next up previous
Next: Monte Carlo Simulation Up: Statistical Mechanics: A Brief Previous: Entropy and Temperature

Classical Statistical Mechanics

Section 2.2 of Frenkel & Smit [1] discusses a derivation of the ``quasi-classical'' representation of the canonical partition function, $Q_{\rm classical}$:

\begin{displaymath}
Q_{\rm classical} = \frac{1}{h^{dN}N!}\int \int d{\bf r}^N d...
...\left[-\beta\mathscr{H}\left({\bf r}^N,{\bf p}^N\right)\right]
\end{displaymath} (42)

$\mathscr{H}\left({\bf r}^N,{\bf p}^N\right)$ is the Hamiltonian function which computes the energy of a point in phase space. The derivation of Eq. 42 is not repeated here. What is important is that the probability of a point in phase space is represented as
\begin{displaymath}
P\left({\bf r}^N,{\bf p}^N\right) = \left(Q_{\rm classical}\...
...left[-\beta\mathscr{H}\left({\bf r}^N,{\bf p}^N\right)\right].
\end{displaymath} (43)

So, the general ``sum-over-states' ensemble average of quantum statistical mechanics, first presented in Eq. 1, becomes an integral over phase space in classical statistical mechanics:
\begin{displaymath}
\left<G\right> = \frac{
\mbox{
\begin{minipage}{8cm}
\beg...
...,{\bf p}^N\right)\right]
\end{displaymath} \end{minipage} }},
\end{displaymath} (44)

where $G\left({\bf r}^N,{\bf p}^N\right)$ is the value of the observable $G$ at phase space point $\left({\bf r}^N,{\bf
p}^N\right)$. Before moving on, it is useful to recognize that we normall simplify this ensemble average by noting that, for a system of classical particles, the usual choice for the Hamiltonian has the form
\begin{displaymath}
\mathscr{H}\left({\bf r}^N, {\bf p}^N\right) = \mathscr{K}\left({\bf p}^N\right)
+ \mathscr{U}\left({\bf r}^N\right)
\end{displaymath} (45)

where $\mathscr{K}$ is the kinetic energy, which is only a function of momenta, and $\mathscr{U}$ is the potential energy, which is only a function of position. The canonical partition function, $Q$, can in this case be factorized:
$\displaystyle Q\left(N,V,T\right)$ $\textstyle =$ $\displaystyle \frac{1}{h^{3N}N!} \left\{\int d{\bf p}^N \exp\left[-\beta\mathsc...
...int d{\bf r}^N \exp\left[-\beta\mathscr{U}\left({\bf r}^N\right)\right]\right\}$ (46)
  $\textstyle =$ $\displaystyle \left\{\frac{V^N}{h^{3N}N!} \int d{\bf p}^N \exp\left[-\beta\math...
...int d{\bf r}^N \exp\left[-\beta\mathscr{U}\left({\bf r}^N\right)\right]\right\}$ (47)
  $\textstyle =$ $\displaystyle Q_{ideal} Z$ (48)

The quantity in the left-hand braces is the ideal gas partition function, because it corresponds to the case when the potential $\mathscr{U}$ is 0. (Note that we have multiplied and divided by $V^N$; this is the equivalent of scaling the positions in the integration over positions.) The quantity in the right-hand braces is called the configurational partition function, $Z$.

Because the kinetic energy $\mathscr{K}$ has the simple form,

\begin{displaymath}
\mathscr{K}\left({\bf p}^N\right) = \sum_i \frac{{\bf p}_i^2}{2m_i},
\end{displaymath} (49)

where $m_i$ is the mass of particle $i$, the integral over particle momenta can be evaluated analytically:
$\displaystyle \int d{\bf p}^N \exp\left[-\beta\mathscr{K}\left({\bf p}^N\right)\right]$ $\textstyle =$ $\displaystyle \prod_{i=1}{3N}\int dp_i \exp\left(-\frac{p_i^2}{2mk_BT}\right)$ (50)
  $\textstyle =$ $\displaystyle \left(2\pi m k_B T\right)^{3N/2}.$ (51)

(We have assumed all particles have the same mass, $m$; in the case of distinct masses, this is just a product of similar factors.)

$Q_{ideal}$ becomes

$\displaystyle \frac{V^N}{N!h^{3N}} \int d{\bf p}^N \exp\left[-\beta\mathscr{K}\left({\bf p}^N\right)\right]$ $\textstyle =$ $\displaystyle \frac{V^N}{N!}\left(\sqrt{\frac{2\pi m k_B T}{h^2}}\right)^{3N}$ (52)
$\displaystyle = Q_{ideal}\left(N,V,T\right)$ $\textstyle =$ $\displaystyle \frac{V^N}{N!\Lambda^{3N}}$ (53)

where $\Lambda$ is the de Broglie wavelength.

So, when the observable $G$ is a function of positions only, the ensemble average becomes a configurational average:

\begin{displaymath}
\left<G\right> = Z^{-1}\int d{\bf r}^N \exp\left[-\beta\mathscr{U}\left({\bf r}^N\right)\right] G\left({\bf r}^N\right).
\end{displaymath} (54)

Note that the integation over momentum yields a factor $Q_{ideal}$ in both the numerator and denominator, and thus divides out. We can write this configurational average using a probability distribution, $\rho_{NVT}$, as
\begin{displaymath}
\left<G\right> \int d{\bf r}^N G\left({\bf r}^N\right)\rho_{NVT}\left({\bf r}^N\right)
\end{displaymath} (55)

where
\begin{displaymath}
\rho_{NVT}\left({\bf r}^N\right) \equiv Z^{-1}e^{-\beta U\left({\bf r}^N\right)}
\end{displaymath} (56)

is called the ``canonical probability distribution.'' As pointed out on p. 15 of Frenkel & Smit [1], Eq. 54 is ``the starting point for virtually all classical simulations of many-body systems''; that is, it is the starting point for all simulations discussed in this course.


next up previous
Next: Monte Carlo Simulation Up: Statistical Mechanics: A Brief Previous: Entropy and Temperature
cfa22@drexel.edu