1 |
chuckv |
3483 |
%!TEX root = /Users/charles/Documents/chuckDissertation/dissertation.tex |
2 |
chuckv |
3496 |
\chapter{\label{chap:metalglass}COMPARING MODELS FOR DIFFUSION IN SUPERCOOLED LIQUIDS: THE EUTECTIC COMPOSITION OF THE Ag-Cu ALLOY} |
3 |
chuckv |
3483 |
|
4 |
|
|
Solid solutions of silver and copper near the eutectic point have been |
5 |
|
|
historical curiosities since Roman emperors used them to cut the |
6 |
|
|
silver content in the ubiquitous {\em denarius} coin by plating copper |
7 |
|
|
blanks with the eutectic mixture.\cite{Pense92} Curiosity in this |
8 |
|
|
alloy continues to the current day since the discovery that it forms a |
9 |
|
|
glassy material when rapidly quenched from the melt. The eutectic |
10 |
|
|
composition (60.1\% Ag, 39.9\% Cu) was the first glassy metal reported |
11 |
|
|
by Duwez in 1960.\cite{duwez:1136} The Duwez experiments give us an upper |
12 |
|
|
bound on the cooling rate required ($\approx 10^{6} K/s$) in this |
13 |
|
|
alloy, while other metallic glasses (based on Ti-Zr alloys) can be |
14 |
|
|
formed with cooling rates of only $1 K/s$.\cite{Peker93} There is |
15 |
|
|
clearly a great wealth of information waiting to be discovered in the |
16 |
|
|
vitrification and crystallization in liquid alloys. |
17 |
|
|
|
18 |
|
|
The Ag-Cu alloy is also of great theoretical interest because it |
19 |
|
|
resembles a model glass-forming Lennard-Jones system which has |
20 |
|
|
correlation functions that decay according to the the famous |
21 |
|
|
Kohlrausch-Williams-Watts ({\sc kww}) law, |
22 |
|
|
\begin{equation} |
23 |
chuckv |
3496 |
C(t) \approx A \exp\left[-(t/\tau)^{\beta}\right], |
24 |
chuckv |
3483 |
\label{eq:kww2} |
25 |
|
|
\end{equation} |
26 |
chuckv |
3496 |
where $\tau$ is the characteristic relaxation time for the system. |
27 |
|
|
|
28 |
chuckv |
3483 |
Kob and Andersen observed stretched exponential decay of the van Hove |
29 |
|
|
correlation function in a system comprising an 80-20 mixture of |
30 |
|
|
particles with different well depths $(\epsilon_{AA} \neq |
31 |
|
|
\epsilon_{BB})$ and different range parameters $(\sigma_{AA} \neq |
32 |
|
|
\sigma_{BB})$.\cite{Kob95a,Kob95b} They found that at low |
33 |
|
|
temperatures, the best fitting value for the stretching parameter |
34 |
|
|
$\beta$ was approximately $0.8$. The same system was later |
35 |
|
|
investigated by Sastry, Debenedetti and Stillinger.\cite{Stillinger98} |
36 |
|
|
They observed nearly identical stretching parameters in the time |
37 |
|
|
dependence of the self intermediate scattering function |
38 |
|
|
$F_{s}(k,t)$.\cite{Hansen86} |
39 |
|
|
|
40 |
|
|
In single component Lennard-Jones systems, the stretching parameters |
41 |
chuckv |
3496 |
appear to be somewhat lower. Angelani {\em et al.}\cite{Angelani98} have reported |
42 |
chuckv |
3483 |
$\beta \approx 1/2$ for relatively low temperature Lennard-Jones |
43 |
|
|
clusters, and Rabani {\em et al.} have reported similar values for the |
44 |
chuckv |
3496 |
decay of correlation functions in defective Lennard-Jones crystals.\cite{Rabani99} |
45 |
chuckv |
3483 |
|
46 |
|
|
There have been a few recent studies of amorphous metals using fairly |
47 |
|
|
realistic potentials and molecular dynamics methodologies. Gaukel and |
48 |
|
|
Schober studied diffusion in |
49 |
|
|
$\mbox{Zr}_{67}\mbox{Cu}_{33}$,\cite{Gaukel98} and Qi {\em et al.} |
50 |
|
|
have looked at the static properties ($g(r)$, $s({\bf k})$, etc.) for |
51 |
|
|
alloys of Cu with both Ag and Ni.\cite{Qi99} None of the metallic |
52 |
|
|
liquid simulation studies have observed time correlation functions in |
53 |
|
|
the super-cooled liquid near $T_{g}$. It would therefore be |
54 |
|
|
interesting to know whether these realistic models have the same |
55 |
|
|
anomalous time dependence as the Lennard-Jones systems, and if so, |
56 |
|
|
whether their stretching behavior is similar to that observed in the |
57 |
|
|
two-component Lennard-Jones systems. |
58 |
|
|
|
59 |
chuckv |
3496 |
Additionally, bi-metallic alloys present an ideal opportunity |
60 |
|
|
to apply the cage-correlation function methodology which was |
61 |
chuckv |
3483 |
developed to study the hopping rate in supercooled |
62 |
|
|
liquids.\cite{Rabani97,Gezelter99,Rabani99,Rabani2000} In particular, |
63 |
chuckv |
3496 |
it will be used to test two models for diffusion, both of which use |
64 |
chuckv |
3483 |
hopping times which are easily calculated by observing the long-time |
65 |
|
|
decay of the cage correlation function. The two models are Zwanzig's |
66 |
|
|
model,\cite{Zwanzig83} which is based on the periodic interruption of |
67 |
|
|
harmonic motions around inherent structures, and the Continuous Time |
68 |
|
|
Random Walk ({\sc ctrw}) |
69 |
|
|
model,\cite{Blumen83,Klafter94,Klafter96,Shlesinger99} which can be |
70 |
|
|
used to derive transport properties from a random walk on a regular |
71 |
|
|
lattice where the time between jumps is non-uniform. |
72 |
|
|
|
73 |
chuckv |
3496 |
\section{Theory} |
74 |
chuckv |
3483 |
|
75 |
chuckv |
3496 |
In this section a brief introduction will be given to the models for |
76 |
chuckv |
3483 |
diffusion that we will be comparing, as well as a brief description of |
77 |
|
|
how one can use the the cage-correlation function to obtain hopping |
78 |
|
|
rates in liquids. |
79 |
|
|
|
80 |
chuckv |
3496 |
\subsection{Zwanzig's Model} |
81 |
chuckv |
3483 |
In his 1983 paper~\cite{Zwanzig83} on self-diffusion in liquids, |
82 |
|
|
Zwanzig proposed a model for diffusion which consisted of ``cells'' or |
83 |
|
|
basins in which the liquid's configuration oscillates until it |
84 |
|
|
suddenly finds a saddle point on the potential energy surface and |
85 |
|
|
jumps to another basin. This model was based on and supported by |
86 |
|
|
simulations done by Stillinger and Weber |
87 |
|
|
\cite{Stillinger82,Stillinger83,Weber84,Stillinger85} in which the |
88 |
|
|
liquid configurations generated by molecular dynamics were quenched |
89 |
|
|
periodically by following the steepest descent path to the nearest |
90 |
|
|
local minima on the potential energy surface. Stillinger and Weber |
91 |
|
|
found that as their simulations progressed, the quenched |
92 |
|
|
configurations were stable for short periods of time and then suddenly |
93 |
|
|
jumped (with some re-crossing) to other |
94 |
|
|
configurations.\cite{Stillinger83} |
95 |
|
|
|
96 |
|
|
The starting point in Zwanzig's theory is the Green-Kubo |
97 |
|
|
formula~\cite{Berne90,Hansen86} for the self-diffusion constant, |
98 |
|
|
\begin{equation} |
99 |
|
|
D=\frac{1}{3} \int_{0}^{\infty}dt \langle {\bf v}(t) \cdot {\bf v}(0) |
100 |
|
|
\rangle, |
101 |
|
|
\label{eq:kubo} |
102 |
|
|
\end{equation} |
103 |
|
|
where $\langle {\bf v}(t) \cdot {\bf v}(0) \rangle$ is the |
104 |
|
|
time-dependent velocity autocorrelation function. Zwanzig's model |
105 |
|
|
predicts the diffusion constant using $\tau$, the lifetime which |
106 |
|
|
characterizes the distribution ($\exp(-t/\tau)$) of residence times in |
107 |
|
|
the cells. |
108 |
|
|
|
109 |
|
|
Following a jump, the coherence of the harmonic oscillations is |
110 |
|
|
disrupted, so all correlations between velocities will be destroyed |
111 |
|
|
after each jump. Zwanzig writes the velocity autocorrelation function |
112 |
|
|
in terms of the velocities of the normal modes in the nearest cell. |
113 |
|
|
The normal mode frequencies are characterized by the normalized |
114 |
|
|
distribution function $\rho(\omega)$, and the time integral can be |
115 |
|
|
solved assuming the time dependence of a damped harmonic oscillator |
116 |
|
|
for each of the normal modes. In the continuum limit of normal mode |
117 |
|
|
frequencies, one obtains |
118 |
|
|
\begin{equation} |
119 |
|
|
D = \frac{kT}{M}\int_{0}^{\infty} d\omega \rho(\omega) |
120 |
|
|
\frac{\tau}{(1 + \omega^{2}\tau^{2})}, |
121 |
|
|
\label{eq:zwanz} |
122 |
|
|
\end{equation} |
123 |
|
|
where $M$ is the mass of the particles. |
124 |
|
|
|
125 |
|
|
Zwanzig does not explicitly derive the inherent structure normal modes |
126 |
|
|
from the potential energy surface (he used the Debye spectrum for |
127 |
chuckv |
3496 |
$\rho(\omega)$.) Moreover, the theory avoids the |
128 |
chuckv |
3483 |
problem of how to estimate the lifetime $\tau$ for cell jumps that |
129 |
|
|
destroy the coherent oscillations in the sub-volume. Nevertheless, the |
130 |
|
|
model fits the experimental results quite well for the self-diffusion |
131 |
|
|
of tetramethylsilane (TMS) and benzene over large ranges in |
132 |
|
|
temperature.\cite{Parkhurst75a,Parkhurst75b} |
133 |
|
|
|
134 |
chuckv |
3496 |
\subsection{The {\sc ctrw} Model} |
135 |
chuckv |
3483 |
In the {\sc ctrw} |
136 |
|
|
model,\cite{Blumen83,Klafter94,Klafter96,Shlesinger99} random walks |
137 |
|
|
take place on a regular lattice but with a distribution of waiting |
138 |
|
|
times, |
139 |
|
|
\begin{equation} |
140 |
|
|
\psi(t) \approx \left\{ \begin{array}{ll} |
141 |
|
|
(t/\tau)^{-1-\gamma}, & 0 < \gamma < 1 \\ |
142 |
|
|
\frac{1}{\tau} e^{-t/\tau}, & \gamma = 1 |
143 |
|
|
\end{array} |
144 |
|
|
\right., |
145 |
|
|
\label{eq:ctrw} |
146 |
|
|
\end{equation} |
147 |
|
|
between jumps. $\gamma=1$ represents normal transport, while $\gamma |
148 |
|
|
< 1$ represents transport in a ``fractal time'' regime. In this |
149 |
|
|
regime, the number of jumps grows in time $t$ as $t^\gamma$ and not |
150 |
|
|
linearly with time. This dependence implies that the jumps do not |
151 |
|
|
possess a well-defined mean waiting time, and that there is no way to |
152 |
|
|
define a hopping rate for systems that are operating in the fractal |
153 |
|
|
time regime. |
154 |
|
|
|
155 |
|
|
Klafter and Zumofen have derived probability distributions for |
156 |
|
|
transport in these systems.\cite{Klafter94} In the $\gamma=1$ limit, |
157 |
|
|
they obtain the standard diffusive behavior in which the diffusion |
158 |
|
|
constant is inversely proportional to $\tau$ and proportional to the |
159 |
|
|
square of the lattice spacing for the random walk ($\sigma_{0}$): |
160 |
|
|
\begin{equation} |
161 |
|
|
\label{eq:CTRW_diff} |
162 |
|
|
D = \frac {\sigma_{0}^{2}} {6\tau} |
163 |
|
|
\end{equation} |
164 |
chuckv |
3496 |
This behavior is also suggested by estimates of $\tau$ in |
165 |
chuckv |
3483 |
$\mbox{CS}_{2}$ and in Lennard-Jones systems using the cage |
166 |
|
|
correlation function.\cite{Gezelter99,Rabani99} |
167 |
|
|
|
168 |
|
|
When $\gamma < 1$, the situation is somewhat more complex. The |
169 |
|
|
mean-square displacement can be approximated at long times as |
170 |
|
|
\begin{equation} |
171 |
|
|
\label{eq:ctrw_msd} |
172 |
|
|
\langle r^{2}(t) \rangle = |
173 |
|
|
\frac{4 \sigma_{0}^{2}}{3 \sqrt{\pi}} |
174 |
|
|
\left(\frac{t}{\tau}\right)^{\gamma} |
175 |
|
|
\left(\frac{2-\gamma}{\gamma}\right)^{\gamma} |
176 |
|
|
\frac{1}{\left(2-\gamma\right)^{2}} \Gamma(\frac{7}{2} - \gamma). |
177 |
|
|
\end{equation} |
178 |
|
|
Note that there is no well-defined diffusion constant for transport |
179 |
|
|
that behaves according to this expression. |
180 |
|
|
|
181 |
|
|
The waiting time distribution in Eq. (\ref{eq:ctrw}) can be used to |
182 |
|
|
derive a sticking probability, |
183 |
|
|
\begin{equation} |
184 |
|
|
\label{eq:ctrw_phi} |
185 |
|
|
\Phi(t) \approx 1 - \int_{0}^{t} dt' \psi(t'), |
186 |
|
|
\end{equation} |
187 |
|
|
which is the probability of not having made a jump until time t. The |
188 |
|
|
long-time behavior of this function should be directly related to the |
189 |
|
|
long-time behavior of the cage correlation function, which is |
190 |
|
|
essentially a measurement of the fraction of atoms that are still |
191 |
|
|
within their initial surroundings. For the $\gamma < 1$ case, the |
192 |
|
|
distribution of waiting times in Eq. (\ref{eq:ctrw}) results in |
193 |
|
|
sticking probability that decays as $t^{-\gamma}$. This is a very |
194 |
|
|
different behavior than the {\sc kww} law (Eq. (\ref{eq:kww2}) that |
195 |
|
|
has been observed in the defective Lennard-Jones |
196 |
|
|
crystals.\cite{Rabani99} |
197 |
|
|
|
198 |
|
|
Ngai and Liu have asserted that the {\sc kww} decay law |
199 |
|
|
(Eq. (\ref{eq:kww2}) cannot be explained through the use of the |
200 |
|
|
waiting time distribution in Eq. (\ref{eq:ctrw}).\cite{Ngai81} |
201 |
|
|
Instead, they propose a distribution, |
202 |
|
|
\begin{equation} |
203 |
|
|
\psi(t) = c \alpha t^{\alpha -1} e^{-c t^\alpha} |
204 |
|
|
\label{eq:ngai} |
205 |
|
|
\end{equation} |
206 |
|
|
which can explain the experimentally-observed decay. The form of the |
207 |
|
|
waiting time distribution in Eq. (\ref{eq:ctrw}) is more amenable to |
208 |
|
|
analytical approaches to calculating transport behavior. Blumen, {\it |
209 |
|
|
et al.} have investigated both proposed functional forms for $\psi(t)$ |
210 |
|
|
for recombination phenomena,\cite{Blumen83} and found that |
211 |
|
|
Eq. (\ref{eq:ngai}) leads to time-independent rates at longer times. |
212 |
|
|
They conclude that in order to observe anomalous transport, waiting |
213 |
|
|
time distributions with pathological long-time tails are required. |
214 |
|
|
|
215 |
chuckv |
3496 |
\subsection{The Cage Correlation Function} |
216 |
chuckv |
3483 |
|
217 |
|
|
In a recent series of |
218 |
|
|
papers,\cite{Gezelter97,Rabani97,Gezelter98a,Gezelter99,Rabani99,Rabani2000} |
219 |
chuckv |
3496 |
Gezelter {\it et al.} have investigated approaches to calculating the hopping |
220 |
chuckv |
3483 |
rate ($k_{h} = 1/\tau$) in liquids, supercooled liquids and defective |
221 |
chuckv |
3496 |
crystals. To obtain this estimate, they introduced the {\it cage |
222 |
chuckv |
3483 |
correlation function} which measures the rate of change of atomic |
223 |
|
|
surroundings, and associate the long-time decay of this function with |
224 |
|
|
the basin hopping rate for diffusion. The details on calculating the |
225 |
chuckv |
3496 |
cage correlation function can be found in Refs. \cite{Rabani97} and |
226 |
|
|
\cite{Gezelter99}, but the concept will be briefly reviewed here. |
227 |
chuckv |
3483 |
|
228 |
|
|
An atom's immediate surroundings are best described by the list of |
229 |
|
|
other atoms in the liquid that make up the first solvation shell. |
230 |
|
|
When a diffusive barrier crossing involving the atom has occurred, the |
231 |
|
|
atom has left its immediate surroundings. Following the barrier |
232 |
|
|
crossing, it will have a slightly different group of atoms surrounding |
233 |
|
|
it. If one were able to paint identifying numbers on each of the |
234 |
|
|
atoms in a simulation, and kept track of the list of numbers that each |
235 |
|
|
atom could see at any time, then the barrier crossing event would be |
236 |
|
|
evident as a substantial change in this list of neighbors. |
237 |
|
|
|
238 |
|
|
The cage correlation function uses a generalized neighbor list to keep |
239 |
|
|
track of each atom's neighbors. If the list of an atom's neighbors at |
240 |
|
|
time $t$ is identical to the list of neighbors at time $0$, the cage |
241 |
|
|
correlation function has a value of $1$ for that atom. If any of the |
242 |
|
|
original neighbors are {\em missing} at time $t$, it is assumed that |
243 |
|
|
the atom participated in a hopping event, and the cage correlation |
244 |
|
|
function is $0$. The atom's surroundings can also change due to |
245 |
|
|
vibrational motion, but at longer times, the cage will reconstitute |
246 |
|
|
itself to include the original members. Only those events which |
247 |
|
|
result in irreversible changes to the surroundings will cause the cage |
248 |
|
|
to de-correlate at long times. The mathematical formulation of the |
249 |
chuckv |
3496 |
cage correlation function is given in Refs. \cite{Rabani97} and |
250 |
|
|
\cite{Gezelter99}. |
251 |
chuckv |
3483 |
|
252 |
|
|
Averaging over all atoms in the simulation, and studying the decay of |
253 |
|
|
the cage correlation function gives us a way to measure the hopping |
254 |
chuckv |
3496 |
rates directly from relatively short simulations. Gezelter {\it et al.} have used the |
255 |
chuckv |
3483 |
cage correlation function to predict the hopping rates in |
256 |
|
|
atomic~\cite{Rabani97} and molecular~\cite{Gezelter99} liquids, as |
257 |
|
|
well as in defective crystals.\cite{Rabani99,Rabani2000} In the |
258 |
chuckv |
3496 |
defective crystals, they found that the cage correlation function, after |
259 |
chuckv |
3483 |
being corrected for the initial vibrational behavior at short times, |
260 |
|
|
decayed according to the {\sc kww} law (Eq. (\ref{eq:kww2})) with a |
261 |
|
|
stretching parameter $\beta \approx 1/2$. Angelani {\em et al.} have |
262 |
|
|
also reported $\beta \approx 1/2$ for relatively low temperature |
263 |
|
|
Lennard-Jones clusters. This is notably different behavior from the |
264 |
|
|
correlation functions calculated for the Lennard-Jones mixtures that |
265 |
|
|
have been studied by Kob and Andersen,\cite{Kob95a,Kob95b} and Sastry |
266 |
|
|
{\em et al.}\cite{Stillinger98} In this system, the stretching |
267 |
|
|
parameter was closer to 0.8. |
268 |
|
|
|
269 |
|
|
This paper will concentrate on the use of the cage correlation |
270 |
|
|
function to obtain hopping times for use by the Zwanzig and {\sc ctrw} |
271 |
|
|
models for diffusion in the $\mbox{Ag}_{6}\mbox{Cu}_{4}$ melt. We are |
272 |
|
|
also looking for the presence of anomalous dynamical behavior in the |
273 |
|
|
supercooled liquid at temperatures just above the glass transition to |
274 |
|
|
confirm the behavior observed in the simpler Lennard-Jones system. |
275 |
|
|
Section \ref{metglass:sec:details} will outline the computational methods used |
276 |
|
|
to perform the simulations. Section \ref{metglass:sec:results} contains our |
277 |
|
|
results, and section \ref{metglass:sec:discuss} concludes. |
278 |
|
|
|
279 |
chuckv |
3496 |
\section{Computational Details} |
280 |
chuckv |
3483 |
\label{metglass:sec:details} |
281 |
|
|
|
282 |
chuckv |
3496 |
The Sutton-Chen potential with the same |
283 |
|
|
parameterization as Qi {\it et al.}\cite{Qi99} has been chosen for this set of simulations. Details of this potential can be found in section \ref{introSec:tightbind}. This model was chosen over {\sc EAM} because of better experimental agreement with the heat of solution for Ag and Cu alloys and for better prediction of melting point and liquid state properties. This particular parametarization has been shown to reproduce the experimentally available heat of mixing data for both FCC solid solutions of Ag-Cu and the high-temperature liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does not reproduce the experimentally observed heat of mixing for the liquid alloy.\cite{MURRAY:1984lr} |
284 |
chuckv |
3483 |
|
285 |
|
|
In order to study the long-time portions of the correlation functions |
286 |
|
|
in this system without interference from the simulation methodology, |
287 |
chuckv |
3496 |
a set of molecular dynamics simulations in the constant-NVE |
288 |
|
|
ensemble was conducted. The density of the system was taken to be $8.742 |
289 |
chuckv |
3483 |
\mbox{g}/\mbox{cm}^3$. This density was chosen immediately to the |
290 |
|
|
liquid side of the melting transition from the constant thermodynamic |
291 |
chuckv |
3496 |
tension simulations of Qi {\it et al.}.\cite{Qi99} Their simulations gave |
292 |
chuckv |
3483 |
excellent estimates of phase and structural behavior, and should be |
293 |
|
|
seen as a starting point for investigations of these materials. |
294 |
|
|
|
295 |
|
|
The equations of motion were integrated using the velocity Verlet |
296 |
|
|
integrator with a timestep of 1 fs. A cutoff radius was used in |
297 |
|
|
which $r^{cut}_{ij}=2\alpha_{ij}$. Nine independent configurations of |
298 |
|
|
500 atoms were generated by starting from an FCC lattice and randomly |
299 |
|
|
choosing the identities of the particles to match the |
300 |
|
|
$\mbox{Ag}_{6}\mbox{Cu}_{4}$ composition which is near the eutectic |
301 |
|
|
composition for this alloy.\cite{Banhart:1992sv} |
302 |
|
|
|
303 |
|
|
Each configuration was started at a temperature of 1350 K (well in |
304 |
|
|
excess of the melting temperature at this density) and then cooled in |
305 |
|
|
50 K increments to 400 K. At each temperature increment, the systems |
306 |
|
|
re-sampled velocities from a Maxwell-Boltzmann distribution every ps |
307 |
|
|
for 20 ps, then were allowed to equilibrate for 50 ps. Following the |
308 |
chuckv |
3496 |
equilibration period, particle positions and velocities were collected |
309 |
chuckv |
3483 |
every ps for 250 ps. The lower temperature runs (375 K, 350 K, 325 K, |
310 |
|
|
and 300 K) were sampled for 500 ps, 1 ns, 1 ns and 7 ns respectively |
311 |
|
|
to accumulate more accurate long-time statistics. Cooling in this |
312 |
|
|
manner leads to a effective quenching rate of approximately 10$^{11}$ |
313 |
|
|
K/s. This quenching rate is much larger then those that are achieved |
314 |
|
|
through experimental methods which typically are on the order of |
315 |
|
|
10$^{6}$ or 10$^{7}$ K/s depending on the quenching method.\cite{duwez:1136} |
316 |
|
|
|
317 |
chuckv |
3496 |
\section{Results} |
318 |
chuckv |
3483 |
\label{metglass:sec:results} |
319 |
|
|
|
320 |
|
|
The simulation results were analyzed for several different structural |
321 |
|
|
and dynamical properties. Fig \ref{fig:gofr}. shows the pair |
322 |
|
|
correlation function, |
323 |
|
|
\begin{equation} |
324 |
|
|
\label{eq:gofr} |
325 |
|
|
g(r) = \frac {V} {N^2} \left \langle \sum _i \sum _{j\ne i}\delta(\mathbf |
326 |
|
|
{r-r}_{ij}) \right \rangle, |
327 |
|
|
\end{equation} |
328 |
|
|
at seven temperatures ranging from liquid to supercooled liquid. The |
329 |
|
|
appearance of a split second peak in the radial distribution function |
330 |
|
|
has been proposed as a signature of the glass transition.\cite{Nagel96} This |
331 |
|
|
behavior is evident in Fig. \ref{fig:gofr} for temperatures below 500 |
332 |
|
|
K. |
333 |
|
|
|
334 |
|
|
\begin{figure} |
335 |
|
|
|
336 |
|
|
\centering |
337 |
|
|
\includegraphics[height=3in]{images/gofr_all.pdf} |
338 |
|
|
\caption{Radial distribution functions for Ag$_6$Cu$_4$. The appearance |
339 |
|
|
of the split second peak at 500 K indicates the onset of a structural |
340 |
|
|
change in the super-cooled liquid at temperatures above $T_{g}$.} |
341 |
|
|
\label{fig:gofr} |
342 |
|
|
\end{figure} |
343 |
|
|
|
344 |
|
|
Wendt and Abraham have proposed another structural feature to estimate |
345 |
|
|
the location of the glass transition. Their measure, |
346 |
|
|
\begin{equation} |
347 |
|
|
\label{eq:wendt} |
348 |
|
|
R_{WA} = \frac {g_{min}} {g_{max}}, |
349 |
|
|
\end{equation} |
350 |
|
|
is the ratio of the magnitude of the first minimum, $g_{min}$ to |
351 |
|
|
the first maximum, $g_{max}$ in the radial distribution function. |
352 |
|
|
According to their estimates, when the value of $R_{WA}$ reaches 0.14, |
353 |
chuckv |
3496 |
the system has passed through the glass transition.\cite{Wendt78} A $T_{g}^{WA}$ of 547 K was observed given a cooling rate of $1.56 \times |
354 |
|
|
10^{11}$ K/s. Goddard, {\it et al.} observed a $T_g^{WA}\approx 500 |
355 |
|
|
K$ for a cooling rate of $2\times 10^{12}$ K/s in constant temperature, |
356 |
chuckv |
3483 |
constant thermodynamic tension (TtN) simulations\cite{Qi99}. |
357 |
|
|
|
358 |
chuckv |
3496 |
It should be noted that the split second peak in $g(r)$ appears at $T_g^{WA}$ (Figure \ref{fig:Wendt_abraham}), a temperature for which the diffusion constant is still an appreciable |
359 |
|
|
fraction of it's value in the liquid phase. In addition, the |
360 |
chuckv |
3483 |
diffusive behavior continues at the lowest simulated temperature of |
361 |
|
|
300 K indicating that glass transition for our system lies below 300 |
362 |
|
|
K. Taking the operational definition of the glass transition to be the |
363 |
|
|
temperature at which the viscosity exceeds $10^{13}$ |
364 |
|
|
poise,\cite{Nagel96} it is obvious that neither of these structural |
365 |
|
|
factors should be considered a definitive marker of $T_{g}$. It has |
366 |
|
|
also been noted in previous papers that changes in these structural |
367 |
|
|
factors can occur in certain supercooled metallic liquids |
368 |
|
|
independently of glassy behavior.\cite{Lewis91,Liu92} However, both |
369 |
|
|
structural features seem to mark the onset of anomalous thermodynamic |
370 |
|
|
and dynamical behavior, and should at best be taken as evidence of the |
371 |
|
|
location of the mode coupling theory ({\sc mct}) critical temperature, |
372 |
|
|
$T_{c}$, and not $T_{g}$ itself. |
373 |
|
|
\begin{figure}[htbp] |
374 |
|
|
\centering |
375 |
|
|
\includegraphics[height=3in]{images/w_a.pdf} |
376 |
|
|
\caption{Wendt-Abraham parameter ($R_{WA}$) as a |
377 |
|
|
function of temperature. $T^{WA}_{g} \approx$ 540 K for a cooling rate |
378 |
|
|
of $1.56 \times 10^{11}$ K/s.} |
379 |
|
|
\label{fig:Wendt_abraham} |
380 |
|
|
\end{figure} |
381 |
|
|
|
382 |
|
|
In order to compute diffusion constants using Zwanzig's model |
383 |
chuckv |
3496 |
(Eq. (\ref{eq:zwanz})), one must obtain an estimate of the density of |
384 |
|
|
vibrational states ($\rho(\omega)$) of the liquid. $\rho(\omega)$ has been obtained in two different ways, first by quenching twenty |
385 |
chuckv |
3483 |
high-temperature (1350 K) configurations to the nearest local minimum |
386 |
|
|
on the potential energy surface. These structures correspond to |
387 |
|
|
inherent structures on the liquid state potential energy surface. |
388 |
|
|
Normal modes were then calculated for each of these inherent |
389 |
|
|
structures by diagonalizing the mass-weighted Hessian to give the |
390 |
|
|
squares of the normal mode frequencies. The second method we used for |
391 |
|
|
determining the vibrational density of states was to compute the |
392 |
|
|
normalized power spectrum of the velocity autocorrelation function |
393 |
|
|
\begin{equation} |
394 |
|
|
\label{eq:vcorr} |
395 |
|
|
\rho_0(\omega) = \int_{-\infty}^\infty \left \langle \mathbf{v}(t) \cdot |
396 |
|
|
\mathbf{v}(0) \right \rangle e^{-i \omega t} dt |
397 |
|
|
\end{equation} |
398 |
chuckv |
3496 |
for trajectories which hover just above the inherent structures. These are displayed in Figure \ref{fig:rho} |
399 |
|
|
To calculate the power spectrum, a small amount of kinetic energy (8 K) |
400 |
chuckv |
3483 |
was given to each of the twenty inherent structures and the system was |
401 |
|
|
allowed to equilibrate for 30 ps. After equilibration, the velocity |
402 |
|
|
autocorrelation functions were calculated from relatively short (30 |
403 |
|
|
ps) trajectories and the density of states was obtained from a simple |
404 |
|
|
discrete Fourier transform. At this low temperature, the particles do |
405 |
|
|
not diffuse, so the velocity correlation function doesn't decay due to |
406 |
|
|
hopping. This implies that the Fourier transform of $\langle |
407 |
|
|
\mathbf{v}(t) \cdot \mathbf{v}(0) \rangle$ is $0$ at $\omega=0$ |
408 |
|
|
|
409 |
|
|
\begin{figure}[htbp] |
410 |
|
|
\centering |
411 |
|
|
\includegraphics[height=3in]{images/rho.pdf} |
412 |
|
|
\caption{Density of states calculated from quenched normal modes, |
413 |
|
|
$\rho_q(\omega)$, and from the Fourier transform of the |
414 |
|
|
velocity autocorrelation function $\rho_o(\omega)$.} |
415 |
|
|
\label{fig:rho} |
416 |
|
|
\end{figure} |
417 |
|
|
|
418 |
|
|
|
419 |
|
|
|
420 |
|
|
|
421 |
|
|
Both of these methods result in similar gross features for the density |
422 |
|
|
of states, but there are some important differences in their estimates |
423 |
|
|
at low frequencies. Notably the normal mode density of states |
424 |
|
|
($\rho_{q}(\omega)$) is missing some of the low frequency modes at |
425 |
|
|
frequencies below $10 \mbox{cm}^{-1}$. The power spectrum |
426 |
|
|
($\rho_{0}(\omega)$) recovers these modes, but gives a much noisier |
427 |
|
|
estimate for the density of states. |
428 |
|
|
|
429 |
chuckv |
3496 |
It was determined that with a radial cutoff of $2 \alpha_{ij}$ the |
430 |
chuckv |
3483 |
potential (and forces) exhibited a large number of discontinuities |
431 |
|
|
which made the minimization into the inherent structures somewhat |
432 |
|
|
challenging. In this type of potential, the discontinuities at the |
433 |
|
|
cutoff radius {\em cannot} be fixed by shifting the potential as is |
434 |
|
|
done with the Lennard-Jones potential. This limitation is due to the |
435 |
|
|
non-additive nature of the density portion of the potential function. |
436 |
|
|
In order to address this problem, the minimizations and frequencies |
437 |
|
|
were obtained with a radial cutoff of $3 \alpha_{ij}$, which provided |
438 |
|
|
a surface of adequate smoothness. The larger cutoff radius requires a |
439 |
|
|
larger system size, and so for the minimizations and the calculation |
440 |
|
|
of the densities of states, we used a total of 1372 atoms. |
441 |
|
|
|
442 |
chuckv |
3496 |
It should be noted that one possible way to provide a surface without |
443 |
chuckv |
3483 |
discontinuities would be to use a shifted form of the density, |
444 |
|
|
\begin{equation} |
445 |
|
|
\rho_{i}=\sum_{j\neq |
446 |
|
|
i}\left(\frac{\alpha_{ij}}{r_{ij}}\right)^{m_{ij}} - |
447 |
|
|
\left(\frac{\alpha_{ij}}{r^{cut}_{ij}}\right)^{m_{ij}}. |
448 |
|
|
\end{equation} |
449 |
|
|
This modification substantially alters the attractive portion of the |
450 |
|
|
potential, and has a deleterious effect on the forces. A similar |
451 |
|
|
shifting in the Lennard-Jones or other pairwise-additive potential |
452 |
|
|
would not exhibit this problem. Although this modification does |
453 |
|
|
provide a much smoother potential, it would properly require a |
454 |
|
|
re-parameterization of $c_{i}$ and $D_{ii}$, tasks which are outside |
455 |
|
|
the purview of the current work. |
456 |
|
|
|
457 |
chuckv |
3496 |
\subsection{Diffusive Transport and Exponential Decay} |
458 |
chuckv |
3483 |
|
459 |
|
|
|
460 |
|
|
Translational diffusion constants were calculated via the Einstein |
461 |
|
|
relation\cite{Berne90} |
462 |
|
|
\begin{equation} |
463 |
|
|
\label{eq:einstein} |
464 |
|
|
D = \lim _{t \rightarrow \infty} \frac {1} {6t} \left \langle \vert |
465 |
|
|
\mathbf {r} _{i} (t) - \mathbf {r} _{i} (0) |
466 |
|
|
\vert ^2 \right \rangle |
467 |
|
|
\end{equation} |
468 |
|
|
using the slope of the long-time portion of the mean square |
469 |
|
|
displacement. We show an Arrhenius plot of the natural logarithm of |
470 |
|
|
the diffusion constant vs. $1/T$ in Fig. \ref{fig:arrhenius} |
471 |
|
|
|
472 |
|
|
The structural features that we noted previously (i.e. the |
473 |
|
|
split-second peak in $g(r)$ and $T_{g}^{WA}$) appear at the same |
474 |
|
|
temperature at which the log of the diffusion constant leaves the |
475 |
|
|
traditional straight-line Arrhenius plot. |
476 |
|
|
|
477 |
|
|
|
478 |
|
|
\begin{figure}[htbp] |
479 |
|
|
\centering |
480 |
|
|
\includegraphics[height=3in]{images/arrhenius.pdf} |
481 |
|
|
\caption{Arrhenius plot of the self-diffusion constants indicating |
482 |
|
|
significant deviation from Arrhenius behavior at temperatures below 450 K.} |
483 |
|
|
\label{fig:arrhenius} |
484 |
|
|
\end{figure} |
485 |
|
|
|
486 |
|
|
|
487 |
|
|
|
488 |
|
|
Truhlar and Kohen have suggested an interpretation of this type of |
489 |
|
|
plot based on Tolman's interpretation of the activation |
490 |
|
|
energy.\cite{Truhlar00,Tolman20,Tolman27} Specifically, positive |
491 |
|
|
convexity in Arrhenius plots requires the average energy of all |
492 |
|
|
diffusing particles rise more slowly then the average energy of all |
493 |
|
|
particles in the ensemble with increasing temperature. This only |
494 |
|
|
occurs if the microcanonical-ensemble diffusion constant, D(E), |
495 |
|
|
decreases with increasing energy. Truhlar and Kohen explain that one |
496 |
|
|
possible explanation of this behavior is if configuration space can be |
497 |
|
|
decomposed into two types of conformation, one which is ``reactive'' |
498 |
|
|
in which diffusive hopping occurs with a rate constant of $K_{R}(E)$, |
499 |
|
|
and one in which there is no diffusion. If the probability of being |
500 |
|
|
found in the reactive region is given by $P_{R}(E)$, then the overall |
501 |
|
|
diffusion constant, |
502 |
|
|
\begin{equation} |
503 |
|
|
\label{eq:truhlar} |
504 |
|
|
D(E) = P_{R}(E) K_{R}(E), |
505 |
|
|
\end{equation} |
506 |
|
|
would allow $P_{R}(E)$ to decrease with $E$ faster than $K_{R}(E)$ can |
507 |
|
|
increase. In other words, higher energy systems explore a larger |
508 |
|
|
segment of phase space and spend a smaller fraction of their time in |
509 |
|
|
regions where there is a probability of diffusive hopping. As the |
510 |
|
|
energy is decreased, more time is spent in regions of configuration |
511 |
|
|
space where diffusive hopping is allowed so $D(E)$ should increase |
512 |
|
|
with decreasing $E$. |
513 |
|
|
|
514 |
|
|
In Fig. \ref{fig:diffplot} we show the diffusion constants calculated |
515 |
|
|
via the Einstein relation Eq.(\ref{eq:einstein}) along with results |
516 |
|
|
for the Zwanzig and {\sc ctrw} ($\gamma=1$) models using hopping times |
517 |
|
|
obtained from simple linear fits to the long-time decay of the cage |
518 |
|
|
correlation function. For Zwanzig's model we show results using the |
519 |
|
|
two different estimates for the vibrational density of states. For |
520 |
|
|
the {\sc ctrw} model, we have used a jump length ($\sigma_0$) of |
521 |
|
|
1.016 {\AA} for {\em all} temperatures shown. |
522 |
|
|
|
523 |
|
|
\begin{figure}[htbp] |
524 |
|
|
\centering |
525 |
|
|
\includegraphics[height=3in]{images/diffplot.pdf} |
526 |
|
|
\caption{Self-Diffusion constant for the two models under |
527 |
|
|
consideration compared to the values computed via standard techniques.} |
528 |
|
|
\label{fig:diffplot} |
529 |
|
|
\end{figure} |
530 |
|
|
|
531 |
|
|
|
532 |
|
|
Zwanzig's model is extremely sensitive to the low-$\omega$ portion of |
533 |
|
|
$\rho(\omega)$,\cite{Gezelter99} so it is no surprise that the choice |
534 |
|
|
of the density of states gives a large variation in the predicted |
535 |
|
|
results. The agreement with the diffusion constants is better at lower |
536 |
chuckv |
3496 |
temperatures, but an obviously incorrect temperature |
537 |
|
|
dependence is observed in the higher temperature liquid regime. |
538 |
chuckv |
3483 |
|
539 |
|
|
The {\sc ctrw} model with $\gamma=1$ and an assumption of fixed jump |
540 |
|
|
distances gives much better agreement in the liquid regime, and the |
541 |
|
|
trend with changing temperature seems to be in excellent agreement |
542 |
|
|
with the Einstein relation. |
543 |
|
|
|
544 |
chuckv |
3496 |
Delving a bit more deeply into the {\sc ctrw} predictions, it can be |
545 |
|
|
assumed that the distribution of hopping times is well behaved |
546 |
chuckv |
3483 |
(i.e. $\gamma=1$) but that the jump {\it distance} is temperature |
547 |
|
|
dependent. To obtain estimates of the jump distance as a function of |
548 |
|
|
temperature, $\sigma_0(T)$, we can invert (Eq.(\ref{eq:CTRW_diff})) by |
549 |
|
|
multiplying the cage-correlation hopping time by the self-diffusion |
550 |
chuckv |
3496 |
constant. The temperature-dependent hopping distances is shown in |
551 |
|
|
Fig. \ref{fig:r0_ctrw}. Note that these assumptions would lead one to |
552 |
chuckv |
3483 |
believe that the average jump distance is increasing sharply as one |
553 |
|
|
approaches the glass transition, which could be an indicator of motion |
554 |
|
|
dominated by Levy flights.\cite{Klafter96,Shlesinger99} |
555 |
|
|
|
556 |
|
|
\begin{figure}[htbp] |
557 |
|
|
\centering |
558 |
|
|
\includegraphics[height=3in]{images/sig0_ctrw.pdf} |
559 |
|
|
\caption{{\sc ctrw} hopping distance, $\sigma_0(T)$, as a function of |
560 |
|
|
temperature assuming $\gamma=1$ and using the cage correlation hopping |
561 |
|
|
times.} |
562 |
|
|
\label{fig:r0_ctrw} |
563 |
|
|
\end{figure} |
564 |
|
|
|
565 |
|
|
|
566 |
chuckv |
3496 |
\subsection{Non-Diffusive Transport and Non-Exponential Decay} |
567 |
chuckv |
3483 |
|
568 |
|
|
|
569 |
|
|
A much more realistic scenario is that the distribution of hopping |
570 |
|
|
{\it times} changes while the jump distance remains the same at all |
571 |
|
|
temperatures. In Figs. \ref{fig:exponent} and |
572 |
chuckv |
3496 |
\ref{fig:tauhop}, the effects of relaxing the linear fits of |
573 |
chuckv |
3483 |
$\langle r^{2}(t) \rangle$ and of the long time portion of |
574 |
chuckv |
3496 |
$ln[C_{cage}(t)]$ are shown. To fit the mean square displacements, a |
575 |
chuckv |
3483 |
weighted non-linear least-squared fits using the {\sc ctrw} ($\gamma < |
576 |
|
|
1$) expression for the mean square displacement |
577 |
chuckv |
3496 |
(Eq. (\ref{eq:ctrw_msd})) was performed. The only free parameter in the fit is the |
578 |
|
|
jump distance, which is fixed at a value of 1.016 \AA, the optimal |
579 |
|
|
jump distance for the $\gamma=1$ case. These fits allow estimates of $\gamma$ and the hopping times $\tau$ to be obtained directly from the |
580 |
chuckv |
3483 |
mean square displacements. |
581 |
|
|
|
582 |
|
|
\begin{figure}[htbp] |
583 |
|
|
\centering |
584 |
|
|
\includegraphics[height=3in]{images/exponent.pdf} |
585 |
|
|
\caption{Comparison of exponential stretching coefficients, $\beta$ |
586 |
|
|
from the cage-correlation function and $\gamma$ from CTRW theory.} |
587 |
|
|
\label{fig:exponent} |
588 |
|
|
\end{figure} |
589 |
|
|
|
590 |
|
|
|
591 |
|
|
|
592 |
|
|
The cage correlation functions were fit to the {\sc kww} law |
593 |
|
|
(Eq. (\ref{eq:kww2})), after correcting for the short-time (i.e. $< 1$ |
594 |
|
|
ps) vibrational decay. The values of $\gamma$ from the $\langle |
595 |
|
|
r^{2}(t) \rangle$ fits are shown with the {\sc kww} stretching |
596 |
|
|
parameters in Fig. \ref{fig:exponent}. Note that fits for $\gamma$ |
597 |
|
|
and the {\sc kww} stretching parameters both begin to show deviation |
598 |
|
|
from their more normal high-temperature behavior at approximately the |
599 |
|
|
same temperature as the appearance of the structural features |
600 |
|
|
($T_{g}^{WA}$, and the split second peak in $g(r)$) noted above. |
601 |
|
|
|
602 |
|
|
There is a small discrepancy between the stretching parameters |
603 |
|
|
($\beta$) for the {\sc kww} fits and the values of $\gamma$ for the |
604 |
|
|
{\sc ctrw} fits to the time dependence of the mean-square |
605 |
|
|
displacement. Although there is not enough data to make a firm |
606 |
|
|
conclusion, it would appear that there is a small constant offset, |
607 |
|
|
\begin{equation} |
608 |
|
|
\beta \approx \gamma - a |
609 |
|
|
\label{eq:exprel} |
610 |
|
|
\end{equation} |
611 |
|
|
with $a \approx 0.06$ for this system. This could be due in part to |
612 |
|
|
the sensitivity of the cage correlation function to intermediate |
613 |
|
|
timescales (i.e. 10-30 ps). Observations at these times will be |
614 |
|
|
susceptible to the effects of short-lived inhomogeneities, while the |
615 |
|
|
{\sc ctrw} fits (done over timescales from 10 ps - 1 ns) will show |
616 |
|
|
only the effects of inhomogeneities which persist for longer times. |
617 |
|
|
Since inhomogeneity can lead to the stretching behavior,\cite{Nagel96} |
618 |
|
|
the values of the stretching parameter obtained from the longer time |
619 |
|
|
fits are the most relevant for understanding the process of |
620 |
|
|
vitrification. |
621 |
|
|
|
622 |
|
|
\begin{figure}[htbp] |
623 |
|
|
\centering |
624 |
|
|
\includegraphics[height=3in]{images/tauhop.pdf} |
625 |
|
|
\caption{The characteristic hopping time, $\tau_{hop}$, which |
626 |
|
|
characterizes the waiting time distribution. The solid circles |
627 |
|
|
represent hopping times predicted from {\sc kww} fits to the |
628 |
|
|
cage-correlation function. The open triangles represent the |
629 |
|
|
characteristic time calculated via the {\sc ctrw} model for $\langle |
630 |
|
|
r^{2}(t) \rangle$.} |
631 |
|
|
\label{fig:tauhop} |
632 |
|
|
\end{figure} |
633 |
|
|
|
634 |
chuckv |
3496 |
In Fig. \ref{fig:tauhop}, the hopping times for both |
635 |
|
|
models are shown. As expected, the hopping times diverge as the temperature is |
636 |
chuckv |
3483 |
lowered closer to the glass transition. There is relatively good |
637 |
|
|
agreement between the hopping times calculated via {\sc ctrw} approach |
638 |
|
|
with those calculated via the cage correlation function, although the |
639 |
|
|
error is as much as a factor of 3 discrepancy at the lowest |
640 |
|
|
temperatures. |
641 |
|
|
|
642 |
chuckv |
3496 |
\section{Discussion} |
643 |
chuckv |
3483 |
\label{metglass:sec:discuss} |
644 |
|
|
|
645 |
|
|
It is relatively clear from using the cage correlation function to |
646 |
|
|
obtain hopping times that the {\sc ctrw} approach to dispersive |
647 |
|
|
transport gives substantially better agreement than Zwanzig's model |
648 |
|
|
which is based on interrupted harmonic motion in the inherent |
649 |
|
|
structures. The missing piece of the {\sc ctrw} theory is the average |
650 |
|
|
hopping distance $\sigma_{0}$. A method for calculating $\sigma_{0}$ |
651 |
|
|
from short trajectories would make it much more useful for the |
652 |
|
|
calculation of diffusion constants in the $\gamma=1$ limit. |
653 |
|
|
|
654 |
|
|
In the low-temperature supercooled liquid, there are substantial |
655 |
|
|
deviations from the $\gamma=1$ limit of the theory at temperatures |
656 |
|
|
below 500 K. Below this temperature, the mean square displacement |
657 |
|
|
cannot be fit well with a linear function in time, and the cage |
658 |
|
|
correlation function no longer has a simple exponential decay. In |
659 |
|
|
this paper, we have attempted to use the waiting time distribution due |
660 |
|
|
to Klafter {\it et |
661 |
|
|
al.}\cite{Blumen83,Klafter94,Klafter96,Shlesinger99} The Laplace |
662 |
|
|
transform of Eq. (\ref{eq:ctrw}) lends itself to the derivation of an |
663 |
|
|
analytical form for the propagator for dispersive |
664 |
|
|
transport.\cite{Klafter94} However, it is somewhat troubling that the |
665 |
|
|
cage correlation function (which has proven itself to be a good |
666 |
|
|
indicator of the average decay of atoms from their initial sites) |
667 |
|
|
cannot be fit with a sticking probability that is commensurate with |
668 |
|
|
Eq. (\ref{eq:ctrw}). |
669 |
|
|
|
670 |
chuckv |
3496 |
Instead, the long-time decay of the cage |
671 |
|
|
correlation function has been fit with the more familiar Kohlrausch-Williams-Watts |
672 |
chuckv |
3483 |
law (Eq. (\ref{eq:kww2})), which appears to be a more accurate model |
673 |
|
|
for the sticking probability. This point has been addressed by Ngai |
674 |
|
|
and Liu.\cite{Ngai81} It is of some interest that there is a |
675 |
|
|
relatively simple relationship between the stretching parameter in the |
676 |
|
|
fit to the cage correlation function and the value of $\gamma$ from |
677 |
|
|
the {\sc ctrw} model. One hopes that there is some fundamental |
678 |
|
|
relationship waiting to be discovered between {\sc kww}-like decay and |
679 |
|
|
a {\sc ctrw}-analysis of the propagator for dispersive transport. |