Forcefield Based Simulations



5       Molecular Dynamics

While minimization computes the forces on the atoms and changes their positions to minimize the interaction energies, dynamics computes forces and moves atoms in response to the forces.

Molecular dynamics solves the classical equations of motion for a system of N atoms interacting according to a potential energy forcefield as described in Forcefields. Dynamics simulations are useful in studies of the time evolution of a variety of systems at nonzero temperatures, for example, biological molecules, polymers, or catalytic materials, in a variety of states, for example, crystals, aqueous solutions, or in the gas phase.

This chapter explains

To perform the most reasonable and realistic dynamics simulations, you should read this entire chapter, which includes information on:

Integration algorithms

The choice of timestep

Integration errors

Statistical ensembles

Temperature

Pressure and stress

Types of dynamics simulations

Constraints during dynamics simulations

Dynamics trajectories

General methodology for dynamics calculations

Related information

Forcefields and Preparing the Energy Expression and the Model focus on the representation of the potential energy surface and the useful ways that it can be biased through the addition of restraints and constraints, as well as other information on preparing the model system for calculations.

Use of simulation engines for calculating normal modes is found under Vibrational calculation.

Specific information

For specific information on setting up and running dynamics calculations with the various MSI simulation engines, please see the relevant documentation (see Available documentation).

Table 16 . Finding information in Chapter 5

If you want to know about: Read:
Repeating a dynamics simulation.  
Defined initial coordinates and random initial velocities.  
Integrators used by MSI simulation engines.  
Table 17.  
Thermodynamic ensembles handled by MSI simulation engines.  
Table 19.  
Obtaining thermodynamic properties of a model.  
Equilibrium thermodynamic properties.  
Temperature-control methods used by MSI simulation engines.  
Table 20.  
Producing true canonical ensembles.  
Nosé and Nosé-Hoover dynamics.  
Pressure- and stressecontrol methods used by MSI simulation engines.  
Table 21.  
Types of dynamics simulations readily set up with MSI simulation engines.  
Table 22.  
Dynamics runs with periodic minimization.  
Quenched dynamics.  
Controlled temperature change during dynamics.  
Simulated annealing.  
Finding a common configuration of related models.  
Consensus dynamics.  
Impulse dynamics.  
Impulse dynamics.  
Simulated dynamics in a viscous fluid.  
Langevin dynamics.  
Dynamics of localized parts of a model.  
Stochastic boundary dynamics.  
SHAKE and RATTLE algorithms.  
Constraints during dynamics simulations.  
Equilibration and data-collection stages of a dynamics simulation.  
Stages and duration of dynamics simulations.  
How long is long enough for a dynamics simulation.  
Has equilibrium been achieved?; How long should the simulation be?.  
Setting up a dynamics simulation.  
Dynamics with MSI simulation engines.  
Continuing or restarting a dynamics run.  
Restarting a dynamics simulation.  
Using a previous dynamics run to start a new simulation.  
Restarting a dynamics simulation.  

Some uses of dynamics calculations

The major applications of molecular dynamics are:

During dynamics simulations, a system undergoes conformational and momentum changes so that different parts of the phase space accessible to the model can be explored. The conformational search capability of dynamics is one of its most important uses.

By providing several mechanisms for controlling the temperature and pressure of simulated systems, molecular dynamics allows you to generate statistical ensembles from which various energetic, thermodynamic, structural, and dynamic properties can be calculated. For such studies, it is important that the calculation visit various conformational states with the correct statistical frequency.

Although modern crystallography has provided a window into the static structure of molecules both small and large, the thought of intermolecular collisions and conformational variation is always present. After all, binding of substrates by proteins, folding of proteins and peptides into unique shapes, the dynamic behavior of polymers, and chemical reactions themselves would be inconceivable without the concept of molecular motion.

Studies of model motions can be used to derive properties such as diffusion coefficients.

Non-dynamics approaches are also used

Other approaches to simulating molecular motion and generating conformational searches exist.

For example, a dynamics trajectory can be constructed from a set of normal modes to represent the vibrations of a model. While this is a fast method, it is restricted to harmonic motion about a single energy minimum.

An approach to doing conformational searches is the Monte Carlo method. While this method can sample conformational space so as to produce meaningful statistical ensembles, it does not provide dynamic information about the model, since particles of the model system are simply moved randomly according to some statistical rules.


Integration algorithms

This section includes:

Introduction

Criteria of good integrators in molecular dynamics

Integrators in MSI simulation engines

Introduction

Newton's equation of motion applied to atoms

At its simplest, molecular dynamics solves Newton's familiar equation of motion:

Eq. 83        

where Fi is the force, mi is the mass, and ai is the acceleration of atom i.

The force on atom i can be computed directly from the derivative of the potential energy V with respect to the coordinates ri:

Eq. 84        

What is a trajectory?

Notice that classical equations of motion are deterministic. That is, once the initial coordinates and velocities are known, the coordinates and velocities at a later time can be determined. The coordinates and velocities for a complete dynamics run are called the trajectory. (However, trajectories are sensitive to initial conditions, so the same simulation run with a different simulation engine or on a different computer does not produce an identical trajectory.)

The finite-difference method

A standard method of solving an ordinary differential equation such as Eq. 84 numerically is the finite-difference method. The general idea is as follows. Given the initial coordinates and velocities and other dynamic information at time t, the positions and velocities at time t + t are calculated. The timestep t depends on the integration method as well as the system itself.

Defined initial coordinates and random initial velocities

Although the initial coordinates are determined in the input file or from a previous operation such as minimization, the initial velocities are randomly generated at the beginning of a dynamics run, according to the desired temperature. Therefore, dynamics runs cannot be repeated exactly, except for forcefield engines (CHARM standalone, Discover) that allow you to set the random number seed to the value that was used in a previous run.

More details on the initial velocities are provided under Temperature.

Criteria of good integrators in molecular dynamics

Molecular dynamics is usually applied to a large model. Energy evaluation is time consuming and the memory requirement is large. To generate the correct statistical ensembles, energy conservation is also important.

Thus, the basic criteria for a good integrator for molecular simulations are as follows:

Integrators in MSI simulation engines

Integrators provided in MSI simulation engines were chosen according to the above criteria. Only dynamics algorithms used in MSI's simulation engines (Table 17) are considered here:

Verlet leapfrog integrator

Verlet velocity integrator

ABM4 integrator

Runge-Kutta-4 integrator

Table 17 . Dynamics integrators used by MSI simulation engines

integrator simulation engine
CHARMm Discover OFF
Verlet leapfrog  
 
1  
 
Verlet velocity  
 
2  
 
ABM4  
 
2  
 
Runge-Kutta-4  
 
2  
 
1 FDiscover only, not in CDiscover.

2 CDiscover only, not in FDiscover.

Choosing the dynamics algorithm(s)

To specify the dynamics integrator:

Alternatively (in Insight·Discover_3), select the Language_Control/Command_Comment command. Set the Comment Type to Command. Enter integration_method = ABM4 or integration_method = Runge_Kutta in the Command/Comment entry box and select Execute. Be sure that you insert this stage at the correct point in your command input file.

Verlet leapfrog integrator

Advantages of Verlet methods

Variants of the Verlet (1967) algorithm of integrating the equations of motion (Eq. 84) are perhaps the most widely used method in molecular dynamics. The advantages of Verlet integrators is that these methods require only one energy evaluation per step, require only modest memory, and also allow a relatively large timestep to be used.

The leapfrog algorithm

The Verlet leapfrog algorithm is as follows:

Given r(t), v(t -t/2), and a(t), which are (respectively) the position, velocity, and acceleration at times t, t -t/2, and t, compute:

where f (t + t) is evaluated from -dV/dr at r (t + t).

Disadvantage of Verlet leapfrog method

The Verlet leapfrog method has one major disadvantage: the positions and velocities calculated are half a timestep out of synchrony.

Verlet velocity integrator

The Verlet velocity algorithm overcomes the out-of-synchrony shortcoming of the Verlet leapfrog method. The Verlet velocity algorithm is as follows:

The velocity algorithm

Given r(t), v(t), and a(t), which are (respectively) the position, velocity, and acceleration at time t, compute:

ABM4 integrator

ABM4, which stands for Adams-Bashforth-Moulton fourth order, is a predictor and corrector method. It is a fourth-order method, meaning that the truncation error is to the fifth order of the timestep used.

This method requires two energy evaluations per step and has to make use of the results of the previous three steps. It is thus not self starting--the first three steps are generated by the Runge-Kutta method. More memory has to be used, because previous information has to be stored.

ABM4's algorithm

The algorithm is as follows:

Let:

where subscripts (not shown above) 0, 1, -1, -2, and -3 indicate y at times t, t + t, t - t, t - 2t, and t - 3t.

Predictor step

The predictor is:

Now evaluate y1´, using y1predicted, which involves one energy evaluation.

Corrector step

The corrector is:

Now evaluate y1´, using y1corrected, which involves another energy evaluation.

Runge-Kutta-4 integrator

Robust, but disadvantages

Runge-Kutta-4 stands for the fourth-order Runge-Kutta method, which is one of the oldest numerical methods for solving ordinary differential equations. The method is self starting but requires four energy evaluations per step.

From testing done at MSI, the timestep has to be very small. This is thus not a very suitable integrator for molecular simulation.

However, the method is very robust, meaning that it can deal with almost all kinds of equations, including stiff ones. This integrator is used to generate the trajectory for the first three steps for ABM4.

Since we do not recommend using this integrator, the algorithm is not presented here. Details can be found in Press et al. 1986.


The choice of timestep

A key parameter in the integration algorithms is the integration timestep t. To make the best use of the computer time, a large timestep should be used. However, too large a timestep causes instability and inaccuracy in the integration process.

Relation of timestep to molecular vibration

The timestep used depends on the model as well as the integrators. The main limitation imposed by the model is the highest-frequency motion that must be considered. A vibrational period must be split into at least 8-10 segments for models to satisfy the Verlet assumption that the velocities and accelerations are constant over the timestep used.

In most organic models, the highest vibrational frequency is that of C-H bond stretching, whose period is on the order of 10-14 s (10 fs). The integration timestep should therefore be about 0.5-1 fs. If you use the SHAKE or RATTLE constraint algorithm (Constraints during dynamics simulations), a longer timestep is possible.

If you are studying simple model liquids or solids and are not interested in internal modes, much longer timesteps may be used, e.g., up to 20 fs. A timestep of about 5 fs should be adequate for ionic material models.

Appropriate for the integrator

The timestep must also be appropriate to the integrator. For the ABM4 method, the timestep should be about half that needed for the Verlet algorithm. The Runge-Kutta-4 method seems to require a much smaller timestep than the other methods.

Setting the timestep

To specify the length of the timestep:


Integration errors

If the chosen timestep is too small, no harm is done, except for the waste of computer time. However, if the timestep is too large for the calculation conditions, the simulation can "blow up".

Two examples illustrate how the timestep, temperature, and integration algorithm affect the results. The first example is a simulation of the collision of two hydrogen atoms travelling towards each other. The second is an examination of energy conservation in a simple harmonic oscillator when different integrators and timesteps are used.

Example 1--Two colliding hydrogen atoms

The stability of the numerical integration with respect to the time step can be tested directly by integrating (over distance) the forces used by dynamics and comparing the integral with the analytical energy. The error in this integral as a function of time step is an indication of the intrinsic limitations of molecular dynamics.

Timestep slightly too large

To illustrate this, consider the collision of two atoms. Figure 26 plots the true potential energy of the van der Waals potential between two hydrogens along with the energy integrated numerically from the forces. The time step used is 1 fs at a temperature of 300 K. The temperature is set by assigning an initial velocity of 1500 m s-1 (0.015Å fs-1) to one of the hydrogens along the vector connecting them. This velocity is the most probable velocity of a hydrogen atom at 300 K.

Figure 26 . Numerical integration of energy from molecular dynamics hydrogen-collision trajectory, 1 fs timestep

The integrated energy calculated numerically from a dynamics trajectory of two colliding hydrogen atoms (circles) is compared with the analytical energy curve (thick line). Simulation done with FDiscover.  

As the two atoms approach each other, the integration agrees well with the analytic curve. However, after the atoms collide, the integrated energy is significantly higher than the true energy. This behavior is due to the atom moving too quickly through a rapidly changing energy function. When the atoms are far apart, the energy change is smallest and therefore, the forces are smallest and most linear. As the atoms approach, they speed up and take larger and larger steps, until they reach their highest velocity at the energy minimum. Unfortunately, this is precisely where the forces start changing the fastest. Thus, when the atoms should be taking large steps (far apart), they are taking the smallest steps, and when they should take small steps (near the minimum), they take the largest steps. The consequence is that the particles "step through" the energy barrier momentarily. Of course, once a new force is calculated at the extrapolated coordinate, the trajectory is rapidly corrected. However, it is too late for the energy integral--some energy has been gained.

In this example, the total energy rose by about 0.02 kcal mol-1. Whether this is a reasonable error depends on how closely the exact motions need to be reproduced.

Shorter timesteps mean more computational cost

Figure 27 shows the same curve as Figure 26, but with timesteps of 0.33 and 0.10 fs. Both give better results than 1.0 fs, but it is not clear whether the extra time required to calculate the smaller timesteps is worthwhile.

Figure 27 . Numerical integration of energy from molecular dynamics hydrogen-collision trajectory, 0.33 and 0.1 fs timesteps

Energy integration errors decrease with smaller time steps. Compared to 0.016 kcal error with 1 fs time steps, the 0.33 fs time step has a 0.006 error, and the 0.1 fs time step a 0.001 error. The cost for increased accuracy is the computational burden to compute more steps. For most simulations, a 1-fs time step is a good compromise between numerical accuracy and computational efficiency. Simulation done with FDiscover.  

A very large timestep leads to artifactual behavior

The consequences of too large a time step are much more dramatic. Doubling the time step from 1 to 2 fs results in an unusual artifact for the hydrogen system (Figure 28). In this case, the integration error affects not only the potential energy, but the kinetic energy as well. Momentum is removed, so that the hydrogens no longer have the velocity needed to escape after the collision. The atoms are trapped forever (barring an inverse error that could impart momentum). The atoms now vibrate back and forth (only 2 cycles are plotted in Figure 28), and each cycle incurs an additional integration error.

Figure 28 . Numerical integration of energy from molecular dynamics hydrogen-collision trajectory, 2 fs timestep

A 2-fs time step causes the rebounding force to be underestimated, robbing the colliding Hs of sufficient escape velocity and resulting in the two Hs being "bound" by their van der Waals forces. Simulation done with FDiscover.  

An even higher timestep can cause the system to explode

Increasing the time step another factor of 2 to 4 fs (Figure 29) finally causes the system to "blow up". The timestep is so long that the atoms deeply interpenetrate each other's steeply repulsive wall between steps. The resulting force is now so large that the atoms fly off at a speed of about 1 Å fs-1 (or 105 m s-1). An equivalent temperature would be hundreds of thousands of degrees.

Figure 29 . Numerical integration of energy from molecular dynamics hydrogen-collision trajectory, 4 fs timestep

With a time step of 4 fs, the Hs travel too far in a single step, interpenetrating each other's van der Waals radii before the forces are recalculated. By this time, the forces are so large that the Hs are flung apart at a temperature equivalent to hundreds of thousands of degrees. Simulation done with FDiscover.  

Effect of temperature

To complete the analysis of integration errors, it is instructive to compare the effects of increasing kinetic energy on the stability of the numerical integration. Figure 30 shows that there is essentially no difference in the error between 300 and 1200 K. Going as high as 30,000 K merely almost doubles the error, indicating that dynamics simulations are not as sensitive to the temperature as to the timestep.

Figure 30 . Integration errors for hydrogen-collision trajectories at several temperatures

Integration errors observed for an H-H collision for an initial velocity equal to the mean velocity appropriate for 300 K (squares), 1200 K (circles) and 30,000 K (triangles). Simulation done with FDiscover.  

Example 2--Energy conservation of a harmonic oscillator

Verlet velocity vs. ABM4

To test the conservation of energy in a simulation, 10 ps of molecular dynamics was performed on a harmonic oscillator having an equilibrium length of 0.75 Å and period of 7.5 fs. As can be seen in Table 18, the Verlet velocity method can use a larger timestep than the ABM4 method. Although the Verlet algorithm starts to show instability at 1 fs, ABM4 starts to fail at 0.25 fs.

For the Verlet velocity integrator, the standard deviation in the total energy is proportional to t2, as predicted by the theory. This is a simple verification that the integrator has been implemented correctly.

Table 18 . Energy conservation for different timesteps with the Verlet velocity and ABM4 integrators

Run conditions: 10,000 fs, constant-energy (NVE), harmonic oscillator, initial energy 0.296 kcal mol-1, CDiscover 94.0.

integrator timestep fs final energy kcal mol-1 average energy kcal mol-1 standard deviation kcal mol-1
Verlet velocity  
1.0  
0.327  
0.325  
0.021  
Verlet velocity  
0.5  
0.296  
0.302  
0.005  
Verlet velocity  
0.25  
0.298  
0.298  
0.001  
ABM4  
0.25  
0.693  
0.467  
0.114  
ABM4  
0.10  
0.299  
0.297  
0.001  


Statistical ensembles

You can control the temperature and pressure

Integrating Newton's equations of motion allows you to explore the constant-energy surface of a system. However, most natural phenomena occur under conditions where a system is exposed to external pressure and/or exchanges heat with the environment. Under these conditions, the total energy of the system is no longer conserved, and extended forms of molecular dynamics are required.

Purpose of the calculation

Several methods are available for controlling temperature and pressure. Depending on which state variables (the energy E, enthalpy H (i.e., E + PV), number of particles N, pressure P, stress S, temperature T, and volume V) are kept fixed, different statistical ensembles can be generated. A variety of structural, energetic, and dynamic properties can then be calculated from the averages or the fluctuations of these quantities over the ensemble generated.

Available thermodynamic ensembles

Both isothermal (exchange heat with a temperature bath to maintain a constant thermodynamic [not kinetic] temperature) and adiabatic (do not exchange heat) ensembles are available:

Table 19 . Thermodynamic ensembles handled by MSI simulation engines

ensemble1 simulation engine
CHARMm Discover OFF
Constant temperature, constant volume (NVT)  
 
 
 
Constant temperature, constant pressure (NPT)2  
3  
 
 
Constant temperature, constant stress (NST)2  
 
4  
 
Constant energy, constant volume (NVE)  
 
 
 
Constant pressure, constant enthalpy (NPH)2  
 
4  
 
Constant stress, constant enthalpy (NSH)2  
 
4  
 
1 In all ensembles, the number of particles is conserved.

2 Only for periodic systems, because volume is undefined in nonperiodic systems. For all space-group symmetries unless otherwise noted.

3 Only for cubic, orthorhombic, and triclinic unit cells.

4 CDiscover only, not in FDiscover.

Choosing the thermodynamic ensemble

To access the controls used for specifying the thermodynamic ensemble:

NVE ensemble

Some energy drift

The constant-energy, constant-volume ensemble (NVE), also known as the microcanonical ensemble, is obtained by solving the standard Newton equation without any temperature and pressure control. Energy is conserved when this (adiabatic) ensemble is generated. However, because of rounding and truncation errors during the integration process, there is always a slight fluctuation or drift in energy.

When the Verlet leapfrog integrator is used, only r (t) and v (t - 1 \xda 2 t) are known at each timestep. Thus, the potential and kinetic energies at each timestep are also half a step out of synchrony. Although the difference between the kinetic energies half a timestep apart is small, this can also contribute to the fluctuation in the total energy.

Although the temperature is not controlled during true NVE dynamics, you might want to use NVE conditions during the equilibration phase (Stages and duration of dynamics simulations) of your simulation. For this purpose, Cerius2·Discover and Cerius2·Dynamics Simulation allow you to hold the temperature within specified tolerances by periodic scaling of the velocities.

When to use it

True constant-energy conditions (i.e., without temperature contol) are not recommended for equilibration because, without the energy flow facilitated by temperature control, the desired temperature cannot be achieved.

However, during the data collection phase, if you are interested in exploring the constant-energy surface of the conformational space, or for other reasons do not want the perturbation introduced by temperature- and pressure-bath coupling, this is a useful ensemble.

Results

The results can be used (Equilibrium thermodynamic properties) to calculate the thermodynamic response function (Ray 1988).

NVT ensemble

The constant-temperature, constant-volume ensemble (NVT), also referred to as the canonical ensemble, is obtained by controlling the thermodynamic temperature. Direct temperature scaling should be used only during the initialization stage (Stages and duration of dynamics simulations), since it does not produce a true canonical ensemble (it is not truly isothermal). Any of the other temperature-control methods available (How temperature is controlled) is used during the data collection phase.

When to use it

This is the appropriate choice when conformational searches of models are carried out in vacuum without periodic boundary conditions. (Without periodic boundary conditions, volume, pressure, and density are not defined and constant-pressure dynamics cannot be carried out.)

Even when periodic boundary conditions are used, if pressure is not a significant factor, the constant-temperature, constant-volume ensemble provides the advantage of less perturbation of the trajectory, due to the absence of coupling to a pressure bath.

NPT and NST ensembles

Periodic systems

The constant-temperature, constant-pressure ensemble (NPT) allows control over both the temperature and pressure. The unit cell vectors are allowed to change, and the pressure is adjusted by adjusting the volume (i.e., the size and also, in some programs, the shape of the unit cell). This method applies only to periodic systems.

The constant-temperature, constant-stress ensemble (NST) is an extension of the constant-pressure ensemble. In addition to the hydrostatic pressure which is applied isotropically, the constant-stress ensemble allows you to control the xx, yy, zz, xy, yz, and zx components of the stress tensor.

Control of run conditions

Pressure can be controlled by the Berendsen, Andersen, or Parrinello-Rahman method (How pressure and stress are controlled). However, only the size, and not the shape, of the unit cell can be changed with the Berendsen and Anderson methods (Berendsen method of pressure control and Andersen method of pressure control).

Stress can be controlled by the Parrinello-Rahman method (Parrinello-Rahman method of pressure and stress control), since it allows both the cell volume and its shape to change.

Temperature can be controlled by any method available (How temperature is controlled) (except, of course, the temperature scaling method, since it is not truly isothermal).

When to use it

NPT is the ensemble of choice when the correct pressure, volume, and densities are important in the simulation. This ensemble can also be used during equilibration to achieve the desired temperature and pressure before changing to the constant-volume or constant-energy ensemble when data collection starts.

Results

The NST ensemble is particularly useful if you want to run a simulation at incremented tensile loads to study the stress-strain relationship in polymeric or metallic materials.

If the forcefield being used yields a high pressure at the experimental volume, it may be more realistic to simulate at the experimental pressure rather than the experimental volume. High simulated pressure is a sign that the system is unduly compressed, which restricts atomic motions, artefactually slowing down the dynamic relaxations.

NPH and NSH ensembles

Periodic systems

The constant-pressure, constant-enthalpy ensemble (NPH, Andersen 1980, see also Andersen method of pressure control) is the analogue of constant-volume, constant-energy ensemble, where the size of the unit cell is allowed to vary. In the constant-pressure (or -stress), constant-enthalpy ensemble (NPH or NSH, Parrinello and Rahman 1981, see also Parrinello-Rahman method of pressure and stress control), both the size and shape of the unit cell are allowed to vary (meaning that external stress can be applied). These methods apply only to 3D periodic systems.

Enthalpy H, which is the sum of E and PV, is constant when the pressure is kept fixed without any temperature control. Although the temperature is not controlled during true (adiabatic) NPH or NSH dynamics, you might want to use these conditions during the equilibration phase (Stages and duration of dynamics simulations) of your simulation. For this purpose, Cerius2·Discover and Cerius2·Dynamics Simulation allow you to hold the temperature within specified tolerances by periodic scaling of the velocities.

Results

The natural response functions (specific heat at constant pressure, thermal expansion, adiabatic compressibility, and adiabatic compliance tensor) are obtained (Equilibrium thermodynamic properties) from the proper statistical fluctuation expressions of kinetic energy, volume, and strain (Ray 1988).

Equilibrium thermodynamic properties

Precautions

Since the ensembles are artificial constructs, they produce averages that are consistent with one another when they represent the same state of the model. Nevertheless, the fluctuations vary in different ensembles. Some of the fluctuations are related to thermodynamic derivatives, such as the specific heat or the isothermal compressibility.

Caution

In practice, obtaining accurate fluctuations to calculate physical quantities is difficult, and this approach should be used with caution.

The transformation and relation between different ensembles has been discussed in greater detail by Allen and Tildesley (1987).

Obtaining equilibrium thermodynamic properties

One of the objectives of molecular dynamics is to obtain the equilibrium thermodynamic properties of a model. If a microscopic dynamic variable A takes on values A(t) along a trajectory, then the following time average:

Eq. 85        

yields the thermodynamic value for the selected variable. This dynamic variable can be any function of the coordinates and momenta of the particles of the model.

Time averaging for first-order properties

Through time averaging, you can calculate the first-order properties of a system (such as the internal energy, kinetic energy, pressure, and virial). Similarly, using microscopic expressions in the form of fluctuations of these first-order properties, you can also calculate thermodynamic properties of a system. These include the specific heat, thermal expansion, and bulk modulus.

In the thermodynamic limit, the first-order properties obtained in one ensemble are equivalent to those obtained in other ensembles (differences are on the order of 1/N).

Ensemble-dependent second-order properties

However, second-order properties such as specific heats, compressibilities, and elastic constants differ between ensembles. For example, the specific heat at constant pressure differs from the specific heat at constant volume.

Therefore, it is important to use the appropriate ensemble when performing simulations to obtain these properties.


Temperature

This section includes:

How temperature is calculated

How temperature is controlled

Relation of temperature and velocity

Temperature is a state variable that specifies the thermodynamic state of the system and is also an important concept in dynamics simulations. This macroscopic quantity is related to the microscopic description of simulations through the kinetic energy, which is calculated from the atomic velocities.

The temperature and the distribution of atomic velocities in a system are related through the Maxwell-Boltzmann equation:

Eq. 86        

This well known formula expresses the probability f (v) that a molecule of mass m has a velocity of v when it is at temperature T. Figure 31 shows this distribution at various temperatures.

Figure 31 . Maxwell-Boltzmann distribution of velocity of water at various temperatures

Distribution of model velocities at equilibrium as predicted by the Maxwell-Boltzmann equation. The simulation program assigns random initial velocities to a system of atoms such that the overall distribution of velocities matches a Maxwell-Boltzmann distribution for the desired temperature.  

The x, y, z components of the velocities, on the other hand, have Gaussian distributions:

Eq. 87        

How initial velocities are generated

The initial velocities are generated from the Gaussian distribution of vx, vy, and vz. The Gaussian distribution is generated from a random number generator and a random number seed.

How temperature is calculated

Temperature is a thermodynamic quantity, which is meaningful only at equilibrium. It is related to the average kinetic energy of the system through the equipartition principle. This principle states that every degree of freedom (either in momenta or in coordinates), which appears as a squared term in the Hamiltonian, has an average energy of kT/2 associated with it. This is true for momenta pi which appear as pi2/2m in the Hamiltonian.

Relation of kinetic energy, degrees of freedom, and temperature

Hence we have:

Eq. 88        

The left side of Eq. 88 is also called the average kinetic energy of the system, Nf is the number of degrees of freedom, and T is the thermodynamic temperature. In an unrestricted system with N atoms, Nf is 3N because each atom has three velocity components (i.e., vx, vy, and vz).

Instantaneous kinetic temperature

It is convenient to define an instantaneous kinetic temperature function:

Eq. 89        

The thermodynamic temperature

The average of the instantaneous temperature Tinstan is the thermodynamic temperature T.

Temperature is calculated from the total kinetic energy and the total number of degrees of freedom. For a nonperiodic system:

Temperature in nonperiodic systems...

Eq. 90        

Six degrees of freedom are subtracted because both the translation and rotation of the center of mass are ignored.

...and in periodic systems

And for a periodic system:

Eq. 91        

Only the three degrees of freedom corresponding to translational motion can be ignored, since rotation of a central cell imposes a torque on its neighboring cells.

How temperature is controlled

Although the initial velocities are generated so as to produce a Maxwell-Boltzmann distribution at the desired temperature, the distribution does not remain constant as the simulation continues. This is especially true when the system does not start at a minimum-energy configuration of the model. This occurs often, since the model is commonly minimized only enough to eliminate any hot spots.

During dynamics, kinetic energy is changed to potential energy as the minimized structure changes to the thermal equilibrium structure, and the temperature also changes.

Need to control temperature

To maintain the correct temperature, the computed velocities have to be adjusted appropriately. In addition to maintaining the desired temperature, the temperature-control mechanism must produce the correct statistical ensemble. This means that the probability of occurrence of a certain configuration obeys the laws of statistical mechanics.

For example, in order for constant-temperature, constant-volume dynamics to generate the canonical ensemble, P(E) (i.e., the probability that a configuration with energy E will occur) must be proportional to exp(-E/kBT), also called the Boltzmann factor.

Methods of controlling temperature

Only temperature-control methods used in MSI's simulation engines (Table 20) are considered here:

Direct velocity scaling

Berendsen method of temperature-bath coupling

Nosé and Nosé-Hoover dynamics

Andersen method

Table 20 Temperature-control methods used by MSI simulation engines

method simulation engine
CHARMm Discover OFF
Velocity scaling1  
 
 
 
Berendsen temperature bath2  
 
 
 
Nosé  
 
 
 
Nosé-Hoover3  
 
4  
 
Andersen  
 
4  
 
1 Temperature scaling is not generally used to control temperature during a simulation, but to quickly change the simulated temperature to the desired value. It does not produce the correct statistical ensemble.

2 Referred to as T_DAMPING in Cerius2·Dynamics Simulation.

3 Referred to as Nosé in the Cerius2·Discover and Insight·Discover_3 modules and Hoover in Cerius2·Dynamics Simulation.

4 CDiscover only, not in FDiscover.

Choosing the temperature-control method(s)

To access the controls used for specifying the temperature-control method:

Direct velocity scaling

Direct velocity scaling is a drastic way to change the velocities of the atoms so that the target temperature can be exactly matched whenever the system temperature is higher or lower than the target by some user-defined amount.

Important

Direct velocity scaling can not be used to generate realistic thermodynamic ensembles, since it suppresses the natural fluctuations of a system.

Implementation

In Discover, the velocities of all atoms are scaled uniformly as follows:

Eq. 92        

In the Cerius2·Dynamics Simulation module, the rescale factor is:

Eq. 93        

where Tinst = the instantaneous temperature, Tr = the required temperature, Tavg = the average kinetic temperature, and the term under the root is positive. Otherwise, rescaling is not done.

CHARMm allows you to assign velocities based on values in a comparison coordinate file or to assign either a uniform or a Gaussian distribution of velocities to the atoms. The Gaussian distribution is recommended.

Achieving equilibrium

Direct temperature scaling adds (or subtracts) energy from the system efficiently, but it is important to recognize that the fundamental limitation to achieving equilibrium is how rapidly energy can be transferred to, from, and among the various internal degrees of freedom of the model. The speed of this process depends on the energy expression, the parameters, and the nature of the coupling between the vibrational, rotational, and translational modes. It also depends directly on the size of the system, larger systems taking longer to equilibrate.

Berendsen method of temperature-bath coupling

After equilibration, a more gentle exchange of thermal energy between the system and a heat bath can be introduced through the Berendsen et al. (1984) method (referred to as temperature damping in Cerius2·Dynamics Simulation), in which each velocity is multiplied by a factor given by:

Eq. 94        

where t is the timestep size, is a characteristic relaxation time, T0 is the target temperature, and T the instantaneous temperature.

To a good approximation, this treatment gives a constant-temperature ensemble that can be controlled, both by adjusting the target temperature T0 and by changing the relaxation time (generally between 0.1 and 0.4 ps).

This is a simple approach that does not use Hamiltonians.

Nosé and Nosé-Hoover dynamics

Produces true canonical ensembles...

Nosé dynamics (Nosé 1984a, 1984b, 1991) is a method for performing constant-temperature dynamics that produces true canonical ensembles both in coordinate space and in momentum space. The Nosé-Hoover formalism (referred to as Nosé in Discover and as Hoover in Cerius2·Dynamics Simulation) is based on a simplified reformulation by Hoover (1985), which eliminates time scaling and therefore yields trajectories (Dynamics trajectories) in real time and with evenly spaced time points. The method is also called the Nosé-Hoover thermostat.

...through use of a fictitious mass

The main idea behind Nosé-Hoover dynamics is that an additional (fictitious) degree of freedom is added to the model, to represent the interaction of the model with the heat bath. This fictitious degree of freedom is given a mass Q. The equations of motion for the extended (i.e., model plus fictitious) system are solved. If the potential chosen for that degree of freedom is correct, the constant-energy dynamics (or the microcanonical dynamics, NVE) of the extended system produces the canonical ensemble (NVT) of the real model.

Hamiltonian and equations of motion for this extended system

The Hamiltonian H* of the extended system is:

Eq. 95        

Equations of motion for the real-atom coordinates q and moments p, as well as for the fictitious coordinates S and momentum (where is the interaction potential) are:

Eq. 96        

Eq. 97        

Eq. 98        

where      Q = the user-defined q_ratio X a constant X g X T;
      g = number of degrees of freedom;
      T = temperature.

Magnitude of the fictitious mass and computational efficiency

The choice of the fictitious mass Q of that additional degree of freedom is arbitrary but is critical to the success of a run. If Q is too small, the frequency of the harmonic motion of the extended degree of frequency is too high. This forces a smaller timestep to be used in integration. However, if Q is too large, the thermalization process is not efficient--as Q approaches infinity, there is no energy exchange between the heat bath and the model.

The choice of Q should therefore be based on a balance between the stability of the solution and the highest-frequency motions of the model.

Suggested value for fictitious mass...

Q should be different for different models--Nosé (1991) suggests that Q should be proportional to gkBT, where g is the number of degrees of freedom in the model, kB is the Boltzmann constant, and T is the temperature.

To determine the proportionality constant, studies were done with a model consisting of a box of liquid argon containing 343 atoms at 87 K. Setting Q to 2.5 10-5 kcal mol-1 fs-2 for this model yielded good results. This proportionality constant, together with the gkBT term, was then used to generate Q for a box of water and an amorphous cell of polypropylene, which also yielded satisfactory results.

Discover ordinarily chooses Q automatically. However, you may choose to multiply this calculated Q by a user-defined Q ratio.

...or for the relaxation time

In the Cerius2·Dynamics Simulation module, the factor that you can control directly is , a relaxation time for the model (2 is directly proportional to Q).

For simple fluid models, can be chosen as the second moment of the velocity autocorrelation function. For covalently bonded models, however, it is better to choose based on the frequencies involved in the model.

A rule of thumb, therefore, is to choose on the same order as the smallest time scale (highest frequency) of your model.

Timestep to use

Tests on polypropylene indicate that Nosé-Hoover dynamics needs a timestep of 0.5 fs with the velocity Verlet method in order to approach within 3% of the target temperature of 298 K. In contrast, 1.0 fs is sufficient for the direct velocity scaling method of temperature control.

For the Nosé-Hoover method, the smaller the timestep, the closer it approaches the target temperature. Apparently, the need for a smaller timestep to achieve accuracy is unrelated to Q, as long as Q is in the appropriate range (because the error in reaching the target temperature remains the same when Q is increased or decreased by a factor of 2).

Accurate integrator

For energy to be conserved, the Nosé-Hoover method also requires an accurate integrator. Thus, the ABM4 integrator, if available, with 0.5 fs as the timestep should be used.

Andersen method

One version of the Andersen method of temperature control involves randomizing the velocities of all atoms at a predefined "collision frequency". (The other version involves choosing one atom at each timestep and changing its velocity according to the Boltzmann distribution.)

The CDiscover program implements the first version. The predefined frequency is proportional to N 2/3, where N is the number of atoms in the system. Although this frequency is calculated by the program, you can change it.


Pressure and stress

This section includes:

Units and sign conventions for pressure and stress

How pressure and stress are calculated

How pressure and stress are controlled

What are pressure and stress?

Pressure is another basic thermodynamic variable that defines the state of the system. It is a familiar concept, defined as the force per unit area. Standard atmospheric pressure is 1.013 bar, where 1 bar = 105 Pa. A single number for the pressure implies that pressure is a scalar quantity, but in fact, pressure is a tensor of the more general form (McQuarrie 1976):

Eq. 99        

Each element of the tensor is the force that acts on the surface of an infinitesimal cubic volume that has edges parallel to the x, y, and z axes. The first subscript refers to the direction of the normal to the plane on which the force acts, and the second subscript refers to the direction of the force.

In an isotropic situation, where forces are the same in all directions and there is no viscous force, the pressure tensor is diagonal. With the same diagonal elements, the tensor can be written as:

Eq. 100        

where the scalar quantity p is the equivalent hydrostatic pressure.

Sometimes, (especially in materials science studies), the the stress tensor, or stress, is used in preference to the pressure tensor (its negative). The diagonal elements are known as the tensile stress, and the nondiagonal elements are the shear stress.

The changes in unit-cell lattice parameters and volume resulting from the stress can be obtained from an analysis of dynamics trajectory output file data (Dynamics trajectories). Multiple dynamics runs can be performed at varying stress or pressure values, and the strains obtained can be used to plot a stress-strain curve.

Units and sign conventions for pressure and stress

Pressure and stress can be expressed in many different units. The most common ones are bar and GPa. In SI units, 1 bar = 105 N m-2 and 1 GPa = 109 N m-2. Hence, 1 GPa = 104 bar. Pressure is usually expressed in bars, but in materials science, stress is often expressed in terms of GPa.

In the CDiscover program, constant-stress dynamics has GPa as the default unit. Since the FDiscover program is geared towards controlling pressure, the unit used is bar. In the Cerius2·Dynamics Simulation module, GPa are the only units used for pressure and stress.

Although pressure and stress are defined with the same physical quantities, they have opposite sign conventions. Positive pressure implies a compressive force pushing the system inward, but positive stress means a force acting outward to expand the system.

How pressure and stress are calculated

Pressure is calculated through the use of the virial theorem (Allen and Tildesley 1987). Like temperature, pressure is a thermodynamic quantity and is, strictly speaking, meaningful only at equilibrium.

Relation of pressure, temperature, volume and internal virial

Thermodynamic pressure, thermodynamic temperature, volume, and internal virial can be related in the following way:

Eq. 101        

And W is defined as:

Eq. 102        

Volume (and pressure) is known only for periodic models

Note that pressure is defined only when the system is placed in a container having a definite volume. In a computer simulation, the unit cell under periodic boundary conditions is viewed as the container. Volume, pressure, and density can be calculated only when the model is recognized as periodic.

Instantaneous pressure

Analogous to the temperature, an instantaneous pressure function P can be defined so that thermodynamic pressure is the average of the instantaneous values:

Eq. 103        

where P is the instantaneous pressure and T is the instantaneous kinetic temperature, which is related to the instantaneous kinetic energy K of the system as:

Eq. 104        

The instantaneous pressure function can be written as:

Eq. 105        

Instantaneous pressure tensor

As mentioned above, pressure is a tensor and its components can also be expressed in tensorial form. Eq. 101 can be recast in the form of:

Eq. 106        

In detail, the two terms on the right-hand side of the equation are:

Eq. 107        

Eq. 108        

Instantaneous hydrostatic pressure

where rix, vix, and f ix indicate the x components of the position, velocity, and force vectors of the ith atom, respectively.

From the definition of the instantaneous pressure tensor, the instantaneous hydrostatic pressure is calculated as 1/3 the trace of the pressure tensor (Pxx + Pyy + Pzz).

Forces on the image atoms

When periodic boundary conditions are used, atoms in the unit cell interact not only with the other atoms in the unit cell but also with their translated images. Forces on the images in the virial W must be included correctly. If the interaction is pairwise, using Newton's third law, W can be written as:

Eq. 109        

instead of:

Eq. 110        

Berendsen et al. (1984) use the rij · fij formalism by evaluating the virial and the kinetic energy tensor based on the centers of mass, which is valid because the internal contribution to the virial is canceled (on the average) by the internal kinetic energy. Because of the way forces are evaluated, rescaling of coordinates is also done on the basis of the centers of mass of the models.

How coordinates are scaled in response to pressure changes

The Discover program, however, calculates pressure on an atomic basis and performs atomic scaling. The major advantage of using atomic scaling of the coordinates is that atom overlapping can be avoided. Such overlap can occur if centers of mass are moved instead of individual atoms. In addition, for large models having internal flexibility, atomic scaling yields a smoother response to pressure changes.

Explicit or implicit images

In the Discover program, when explicit images (also called ghost atoms) are used under periodic boundary conditions, the rij · fij formalism is used, with the explicit images included.

In the FDiscover program when minimum images (implicit images) are used under periodic boundary conditions, the rij · fij formalism is used so that forces between real and image atoms can be properly accounted for.

How pressure and stress are controlled

As with temperature (How temperature is controlled), the pressure (and stress) -control mechanism must produce the correct statistical ensemble. This means that the probability of occurrence of a certain configuration obeys the laws of statistical mechanics.

Methods of controlling pressure

Only pressure-control methods used in MSI's simulation engines (Table 21) are considered here:

Berendsen method of pressure control

Andersen method of pressure control

Parrinello-Rahman method of pressure and stress control

Table 21 . Pressure- and stress-control methods used by MSI simulation engines

method simulation engine
CHARMm Discover OFF
Berendsen pressure "bath"  
 
 
 
Andersen  
 
1  
 
Parrinello-Rahman  
 
1  
 
1 CDiscover only, not in FDiscover.

Important

With the Berendsen and Andersen methods the volume can change, but there is no change in the shape of the cell; thus only the pressure is controlled. With the Parrinello-Rahman method, the cell's shape can change, and therefore both pressure and stress can be controlled.

Choosing the pressure- and/or stress-control method(s)

To access the controls used for specifying the pressure- and stress-control methods:

Alternatively (for Insight·Discover_3), select the Language_Control/Command_Comment command. Set the Comment Type to Command. Enter pressure_control_method = Andersen_cp or pressure_control_method = Berendsen_pc in the Command/Comment entry box and select Execute. Be sure that you insert this stage at the correct point in your command input file.

Berendsen method of pressure control

Pressure changes can be accomplished by changing the coordinates of the particles and the size of the unit cell in periodic boundary conditions.

How it works

The Berendsen method (Berendsen et al. 1984) couples the system to a pressure "bath" to maintain the pressure at a certain target. The strength of coupling is determined both by the compressibility of the system (using a user-defined variable ) and by a relaxation time constant (a user-defined variable ). At each step, the x, y, and z coordinates of each atom are scaled by the factor:

Eq. 111        

where t is the time step, P is the instantaneous pressure, and P0 is the target pressure. The Cartesian components of the unit cell vectors are scaled by the same factor µ.

Change in cell size but not shape

Note that this method (as implemented) changes the cell uniformly, so that the size of the cell is changed, but not its shape. Therefore, for simulations such as crystal phase transitions, where both the cell size and shape are expected to change, this method is not appropriate.

Disadvantages of the method

The pressure fluctuation has been observed to be large during test runs with the Discover program and in studies in the literature (Brown and Clark 1984).

Negative pressures sometimes occur because the virial can be negative, even though this defies the usual sense that pressure is a positive number. The calculated pressure depends on the cutoff distances used in the simulation.

Precautions

To compensate for the missing long-range part of the potential contributions to the energy, pressure at r > rc can be estimated by assuming the radial distribution g (r) ~ 1 (uniform distribution) in that region. With this assumption and the known form of nonbond potentials, the correction can be estimated analytically (Allen and Tildesley 1987).

However, this correction is not built into the calculation. If the cutoff distance is too short, the calculated pressure may be wrong. Therefore in practice, you should test the effect of cutoff on pressure by gradually increasing the cutoff and choosing the appropriate cutoff accordingly.

Andersen method of pressure control

Change in cell size but not shape

With the Andersen method (1980) of pressure control, the volume of the cell can change, but its shape is preserved by allowing the cell to change isotropically.

Uses

The Anderson method is useful for liquid simulations since the box could become quite elongated in the absence of restoring forces if the shape of the cell were allowed to change. A constant shape also makes the dynamics analysis easier.

However, this method is not very useful for studying materials under nonisotropic stress or phase transitions, which involve changes in both cell lengths and cell angles (for these conditions, the Parrinello-Rahman method should be used).

How it works

The basic idea of the method is to treat the volume V of the cell as a dynamic variable in the system. The Lagrangian of the system is modified so that it contains a term in the kinetic energy with a user-defined mass M and a potential term which is the pV potential derived from an external pressure Pext acting on volume V of the system.

Parrinello-Rahman method of pressure and stress control

Both size and shape of cell can change

The Parrinello-Rahman method of pressure and stress control can allow simulation of a model under externally applied stress. This is useful for studying the stress-strain relationship of materials. Both the shape and the volume of the cell can change, so that the internal stress of the system can match the externally applied stress.

Summary of derivation

The method is presented in detail by Parrinello and Rahman (1981) and is only summarized here.

The Lagrangians of the system are modified such that a term representing the kinetic energy of the cell depends on a user-defined mass-like parameter W. An elastic energy term p is related to the pressure and volume or the stress and strain of the system. The equations of motion for the atoms and cell vectors can be derived from this Lagrangian. The motion of the cell vectors, which determines the cell shape and size, is driven by the difference between the target and internal stress.

With hydrostatic pressure only:

Eq. 112        

where h = {a · b · c} is the cell vector matrix, G = h´h, ri = hsi, is the interaction potential, and is the volume of the cell. The dots above some symbols indicate time derivatives and the primes indicate matrix transposition. Tr is the trace of a matrix.

With stress, the elastic term p is replaced by:

Eq. 113        

where S is the applied stress, 0 is the initial volume, and is the strain; is also a tensor, defined as (h0´-1 Gh0-1 - 1)/2.

Choice of the mass variable

Note the user-defined variable W, which determines the rate of change of the volume/shape matrix.

A large W means a heavy, slow cell. In the limiting case, infinite W reverts to constant-volume dynamics. A small W means fast motion of the cell vectors. Although that may mean that the target stress can be reached faster, there may not be enough time for equilibration.

In tests, too small a W also induced artificial periodic motions of the cell. A value of 20 seems to give satisfactory results for a cell of 76mers of polyethylene.


Types of dynamics simulations

In addition to relatively simple dynamics simulations, it is also possible to bias or control the dynamics run in several ways and/or to combine dynamics and minimization in one simulation. These types of dynamics runs can be used in conjunction with one or more of the various thermodynamic statistical ensembles (as appropriate, see Statistical ensembles).

Several such types of dynamics simulations are discussed in this section:

Quenched dynamics

Simulated annealing

Consensus dynamics

Impulse dynamics

Langevin dynamics

Stochastic boundary dynamics

Multibody order-N dynamics

Table 22 . Types of dynamics simulations easily set up with MSI simulation engines

method simulation engine
CHARMm Discover OFF
Quenched dynamics  
 
1  
 
Simulated annealing  
2  
1  
 
Consensus dynamics  
 
3  
 
Impulse dynamics  
 
4  
 
Langevin dynamics  
 
 
 
Stochastic boundary dynamics  
 
 
 
1 Available through Insight·Discover_3 module for CDiscover, via DSL only for FDiscover; not yet available in Cerius2·Discover.

2 Truncated (only one T-to-0 K half-cycle), and referred to as "quenched dynamics" in the CHARMm documentation.

3 Available in standalone mode only.

4 Available in standalone mode of CDiscover only.

Choosing the type of dynamics

To access the controls used for specifying types of dynamics simulations:

Alternatively, to produce a phi/psi map (showing energy as a function of rotation about two torsions of a model), you can use the Strategy/Phi_Psi_Map command.

Langevin dynamics is set up by clicking the Setup Detailed Dynamics button. You can combind Langevin dynamics with simulated annealing by setting the Bath Temperature to 0.

Stochastic boundary conditions are set by selecting the CHARMm/Constraints Options/Stochastic Bdry Settings menu item to specify these constraints and then selecting the CHARMm/Constraints Options/Stochastic Bdry On menu item to activate them.

Exploring conformational space using dynamics

A common limitation of classical minimization algorithms is that they usually locate a local minimum close to the starting configuration, which is not necessarily the global minimum (Local or global minimum?). This is because the minimizers discussed under Minimization algorithms are specifically designed to ignore configurations if the energy increases.

By using the available thermal energy to climb and cross conformational energy barriers, dynamics provides insight into the accessible conformational states of the molecule that would be inaccessible to classical minimization.

The results can be assessed by plotting selected data in 3D format (e.g., a phi/psi map for rotation about the peptide bond) or by minimizing selected structures that are generated during the dynamics run (see Quenched dynamics).

Quenched dynamics

How it works

In quenched dynamics, periods of dynamics are followed by a quench period in which the structure is minimized. You can specify the simulation time between quenches and the number of minimization steps. The quenched structure can then be written to a trajectory file (Dynamics trajectories), and dynamics continues with the prequenched structure.

Uses

Quenched dynamics is a way to search conformational space for low-energy structures.

Simulated annealing

How it works

In simulated annealing, the temperature is altered in time increments from an initial temperature to a final temperature and back again. This cycle can be repeated. The temperature is changed by adjusting the kinetic energy of the structure (by rescaling the velocities of the atoms).

Simulated annealing can be combined with quenched dynamics. That is, at the end of each temperature cycle, the lowest-energy structure of that cycle can be minimized and saved in a trajectory file (Dynamics trajectories). Annealing continues using the last structure and velocities from the previous cycle.

Tip

Use the quartic form of the nonbond term for simulated annealing studies in which the initial coordinates are unknown. These simulations can involve atoms moving through other atoms and for this it is essential that the nonbond term not go to infinity.

Simulated annealing cannot be combined with impulse dynamics (Impulse dynamics).

Uses

Annealing allows the energy of the structure to be changed gradually, without allowing the structure to become trapped in a conformation that has a lower energy than nearby conformations but a higher energy than more distant conformations (that is, in a local energy minimum).

Consensus dynamics

How it works

Consensus dynamics can be thought of as an extension of the tethering restraint technique (Template forcing, tethering, quartic droplet restraints, and consensus conformations). In tethering, a model system serves as a fixed template. The "moving" system, driven by molecular mechanics or dynamics, is then forced to conform to the template by applying restraints.

In contrast, the consensus technique allows the "template" to respond to changes in the "moving" system, by treating all the models as "moving templates". The net result is that both models change so that their structures become similar. Consensus dynamics can be applied to more than two models simultaneously.

Uses

Consensus restraints have several uses. One example is the determination of structural similarities among a set of homologous compounds--perhaps several similar compounds that bind to a particular receptor. If you hypothesize that an apparently homologous region of the compounds is responsible for the binding, then you would want to find a configuration for this region that is compatible with relatively low-energy conformations for all the compounds. By applying consensus restraints to the homologous region of each compound, you can use dynamics followed by minimization to find a likely binding configuration.

Impulse dynamics

How it works

Impulse dynamics allows you to assign initial directional velocities to selected atoms before carrying out dynamics.

Impulse dynamics can be used only with constant-energy, constant-volume dynamics (NVE ensemble) and cannot be combined with simulated annealing (Simulated annealing).

Uses

You can use impulse dynamics to push interacting molecules over energy barriers before allowing the structure to relax.

Langevin dynamics

How it works

The Langevin dynamics method (McCammon et al. 1976, Levy et al. 1979) approximates a full molecular dynamics simulation of a system by eliminating unimportant or uninteresting degrees of freedom. The effects of the eliminated degrees of freedom are simulated by mean and stochastic forces.

Friction coefficients must be assigned to selected atoms before starting the dynamics run. Langevin dynamics includes a constant-temperature bath.

Uses

For example, instead of simulating hundreds of solvent molecules surrounding the solute molecules, the solvent can be ideally represented as a viscous fluid described in terms of dissipative and fluctuative equations.

Stochastic boundary dynamics

How it works

The stochastic boundary molecular dynamics method (Brooks et al. 1985a) uses a combination of both Langevin dynamics and Newtonian dynamics. With this method, the model is partitioned into a reaction region where Newtonian dynamics simulation is run, a buffer region where Langevin dynamics is run, and a reservoir region.

In this way, atoms distant from the specific interactive sites in a large macromolecular system can be effectively eliminated from extensive analysis.

Uses

This allows detailed studies of spatially localized portions of interacting models. Enzyme-substrate interactions at the active site can be effectively studied using this technique.

Multibody order-N dynamics

The Insight·MBOND module and the standalone MBOND/CHARMm program (documented separately) is used for multibody order-N dynamics, in which models are substructured into a set of interconnected rigid and flexible bodies, as well as atomistic regions. The rigid bodies move as units, and the deformations of flexible bodies are represented by sets of low-frequency component modes. The atomistic regions are simulated by conventional dynamics. This method is another way of eliminating unimportant or uninteresting degrees of freedom and thus allowing a longer timestep to be used.


Constraints during dynamics simulations

Constraints and restraints (Applying constraints and restraints) can be selectively applied during a dynamics run to save computing time and/or focus the simulation on more interesting parts of your model. For example, atoms can be specified as fixed or movable, and restraints can be applied in order to pull certain atoms towards one another.

Improving computational efficiency

In addition, certain types of constraints are applied only during dynamics, to increase computational efficiency. This section includes information only on constraints that are used only in dynamics simulations:

The SHAKE algorithm

The RATTLE algorithm

Setting constraints during dynamics

To access the controls used for setting up SHAKE or RATTLE dynamics:

Note

For full access to the RATTLE functionality, you need to use BTCL statements. You can write out a command input file from Cerius2 or Insight, edit the BTCL rattle command statement(s) with a text editor, and then use that file for your run. Alternatively (in Insight), select the Language_Control/Command_Comment command. Set the Comment Type to Command. Enter the desired rattle command statement(s) in the Command/Comment entry box and select Execute.

Effect on timestep and energy conservation

Bond vibrations constitute the highest frequencies in the system and thus determine the largest timestep that can be used during dynamics. If the bonds are constrained, longer timesteps can be used during dynamics, because of the absence of these high-frequency bond vibrations.

In a test run on crambin, a timestep as long as 3 fs could be used with the RATTLE algorithm, but dynamics without RATTLE or SHAKE can have a timestep of no more than 1 fs. Furthermore, energy conservation for the 3 fs timestep with RATTLE was as good as that at 1 fs without RATTLE or SHAKE.

Computational costs

Although constraining bonds allows the timestep to be increased without incurring significant overhead for carrying out RATTLE iteratively, constraining angles does not really help to reduce the computational cost, because it takes a long time for the iterative procedure to converge. In some circumstances, it actually takes much longer than without RATTLE or SHAKE.

Tip

We do not recommend that angle constraints be used even though this functionality is available. However, if angle constraints have to be used, you should use a larger tolerance than with bonds.

The SHAKE algorithm

Availability

The SHAKE method (Ryckaert et al. 1977), which is a popular algorithm for introducing distance constraints during molecular dynamics simulations, is used in CHARMm to constrain the harmonic stretching of bonds.

Allowed constraints during dynamics

SHAKE may be applied to all bonds or only to bonds containing hydrogens, and to no angles, all angles, or only angles containing hydrogens.

Tip

To maintain a high degree of accuracy during the dynamics simulation, as well as to provide a larger integration time step, it is best to apply SHAKE only to bonds containing hydrogens.

Effect on timestep

A twofold increase in the time step (to 0.001 ps) is enabled.

The RATTLE algorithm

What is the RATTLE algorithm

Constraints can be applied during dynamics runs via the RATTLE algorithm, which is the velocity version of SHAKE. The RATTLE procedure is to go through the constraints one by one, adjusting the coordinates so as to satisfy each in turn. The procedure is iterated until all the constraints are satisfied to within a given tolerance. Unlike SHAKE, RATTLE (Andersen 1983) makes sure that velocities are adjusted also, to satisfy the constraints, and is suitable for use with the velocity version of Verlet integrators.

Allowed constraints during dynamics

RATTLE can be used to constrain bonds, angles spanned by two constrained bonds, as well as the distance between any pair of atoms in periodic and nonperiodic systems. It can be used with the constant-volume ensemble.

Note

For the time being, RATTLE cannot be used with constant-pressure dynamics, since that involves calculating the pressure on a molecule basis, which is not currently available.

Dynamics in fixed-geometry water

Another major use of RATTLE is to enable the use of a fixed-geometry water model. With RATTLE, you can use the SPC, TIP3P, or current water models. The initial configurations of the water molecules are set to the equilibrium geometry of the SPC or TIP3P model or to the current geometry of water in your model.


Dynamics trajectories

What are trajectories?

The results of dynamics simulations can be saved in trajectory files, which are essentially a series of snapshots of the simulation taken at regular intervals. Trajectories can include data such as model structure, minimized-model structure (in quenched dynamics, see Quenched dynamics), temperature, energies, volume, pressure, cell parameters, and stress.

Uses

Trajectory files can be used for analysis of the results and also for continuing an interrupted dynamics simulation.

You can use trajectory files for computing the average structure and for analyzing fluctuations in geometric parameters, thermodynamic properties, and time-dependent processes. The time series can be evaluated for global properties such as the radius of gyration and the number density of the model. Examples of experimental quantities that can be calculated using correlation functions are frictional coefficients, IR line widths, fluorescence depolarization rates, spectral densities (the Fourier transform of the correlation function), and NMR relaxation times.

Animations

You can also animate trajectories, to view how the model behaved during a dynamics run.

Typically, the displayed model's conformation is updated during a run so you can monitor progress visually. The frequency with which the display is updated can affect the time needed to perform a dynamics simulation, especially with large models.

With the Cerius2·Dynamics Simulation module, some information also can be displayed in the form of graphs that are updated as the run proceeds.


General methodology for dynamics calculations

To perform dynamics simulations, you need to know something about the general strategy of all types of dynamics simulations, and you need to know the general procedure for setting up a dynamics run.

This section includes information on:

Stages and duration of dynamics simulations

Dynamics with MSI simulation engines

Restarting a dynamics simulation

Prerequisites

One of the most important steps in any simulation is properly preparing the system to be simulated. Calculations on the fastest computer running the most efficient dynamics algorithm may be worthless if the hydrogen is put on the wrong nitrogen or an important water molecule is omitted.

Unfortunately, it is impossible to provide a single recipe for a successful model--too much depends on the objectives and expectations of each calculation. Are conformational changes that are far removed in conformational space from the area being focused on expected or interesting? What is the hypothesis being tested? The effects of specific dynamics variables, as well as tethering, fixing, energy cutoffs, etc., on the results can be answered only by controlled preliminary experiments.

Stages and duration of dynamics simulations

Dynamics simulations are usually carried out in two stages, equilibration and data collection (or "production"). The duration of each stage depends on the system as well as on the purpose of the run.

Equilibration stage

For the equilibration stage, you typically assign random velocities to atoms in the model according to the Maxwell-Boltzmann distribution around the desired target temperature. Temperature control during the equilibration stage is usually by direct velocity scaling (Direct velocity scaling). Depending on your model and the purposes of your simulation, you may bring the simulated system up to the target temperature relatively quickly or in gradual steps.

The purpose of equilibration is to prepare the system so that it comes to the most probable configuration consistent with the target temperature and pressure.

When a system is large, it may take a long time to equilibrate because of the vast conformational space it has to search. The contour of the energy surface is another factor to consider. If energy barriers between various local minima and the global minimum are high, barrier crossing is difficult and may take longer.

Has equilibrium been achieved?

One way to judge whether a model has equilibrated is to plot the various thermodynamic quantities, such as energy, temperature and pressure, versus time. When equilibration has been achieved, these quantities fluctuate around their averages, which remain constant over time. This is a necessary but not sufficient test, because it is not unusual for a sudden conformational change to occur after a long period of time.

Another way to check equilibration is to start the calculation with different initial conformations and different initial velocities. Convergence to similar conformations and properties from different initial values is a good indicator that equilibration has occurred.

Production (data-collection) stage

After equilibrating the system at the target temperature and pressure, you can begin the production stage, during which data and statistics are collected. Temperature control during the production stage is generally by a more gentle, realistic method than direct velocity scaling (How temperature is controlled). You may use different thermodynamic ensembles (Statistical ensembles) during the equilibration and production stages.

Depending on the purpose of your study (Types of dynamics simulations), you may run several data-collection simulations under different conditions after a single equilibration run or stage.

How long should the simulation be?

One commonly asked question is how long to collect the statistics. The answer depends on both the properties being calculated and the model being simulated.

Some quantities, such as internal energy, temperature, and pressure, converge readily, since they are calculated from averages and the dominant contributions are from the most probable states. Other quantities, like specific heat or isothermal compressibilities, are more difficult to obtain, since they depend on the fluctuations.

One empirical way to find out how long is long enough for a dynamics simulation is to monitor the change in the desired quantities over time.

The length of a trajectory needed to calculate a property depends on the time variation of the property under consideration. If the property is a slowly varying function, the dynamics integration should be extended to cover several periods.

Characteristic durations for common events in real molecules are listed in Table 23.

Table 23 . Durations of some real molecular events

Event Approximate duration
Bond stretching.  
1-20 fs  
Elastic domain modes.  
100 fs to several ps  
Water reorientation.  
4 ps  
Inter-domain bending.  
10 ps-100 ns  
Globular protein tumbling.  
1-10 ns  
Aromatic ring flipping.  
100 µs to several seconds  
Allosteric shifts.  
2 µs to several seconds  
Local denaturation.  
1 ms to several seconds  

If a property varies randomly about a mean value with a decay time of t and the simulation is run for length T, the variance in the estimate is proportional to (t/T)0.5. When multiple independent fragments are present (for example, each water molecule in a solvent simulation), averaging can be done over the fragments to improve sampling.

Dynamics with MSI simulation engines

Prerequisites

To set up a dynamics run, first:

1.   Choose the desired forcefield if you don't want to use the default forcefield (Forcefields).

2.   Set up the forcefield and prepare your model (Preparing the Energy Expression and the Model).

3.   Run a preliminary minimization

The model usually needs to be minimized (Minimization) to remove strains that might cause abnormally large forces on some atoms and therefore result in unrealistic dynamics simulations.

Purpose of the run

Next:

4.   Specify items such as the dynamics algorithm(s) (Integration algorithms), time step (The choice of timestep), and temperature- and pressure-control methods (How temperature is controlled and How pressure and stress are controlled) (unless you want your calculation to run under the default conditions).

You generally run two dynamics simulations in sequence (or a single two-stage dynamics run). The first run or stage is for equilibrating the model under the desired conditions and the second, for collecting data and statistics (see Stages and duration of dynamics simulations. In Discover and CHARMm, these two runs are generally both set up at the same time, by repeating Steps 4 and 5 for each stage before starting the run.

Additional runs or stages may be required, depending on the purpose of the simulation (see Types of dynamics simulations). Depending on the type of run and the simulation engine, you may be able to set up all stages at the same time, by repeating Steps 4 and 5 for each stage before starting the run. If not, the procedure for restarting runs that have ended is outlined under Restarting a dynamics simulation.

Accessing dynamics controls

To find the relevant controls in the molecular modeling programs:

In addition, you may select Dynamics Controls on the DYNAMICS SIMULATION card. These controls, however, are typically left at their default values.

Additional control over the equilibration stage is available by clicking the More... pushbutton in the Equilibration section to open the Equilibration Options control panel.

If you want a minimization stage to precede the equilibration stage, check the Pre-Minimize check box. The associated More... pushbutton opens the Discover Minimize control panel (Minimization).

Alternatively, if you want to run only a simple minimization followed by dynamics, select the Strategy/Simple_Min_Dyn command and change any desired parameters.

Discover and CHARMm offer additional functionality when run in standalone mode. (How to run Discover and CHARMm in standalone mode is documented separately--see Available documentation.)

Specifying output

5.   Specify the desired output:

Click the Trajectory... button in the Dynamics Simulation control panel or select the Trajectory/Output menu item on the DYNAMICS SIMULATION card to write out various types of trajectory files.

Discover and CHARMm offer additional functionality when run in standalone mode.

Verifying the run instructions

6.   Review what you have requested for the run:

Since dynamics simulations are often quite complex and time consuming, you should review your run specifications before starting the simulation.

To view the complete command input file, select the Run/Run command. Set the controls as desired, being sure that Run_Dynamics is on and Computation Mode is set to Command File, and select Execute. (Run_Minimization can be on or off, depending on whether you want to specify that minimization be performed before the dynamics equilibration stage.) Read the input file with any text editor or the UNIX more command.

Starting a dynamics run

Finally:

7.   Start the dynamics run:

Alternatively, if you want to run only a simple minimization followed by dynamics, select Execute in the Strategy/Simple_Min_Dyn parameter block.

Specific information

For specific information on setting up and running dynamics with the various MSI simulation engines, and on analyzing the results, please see the relevant documentation (see Available documentation).

Restarting a dynamics simulation

You can continue a dynamics run without a break from the previous stage of the simulation, restart an interrupted run from where it ended, or start a new dynamics run from a particular point in a previous run.

When needed

With the Insight·Discover_3 module, continuing a run from one stage to the next (e.g., switching from equilibration to data collection or setting up complicated simulations) is straightforward: all stages can be specified when you set up the simulation, before starting the run. However, you may need to continue or restart a run, for any number of reasons.

Cerius2·Dynamics Simulation and CHARMm enable stages only for limited types of simulations, and Cerius2·Discover and Insight·Discover allow you to specify only a simple two-stage equilibrium and data-collection run (with or without preliminary minimization): to set up some other type of simulation, you must wait until a run has ended before setting up subsequent runs.

Read this section if you need to continue or restart a simulation that has already ended.

Files and precautions

The trajectory and other input file(s) that are required depend on what simulation engine is being used, and the exact procedure depends on whether you a continuing a run without a break from the last conformation of an immediately preceding run, restarting an interrupted run from the last conformation, or using an intermediate point of an earlier run as the basis for a new simulation.

If you changed the model's conformation (for example, by minimization) since your previous dynamics simulation ended, you need to reassign the initial (random) velocities, since the old velocities do not apply to the new coordinates. The Insight·Discover_3 and Cerius2·Dynamics Simulation modules can detect this situation and reinitialize velocities automatically.

You may need to analyze the data from a dynamics run before doing a restart from other than its final conformation. You may, for example, want to find and use the lowest-energy conformer as the starting point for a new simulation or may want to restart a run with a particular set of velocities.

Please see the relevant specific documentation (Available documentation).

Restarting a dynamics simulation

Briefly, the general procedure with each simulation engine is:

To restart the simulation at the last step of an immediately preceding run, click the RUN DYNAMICS button in the Dynamics Simulation control panel.

To start a new simulation from the end or some point before the end of a previous run, specify a new dynamics method, algorithm, output, etc., if desired (see Purpose of the run). Then click the RUN DYNAMICS button.

To re-randomize the atomic velocities before starting a new run, click the Reset button in the Dynamics Simulation control panel before clicking the RUN DYNAMICS button.

If you want to start the new simulation from some point before the end of a previous run, next set Select File Type to Archive_File (or History_File) and set the Frame No to a constant or a variable. (A variable can be can be initialized with the Language_Control/Command_Comment command and used, for example, in a Loop statement set up with the Language_Control/Looping_Control command.) Select Execute to set up a stage to control reading frame(s) from the desired file.

Specify any other dynamics parameters if desired, as outlined above in Steps 4 and 5. Velocity (in the Dynamics Calculate parameter block) should be set to Current if you want to continue or restart the previous run. Set it to Create if you want to re-randomize the atomic velocities before starting a new run.

Review the command input file and start the run as outlined in Steps 6 and 7.

In Cerius2·Discover, use the prepared command input file by selecting the Run menu item in the DISCOVER card and clicking the Input... pushbutton in the Run Discover control panel. Then select the desired file using the browser controls in the Discover Input File control panel and click the Run .inp File action button.

In Insight·Discover, use the prepared command input file by selecting the Run/Run command, toggling Strategy on, and entering that file's name in the Input_File_Name entry box. Select Execute.

Discover and CHARMm offer additional functionality when run in standalone mode. (How to run Discover and CHARMm in standalone mode is documented separately--see Available documentation.)




Last updated September 04, 1998 at 12:14PM Pacific Daylight Time.
Copyright © 1997, 1998 Molecular Simulations, Inc. All rights reserved.