next up previous contents
Next: Thermodynamic Properties Up: Anharmonicity Previous: The effect of transition Home: Return to my homepage

The effect of well anharmonicity

Here we follow the method of Haarhoff[164,166] to calculate an anharmonic correction term to the density of states. The energy levels for a Morse oscillator are given by
\begin{displaymath}
E=(n+\half)h\nu-(n+\half)^2{\left(h\nu\right)^2\over 4D_e}.\end{displaymath} (3.17)
This quadratic can be solved for n. The root corresponding to a bound state is
\begin{displaymath}
n+\half={2D_e\over h\nu}\left(1-\sqrt{1-{E\over D_e}}\right).\end{displaymath} (3.18)
Assuming n is continuous and differentiating with respect to E gives a classical density of states,

 \begin{eqnarray}
\Omega(E)&=&{dn\over dE}={1\over h\nu\sqrt{1-\displaystyle{E\ov...
 ...E\over 2D_e} +{3\over 8}\left({E\over D_e}\right)^2+\cdots\right),\end{eqnarray}



(3.19)



where the square root has been expanded binomially. The expansion is valid if E/De is small. The first term in the series is the harmonic term. The full series will diverge as $E\rightarrow D_e$. This divergence is the correct behaviour for the Morse oscillator, but for the case of a cluster isomerization $\Omega(E)$ remains finite as $E\rightarrow \Delta$ from below, where $\Delta$ is the barrier height. As we are seeking a correction term for well anharmonicity, it seems reasonable to truncate equation 3.19 and examine the effect of the first term in this series. Laplace transformation then gives for the partition function
\begin{displaymath}
z={1\over h\nu}\left({1\over\beta}+{1\over 2\Delta\beta^2} \right).\end{displaymath} (3.20)
The multi-dimensional partition function is then
\begin{displaymath}
Z=\prod_{j=1}^\kappa{1\over\beta h\nu_j}\left(1+{1\over 2\Delta_j\beta}\right).\end{displaymath} (3.21)
Making the approximation that an average value of $1/\Delta$ can be used then gives
\begin{displaymath}
Z={1\over\prod_{j=1}^\kappa h\nu_j}\sum_{l=0}^\kappa {C^\kappa_l a^l\over\beta^{\kappa+l}},\end{displaymath} (3.22)
where $a=\langle 1/2\Delta\rangle$ is an anharmonicity parameter and $C^\kappa_l$is a binomial coefficient. Inverse Laplace transforming gives
\begin{displaymath}
\Omega(E)={1\over\prod_{j=1}^\kappa h\nu_j}\sum_{l=0}^\kappa {C^\kappa_l
a^l E^{\kappa+l-1}\over\Gamma(\kappa+l)}.\end{displaymath} (3.23)
The total density of states is found by summing over all the minima, giving
\begin{displaymath}
\Omega(E)=\sum_{E^0_s<E}{n_s^*\over \prod^{\kappa}_{j=1} h\n...
 ...{C^\kappa_l a_s^l (E-E^0_s)^{\kappa+l-1}\over\Gamma(\kappa+l)}.\end{displaymath} (3.24)
Converting the above equation to the $\gamma$ formulation using equation 3.7 gives
\begin{displaymath}
\Omega(E)\propto\sum_{E^0_s<E}\gamma(E')_s\left[\sum_{l=0}^\...
 ...a_l a_s^l (E'-E^0_s)^{\kappa+l-1}\over\Gamma(\kappa+l)}\right].\end{displaymath} (3.25)
The temperature follows from equation 3.10  
 \begin{displaymath}
T_\mu={\displaystyle{\sum_{E^0_s<E}\gamma(E')_s\left[\sum_{l...
 ..._l a_s^l(E'-E^0_s)^{\kappa+l-1}\over\Gamma(\kappa+l)}\right]}}.\end{displaymath} (3.26)
This anharmonic correction has a similar form to that used by Chekmarev and Urmirzakov[165] but has been derived in a different way.

We consider two possible ways of calculating as from the potential energy surface of LJ55. Firstly, we consider using the third derivatives of the potential. The dissociation energy of a one-dimensional Morse oscillator can be found from the second and third derivatives of the potential by
\begin{displaymath}
D_e={(V'')^3\over (V''')^2}.\end{displaymath} (3.27)
If off-diagonal elements in the multidimensional case are ignored, the above equation can be used to calculate the barrier heights associated with each normal mode. Using analytical third derivatives of the LJ potential, the Cartesian third derivatives for the LJ55 icosahedron were calculated and transformed to obtain the diagonal elements in the Hessian eigenvector basis. This scheme underestimates the anharmonicity and gives values for as which when substituted into equation 3.26 have an insignificant effect on $\Omega(E)$, because the off-diagonal elements, which far outnumber the diagonal elements, should not be neglected. However, there is no obvious way to calculate the effect of the off-diagonal elements and for a system such as LJ55 the transformation of the third derivatives into the Hessian eigenvector basis is too computationally expensive for the off-diagonal elements.

The second method considered was to use the barrier heights of our transition state sample. An average barrier height was calculated for each minimum, but for some minima it led to a gross overestimation of the anharmonicity and unphysical caloric curves. The barrier heights are not a good quantitative measure of the anharmonicity of an individual minimum. This method was therefore not considered further.

Therefore, instead of trying to obtain as from the PES, we considered how it could be obtained by comparison with the simulation results. If as is assumed to be the same for all minima, a value of a can be found which leads to the reproduction of the Van der Waals loop in the caloric curve at the observed temperature and energy, but the temperature difference between the two turning points is still too small. However, one would expect the anharmonicity to depend upon the energy of the minimum: the higher energy minima are likely to have a greater anharmonicity than the solid-like minima.

  
Table 3.5: Values of the anharmonicity parameter, ai, for LJ55 in the six energy ranges given in Table 3.1. Values for I, II, III and VI were found by fitting the temperatures for these regions to the simulation results.
\begin{table}
% latex2html id marker 3003
\begin{center}

\begin{displaymath}
\b...
 ...$\space & 0.73\\ \hline\hline\end{array}\end{displaymath}\end{center}\end{table}

In §3.3, we saw an example of how the properties of different regions of configuration space could be obtained using suitable order parameters. Indeed, for LJ55, the caloric curves associated with minima in different energy ranges were found. Using the superposition method, it is easy to calculate the properties of a particular region of configuration space simply by restricting the summation to those minima with suitable values of an order parameter, i.e. 
\begin{displaymath}
\Omega_i=\sum_{E^0_s<E, s\in i} \Omega_s,\end{displaymath} (3.28)
where $\Omega_i$ is the densities of states associated with region i. From the definition, $1/kT_i=(\partial \ln \Omega_i/\partial E)_{N,V}$, where Ti is the temperature of region i, it follows that  
 \begin{displaymath}
T_i={\displaystyle{\sum_{E^0_s<E, s\in i}\gamma(E')_s\left[\...
 ..._l a_s^l(E'-E^0_s)^{\kappa+l-1}\over\Gamma(\kappa+l)}\right]}}.\end{displaymath} (3.29)
We assigned values of the anharmonicity parameter, ai, to minima in the six different energy ranges (Table 3.1). Values were chosen for regions I, II, III and VI which reproduced the simulation results of Figure 3.4. The caloric curves for each region accurately fitted the simulation curves showing that our anharmonic correction term has an appropriate form. Values of ai for regions IV and V were chosen that were intermediate between the values for regions III and VI, but as these regions only make a small contribution to $\Omega(E)$ the exact value chosen will only have a very small effect on the overall caloric curve. The values of ai are given in Table 3.5. As would be expected the anharmonicity increases with the potential energy of the minima.


  
Figure 3.7: Microcanonical caloric curves for LJ55 using the $\gamma$ formula including anharmonic terms with minima samples A (solid line) and B (dashed line). The calculated caloric curves are compared to simulation (dashed line with diamonds).
\begin{figure}
\epsfxsize=10cm
\centerline{\epsffile{t.ahcal.eps}}\end{figure}

The microcanonical caloric curve was calculated using these values of ai. Figure 3.7 shows that for sample A the calculated curve is in remarkable agreement with the simulation results. The calculated Van der Waals loop now has the correct depth, because the effect of the greater anharmonicity of the liquid-like state is to increase the difference in temperature between the two branches of the caloric curve. The success of this method can be understood from the equation  
 \begin{displaymath}
{1\over T}=\sum_i{p_i\over T_i},\end{displaymath} (3.30)
where pi is the probability that the cluster is in region i. The probabilities at E', pi(E') are fixed by the quench frequencies,  
 \begin{displaymath}
p_i(E')={\sum_{s\in i}\gamma_s(E')\over\sum_s\gamma_s(E')}\end{displaymath} (3.31)
Furthermore, the pi calculated from sample A are in good agreement with simulation results over a wide range of energy (see §3.6). The values of ai have been chosen to reproduce the simulation temperatures, Ti, and so it is unsurprising that the overall temperature, T, is very accurate. Sample B produces worse results because there are fewer quenches to regions 1-3, and so the quench frequencies are subject to larger statistical errors than for sample A.

For LJ13, the STA temperature distribution is bimodal. The high temperature peak corresponds to solid-like clusters associated with the icosahedral global minimum, and the low temperature peak to liquid-like clusters. The residence times in the solid-like and liquid-like states are much shorter than for LJ55 and so the information provided by short-time averaging for LJ13 is less well-resolved. We partitioned the minima distribution into these two regions, and assigned values of ai to each. The values of ai were higher than for LJ55 and the curvature of the Ti curves differed from the simulation results. The apparently greater anharmonicity can be explained by the larger number of surface atoms for LJ13. The incorrect curvature suggests that the energy dependence of the density of states is inappropriate and so the 2nd-order Morse correction term to the density of states was included. The resulting Ti curves fitted the simulation results much more accurately. The partition function for a single minimum including this term is

\begin{eqnarray}
Z&=&\prod_{j=1}^\kappa{1\over\beta h\nu_j}\left(1+{a\over\beta}...
 ...{\kappa-l}
D^\kappa_{l, m} {3^m a^{l+2m}\over\beta^{\kappa+l+2m}},\end{eqnarray}



(3.32)


(3.33)


where
\begin{displaymath}
D^\kappa_{l, m}={\kappa!\over l!m!(\kappa-l-m)!}.\end{displaymath} (3.34)
Inverse Laplace transforming and summing over all minima gives for the total density of states,
\begin{displaymath}
\Omega(E)=\sum_{E^0_s<E}{n_s^*\over \prod^{\kappa}_{j=1} h\n...
 ...a_s^{l+2m} (E-E^0_s)^{\kappa+l+2m-1}
\over\Gamma(\kappa+l+2m)}.\end{displaymath} (3.35)

In an MD simulation one calculates the kinetic temperature, TK. For LJ55 the difference between the thermodynamic and kinetic temperatures is negligible. For LJ13 it is still small but not insignificant, and so TK rather than $T_\mu$ is fitted to the simulation results. Bixon and Jortner found both temperatures produced very similar results in their model calculations of the microcanonical caloric curve[152]. TK has been obtained for the harmonic superposition method by Franke using the equipartition theorem[154], however this approach cannot be extended to include anharmonicity. We obtain TK by a different method. Firstly we note that

 \begin{eqnarray}
\langle E_K\rangle &=&\int_0^E p(E_K) E_K dE_K \\  &=& {1 \over \Omega(E)} \int_0^E E_K \Omega_c(E-E_K) \Omega
_K(E_K) dE_K,\end{eqnarray}



(3.36)


(3.37)


where $\Omega_K(E_K)$ is the kinetic density of states, $\Omega_c(E_c)$ is the configurational density of states, and the potential energy Ec is given by Ec=E-EK. Deconvoluting $\Omega_K(E_K)$ from $\Omega(E)$ gives
\begin{displaymath}
\Omega_c(E_c)={1\over (2\pi m)^{\kappa/2} \prod^\kappa_{j=1}...
 ...+2m} (E_c-E^0_s)^{\kappa/2+l+2m-1}
\over\Gamma(\kappa/2+l+2m)}.\end{displaymath} (3.38)
Integrating equation 3.37, summing over all minima and substituting into equation 3.3 gives
\begin{displaymath}
T_K={\displaystyle{\sum_{E^0_s<E} {n_s^*\over\prod^{\kappa}_...
 ...s^{l+2m} (E-E^0_s)^{\kappa+l+2m-1}
\over\Gamma(\kappa+l+2m)}}}.\end{displaymath} (3.39)

 
Table 3.6: Details of the two samples of minima for LJ13. E' is measured with respect to the global minimum icosahedron.
  $E'/\epsilon$ Number of minima
C 13.7768 95
D 18.3268 295


 
Table 3.7: Values of the anharmonicity parameter, ai, for the solid-like (1) and liquid-like (2) regions of the LJ13 PES. Region I is the potential well of the icosahedron and region II consists of all other minima.
Region $a_i/\epsilon^{-1}$
I 0.470
II 0.625

Sample C (Table 3.6) was used to calculate the caloric curve of LJ13 for the $\gamma$ formulation because it was produced from an MD run in the coexistence region and so should have more accurate quench statistics for the low energy minima, and sample D was used for the n* formulation because it is from the liquid-like region and so has a larger sample of minima. The values of ai assigned to solid-like and liquid-like clusters are given in Table 3.7. From Figure 3.8 it can be seen that the caloric curve given by the n* formulation agrees very well with simulation, because it is possible to obtain a near-complete set of minima for LJ13. (Tsai and Jordan found 1328 minima in a comprehensive search of the LJ13 PES[162]. Many of these, though, are too high in energy to be thermodynamically significant.) The $\gamma$ formulation, however, has too high a transition temperature. This error arises because the probability that the cluster is in the solid-like state derived from quenching is higher than the corresponding probability derived from the STA temperature distributions. In the $\gamma$ formulation the probabilities, pi, at E' are fixed by the quench statistics (equation 3.31), and so this constraint leads to the higher transition temperature. The difference between the probabilities derived from quenching and the STA temperature distributions may be because there are regions of the PES which are in the basin of attraction of the icosahedron but which have thermodynamic properties similar to a liquid-like well, or because the short-time averaging does not distinguish between the solid-like and liquid-like states with the same resolution as for LJ55. This difference does not stem from the quench method since we obtained similar quench statistics with steepest-descent[170,171], conjugate gradient[87] and eigenvector-following[88] methods.

  
Figure 3.8: Microcanonical caloric curves for LJ13 using the n* formula (solid line), the $\gamma$ formula (dashed line) and from simulation (dashed line with diamonds). Both the n* and $\gamma$ formula curves include anharmonic terms.
\begin{figure}
\epsfxsize=10cm
\centerline{\epsffile{t.ahcal13.eps}}\end{figure}


next up previous contents
Next: Thermodynamic Properties Up: Anharmonicity Previous: The effect of transition Home: Return to my homepage
Jon Doye
8/27/1997