ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/tags/start/chuckDissertation/metallicglass.tex
Revision: 3483
Committed: Tue Jan 13 14:39:50 2009 UTC (15 years, 8 months ago) by chuckv
Content type: application/x-tex
Original Path: trunk/chuckDissertation/metallicglass.tex
File size: 34535 byte(s)
Log Message:
Chuck's dissertation for PhD Jan 2009

File Contents

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