Franz J. Vesely > MolSim Tutorial > MD in other ensembles  
 
 





 
< >
Vesely MolSim Tutorial




3.4 Molecular Dynamics in other Ensembles

NVT, NPH, NPT, ...

In the basic MD setting the macroscopic observables $N,V,E,\vec{p}$ are held constant (where $\vec{p}$ is the total linear momentum.) This defines the so-called "MD ensemble" which can be shown to be thermodynamically equivalent to the other ensembles.

Other ensembles may be of advantage for specific problems: $NVT$ is useful for comparisons with theory and MC simulations; non-equilibrium simulations require a thermostat, thus $T=const$; phase transitions are best studied in an $NPT$ ensemble, etc.

Subsections


NVT Molecular Dynamics

Initially (around 1980) stochastic procedures were used, in which the particle velocities were regularly reset to random values picked from the appropriate Maxwell-Boltzmann distribution. [Andersen 1980, Heyes 1983]

Later, deterministic and even reversible equations of motion were developed that nevertheless serve to roam phase space in a more or less "canonical" manner. These are known as Gaussian and Nose-Hoover dynamics, respectively.

Gauss dynamics

Gaussian, or isokinetic, dynamics was introduced by Hoover and Evans in 1982. This is the idea:

Let
$ \sigma\{ \vec{v}\} \equiv \sum_{i=1}^{N} \frac{\textstyle m}{\textstyle 2}v_{i}^{2}-E_{0} $

be a (non-holonomic) constraint to the Newtonian motion of the $N$ -particle system. This is only one constraint equation for $3N$ variables, thus we have some freedom of choice.

A dynamical constraint may always be put in terms of constraint forces. A particularly economical and "impartial" prescription for the definition of these forces was given by C. F. Gauss in 1829:
  • From $\sigma\{ \vec{v}\}=0$ for all times we have
    $ 0=\dot{\sigma} =\sum_{i}\left( \frac{\textstyle \partial}{\textstyle \partial \vec{v}_{i}} \sigma \right) \cdot \dot{\vec{v}_{i}} = m \sum_{i} \vec{v}_{i} \cdot \dot{\vec{v}_{i}} \equiv m \vec{v} \cdot \dot{\vec{v}} $
    Since $\dot{\vec{v}}=\left( \vec{K}+\vec{Z}\right)/m$ (with the Newtonian forces $\vec{K}$ and the yet unknown constraint forces $\vec{Z}$) we have
    $ \vec{v} \cdot \left( \vec{K}+\vec{Z}\right) = 0 $

  • Among all possible sets of constraint forces let us choose the one with the smallest norm. Requiring $\left\vert \vec{Z}^{0}\right\vert^{2}=min$ we find, upon variation, that $\vec{Z}^{0} \cdot \delta \vec{Z}=0$. Since for all physically allowed variations we have $\delta \vec{Z} \cdot \vec{v}=0$ we conclude that $\vec{Z}^{0}=-\lambda \vec{v}$. Inserting this in $\vec{v} \cdot \left( \vec{K}+\vec{Z}^{0}\right) = 0$ we find
    $ \vec{Z}_{i}^{0}= - \frac{\textstyle \sum_{j}\vec{v}_{j}\cdot \vec{K}_{j}}{\textstyle \sum_{j}v_{j}^{2}} \vec{v}_{i} = - \frac{\textstyle \vec{v}\cdot\vec{K}}{\textstyle |\vec{v}|^{2}} \vec{v} $

  • The equations of motion according to Gauss and Hoover are thus

    $ \begin{eqnarray} \dot{\vec{r}_{i}} &=& \vec{v}_{i} \\ \dot{\vec{v}_{i}}&=& \frac{1}{m} \left[ \vec{K}_{i} - \frac{\sum_{j}\vec{v}_{j}\cdot \vec{K}_{j}}{\sum_{j}v_{j}^{2}} \vec{v}_{i} \right] \end{eqnarray} $

    These e.o.m. lead to a deterministic trajectory with $E_{kin}=const$.

    Note that $\vec{v}$ and $\vec{K}$ are needed at the same time step; thus the usual Verlet algorithm is not applicable here. Instead, a Predictor-Corrector algorithm or Runge-Kutta should be used.


Nosé-Hoover dynamics

Thermostats:
The constraint force used in isokinetic dynamics may be thought of as the action of a particularly efficient thermostat, in which not only the average kinetic energy (i. e. the temperature) but even the momentary value of $E_{kin}$ is kept constant at all times.

In a nonequilibrium process work must be invested which systematically heats up the system. In lab experiments a thermostat takes care of this; in simulations the problem of thermostating is non-trivial. The temperature of a molecular dynamics sample is not an input parameter to be manipulated at will; rather, it is a quantity to be "measured" in terms of an average of the kinetic energy of the particles,
$ \langle E_{kin}\rangle \equiv \langle \sum_{i}\frac{\textstyle mv_{i}^{2}}{\textstyle 2}\rangle = d \frac{\textstyle NkT}{\textstyle 2} $

($d \dots $ dimension). Suggestions as to how one can maintain a desired temperature in a dynamical simulation range from repeatedly rescaling all velocities ("brute force thermostat") to adding a suitable stochastic force acting on the molecules. Such recipes are unphysical and introduce an artificial trait of irreversibility and/or indeterminacy into the microscopic dynamics.

A interesting deterministic method of thermostating a simulation sample was introduced by Shuichi Nosé and William Hoover. Nosé had proposed a set of equations of motion that allow for a certain degree of variation in the total energy of the system. He could show that with these dynamics, and under a few very general conditions, the spread around the given $E_{0}$ is such that a canonical distribution is approached.

Hoover rewrote Nosé's original equations of motion in the following simple way:

$ \begin{eqnarray} \dot{\vec{r}} &=& \frac{\vec{p}}{m} \\ \dot{\vec{p}}&=& \vec{K}_{i} - \zeta \vec{p} \\ \dot{\zeta}&=& \frac{2}{Q} \left[ E_{kin} - E_{0} \right] \end{eqnarray} $

where we have omitted the particle index $i$. The coupling parameter $Q^{-1}$ represents the efficiency of the thermal bath holding the temperature constant. $\zeta$ is a kind of "viscosity" which, however, may take on negative values.

For systems of many degrees of freedom (particles) this is sufficient to produce a pseudo-canonical sequence of states. (See also Posch, Hoover, Vesely Phys.Rev.A 33(1986)4253.) For low-dimensional systems it may be necessary to use two NH thermostats in tandem [MARTYNA 92].

For $Q \rightarrow 0$ the spread of $E_{kin}$ around $E_{0}$ approaches zero, and the Nosé-Hoover dynamics becomes identical to Gauss dynamics.


NPH Molecular Dynamics

Andersen (1980) introduced an additional ("synthetic") energy that was coupled to changes of the system volume. Putting $\Delta E_{pot}=P_{0}V$ and $\Delta E_{kin}=Q\dot{V}^{2}/2$ (with a generalized mass $Q$ which may be visualized as the mass of some piston) he wrote the Hamiltonian as

$ H=E_{pot}+\Delta E_{pot}+E_{kin}+\Delta E_{kin} $

Introducing scaled position vectors $\vec{s}_{i} \equiv \vec{r}_{i} V^{-1/3}$ he derived the equations of motion

$ \begin{eqnarray} \ddot{\vec{s}_{i}}&=&\frac{\vec{K}_{i}}{m V^{1/3}} - \frac{2}{3} \frac{\dot{V}}{V}\dot{\vec{s}_{i}} \\ \ddot{V}&=&\left( P-P_{0}\right)/Q \end{eqnarray} $


These equations of motion conserve the enthalpy, as appropriate in an isobaric ensemble.

Parrinello and Rahman (1980) extended the NPH method of Andersen to allow for non-isotropic stretching and shrinking of box sides. Important applications are structural phase transitions in solids.

Scaling of the position vectors now follows the equation $\vec{r}_{i}= H \cdot \vec{s}_{i}$ instead of $=V^{1/3} \vec{s}_{i}$.

The matrix $ H\equiv \left\{ \vec{h}_{1},\vec{h}_{2},\vec{h}_{3}\right\}$ describes the anisotropic transformation of the basic cell. The cell volume is

$ V \equiv \left| H\right| = \vec{h}_{1} \cdot \left[\vec{h}_{2} \times \vec{h}_{3}\right] $

The additional terms in the Hamiltonian are
$ \Delta E_{pot} = P_{0} V\;\;\;\;{\rm and} \;\;\;\; \Delta E_{kin} = \frac{\textstyle Q}{\textstyle 2}\sum_{\alpha}\sum_{\beta}\dot{H}_{\alpha\beta}^{2} $

and the equations of motion are

$ \begin{eqnarray} \ddot{s}_{i}&=&\frac{1}{m} H^{-1} \cdot \vec{K}_{i} - G^{-1}\cdot \dot{ G} \cdot \dot{\vec{s}} \\ \ddot{ H}&=&\frac{1}{Q} \left[ P- I P_{0}\right] V ( H^{-1})^{T} \end{eqnarray} $


where $ G= H^{T} \cdot H$ is a metric tensor, $Q$ is a virtual mass (of dimension mass), and the pressure/stress tensor is defined as

$ ( P)_{\alpha \beta} = \frac{1}{V} \left[ \sum_{i}m\left( H \cdot \vec{s}_{i}\right)_{\alpha}^{T} \left( H \cdot \vec{s}_{i}\right)_{\beta} + \frac{1}{2}\sum_{i}\sum_{j} \left( H\cdot \vec{s}_{ij}\right)_{\alpha} \left( \vec{K}_{ij}\right)_{\beta} \right] $


Morriss and Evans (1983) devised another type of NPH dynamics. They suggested to constrain the pressure not by an inert piston but by a generalized constraint force in the spirit of Gaussian dynamics.

The same idea may be carried over to constrain $E_{kin}$ in addition to the pressure. In this way one arrives at a $NPT$ molecular dynamics procedure. (See Allen-Tildesley, Ch. 7.)


NPT Molecular Dynamics

The $NPT$ dynamics of Morriss and Evans (see end of previous subsection) was actually a $NPE_{kin}$ dynamics, with zero fluctuation of the kinetic energy. A genuine $NPT$ dynamics would constrain $E_{kin}$ to stay near a desired average but with - hopefully gaussian distributed - fluctuations around that value.

Such a method was conceived by Hoover (1985). Extending the Hoover-Nosé idea to keep both the pressure and the kinetic energy near a desired value $P_{0}$, he wrote the equations of motion as

$ \begin{eqnarray} \dot{\vec{s}_{i}} &=& \vec{v}_{i} V^{-1/3} \\ \dot{\vec{v}_{i}} &=& \frac{1}{m}\vec{K}_{i}-(\chi+\zeta) \vec{v}_{i} \\ \dot{\zeta} &=& \frac{2}{Q} \left[ E_{kin}-E_{0} \right] \\ \chi &=& \frac{\dot{V}}{3V} \\ \dot{\chi} &=& \frac{1}{t_{P}^{2} kT} \left[ P-P_{0} \right] \end{eqnarray} $


where $t_{P}$ and $Q$ are relaxation times of the pressure and temperature regulators, respectively.



Exercises

Gauss Dynamics:
Write a Gaussian MD program to simulate the following system:
3 particles without pairwise interaction are moving in a one-dimensional harmonic potential. The sum of their kinetic energies is kept constant. Draw the $x,\dot{x}$ trajectory of the first particle. Use Runge-Kutta.

Nosé-Hoover thermostat:
The equations of motion of the Nosé oscillator with $k=m=E_{0}=1$ are

$ \begin{eqnarray} \dot{x}&=&v \\ \dot{v}&=&-x-\zeta v \\ \dot{\zeta}&=&\frac{\textstyle 1}{\textstyle Q} \left[ v^{2}-2\right] \end{eqnarray} $

Integrate this set of equations, using Runge-Kutta, and draw the $x,v$ trajectory.

vesely may-06

< >