $\mbox{\AA}$ & Sets the diameter of the beads to use |
2626 |
+ |
when the {\tt RoughShell} model is used to approximate the resistance |
2627 |
+ |
tensor. |
2628 |
+ |
\label{table:ldParameters} |
2629 |
+ |
\end{longtable} |
2630 |
+ |
|
2631 |
+ |
\section{\label{sec:constraints}Constraint Methods} |
2632 |
+ |
|
2633 |
+ |
\subsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond |
2634 |
+ |
Constraints} |
2635 |
+ |
|
2636 |
+ |
In order to satisfy the constraints of fixed bond lengths within {\sc |
2637 |
+ |
oopse}, we have implemented the {\sc rattle} algorithm of |
2638 |
+ |
Andersen.\cite{andersen83} {\sc rattle} is a velocity-Verlet |
2639 |
+ |
formulation of the {\sc shake} method\cite{ryckaert77} for iteratively |
2640 |
+ |
solving the Lagrange multipliers which maintain the holonomic |
2641 |
+ |
constraints. Both methods are covered in depth in the |
2642 |
+ |
literature,\cite{leach01:mm,Allen87} and a detailed description of |
2643 |
+ |
this method would be redundant. |
2644 |
+ |
|
2645 |
+ |
\subsection{\label{oopseSec:zcons}The Z-Constraint Method} |
2646 |
+ |
|
2647 |
+ |
A force auto-correlation method based on the fluctuation-dissipation |
2648 |
+ |
theorem was developed by Roux and Karplus to investigate the dynamics |
2649 |
+ |
of ions inside ion channels.\cite{Roux91} The time-dependent friction |
2650 |
+ |
coefficient can be calculated from the deviation of the instantaneous |
2651 |
+ |
force from its mean value: |
2652 |
+ |
\begin{equation} |
2653 |
+ |
\xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T, |
2654 |
+ |
\end{equation} |
2655 |
+ |
where% |
2656 |
+ |
\begin{equation} |
2657 |
+ |
\delta F(z,t)=F(z,t)-\langle F(z,t)\rangle. |
2658 |
+ |
\end{equation} |
2659 |
+ |
|
2660 |
+ |
If the time-dependent friction decays rapidly, the static friction |
2661 |
+ |
coefficient can be approximated by |
2662 |
+ |
\begin{equation} |
2663 |
+ |
\xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt. |
2664 |
+ |
\end{equation} |
2665 |
+ |
|
2666 |
+ |
This allows the diffusion constant to then be calculated through the |
2667 |
+ |
Einstein relation:\cite{Marrink94} |
2668 |
+ |
\begin{equation} |
2669 |
+ |
D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty |
2670 |
+ |
}\langle\delta F(z,t)\delta F(z,0)\rangle dt}.% |
2671 |
+ |
\end{equation} |
2672 |
+ |
|
2673 |
+ |
The Z-Constraint method, which fixes the $z$ coordinates of a few |
2674 |
+ |
``tagged'' molecules with respect to the center of the mass of the |
2675 |
+ |
system is a technique that was proposed to obtain the forces required |
2676 |
+ |
for the force auto-correlation calculation.\cite{Marrink94} However, |
2677 |
+ |
simply resetting the coordinate will move the center of the mass of |
2678 |
+ |
the whole system. To avoid this problem, we have developed a new |
2679 |
+ |
method that is utilized in {\sc oopse}. Instead of resetting the |
2680 |
+ |
coordinates, we reset the forces of $z$-constrained molecules and |
2681 |
+ |
subtract the total constraint forces from the rest of the system after |
2682 |
+ |
the force calculation at each time step. |
2683 |
+ |
|
2684 |
+ |
After the force calculation, the total force on molecule $\alpha$ is: |
2685 |
+ |
\begin{equation} |
2686 |
+ |
G_{\alpha} = \sum_i F_{\alpha i}, |
2687 |
+ |
\label{oopseEq:zc1} |
2688 |
+ |
\end{equation} |
2689 |
+ |
where $F_{\alpha i}$ is the force in the $z$ direction on atom $i$ in |
2690 |
+ |
$z$-constrained molecule $\alpha$. The forces on the atoms in the |
2691 |
+ |
$z$-constrained molecule are then adjusted to remove the total force |
2692 |
+ |
on molecule $\alpha$: |
2693 |
+ |
\begin{equation} |
2694 |
+ |
F_{\alpha i} = F_{\alpha i} - |
2695 |
+ |
\frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}. |
2696 |
+ |
\end{equation} |
2697 |
+ |
Here, $m_{\alpha i}$ is the mass of atom $i$ in the $z$-constrained |
2698 |
+ |
molecule. After the forces have been adjusted, the velocities must |
2699 |
+ |
also be modified to subtract out molecule $\alpha$'s center-of-mass |
2700 |
+ |
velocity in the $z$ direction. |
2701 |
+ |
\begin{equation} |
2702 |
+ |
v_{\alpha i} = v_{\alpha i} - |
2703 |
+ |
\frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}, |
2704 |
+ |
\end{equation} |
2705 |
+ |
where $v_{\alpha i}$ is the velocity of atom $i$ in the $z$ direction. |
2706 |
+ |
Lastly, all of the accumulated constraint forces must be subtracted |
2707 |
+ |
from the rest of the unconstrained system to keep the system center of |
2708 |
+ |
mass of the entire system from drifting. |
2709 |
+ |
\begin{equation} |
2710 |
+ |
F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha} G_{\alpha}} |
2711 |
+ |
{\sum_{\beta}\sum_i m_{\beta i}}, |
2712 |
+ |
\end{equation} |
2713 |
+ |
where $\beta$ denotes all {\it unconstrained} molecules in the |
2714 |
+ |
system. Similarly, the velocities of the unconstrained molecules must |
2715 |
+ |
also be scaled: |
2716 |
+ |
\begin{equation} |
2717 |
+ |
v_{\beta i} = v_{\beta i} + \sum_{\alpha} \frac{\sum_i m_{\alpha i} |
2718 |
+ |
v_{\alpha i}}{\sum_i m_{\alpha i}}. |
2719 |
+ |
\end{equation} |
2720 |
+ |
|
2721 |
+ |
This method will pin down the centers-of-mass of all of the |
2722 |
+ |
$z$-constrained molecules, and will also keep the entire system fixed |
2723 |
+ |
at the original system center-of-mass location. |
2724 |
+ |
|
2725 |
+ |
At the very beginning of the simulation, the molecules may not be at |
2726 |
+ |
their desired positions. To steer a $z$-constrained molecule to its |
2727 |
+ |
specified position, a simple harmonic potential is used: |
2728 |
+ |
\begin{equation} |
2729 |
+ |
U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},% |
2730 |
+ |
\end{equation} |
2731 |
+ |
where $k_{\text{Harmonic}}$ is an harmonic force constant, $z(t)$ is |
2732 |
+ |
the current $z$ coordinate of the center of mass of the constrained |
2733 |
+ |
molecule, and $z_{\text{cons}}$ is the desired constraint |
2734 |
+ |
position. The harmonic force operating on the $z$-constrained molecule |
2735 |
+ |
at time $t$ can be calculated by |
2736 |
+ |
\begin{equation} |
2737 |
+ |
F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}= |
2738 |
+ |
-k_{\text{Harmonic}}(z(t)-z_{\text{cons}}). |
2739 |
+ |
\end{equation} |
2740 |
+ |
|
2741 |
+ |
The user may also specify the use of a constant velocity method |
2742 |
+ |
(steered molecular dynamics) to move the molecules to their desired |
2743 |
+ |
initial positions. Based on concepts from atomic force microscopy, |
2744 |
+ |
{\sc smd} has been used to study many processes which occur via rare |
2745 |
+ |
events on the time scale of a few hundreds of picoseconds. For |
2746 |
+ |
example,{\sc smd} has been used to observe the dissociation of |
2747 |
+ |
Streptavidin-biotin Complex.\cite{smd} |
2748 |
+ |
|
2749 |
+ |
To use of the $z$-constraint method in an {\sc oopse} simulation, the |
2750 |
+ |
molecules must be specified using the {\tt nZconstraints} keyword in |
2751 |
+ |
the meta-data file. The other parameters for modifying the behavior |
2752 |
+ |
of the $z$-constraint method are listed in table~\ref{table:zconParams}. |
2753 |
+ |
|
2754 |
|
\begin{longtable}[c]{ABCD} |
2755 |
< |
\caption{Meta-data Keywords: Langevin Dynamics Parameters} |
2755 |
> |
\caption{Meta-data Keywords: Z-Constraint Parameters} |
2756 |
|
\\ |
2757 |
|
{\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline |
2758 |
|
\endhead |
2759 |
|
\hline |
2760 |
|
\endfoot |
2761 |
< |
{\tt viscosity} & centipoise & Sets the value of viscosity of implicit |
2762 |
< |
solvents & \\ {\tt HydroPropFile} & string & specifies the resistance |
2763 |
< |
tensor file & usually a {\tt .diff} file which is precalculated by |
2764 |
< |
{\sc Hydro}. Not necessory for simple shaped particles (spheres and |
2765 |
< |
ellipsoids) \\ |
2766 |
< |
{\tt beadSize} & $\mbox{\AA}$ & Sets the diameters of |
2767 |
< |
beads when {\sc Rough Shell Model} is used to generate the resistance |
2768 |
< |
tensor file. \\ |
2769 |
< |
\label{table:ldParameters} |
2761 |
> |
{\tt zconsTime} & fs & Sets the frequency at which the {\tt .fz} file |
2762 |
> |
is written & \\ |
2763 |
> |
{\tt zconsForcePolicy} & string & The strategy for subtracting |
2764 |
> |
the $z$-constraint force from the {\it unconstrained} molecules & Possible |
2765 |
> |
strategies are {\tt BYMASS} and {\tt BYNUMBER}. The default |
2766 |
> |
strategy is {\tt BYMASS}\\ |
2767 |
> |
{\tt zconsGap} & $\mbox{\AA}$ & Sets the distance between two adjacent |
2768 |
> |
constraint positions&Used mainly to move molecules through a |
2769 |
> |
simulation to estimate potentials of mean force. \\ |
2770 |
> |
{\tt zconsFixtime} & fs & Sets the length of time the $z$-constraint |
2771 |
> |
molecule is held fixed & {\tt zconsFixtime} must be set if {\tt |
2772 |
> |
zconsGap} is set\\ |
2773 |
> |
{\tt zconsUsingSMD} & logical & Flag for using Steered Molecular |
2774 |
> |
Dynamics to move the molecules to the correct constrained positions & |
2775 |
> |
Harmonic Forces are used by default |
2776 |
> |
\label{table:zconParams} |
2777 |
|
\end{longtable} |
2778 |
|
|
2779 |
|
\chapter{\label{oopseSec:thermInt}Thermodynamic Integration} |
3473 |
|
\end{longtable} |
3474 |
|
|
3475 |
|
\section{\label{oopseSec:Hydro}Hydro} |
3476 |
< |
{\tt Hydro} generates {\tt .diff} file which is required when a Langevin |
3477 |
< |
Dynamics simulation using approximate models (supports Bead Model and |
3478 |
< |
Rough Shell Model) is performed. To generate the {\tt }.diff file, the |
3479 |
< |
meta-data file is needed as the input file. The viscosity of the fluid |
3480 |
< |
flow (solvent) and the temperature of the system have to be defined in |
3481 |
< |
meta-data file. If the approximate model is {\tt Rough Shell Model}, |
3482 |
< |
the {\tt beadSize} which is the diameter of every beads must be |
3483 |
< |
specified in meta-data file. |
3476 |
> |
{\tt Hydro} generates resistance tensor ({\tt .diff}) files which are |
3477 |
> |
required when using the Langevin integrator using complex rigid |
3478 |
> |
bodies. {\tt Hydro} supports two approximate models: the {\tt |
3479 |
> |
BeadModel} and {\tt RoughShell}. Additionally, {\tt Hydro} can |
3480 |
> |
generate resistance tensor files using analytic solutions for simple |
3481 |
> |
shapes. To generate a {\tt }.diff file, a meta-data file is needed as |
3482 |
> |
the input file. Since the resistance tensor depends on these |
3483 |
> |
quantities, the {\tt viscosity} of the solvent and the temperature |
3484 |
> |
({\tt targetTemp}) of the system must be defined in meta-data file. If |
3485 |
> |
the approximate model in use is the {\tt RoughShell} model the {\tt |
3486 |
> |
beadSize} (the diameter of the small beads used to approximate the |
3487 |
> |
surface of the body) must also be specified. |
3488 |
|
|
3489 |
|
The options available for Hydro are as follows: |
3490 |
|
\begin{longtable}[c]{|EFG|} |
3498 |
|
-V& {\tt -{}-version} & Print version and exit\\ |
3499 |
|
-i& {\tt -{}-input=filename} & input MetaData (md) file\\ |
3500 |
|
-o& {\tt -{}-output=STRING} & Output file name\\ |
3501 |
< |
& {\tt -{}-model=STRING} & hydrodynamics model (supports |
3502 |
< |
RoughShell and BeadModel)\\ |
3501 |
> |
& {\tt -{}-model=STRING} & hydrodynamics model (supports both |
3502 |
> |
{\tt RoughShell} and {\tt BeadModel})\\ |
3503 |
|
-b& {\tt -{}-beads} & generate the beads only, |
3504 |
< |
hydrodynamics will be performed (default=off)\\ |
3504 |
> |
hydrodynamic calculations will not be performed (default=off)\\ |
3505 |
|
\end{longtable} |
3506 |
|
|
3507 |
|
|