2.6 First results from the calculation of the state sum

The canonical partition function for a Hamiltonian which completely separates into subspaces can be written in the form

 \begin{equation*} Z_C(T,V,N) = \sum_i \exp(-\beta U_i) = \sum_{[n_\alpha]}^N \exp\left(-\beta \sum_{\alpha}n_{\alpha}\epsilon_{\alpha} \right) \qquad . \end{equation*}(2.40)

The equation on the right side must fulfill the restriction

 \begin{equation*} N = \sum_{\alpha} n_{\alpha} \qquad . \end{equation*}(2.41)

The grand canonical partition function of a Hamiltonian which completely separates into subspaces can be written in the form

 \begin{equation*} Z_{GC}(T,V,\mu) = \sum_{N=0}^{\infty}\sum_{[n_\alpha]}^N \exp\left(-\beta \sum_{\alpha}n_{\alpha} (\epsilon_{\alpha} - \mu)\right) \qquad . \end{equation*}(2.42)

The big advantage of this partition function is the identity

 \begin{equation*} \sum_{N=0}^{\infty}\sum_{[n_\alpha]}^N ... = \sum_{n_1} \sum_{n_2} \sum_{n_3} ... \end{equation*}(2.43)

i.e. states can be occupied independently and within the partition function an independent sum over independent energy subspaces is performed, leading to a product over subspace sums.
Finally we get

 \begin{equation*} \begin{split} Z_{GC}(T,V,\mu) & = \prod_{\alpha} \sum_{n_{\alpha}} \exp(-\beta n_{\alpha} (\epsilon_{\alpha} - \mu))\\ & = \prod_{\alpha} \left( 1 - \exp(-\beta (\epsilon_{\alpha} - \mu) \right)^{-1} \qquad \mbox{for Bosons}\\ & = \prod_{\alpha} \left( 1 + \exp(-\beta (\epsilon_{\alpha} - \mu) \right)^{+1} \qquad \mbox{for Fermions} \end{split} \end{equation*}(2.44)
Just a reminder:

For the grand canonical potential we find

 \begin{equation*} \Omega(T,V,\mu) = \pm k T \sum_{\alpha} \ln\left( 1 \mp \exp\left(- \frac{\epsilon_{\alpha}- \mu}{k T} \right) \right) \label{def_gc_pot} \end{equation*}(2.45)

and

 \begin{equation*} N = - \frac{\partial \Omega}{\partial \mu} = \mp k T \sum_{\alpha} \frac{1}{\exp\left(\frac{\epsilon_{\alpha}- \mu}{k T}\right) \mp 1} \frac{\mp 1}{k T} = \sum_{\alpha} \frac{1}{\exp\left(\frac{\epsilon_{\alpha}- \mu}{k T}\right) \mp 1} \qquad . \end{equation*}(2.46)

If \(\exp\left( - \frac{\epsilon_{\alpha} - \mu}{k T}\right) \ll 1\) holds we get

 \begin{equation*} \ln\left(1 \mp \exp\left( - \frac{\epsilon_{\alpha} - \mu}{k T} \right) \right) \approx \mp \exp\left( - \frac{\epsilon_{\alpha} - \mu}{k T} \right) \qquad . \end{equation*}(2.47)

In this case we find for Fermions as well as for Bosons

 \begin{equation*} N = - \frac{\partial \Omega}{\partial \mu} = \mp k T \sum_{\alpha}\mp \exp\left( - \frac{\epsilon_{\alpha} - \mu}{k T} \right)\frac{1}{k T} = \sum_{\alpha} \exp\left( - \frac{\epsilon_{\alpha} - \mu}{k T} \right) \end{equation*}(2.48)

This is the so called Boltzmann statistics. As we will see later the above assumption is fulfilled for extremely diluted systems and at high temperatures (classical particles).


With frame Back Forward as PDF

© J. Carstensen (Stat. Meth.)