1: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: %%\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb,floatfix]{revtex4}
3: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: \usepackage{epsfig,bm, graphicx, latexsym}
5: \usepackage{amsmath,epsf,psfig}
6:
7: \begin{document}
8: \newcommand{\be}{\begin{equation}}
9: \newcommand{\ee}{\end{equation}}
10: \newcommand{\la}{\langle}
11: \newcommand{\ra}{\rangle}
12:
13: \title{Theory of Aging in Structural Glasses}
14:
15: \author{Vassiliy Lubchenko} \altaffiliation[Current Address:
16: ]{Department of Chemistry, Massachusetts Institute of Technology,
17: Cambridge, MA 02139.} \author{Peter G. Wolynes}
18:
19: \affiliation{Department of Chemistry and Biochemistry and Department
20: of Physics, University of California, San Diego, 9500 Gilman Drive, La
21: Jolla, CA 92093-0371}
22:
23: \date{\today}
24:
25: \begin{abstract}
26:
27: The random first order transition theory of the dynamics of
28: supercooled liquids is extended to treat aging phenomena in
29: nonequilibrium structural glasses. A reformulation of the idea of
30: ``entropic droplets'' in terms of libraries of local energy landscapes
31: is introduced which treats in a uniform way the supercooled liquid
32: (reproducing earlier results) and glassy regimes. The resulting
33: microscopic theory of aging makes contact with the
34: Nayaranaswamy-Moynihan-Tool nonlinear relaxation formalism and the
35: Hodge-Scherer extrapolation of the Adam-Gibbs formula, but deviations
36: from both approaches are predicted and shown to be consistent with
37: experiment. The nonlinearity of glassy relaxation is shown to
38: quantitatively correlate with liquid fragility. The residual
39: nonArrhenius temperature dependence of relaxation observed in quenched
40: glasses is explained. The broadening of relaxation spectra in the
41: nonequilibrium glass with decreasing temperature is quantitatively
42: predicted. The theory leads to the prediction of spatially
43: fluctuating fictive temperatures in the long-aged glassy state, which
44: have non-Gaussian statistics. This can give rise to ``ultra-slow''
45: relaxations in systems after deep quenches.
46:
47: \end{abstract}
48:
49: \maketitle
50:
51: \section{Introduction}
52:
53: The energy landscape metaphor has turned theorists towards viewing the
54: global geometry of the phase space of complex systems. When these
55: systems are mesoscopic in size, for example, small proteins
56: \cite{OLSW} or gas phase clusters \cite{Wales}, a more or less
57: complete mathematical formulation of the idea, capable of treating
58: kinetics and thermodynamics, can be made using statistical tools to
59: characterize minima and saddle points of the entire system. Yet for
60: macroscopic systems, most transitions rearrange particles only
61: locally. This essential aspect of the dynamics is brought home
62: forcefully by noting that a liter of liquid water will move from one
63: energy minimum to another in $10^{-39}$ sec. Such a short time scale
64: cannot be directly relevant to any laboratory measurement on this
65: system! The necessity for using a local description is clearly
66: recognized in the modern theory of supercooled liquids and glasses
67: which is based on the statistical mechanics of random first order
68: transitions \cite{KW_PRA}-\cite{LW_soft}. In the deeply supercooled
69: liquid regime this theory explains not only the phenomenological
70: features of the dynamics \cite{KTW} but quantitatively predicts, on a
71: microscopic basis, the size of cooperative lengths, the precise
72: non-Arrhenius behavior of typical relaxation times \cite{XW,LW_soft}
73: and the non-exponentiality of relaxation \cite{XWbeta}. A quantized
74: version of the theory explains the low temperature thermodynamics of
75: amorphous substances, usually interpreted in terms of two level
76: systems \cite{LW} and the more energetic Boson peak excitations
77: \cite{LW_BP}. The crucial manifestation of the locality concept in
78: this theory (which has many mean field, global, aspects) is the notion
79: that ``entropic droplets'' drive the large-scale activated notions in
80: glass forming liquids \cite{KW_PRB,KTW,XW} and give both liquids and
81: glasses an intrinsic ``dynamical mosaic'' structure.
82:
83: The main purpose of this paper is to describe the predictions of the
84: RFOT theory for the behavior of glasses that have fallen out of
85: equilibrium because of being rapidly cooled from the melt. Dynamics
86: does not cease in rapidly quenched liquids that become glasses; rather
87: motions persist but are frustratingly slow and hard to study
88: experimentally. This slow, far-from-equilibrium dynamics, called
89: ``aging'', is not only of fundamental interest for statistical
90: mechanics \cite{Bouchaud} but also is of great practical interest
91: since so many amorphous materials are used in everyday life for times
92: exceeding their preparation time \cite{Struik}. Sometimes even small
93: changes of properties upon aging are crucial to engineering
94: performance.
95:
96:
97: Like the quantum theory of the low temperature properties of glasses,
98: the description of the nonequilibrium aging regime requires the
99: explicit construction and study of the energy landscape of local
100: regions of the glass. This local energy landscape description turns
101: out to be a microcanonical ensemble reformulation of the entropic
102: droplet concept that was formulated originally in the canonical
103: ensemble. We have found this microcanonical description to be more
104: vivid and somewhat easier to communicate than the original canonical
105: formulation. After introducing this formulation we will show that it
106: indeed reproduces the results for the dynamics of equilibrated
107: supercooled liquids already obtained using the random first order
108: transition theory. More important, this local energy landscape theory
109: makes several very striking predictions about the aging regime and how
110: the aging dynamics in the nonequilibrium glass is related to the
111: kinetics and thermodynamics of the equilibrated liquid. The
112: predictions are consonant with all experiments known to us and make
113: contact with, but are formally distinct from, the phenomenological
114: approaches commonly used to describe aging in structural glasses
115: \cite{Tool,Narayanaswamy,Moynihan,Scherer,Hodge}. The present theory,
116: for example, makes a specific prediction of the so-called nonlinearity
117: parameter, $x_{NMT}$, in the Narayanaswamy-Moynihan-Tool, formalism
118: \cite{Tool,Narayanaswamy,Moynihan}. In addition, the degree of
119: nonexponentiality of relaxation characterized by a $\beta$ exponent is
120: predicted. $\beta$ turns out to be neither precisely fixed at $T_g$
121: nor does it precisely scale like the $\beta$ for an equilibrated
122: liquid with $T$, another assumption often made. The predicted
123: variation of $\beta$ with quench temperature is however modest until
124: very low temperatures are reached. Likewise the Adam-Gibbs expression
125: \cite{AdamGibbs} has been extended phenomenologically notably by Hodge
126: and Scherer into the aging regime below $T_g$ by assuming the
127: configurational entropy to be frozen at $T_g$. While above $T_g$, an
128: Adam-Gibbs like form of the temperature dependence of relaxation time
129: is found from the present theory, below $T_g$, the local energy
130: landscape approach gives a different expression for the relaxation
131: rate in an aging glass. Quantitatively this expression gives
132: relaxation rates close to the Hodge-Scherer-Adam-Gibbs (HSAG) latter
133: extrapolation but the local energy landscape theory predicts a
134: deviation from that formula. This deviation would be interpreted
135: within the HSAG framework as a quench temperature dependence of the
136: apparent configurational entropy. Such a deviation has been found by
137: Alegria et al. in their careful work on aging of polymers
138: \cite{Alegria}.
139:
140:
141: Our most explicit results are obtained for the idealized situation of
142: the aging initially found after a cooling history with a single rapid
143: quench of modest depth. We will also qualitatively discuss
144: modifications of the simple theory expected for very deep quenches.
145: We also discuss the behavior of systems that have significantly
146: relaxed in the quenched state. While, our analysis suggests that, to a
147: first approximation, introducing a single fictive temperature should
148: serve well to describe many quench histories, the local energy
149: landscape theory shows that using a single fictive temperature is not
150: exact. We suggest possible modifications of the usual aging kinematics
151: based on the present theory.
152:
153:
154: The organization of the paper is as follows: we first describe how the
155: local energy landscape view of entropic droplets can be visualized and
156: show how a library of local energy landscapes can be constructed. We
157: next show how to derive in this framework the (previously obtained)
158: Vogel-Fulcher behavior above $T_g$. We then present the results for
159: typical relaxation rate for rapidly quenched, aging glasses and
160: compare these predictions with experimental results. We then discuss
161: the predictions for the stretching exponent both above and below
162: $T_g$. Following this, the generalizations needed for very deeply
163: quenched glasses and glasses that have undergone significant
164: relaxation in the aging state are discussed. Finally we summarize our
165: theory and suggest some further experimental tests of it.
166:
167:
168: \section{The Local Energy Landscape Construction}
169:
170: The energy landscape language is usually applied to a small system
171: (protein or cluster). It is also used for a thermodynamically large
172: system described by mean field theory \cite{MPV,Coluzzi}. In the first
173: situation, the barriers between states are, of course, finite because
174: the system is finite, but barriers may be formally infinite for the
175: mean field system. The ``states'' for a cluster or a protein are often
176: taken to be the basins surrounding minima of the potential energy
177: \cite{inherent}. These are well defined and no transitions can occur,
178: classically, between them at absolute zero. The ``states'' of a mean
179: field system are tied to minima of free energy and again, owing to the
180: $O(N)$ barriers, no transition can occur between them (even at finite
181: $T$).
182:
183:
184: In a supercooled liquid the observed plateau in the time dependent
185: neutron scattering correlation function \cite{Mezei} shows that most
186: molecules spend a great deal of time vibrating about a given
187: location. At least deep in the supercooled regime (where the plateau
188: is well developed) we can, therefore, conceptually imagine
189: constructing an average location for any particle about which it
190: vibrates (for times less than the plateau). The three dimensional
191: structure based on these average locations will be quite close to a
192: potential energy minimum found by removing kinetic energy from the
193: system with infinite speed (``steepest descent'' to an ``inherent
194: structure''). A mean field theory of the glass transition can be
195: obtained by constructing a free energy functional dependent on the
196: (plateau time averaged) density and then noting that this functional
197: has minima for density patterns localized around such inherent
198: structures \cite{SW,SSW,DasguptaValls}. In this way a vibrational
199: component of the free energy can be defined and can be added to the
200: average energy of such a state to give a free energy which determines
201: the thermal probability of being in this state through the Boltzmann
202: law.
203:
204:
205: Free energy functionals only have such aperiodic minima below a
206: temperature $T_A$. This temperature has been shown (within a simple
207: approximation) to be equivalent to the mode coupling dynamical
208: transition temperature \cite{KW_PRA}. The transition has the character
209: of a spinodal for a first order phase transition \cite{KW_PRB}. Above
210: $T_A$ even simple vibrational motions will take the system from one
211: minimum to another, but below $T_A$, the plateau in the time dependent
212: structure factor indicates the persistence of such states and
213: therefore their relevance to dynamics. In fact, as a first
214: approximation, the persistence of these states allows them to act as a
215: ``basis set'' for describing supercooled liquid dynamics. Thus we say
216: a free energy landscape emerges at $T_A$. $T_A$ has been evaluated
217: both for simple liquids \cite{SW,SSW} and for models of network
218: forming liquids with repulsive force \cite{HW} from first principles.
219: The physical meaning of the temperature $T_A$ is simple. When
220: supecooled below its melting temperature $T_m$, the liquid is, of
221: course, in a metastable state as a whole, and therefore must be
222: metastable {\em locally}. $T_A$ can be conveniently interpreted, for
223: example, as a temperature below which every two molecules will spend a
224: particular number (say, 300) of vibrational periods together between
225: the first encounter and the final departure.
226:
227: As part of our programme to describe the microscopics of liquid
228: relaxations, we will show that transitions between the metastable
229: configurations of the liquid {\em as a whole} consists of transitions
230: between {\em local} metastable configurations. We will employ a
231: microcanonical procedure in which only local regions are considered
232: and will find that beyond a certain (relatively small) size $N^*$, the
233: thermal and relaxational properties of the liquid do not depend on the
234: size of the sample. Alternatively said, all liquid properties of
235: interest can be deduced by focusing on a liquid region of size $N^*$
236: and completely disregarding what the rest of the liquid is doing.
237: This is the essence of the {\em locality} of the liquid free energy
238: landscape. The microcanonical procedure is a necessary step in
239: establishing such locality; it is however rarely used, so let us first
240: train our intuition on the very familiar example of a harmonic
241: lattice. Imagine being inside an extremely large, cubic, and purely
242: harmonic lattice and being given the ability to do arbitrary
243: thermodynamic measurement locally. You are further assigned to
244: explain, within a formal model, those thermal measurements, but are
245: allowed to visually inspect and mechanically test the {\em bonds} with
246: an arbitrarily large, but {\em finite} region, limited by how fast you
247: can perform the inspection. Upon checking that all individual bonds
248: are truly harmonic and the lattice is cubic within a certain region
249: (of size $L$), you write down a simple hamiltonian but are left with
250: the issue of the conditions at the boundary - strictly speaking, the
251: assignement is undoable! Being an optimist, you say: let me {\em
252: assume} for today that the rest of the sample is totally rigid (fixed
253: boundary conditions) or does not exist at all (open bondary condition)
254: and diagonalize the resulting Hamiltonian; tomorrow I can expand my
255: horizon, repeat the procedure and see what happens. The first day
256: proves frustrating though, because the Hamiltonian has an energy gap
257: (proportional to $\pi/L$), while the thermal measurements clearly show
258: very low frequency excitations are present. Fortunately, since
259: (unbeknownst to you) the lattice was harmonic and periodic, your
260: consecutive inspections and diagonalizations will yield a smaller and
261: a smaller gap. Clearly, the landscape of a regular harmonic lattice is
262: non-local and the microcanonical procedure offers a definitive test of
263: non-locality. Consider yet another elementary notion: Suppose you have
264: a product of two Hilbert spaces $A \otimes B$ (possibly interacting
265: via $V_{A,B}$) with an energy $\epsilon_{i_A,j_B}$ assigned to
266: configurations $i_A$ and $j_B$ of the subspaces $A$ and $B$. Consider
267: the partition function $\sum_{i_A,j_B} e^{-\beta \epsilon_{i_A,j_B}} =
268: \sum_{i_A} e^{-\phi_{i_A}}$, where $e^{-\beta \phi_{i_A}} \equiv
269: \sum_{j_B} e^{-\beta \epsilon_{i_A,j_B}}$. If $A$ and $B$ are
270: independent: $V_{A,B}= 0 \Rightarrow \epsilon_{i_A,j_B} =
271: \epsilon_{i_A} + \epsilon_{j_B}$, - one may further simplify
272: $\phi_{i_A} = \epsilon_{i_A,j_B} - T S_B$. In either case,
273: $\phi_{i_A}$'s are {\em free} energies of the degree of freedom $B$,
274: but can still be (with advantage) regarded as an energy, as far as the
275: $A$ degree of freedom is concerned. For example, one is allowed to
276: build a {\em microcanonical} construction using $\phi_{i_A}$'s as the
277: label. In the following, we will use a $\phi_{i_A}$-like quantity to
278: describe {\em all} the degree of freedom in the supercooled (or
279: quenched) liquid in excess of the lowest energy crystalline structure
280: corresponding to this substance (at this temperature). (For a polymer
281: that does not form a crystal our results will still apply
282: qualitatively.) While the ``integrated out'' degrees of freedom $B$
283: are clearly related to the vibrations of the corresponding crystal
284: structure, they are easy to conceptualize only in the mean field
285: limit, where they are indeed harmonic vibrations (see
286: shortly). Otherwise, the ``$B$'' motions are strongly anharmonic.
287:
288: We are now ready construct a library of free energy minima for a very
289: large sample of a liquid (or glass). This spectrum is shown
290: schematically in the first column of Figure \ref{library} as a
291: function of the free energy of a state $i$, $\phi_i^{lib}$ which we
292: take as the sum of an energy $\epsilon_i$ and an entropic contribution
293: from vibrations within the basin $- T s_{vib,i}$.
294: \begin{figure}[tb]
295: \includegraphics[width=.95\columnwidth]{library.eps}
296: \caption{\label{library} In the upper panel on the left a global
297: configuration is shown, chosen out of a global energy landscape. A
298: region of $N=5$ particles in this configuration is rearranged in the
299: center illustration. The original particle positions are indicated
300: with dashed lines. A larger rearranged region involving $N=7$
301: particles is connected dynamically to these states and is shown on the
302: right. In the lower panel, the left most figure shows the huge density
303: of states that is possible initially. The density of states found in
304: the local library originating from a given initial state with 5
305: particles being allowed to move locally is shown in the second
306: diagram. These energies are generally higher than the original state
307: owing to the mismatch at the borders. The larger density of states
308: where 7 particles are allowed to move is shown in the right most part
309: of this panel. As the library grows in size, the states as a whole are
310: still found at higher energies but the width of the distribution
311: grows. Eventually with growing $N$, a state within thermal reach of
312: the initial state will be found. At this value of $N^*$ we expect a
313: region to be able to equilibrate.}
314: \end{figure}
315:
316:
317: Because the system is very large this global library of states has an
318: exceedingly dense spectrum, whose density dramatically increases with
319: free energy. Let us imagine the system is presently in one of these
320: basins with (free) energy $\phi_i^{lib}$. We can now construct a set
321: of {\em local} libraries of states. To do this, imagine mentally
322: cutting out a region around a location ${\bf R}$ containing $N$
323: particles, where $N$ is much less than the total number of particles
324: in the sample. Call this region $D({\bf R}, N)$. Next, freeze the
325: molecules outside this region but allow molecules within $D$ to
326: move. (Unlike in the cubic crystal analogy, a supecooled structure has
327: built-in stress, so using open boundary condition is not useful.)
328: With the frozen environment one could (by quenching the potential
329: energy or a free energy functional) find a new set of inherent
330: structures that involve only reconfiguring the particles in
331: $D$. Actually such a set of structures (apart from strains near the
332: surface of $D$) will locally resemble a subset of the original global
333: library. If $N$ is small, however, we expect the subset of the states
334: so sampled to be selected on average from a higher energy part of the
335: global spectrum than where the original $\phi_i^{lib}$ was
336: found. Essentially this is because the density of states rapidly
337: increases with $\phi$ and there is less freedom to readjust particles
338: in the constrained minimization problem because of the fixed
339: environment. In absolute terms, we expect the discrepancy in energy of
340: the mean energy of the states in any local library from the starting
341: state to increase with $N$ since it is the particles at the borders
342: that must be most strained.
343:
344: When $N$ is small the local library is sparse but it grows denser and
345: spreads out over a larger range of free energy as the size of the
346: region increases. The set of such libraries centered around ${\bf R}$
347: is shown also in Figure \ref{library} for $N=1$ up to a large value of
348: $N$. The number of states in the library at free energy $\phi_i$
349: determines the configurational entropy $\Omega_c(\phi_i^{lib}) =
350: e^{S_c(\phi_i^{lib})/k_B}$, where $k_B$ is the Boltzmann constant. The
351: competition between the average energy growth and the spreading of the
352: range of energies with size of the region means that there is a
353: characteristic size $N^*$ where a state will finally be found within
354: thermal energy of the starting state. $N^*$ turns out to be the size
355: of a dynamically correlated region in the liquid or glass. Since the
356: new configuration is within $k_B T$ of the starting state, a
357: transition of such a region from its original configuration to the new
358: state can occur with reasonable probability. The region after
359: reconfiguration will be characterized by a temperature $T$. Elementary
360: transitions in the liquid must leave most of the molecules near their
361: old locations - this is the dynamical essence of locality. Therefore
362: the local landscape libraries are locally connected to each other: in
363: order to re-arrange a large region, smaller regions located in the
364: same place must first be rearranged. The specific law connecting
365: states in neighboring libraries must obey detailed balance (at the
366: vibrational temperature $T$) but otherwise will depend specifically on
367: molecular details for each system. The additional activation energy,
368: for a downhill move in such a library should only be a few times the
369: energy needed to heat a particle to $T_A$, the dynamical transition
370: temperature.
371:
372:
373: Irrespective of the detailed motions allowed, the locality of the
374: dynamics, however, guarantees that the transitions from the initial
375: state to one of the thermally allowed states in the library for
376: $D({\bf R}, N^*)$ will nevertheless be slow: owing to the initial rise
377: of the average energy there will be a bottleneck in the probability
378: flux at an intermediate value of $N$, namely $N^\ddagger$. The
379: activation free energy for reaching this bottleneck state determines
380: the escape rate from the initial configuration by motions in the
381: vicinity of ${\bf R}$.
382:
383:
384: While the local libraries can be constructed explicitly given
385: sufficient computational resources, the RFOT theory suggests a useful
386: set of approximations to the statistics of these libraries when $N^*$
387: and $N^\ddagger$ are moderately large. We now describe these
388: approximations: We ascribe to each state $j$ of the library for a
389: region $D({\bf R}, N)$ a so-called bulk free energy
390: $\Phi_j^{bulk}({\bf R}, N)$. Naively this would be the sum of the
391: vibrational entropies and internal energies of each molecule in $D$
392: and the pair interactions between the molecules within $D$. (Rigor
393: here would require ensuring the appropriate finite ranges of
394: interactions and some precisely specified ways of partitioning
395: vibrational entropies: entanglement entropies of the interface are too
396: subtle for consideration at present). Note that in a system with a
397: globally correlated landscape, such as the harmonic lattice considered
398: earlier, a description in terms of local libraries will not provide
399: the full list of configurations available to the system as a whole.
400:
401:
402: In a similar way we would define a bulk free energy of the initial
403: state relevant to this region $\Phi_{in}^{bulk}({\bf R}, N)$. The
404: actual energy of the complete sample when the state $j$ is inserted in
405: region $D({\bf R}, N)$ will not just reflect the difference in the
406: bulk energies corresponding to the state $j$ and the initial state;
407: instead, it will be higher by an amount $\Gamma_{j,in}$:
408: \begin{equation}
409: \phi_j^{lib} - \phi_{in}^{lib} = \Phi_j^{bulk}({\bf R}, N) -
410: \Phi_{in}^{bulk}({\bf R}, N) + \Gamma_{j,in}.
411: \label{phi_j}
412: \end{equation}
413:
414: As mentioned earlier, the new local structure labelled by
415: $\phi_j^{lib}$, is likely to be higher in energy than the initial
416: configuration. Our construction thus suggests $\Gamma_{j,in}$ will
417: usually be positive and will at most scale with the interface area. We
418: will write $\Gamma_{j,in}=\gamma_{j,in} N^x$. As we shall see later
419: the estimate for the exponent $x=2/3$ in 3 dimensions, based solely on
420: the interface area, is probably naive near $T_g$.
421:
422:
423: We now wish to calculate the equilibration rate of the region $D({\bf
424: R}, N^*)$. If the environment remains frozen, this rate is the escape
425: rate from $N=0$ to $N=N^*$. The probability flux to increase $N$ falls
426: until the bottleneck value $N^\ddagger$ is reached. After
427: $N^\ddagger$, even though the average energy of states in the
428: libraries increases, the growth of the number of states with
429: increasing library size is sufficient so that a rapid path to a
430: thermally equilibrated state at size $N^*$ can be found. Let the
431: typical downhill microscopic rearrangment have a rate
432: $\tau^{-1}_{micro}$. This rate will only be weakly activated. The flux
433: to any state at a size $N$ smaller than $N^*$ will be
434: %\begin{widetext}
435: \begin{eqnarray}
436: k & = & \tau^{-1}_{micro} \int (d \phi_j^{lib}/c_\phi)
437: e^{S_c(\Phi_j^{bulk})/k_B} e^{-(\phi_j^{lib}-\phi_{in}^{lib})/k_B T}
438: \nonumber \\ & \simeq & \tau^{-1}_{micro}
439: e^{S_c(\Phi^{bulk}_{eq})/k_B} e^{-(\phi_{eq}-\phi_{in}^{lib})/k_B T},
440: \label{k}
441: \end{eqnarray}
442: %\end{widetext}
443: where $\phi_{eq}$ maximizes the integrand and $c_\phi$ is some
444: constant of units energy. The quantities $\phi_j^{lib}$ and
445: $\Phi_j^{bulk}$ are related through Eq.(\ref{phi_j}). This
446: maximization means $\phi_{eq}$ will be the internal free energy
447: characteristic of the system at the ambient (i.e. vibrational)
448: temperature $T$. That is, $\phi_{eq}$ is the sum of the energetic and
449: vibrational entropic contributions appropriate to equilibrium at
450: $T$. In essence, the equation above is nothing more than the rate of
451: escape through a transition state with non-zero entropy, hence a
452: structure resembling the canonical ensemble. Aside from a numerical
453: factor, the integration in Eq.(\ref{k}) is indeed a canonical
454: summation in terms of $\phi_j$, which should be regarded as the
455: non-meanfield analog of the free energy of the so called ``pure''
456: state. The concept of the pure state is well developed in the context
457: of frustrated mean-field spin systems \cite{MPV}. In the mean-field
458: limit, the pure states are separated by infinite barriers and thus the
459: vibrations around the metastable free energy minima are purely
460: harmonic. The quantity $\phi_j^{lib}$ is non-meanfield because it can
461: be defined only locally (after paying the ``price'' of the surface
462: energy $\Gamma_{j,in}$). It is the degrees of freedom due to
463: transitions between those (non-meanfield) ``pure'' states of the
464: liquid that give rise to the (measurable) configurational
465: entropy. Note also that owing to the intense and strongly anharmonic
466: motions in the liquid (at this finite $T$), a local liquid state
467: labelled by a particular value of $\phi$ is not a single inherent
468: structure \cite{inherent}, but rather is a weighted superposition of
469: many inherent structures, with more or less harmonic vibrations on
470: top.
471:
472: Note, $\phi_{eq}$ is a function of $T$ and $N$. Since
473: $\Phi^{bulk}_{eq}$ is the equilibrium bulk free energy at temperature
474: $T$, one may replace $S_c(\Phi^{bulk}_{eq})$ by its equilibrium value
475: at that temperature $S_c(N,T)$. Thus we get for the escape rate to $N$
476: the result
477: \begin{equation}
478: k(N) = \tau^{-1}_{micro} \exp\left\{S_c(N,T) -
479: \frac{\phi_{eq}-\phi_{in}^{lib}}{k_B T} \right\}.
480: \end{equation}
481: The location of the bottleneck is determined by the minimum of this
482: expression over $N$. We can thus define an activation free energy
483: \begin{equation}
484: F^\ddagger(N) = \phi_{eq} - \phi_{in}^{lib} - T S_c(N,T),
485: \label{F_phi}
486: \end{equation}
487: whose maximum defines the bottleneck location. Notice this activation
488: barrier depends on the total free energy of having any target state at
489: size $N$ and the initial nonequilibrated particular state free energy
490: $\phi_{in}^{lib}$ in which we only include a vibrational contribution
491: and no configurational entropy. This function is shown in
492: Fig.\ref{profiles}.
493: \begin{figure}[tb]
494: \includegraphics[width=.55\columnwidth]{profiles.eps}
495: \caption{\label{profiles} The free energy to reconfigure the initial
496: configuration is shown as a function of the size of the locally
497: rearranged region. There will be fluctuations in the shape owing to
498: the detailed packing found in the initial configuration but on the
499: average the profile is given by equation (\ref{F(N)fluc}) in the
500: text. The dashed curve shows the profile for an initial state which is
501: much higher than the equilibrium energy at $T$, while the solid curve
502: is the free energy profile for an initially equilibrated state. }
503: \end{figure}
504: Introducing Eq.(\ref{phi_j}) into Eq.(\ref{F_phi})
505: yields:
506: \begin{equation}
507: F^\ddagger(N) = F_{eq}^{bulk}(N,T) -\Phi_{in}^{bulk}(N,T) + \gamma
508: N^x,
509: \label{F(N)fluc}
510: \end{equation}
511: where $F_{eq}^{bulk}(N,T) = \Phi^{bulk}_{eq} - TS_c(N,T)$ is the
512: total equilibrium free energy that includes both the configurational
513: entropy of the region of size $N$ and the internal bulk free energy
514: with its vibrational contribution. In this expression we have
515: substituted the average mismatch energy coefficient $\gamma$, for the
516: state specific values. For dynamically connected states
517: $F_{eq}^{bulk}(N,T)$, $\Phi_{in}^{bulk}(N,T)$, and $S_c(N,T)$ grow
518: linearly with $N$. There will also be $\delta F$, an additional
519: fluctuation that typically scales as $N^{1/2}$. Thus we write
520: \begin{equation}
521: F^\ddagger(N) = [f_{eq}(T) - \phi_{in}(T)] N + \gamma N^x + \delta F,
522: %\label{F(N)fluc}
523: \end{equation}
524: where $f_{eq}(T) = F_{eq}^{bulk}(N,T)/N$ is the total bulk free energy
525: per particle of the final state at temperature $T$ and $\phi_{in}(T)=
526: \Phi_{in}^{bulk}(N,T)/N$ is the internal free energy per particle of
527: the initial state. In the following, we will omit the fluctuation term
528: $\delta F$, whenever computing the most probable barrier, but will
529: consider it explicitly when estimating the degree of
530: non-exponentiality of the relaxation, which is directly related to the
531: barrier fluctuations. Note, the expression above is simply what would
532: be prescribed by a regular nucleation theory for the free energy
533: barrier of conversion from a (usually non-equilibrium) initial state
534: to the other (usually equilibrium) state. Clearly, when the initial
535: state is at equilibrium: $\Phi^{bulk}_{in} = \Phi^{bulk}$, - the
536: ``free energy'' driving force $f_{eq}(T) - \phi_{in}(T)$ is equal to
537: $-T s_c$, i.e. there is still relaxation in the supercooled liquid, as
538: driven by conversion between alternative (local) aperiodic packings of
539: the liquid. In general, the maximum of the typical $F^\ddagger$ occurs
540: at $N^\ddagger$ such that $\partial F^\ddagger/\partial
541: N|_{N=N^\ddagger}$ vanishes giving
542: \begin{equation}
543: N^\ddagger =
544: \left(\frac{\phi_{in}-f_{eq}}{x\gamma}\right)^{\frac{1}{x-1}}
545: \label{N(x)}
546: \end{equation}
547: and a typical (most probable) rate
548: \begin{equation}
549: k = \tau_{micro}^{-1}
550: \exp\left\{-\frac{\gamma}{T}\left(\frac{\phi_{in}-f_{eq}}{x\gamma}
551: \right)^{\frac{x}{x-1}} (1-x) \right\}.
552: \label{k(x)}
553: \end{equation}
554:
555: We finish this Section by elaborating on why the relaxations occuring
556: with the rate from Eq.(\ref{k(x)}) can occur with a zero free energy
557: gap thus leading to a local free energy landscape. Consider first a
558: supercooled liquid at equilibrium (i.e. not quenched):
559: $f_{eq}-\phi_{in} = -Ts_c$. Clearly the size $N^*$ at which
560: $F^\ddagger (N^*)=0$ corresponds to the same liquid state, therefore a
561: region of size $N^*$ can survey {\em all} liquid configurations
562: typical of this temperature. An isoenergetic state exists because the
563: system resides in a metastable state much higher than the lowest
564: energy crystalline arrangement, as reflected in non-zero
565: configurational entropy $s_c$, and the very high density of states
566: $e^{S_c(\phi)}$. (Note, this notion underlies the existence of
567: residual structural degrees of freedom in glasses even at cryogenic
568: temperatures \cite{LW}.) If the liquid is quenched and aging, the
569: driving force $f_{eq}-\phi_{in}$ is still negative (it will be
570: computed shortly), so both quantities $N^\ddagger$ and $N^*$ exist and
571: are comparable to their equilibrium values, although $N^*$ no longer
572: signifies straightforwardly the degree of the landscape locality
573: (consistent with the system being out of equilibrium in the first
574: place). Regardless, relaxations are local during aging too.
575:
576: \section{Relaxation in the Equilibrated Supercooled Liquid}
577:
578: Examining the expressions for the relaxation rate for an equilibrium
579: sample makes explicit the connection to the earlier form of the
580: entropic droplet idea. In the equilibrated sample case, the internal
581: vibrational free energies of the initial and final states are the same
582: so the driving force to distinct equilibrated configurations comes
583: only from the configurational entropy contribution available at the
584: equilibrium temperature. People are sometimes confused how
585: configurational entropy, which is never made available in a strict
586: mean field scenario, can drive a transition event. This is because of
587: the locality of the landscape in the droplet analysis, which goes
588: beyond meanfield thinking. Locally, for a region of space $D({\bf R},
589: N^*)$ there is a ``funnel'' of states leading away from the initial
590: configuration to other equally equilibrated
591: configurations. Configurational entropy drives the activated motions
592: of the local regions of a glass in the same way the entropy arising
593: from the large number of denatured configurations of a protein drives
594: the unfolding of a folded native protein even though the denatured
595: states are individually higher in energy \cite{OLSW}. Any individual
596: escape path is likely to find a big barrier but this barrier is
597: partially cancelled by the growth of the number of escape routes. The
598: free energies along each path fluctuate thus increasing the likelihood
599: of finding a low energy path, when a large number of paths is
600: available. As the entropy per particle, $s_c(T)$ gets smaller the
601: driving force to re-equilibrate falls so the critically activated
602: region of the glass necessary for re-equilibration grows larger and
603: the rate falls, owing to the larger activation barrier from the
604: mismatch contribution. The precise way this happens depends on the
605: mismatch energy and its exponent $x$.
606:
607:
608: Therefore, before proceeding to study the nonequilibrium situation, we
609: digress to discuss aspects of the average mismatch energy $\Gamma$
610: which we have approximated as $\gamma N^x$. First it is clear that the
611: form of $\gamma N^x$ is only a crude approximation to the mismatch
612: energy when $N$ is small. In principle, recall, the mismatch energy
613: could be explicitly computed by carrying out the local library
614: construction on a computer. The only problem is obtaining initial
615: configurations equilibrated to the appropriate low
616: temperatures. Finding such equilibrated configurations deep in the
617: landscape currently requires heroic computer resources. Therefore the
618: mismatch energy must presently be inferred by analytical
619: considerations.
620:
621:
622: The simplest free energy functional calculations of the mismatch
623: energy gives an energy proportional to the interface area and
624: therefore gives the mismatch exponent $x=2/3$. Such simplified
625: calculations also give an explicit value of the prefactor of the
626: scaling relation, $\gamma_0$ at the ideal glass transition
627: temperature, $T_K$. At this temperature, the lost interaction terms of
628: a sharp interface have to balance the entropy cost of localizing each
629: particle to its cage.
630:
631: This gives, at $T_K$ \cite{XW,LW},
632: \begin{equation}
633: \gamma_0 = \frac{2\sqrt{3\pi}}{2} k_B T_K
634: \ln\left[\frac{(a/d_L)^2}{\pi e}\right],
635: \end{equation}
636: where $d_L$ is the mean square fluctuations of particles in a given
637: basin \cite{inherent} and $a$ is the interparticle spacing. The ratio
638: $d_L/a$ is about 0.1 for glassy configurations, just as it is in the
639: Lindemann criterion for melting. Calculations based on free energy
640: functional show that $\gamma$ vanishes as $T_A$ is approached from
641: below \cite{KW_PRB}. This is because $T_A$ resembles a spinodal. From
642: these estimates, $\gamma(T)$ can be obtained \cite{LW_soft}. These
643: naive mean field estimates would give an activation barrier varying as
644: $s_c^{-2}$, as first detailed by Kirkpatrick and Wolynes \cite{KW_PRB}
645: and later discussed by Parisi \cite{Parisi}.
646:
647: Kirkpatrick, Thirumalai and Wolynes pointed out an effect left out in
648: the naive estimate of the mismatch energy \cite{KTW}: One must
649: acknowledge there are numerous solutions of the mean field equations
650: describing minima of the free energy functional, these precisely
651: correspond to our local energy landscape libraries. In principle some
652: of these other configurations can be interpolated between the internal
653: target state to which the region is relaxing and its fixed environment
654: in order to lower the mismatch energy. This interpolation is called
655: ``wetting''.
656:
657:
658: Such wetting only can be rigorously defined for very large
659: $N$. Wetting is a dynamical process that takes some time to
660: develop. We should therefore for greater accuracy ascribe a frequency
661: dependence to the mismatch surface tension $\gamma(T,\omega)$. The
662: mapping between the free energy functional and the random field Ising
663: model allows us to invoke an argument of Villain \cite{Villain} that
664: gives a curvature dependence to the surface tension coefficient
665: $\gamma=\gamma_0 (a/R)^{1/2}$ in three dimensions where $a$ is the
666: microscopic length where the surface tension $\gamma_0$ is established
667: ($R^d \propto N$). Accounting for this wetting correction leads to the
668: exponent $x=1/2$. When this exponent is used in Eq.(\ref{k(x)}), we
669: see that the relaxation rate in the equilibrium system has exactly the
670: Adam-Gibbs form
671: \begin{equation}
672: k=\tau_{micro}^{-1} e^{-A/s_c},
673: \label{k_AG}
674: \end{equation}
675: where $s_c$ is the configurational entropy per particle. If we assume
676: that $s_c$ vanishes linearly at an ideal glass transition temperature
677: $T_K$, this rate agrees with the Vogel-Fulcher law. Distinct from the
678: AG argument, however, in the RFOT theory the critical size
679: $N^\ddagger$ scales as $s_c^{-2}$ not $s_c^{-1}$ and the dynamical
680: correlation length is much bigger than the AG picture implies, but if
681: the simple estimate of $\gamma_0$ based on the vibrational free energy
682: cost is used, explicit values of the typical barrier height are
683: obtained in addition to the experimental scaling of rates, as found by
684: the RFOT theory \cite{XW}. The surface tension $\gamma_0/T_g$, being a
685: logarithmic function of the vibrational amplitude in the glass,
686: depends little on the atomic make up of the glass. Therefore if we use
687: the Vogel-Fulcher expression
688: \begin{equation}
689: k=k_0 \exp\{-D T_g/(T-T_K)\},
690: \end{equation}
691: the D is predicted to depend inversely on $\Delta c_p$ - the change in
692: heat capacity upon vitrification. Specifically one finds
693: \begin{equation}
694: D=32k_B/\Delta c_p
695: \end{equation}
696: The resulting correlation of barrier heights (measured by $D$) and
697: glass thermodynamics is excellent \cite{XW}. The softening of $\gamma$
698: near $T_A$ also explains very well the deviations from the VTF laws
699: that are observed as the temperature is raised to $T_A$ where there is
700: a crossover to collisional dynamics \cite{LW_soft}. We will use the
701: same ``wetted'' form of the surface energy term without softening in
702: the body of the paper to follow. We must note however even at $T_g$,
703: these microscopic calculations give $N^*$ only of the order of
704: hundreds of particles (consistent with experiment \cite{Ediger}), so
705: we are far from asymptopia and other ``ultimate'' scalings (closer to
706: $T_K$) are conceivable.
707:
708:
709: The local library and mismatch energy concepts let us discuss {\it en
710: passant} various ``defect'' pictures of glassy dynamics in landscape
711: terms. Many theories of the glassy state imagine there is a basic
712: undefected structure at the heart of the phenomenon (an
713: ``ur-structure''). Doubtless for many systems, the periodic crystal
714: itself is one such basic ``ur-structure''. For big enough $N^*$ the
715: periodic crystal will indeed be a member of the local landscape
716: library but it will not be entropically favored and as long as the
717: surface tension between liquid and crystal remains sufficient it will
718: not be a major target state. Certainly ``devitrification'' can and
719: does occur in the laboratory but we will leave the study of this
720: transformation for future discussion. Other ur-structures have been
721: discussed as dominating glassy dynamics such as icosahedral crystals
722: \cite{Nelson} or other ``avoided crystalline phases''
723: \cite{Kivelson}. Deep in the local energy landscape, these structures
724: and their defected forms must be found. The dimension of the defects
725: supported by these ur-structures will determine the mismatch energy
726: and thus how the rate will depend on the driving entropy, in the
727: present picture. Point defects such as interstitial-vacancy pairs have
728: an energy cost independent of the region site i.e. giving an exponent
729: $x=0$ and would give a relaxation rate independent of
730: $s_c$. NonArrhenius behavior based on point defects usually must rely
731: on special kinetic constraints, which must be encoded in the
732: transition rules \cite{Andersen}. By construction, the
733: quasi-equilibrium estimate of the rate made above would fail for such
734: models. Of course if there were only a few ``dead-end'' states these
735: could be explicitly subtracted out in the estimates. Such a situation
736: may apply for entangled polymers.
737:
738:
739: Simple estimates of the free energies of such point defects puts their
740: energy cost near to the limit seen in ordinary laboratory glass
741: transitions \cite{Eastwood}. At low temperature they may conceivably
742: short-circuit the generic transitions proposed here. Line-like defects
743: enter prominently into the constellation of approaches based on
744: frustrated icosahedral order \cite{Nelson} and presumably should occur
745: also when other types of frustrated regular ordered systems act as
746: ur-structures. Nussinov has argued that hamiltonians for uniformly
747: frustrated systems should exhibit random first order transitions in
748: the mean field approximation \cite{Nussinov}. It is likely correct to
749: view the states accessible in the high density of states region
750: relevant for real glasses in this way even for models based on
751: frustrated order.
752:
753:
754: Recall, in assessing the relevance of defect based models, that the
755: measured configurational entropy per particle at the laboratory $T_g$
756: is about 1 $k_B$. This means we are usually far from the low defect
757: density regime.
758:
759:
760: At low defect density, nevertheless, line-like defects may either
761: traverse the correlated region directly or wander across the region
762: like a Brownian path. The line length in the first case of
763: ``ballistic'' traversal should scale like $N^{1/3}$ while in the
764: Brownian case the length is proportional to square of the traversed
765: distance so we expect the mismatch energy to scale like $N^{2/3}$,
766: resembling the mean field result without wetting. It is interesting
767: that an interpolation between the ballistic and Brownian scalings
768: would be hard to distinguish from the result we use $\Gamma \sim
769: N^{1/2}$. Even at $T_g$, $N^*$ is only about 150 so all these power
770: laws (with the exception of the point defect case) could likely be
771: fitted to a detailed mismatch construction with similar accuracy for
772: liquids in the laboratory.
773:
774:
775:
776: \section{Relaxation in the Immediately Quenched Glass}
777:
778: The microcanonical local landscape arguments allow us to estimate the
779: relaxation rates once the statistics of the energies of the local
780: regions in the initial nonequilibrium state are known. Once the system
781: has fallen out of equilibrium, in general, these statistics depend on
782: the detailed quenching history of the sample. The complete aging
783: theory should determine these statistics self-consistently. In the
784: present and following sections we will assume the quench involves
785: simple straightforward cooling and that little time has elapsed since
786: the glass transition was passed. In this case the statistics of the
787: initial energies are taken to be those of an equilibrium system at a
788: temperature $T_g$. Presumably $T_g$ will be temperature of the
789: midpoint of the dynamic heat capacity drop upon cooling. Thus we take
790: $\phi_{in}(T)$ to be the bulk energy at $T_g$ augmented by the
791: vibrational part of the free energy at the ambient temperature
792: $T$. The fluctuations of the internal energy of regions, which
793: determine the nonexponentiality of relaxation in the glass will also
794: be taken to be the same as the equilibrium fluctuations at $T_g$, and
795: therefore to be determined by $\Delta c_p(T_g)$.
796:
797:
798: Using the mismatch exponent $x=1/2$ in equations (\ref{N(x)}) and
799: (\ref{k(x)}) we obtained fairly simple results for the critical
800: cluster size
801: \begin{equation}
802: N^\ddagger = \left[\frac{2(\phi_g-f_{eq})}{x\gamma}\right]^{-2}
803: \end{equation}
804: and the typical rate
805: \begin{equation}
806: k = \tau_{micro}^{-1} \exp\left\{-\frac{\gamma^2}{4 k_B T
807: [\phi_g-f_{eq}]} \right\}.
808: \label{k(x)g}
809: \end{equation}
810:
811: The initial and final states are at the same vibrational temperature
812: but not at the same configurational temperature. The initial state is
813: typical of a $T_g$ configuration but the activated state is
814: ``equilibrated'' to the temperature $T$ both in vibrational terms and
815: in locally configurational terms. Because of this the driving force
816: for reconfiguration is not purely entropic in the nonequilibrated case
817: Thus there is a big conceptual difference from the usual extrapolation
818: of the Adam Gibbs formula. In addition, the rates predicted by
819: Eq.(\ref{k(x)g}) differ from the Adam-Gibbs formula extrapolated with
820: a fixed configurational entropy $s_c(T_g)$.
821:
822: At $T_g$, clearly the rates predicted by the equilibrated formula
823: (\ref{k_AG}) and Eq.(\ref{k(x)g}) are the same. Upon further cooling
824: the nonequilibrium rate while decreasing, however, is substantially
825: higher than it would be at equilibrium. This is because the driving
826: force for reconfiguration not only includes the configurational
827: entropy at the ambient temperature but also an energy increment of the
828: initial state $\Delta \epsilon = \epsilon_g - \epsilon_{eq}(T)$. (We
829: assume vibrational free energy contributions are nearly the same in
830: inherent structures typical of $T_g$ and the ambient temperature $T$.)
831: Just such a change of slope is observed in the laboratory: the
832: apparent activation energy in the nonequilibrium glassy state is
833: smaller than the extrapolated value from the equilibrated supercooled
834: liquid. This change is usually quantified in the
835: Nayaranaswany-Moynihan-Tool framework through the expression
836: \begin{equation}
837: k_{n.e.} = k_0 \exp\left\{-x_{NMT}\frac{\Delta E^*}{k_B T} -
838: (1-x_{NMT}) \frac{\Delta E^*}{k_B T_f} \right\}. \hspace{4mm}
839: %\label{k_AG}
840: \end{equation}
841: where $E^*$ is the equilibrated apparent activation energy at $T_g$
842: and $x_{NMT}$ lies between 0 and 1. To compare with our initial quench
843: result we would take $T_f = T_g$, again noting in a realistic cooling
844: history $T_f$ would need to be self-consistently determined. We note
845: that equation (\ref{k(x)g}) gives a gradual transition to Arrhenius
846: behavior. We can find $x_{NMT}$ most easily at low temperatures (near
847: $T_K$), where the typical relaxation rate will follow an Arrhenius law
848: according to Eq.(\ref{k(x)g}). The Arrhenius behavior applies because,
849: for ordinary liquids near to $T_K$ we expect the mismatch free energy
850: to largely be energetic so we take it as a constant. By the definition
851: of the Kauzmann temperature $f_{eq}(T_K)= \phi_K$ is also mostly
852: energetic since the configurational entropy vanishes. The vibrational
853: components of $f_{eq}$ and $\phi_g$ are assumed to cancel so the
854: activation energy is in the nonequilibrium very low temperature regime
855: \begin{equation}
856: \Delta E_{n.e.L.T.}^\ddagger = \frac{\gamma(T_K)^2}{4
857: (\epsilon_g-\epsilon_K)}
858: \label{EneLT}
859: \end{equation}
860: where $\epsilon_g$ is the energy per particle of the frozen glassy
861: state, as prepared, and $\epsilon_K$ is the energy per particle in the
862: putative ideal glassy ground state equilibrated at the Kauzmann
863: temperature. The energy difference $\epsilon_g-\epsilon_K$ is
864: determined by the configurational part of the heat capacity $\Delta
865: c_p(T)$ for intermediate values of $T$ such that $T_g > T > T_K$. The
866: apparent activation energy in the nonequilibrated glass at low
867: temperatures turns out to be comparable to the equilibrium activation
868: {\em free} energy, $\Delta F^\ddagger$ at $T_g$. It therefore should
869: not vary much from substance to substance but depends on the quenching
870: time scale $t_Q$ through the relation
871: \begin{equation}
872: t_Q = \tau_{micro} \exp\left(\frac{\Delta F^\ddagger_g}{k_B T }
873: \right)
874: \label{tQ}
875: \end{equation}
876: To obtain this result of the near equality of $\Delta F_g^\ddagger$
877: and $\Delta E_{n.e.L.T.}^\ddagger$ let us take the configurational
878: heat capacity to have the form $\Delta c_p= \Delta c_p(T_g)T_g/T$, as
879: suggested by Angell \cite{Angell_cp1,Angell_cp2}. This form is based
880: on good laboratory estimates. We can now find by integrating and
881: insert this into Eq.(\ref{EneLT}) to give:
882: \begin{equation}
883: \Delta E_{n.e.L.T.}^\ddagger = \frac{\gamma(T_K)^2}{4 \Delta c_p(T_g)
884: k_B T_g \ln(T_g/T_K)}.
885: %\label{EneLT}
886: \end{equation}
887: The activation free energy at $T_g$, on the other hand is obtained by
888: finding $s_c$ from the integration of $\Delta c_p/T$ and using this in
889: the equilibrium rate expression to give:
890: \begin{equation}
891: \Delta F_g^\ddagger = \frac{\gamma(T_g)^2}{4 \Delta c_p(T_g) k_B T_g
892: (T_g/T_K-1)}.
893: %\label{EneLT}
894: \end{equation}
895: The ratio of the activation energy in the glass to the equilibrated
896: free energy barrier at $T_g$ is therefore
897: \begin{equation}
898: \frac{\Delta E_{n.e.L.T.}^\ddagger}{\Delta F_g^\ddagger} =
899: \left[\frac{\gamma(T_K)}{\gamma(T_g)} \right]^2
900: \frac{(T_g/T_K-1)}{\ln(T_g/T_K)}.
901: \label{DDratio}
902: \end{equation}
903: If $T_K$ and $T_g$ are close, as they are for ``fragile'' systems
904: expanding the logarithm gives a ratio close to one. Even for the
905: strong liquid Si0$_2$ the ratio of $T_g$ (1480K) to $T_K$ (876K) is
906: only 1.7, which gives $\Delta E_{n.e.L.T.}^\ddagger/\Delta
907: F_g^\ddagger=0.8$, if we neglect any temperature dependence of
908: $\gamma$. For a laboratory glass transition on the one hour time
909: scale we should find universally to a good approximation (1 hour
910: quench) ~ 32 to 39 $k_B T_g$. To compare with the NTM phenomenology we
911: note that the RFOT theory predicts universally that the activation
912: energy (not free energy!!!) in the glassy state depends on the quench
913: time
914: \begin{equation}
915: \frac{\Delta E_{n.e.L.T.}^\ddagger}{k_B T_g} = \alpha
916: \ln(t_Q/\tau_{micro})
917: %\label{k(x)g}
918: \end{equation}
919: with a coefficient $\alpha$ very close to one. This is consonant with
920: many experiments.
921:
922: The apparent equilibrium activation energy at $T_g$ in the liquid is
923: much larger than the nonequilibrium value found in the glassy state
924: (being in a sense cancelled in the rate expression by a large positive
925: entropy of activation). Again assuming $\Delta c_p$ depends on
926: temperature, we find
927: \begin{eqnarray}
928: \Delta E_{g,app}^\ddagger &=& \left. \frac{\partial (\Delta
929: F^\ddagger/T)}{\partial (1/T)}\right|_{T_g} \nonumber \\ &=& \Delta
930: F_g^\ddagger \left\{2 \frac{\partial \ln \gamma(T)}{\partial(-\ln T)}
931: + 2 + \frac{1}{T_g/T_K-1}\right\} \hspace{5mm}.
932: %\label{k}
933: \end{eqnarray}
934: Using the condition that the configurational entropy vanishes at
935: $T_K$, this energy can also be expressed in terms of the ratio of
936: $c_p(T_g)$ to $s_c(T_g$):
937: \begin{equation}
938: \Delta E_{g,app}^\ddagger = \Delta F_g^\ddagger \left\{2
939: \frac{\partial \ln \gamma(T)}{\partial(-\ln T)} + 2 + \frac{\Delta
940: c_p(T_g)}{s_c(T_g)} \right\}.
941: \label{DEapp}
942: \end{equation}
943: We then obtain for the inverse of the nonlinearity parameter $x_{NMT}$ at
944: low temperatures from Eqs.(\ref{DDratio}) and (\ref{DEapp}):
945: \begin{widetext}
946: \begin{equation}
947: x_{MNT}^{-1} = \frac{E_{g,app}^\ddagger}{\Delta E_{n.e.L.T.}^\ddagger}
948: = \left\{2 \frac{\partial \ln \gamma(T)}{\partial(-\ln T)} + 2 +
949: \frac{\Delta c_p(T_g)}{s_c(T_g)} \right\}
950: \left[\frac{\gamma(T_g)}{\gamma(T_K)} \right]^2
951: \frac{\ln(T_g/T_K)}{(T_g/T_K-1)}.
952: %\label{DEapp}
953: \end{equation}
954: \end{widetext}
955: $\Delta c_p(T_g)/s_c(T_g)$ is a thermodynamic measure of the liquid
956: fragility. This ratio is large for very fragile liquids and small for
957: ``strong'' liquids. Thus we see ``fragile'' liquids are very nonlinear
958: while ``strong'' liquids, in general, should not be. Just such a
959: correlation has been discussed by McKenna and Angell \cite{McKenna}
960: and can be expressed as a relation between the kinetic fragility
961: parameter
962: \begin{equation}
963: m = \left. \frac{1}{T_g} \frac{\partial \log_{10}
964: \tau}{\partial(1/T)}\right|_{T_g} = \frac{\Delta
965: E_{g,app}^\ddagger}{k_B T_g}\log_{10}e .
966: \end{equation}
967: and the nonlinearity parameter $x_{NMT}$. Using the previously derived
968: relation (\ref{DDratio}) and the equation above we find
969: %\begin{widetext}
970: \begin{eqnarray}
971: x_{MNT}^{-1} &=& \frac{E_{g,app}^\ddagger}{\Delta
972: E_{n.e.L.T.}^\ddagger} \nonumber \\ = &m& \left\{ (\log_{10}e)
973: \frac{\Delta F_g^\ddagger}{k_B T_g}
974: \left[\frac{\gamma(T_K)}{\gamma(T_g)} \right]^2
975: \frac{(T_g/T_K-1)}{\ln(T_g/T_K)} \right\}^{-1} \hspace{5mm}
976: %\label{DEapp}
977: \end{eqnarray}
978: %\end{widetext}
979: Assuming the $\gamma$ ratio is near one and using the glass transition
980: temperature appropriate to 1 hr, so $\Delta F_g^\ddagger/k_B T_g = \ln
981: 10^{17} \simeq 39$, and a generic $T_g/T_K = 1.26$ \cite{LW_soft}, we
982: obtain
983: \begin{equation}
984: m \simeq \frac{19}{x}.
985: \end{equation}
986: This relation is plotted in Figure \ref{m_x_labelled1} along with
987: data for several systems. We see that the estimate agrees reasonably
988: well with experiment. In this estimate we have neglected the $T$
989: dependence of $\gamma$. The microscopic treatment from RFOT shows
990: however that $\gamma_0$ depends both on the proximity to $T_K$ and to
991: $T_A$, as discussed in our paper on the barrier oftening effect
992: \cite{LW_soft}. Thus we will generally have both an entropic and an
993: energetic contribution to $\gamma_0$ which may explain some of the
994: scatter in the curves. Indeed we see more fragile systems lie
995: systematically above the curve as is expected since $T_K$ and $T_A$
996: are closer leading to the longer barrier softening effect.
997: \begin{figure}[tb]
998: \includegraphics[width=.6\columnwidth]{m_x_labelled1.eps}
999: \caption{\label{m_x_labelled1} The fragility parameter $m$ is plotted
1000: as a function of the NTM nonlinearity parameter $x_{NMT}$. The curve
1001: is predicted by the RFOT theory when the temperature variation of
1002: $\gamma_0$ is neglected. The data are taken from
1003: Ref. \cite{app_phys_rev}. A few substances (PVAc = polyvinyl acetate,
1004: PVC = polyvinylchloride, PS = polystyrene, B$_2$O$_3$, and
1005: As$_2$Se$_3$) are labeled. Notice some measured values are not
1006: consistent on multiple measurements; this may reflect a breakdown of
1007: phenomenology for the history dependence discussed in the text or
1008: different material preparation. The more fragile substances lie above
1009: the prediction without barrier softening, which has no adjustable
1010: parameters. }
1011: \end{figure}
1012:
1013: According to the RFOT theory the typical relaxation time in the
1014: nonequilibrium quenched state does {\em not} immediately become
1015: Arrhenius in temperature dependence below the laboratory glass
1016: transition. Thus there should be deviations from the NMT formalism,
1017: which might be crudely fit by allowing the nonlinearity parameter to
1018: be $T$ dependent. Indeed some of the scatter in
1019: Fig.\ref{m_x_labelled1} probably arises also from this cause. Below
1020: $T_g$, the expected nonArrhenius behavior from RFOT theory is much
1021: weaker than the divergently nonArrhenius behavior found above
1022: $T_g$. Within the Adam-Gibbs nonequilibrium extrapolation advocated by
1023: Hodge \cite{Hodge} and Scherer \cite{Scherer}, the RFOT theory result
1024: would appear to involve a temperature dependent ``configurational
1025: entropy''. Just such a behavior has been found by Alegria et al. in
1026: polymer systems \cite{Alegria}. One way to express this connection is
1027: to compare $\Delta F_{n.e.}^\ddagger(T)$ to its value at $T_g$. The
1028: inverse of this ratio would be the apparent configuration entropy in a
1029: Hodge and Scherer-Adam-Gibbs extrapolation. In view of
1030: Eq.(\ref{k(x)g}), the inverse ratio is given by:
1031: %\begin{widetext}
1032: \begin{eqnarray}
1033: \left[\frac{\Delta F_{n.e.}^\ddagger(T)}{\Delta
1034: F_g^\ddagger}\right]^{-1} = \left[\frac{\phi_g-f_{eq}(T)}{T_g
1035: s_c(T_g)} \right] \left[\frac{\gamma(T_g)}{\gamma(T)} \right]^2
1036: \hspace{10mm} \nonumber \\ \hspace{10mm} = \left\{ \frac{T-T_K [1+
1037: \ln(T/T_g)]}{T_g-T_K} \right\} \left[\frac{\gamma(T_g)}{\gamma(T)}
1038: \right]^2, \hspace{5mm}
1039: \label{nonArrh_rat}
1040: \end{eqnarray}
1041: %\end{widetext}
1042: where the second equality is found by computing $f_{eq}$ with the help
1043: of the Angell's form of $\Delta c_p(T)$ (see the Appendix). We plot
1044: the ratio above in Fig.\ref{nonArr}.
1045: \begin{figure}[tb]
1046: \includegraphics[width=.85\columnwidth]{nonArr.eps}
1047: \caption{\label{nonArr} We plot the predicted variation of the
1048: activation free energy versus inverse temperature below the glass
1049: transition, as the ratio $(\Delta F_{n.e.}^\ddagger(T)/\Delta
1050: F_g^\ddagger)^{-1}$ from Eq.(\ref{nonArrh_rat}) versus $T/T_g$. Only
1051: below $T_K$ will the ratio be strictly constant, implying strictly
1052: Arrhenius temperature dependence of the relaxation rate. The
1053: temperature dependence of $\gamma$ is neglected. The figure is plotted
1054: for a fragile material with $T_K/T_g$ (1 hr.) = 0.9, a less fragile
1055: one with $T_K/T_g$= 0.7 and a strong substance with $T_K/T_g$= 0.5. }
1056: \end{figure}
1057: We point out that although we have
1058: highlighted the connection with more traditional treatments of aging
1059: phenomenology, neither of the earlier approaches is exactly
1060: commensurate with our theory. On the other hand we have shown the
1061: expression for the nonequilibrium rate is very explicit once the
1062: average energy of the sample is known. It therefore would not be
1063: terribly difficult to use the full expression (\ref{k(x)g}) in a
1064: dynamic treatment with the nonequilibrium energy as the indicator of
1065: the ``fictive'' temperature. We do not carry out this analysis here
1066: because it involves detailed numerical fitting for each system and the
1067: quench history of each particular experiment.
1068:
1069: \section{Nonexponentiality of Relaxation above and below $T_g$}
1070:
1071: Within the RFOT theory the mosaic structure of the liquid gives rise
1072: to dynamical heterogeneity and nonexponential relaxation. The driving
1073: force for re-equilibration varies from mosaic cell to cell. This leads
1074: to a range of activation barriers, $\delta F^\ddagger$. In the
1075: canonical ensemble formulation, using the usual Landau formula, these
1076: fluctuations depend on $\Delta c_p(T)$. Computing these fluctuations
1077: allowed Xia and Wolynes to predict $\beta$ for a range of substances
1078: \cite{XWbeta}. The same result can be obtained in the microcanonical
1079: formulation, however here the origin is even more transparent: there
1080: is a range of energies of the initial configurations for each mosaic
1081: cell. Since the energy fluctuations also scale with $\Delta c_p(T)$,
1082: above $T_g$ one obtains the same results as Xia and Wolynes previously
1083: derived.
1084:
1085:
1086: We first discuss how $\beta$ varies with temperature below $T_g$ in a
1087: somewhat simplified approximation that makes the issues clear. We will
1088: assume the energy fluctuations are small and the resulting barrier
1089: distribution is Gaussian. We will see later that the results of this
1090: analysis bound the magnitude of the {\em change} expected in $\beta$,
1091: even when a more accurate barrier distribution is used.
1092:
1093:
1094: Let us re-write Eq.(\ref{k(x)g}) for $\Delta F^\ddagger$ now including
1095: a fluctuation for the initial energy per particle
1096: \begin{equation}
1097: \Delta F_{n.e.}^\ddagger = \frac{\gamma^2}{4 [\phi_{in} -
1098: f_{eq}+\delta \phi_{in}]}.
1099: \label{dDFne}
1100: \end{equation}
1101: If $\delta \phi_{in}$ is small, we find the fluctuations in the
1102: barrier height
1103: \begin{equation}
1104: \delta \Delta F_{n.e.}^\ddagger = \frac{\gamma^2 (-\delta
1105: \phi_{in})}{4 [\phi_{in} - f_{eq}]^2} = \frac{\Delta F_{n.e.}^\ddagger
1106: (-\delta \phi_{in})}{\phi_{in} - \phi_{eq}+ T s_c(T)}.
1107: %\label{DDratio}
1108: \end{equation}
1109: Since the structure is frozen at $T_g$ the typical fluctuation in
1110: $\phi$ is the same as at $T_g$ and is determined by $\Delta c_p$ at
1111: $T_g$. We see the ratio of the size of the $\Delta F^\ddagger$
1112: fluctuations in the frozen, cooled state $\delta \Delta
1113: F_{n.e.}^\ddagger$ to those found at $T_g$, $\delta \Delta
1114: F_g^\ddagger$ is
1115: \begin{equation}
1116: \frac{\delta \Delta F_{n.e.}^\ddagger}{\delta \Delta F_g^\ddagger} =
1117: \frac{\Delta F_{n.e.}^\ddagger}{\Delta F_g^\ddagger} \frac{T_g
1118: s_c(T_g)}{\phi_{in} - \phi_{eq}+ T s_c(T)}.
1119: \label{dFdFratio}
1120: \end{equation}
1121: Both factors in this expression increase rather slowly below $T_g$,
1122: and saturate at $T_K$.
1123:
1124: How do these fluctuations translate into stretching exponents? Roughly
1125: speaking, the relaxation function for a Gaussian distribution of
1126: barriers is approximated by a stretched exponential with a value given
1127: by \cite{XWbeta,Castaing}:
1128: \begin{equation}
1129: \beta \simeq \left[1+ (\delta \Delta F^\ddagger/k_B T) \right]^{-1/2}.
1130: \label{beta}
1131: \end{equation}
1132: If the liquid is ``strong'', $\Delta c_p$ is small so there are small
1133: energy fluctuations leading to small $\delta \Delta F^\ddagger$. Thus
1134: for strongly liquids $\beta$ remains near 1 until rather low
1135: temperatures. If the fluctuations in $\phi$ are large (as they are for
1136: very fragile systems) we would find instead
1137: \begin{equation}
1138: \beta \simeq \frac{k_B T}{\delta \Delta F^\ddagger}.
1139: \label{beta_app}
1140: \end{equation}
1141: This formula should thus give an overestimate for the variation of
1142: $\beta$ with temperature. Using this estimate along with
1143: Eqs.(\ref{dDFne}) and (\ref{dFdFratio}), we find
1144: \begin{equation}
1145: \frac{\beta_{n.e.}(T)}{\beta(T_g)} = \frac{T}{T_g} \left[ \frac{\Delta
1146: F_{n.e.}^\ddagger(T)}{\Delta F_g^\ddagger} \right]^{-2}
1147: \left[\frac{\gamma(T)}{\gamma(T_g)} \right]^2.
1148: \label{beta_ratio}
1149: \end{equation}
1150: A plot of this expression is shown in Fig \ref{b_T} (note the $\Delta
1151: F_{n.e.}^\ddagger/\Delta F_g^\ddagger$ ratio has already been computed
1152: in Eq.(\ref{nonArrh_rat})).
1153: \begin{figure}[tb]
1154: \includegraphics[width=.9\columnwidth]{b_T.eps}
1155: \caption{\label{b_T} The variation with temperature of the
1156: nonequilibrium $\beta$ in comparison with $\beta(T_g)$ is shown as a
1157: dashed line for a substance with $T_K/T_g$ (1 hr.) = 0.8,
1158: characteristic of a fragile system and a solid line for a strong
1159: system with $T_K/T_g$=.5. The two other lines indicates how the
1160: equilibrium would vary with $T$. The approximte estimates from
1161: Eq.(\ref{beta_app}) which exaggerate the variation are plotted as the
1162: very thin lines, which nearly coincide with the more accurate
1163: exression Eq.(\ref{beta}). The inset shows a magnified view of the
1164: region near $T_g$, where only factual $\beta$'s, depicted by thick
1165: lines in the main graph, are given.}
1166: \end{figure}
1167: At $T_K$ and below the term in parenthesis saturates and so, we find a
1168: simple expression
1169: \begin{equation}
1170: \frac{\beta_{n.e.}(T)}{\beta(T_g)} = \frac{T}{T_g} \left[ \frac{\Delta
1171: F_{n.e.}^\ddagger(T_K)}{\Delta F_g^\ddagger} \right]^{-2}
1172: \left[\frac{\gamma(T_K)}{\gamma(T_g)} \right]^2.
1173: \label{beta_ratioK}
1174: \end{equation}
1175: Thus we see in general $\beta$ should fall as we cool but the effect
1176: remains modest in the range where $T$ is greater than $T_K$. This
1177: modestly cooled range is where the most detailed aging studies have
1178: been reported \cite{Alegria,Leheny}. Although $\beta$ falls, indeed,
1179: the rate of fall slows from its $T$ dependence above $T_g$ for fragile
1180: substances, since according to the simple RFOT approximation of Xia
1181: and Wolynes the equilibrium $\beta$ would vanish at $T_K$, while this
1182: (approximate) nonequilibrium vanishes only at absolute zero. In
1183: contrast, $\beta$ for strong substances, that was near unity and
1184: $T$-independent above $T_g$, is predicted to show more pronounced
1185: tempearature dependence below the glass transition (see the inset of
1186: Fig.\ref{b_T}). Thus we see the form of the relaxation is not the
1187: same as the equilibrium relaxation at any ``fictive'' temperature. For
1188: fragile liquids $T_K/T_g \simeq .8$, so if $\gamma$ is temperature
1189: independent, a 40\% reduction is expected until $T_K$ is
1190: reached. Often relaxation data in the aging regime have been fit with
1191: the approximation $\beta = \beta(T_g)$. Alegria et al. suggest $\beta$
1192: remains constant below $T_g$. The RFOT indicates this is a reasonable
1193: zeroth order approximation. Alegria et al. have measured $\beta$ in
1194: the regime $T_g >T> T_K$. These are difficult measurements and there
1195: is scatter in the data.
1196:
1197:
1198: We must bear in mind that the theoretical $\beta$ assumes that
1199: measurements can scan over the complete relaxation process (even below
1200: $T_g$!) but part of the relaxation is missed in experiment. Since the
1201: glass only partly relaxes, in most experiments $\beta$ will appear to
1202: be {\em closer} to one than if the full relaxation could be
1203: followed. This effect of missing part of the distribution gives a
1204: positive increment to $\beta$ at the glass transition, as pointed out
1205: by Alegria et al. Such an increment on cooling is found.
1206:
1207:
1208: After this artifactual increment in $\beta$, Alegria et al. actually
1209: do find $\beta$ to slightly decrease upon cooling. We must emphasize
1210: that the arguments leading to Eq.(\ref{beta_ratio}) generally give an
1211: overestimate of the variation of with temperature in the glass
1212: state. First and most simply the approximation made in going from
1213: Eq.(\ref{beta}) to Eq.(\ref{beta_app}) causes an overestimate. Second,
1214: but more important, as discussed by Xia and Wolynes \cite{XWbeta} the
1215: zeroth order Gaussian approximation for the barrier distribution is
1216: not quantitative because the simple argument leading to that result
1217: assumes the environment of a reconfiguring domain is temporally
1218: fixed. Clearly if the domains surrounding a region that may
1219: re-configure have themselves already changed before the
1220: reconfiguration event, the library construction's premise of having a
1221: fixed environment to the mosaic cell fails. This change of
1222: environment effect might be called a ``facilitation''
1223: \cite{Andersen}. In any case this effect means the barrier height
1224: distribution will be cut off on the high barrier side. A simple cutoff
1225: distribution follows from the idea that domains slower than the most
1226: probable rate would actually re-configure when their environmental
1227: neighbors have changed; thus they actually will re-configure at nearly
1228: the most probable rate, which has already been predicted by the RFOT
1229: theory. The resulting cutoff distribution for activation barrier:
1230: \begin{equation}
1231: P(\Delta F^\ddagger) = \left\{ \begin{array}{l} P_f(\Delta
1232: F^\ddagger), \mbox{ for } \Delta F^\ddagger < \Delta F_0^\ddagger \\ C
1233: \delta(\Delta F^\ddagger - \Delta F_0^\ddagger)
1234: \end{array} \right.
1235: \label{Pf}
1236: \end{equation}
1237: - has been shown to reproduce the variation of the equilibrium at
1238: $T_g$ with composition quite well \cite{XWbeta}. It also reproduces
1239: the temperature dependence of $\beta$ in the equilibrated supercooled
1240: liquid. The weight $C$ ensures the normalization of the
1241: distribution. Clearly the cutoff again acts to dampen the variation of
1242: with temperature, below $T_g$. Explicit calculations using
1243: Eqs.(\ref{Pf}) and (\ref{dFdFratio}) for the fluctuation can be used
1244: to study the detailed $T$ variation of $\beta$ in the nonequilibrium
1245: glass, which may be relevant for deep quenches.
1246:
1247: %\vspace{-3mm}
1248:
1249: \section{Relaxation in Quenches below $T_K$}
1250:
1251: %\vspace{-2mm}
1252:
1253: We have emphasized dynamics in the glassy regime just below $T_g$ and
1254: ranging down to $T_K$. In this regime dynamics is fast enough so that
1255: significant relaxation is still accessible to detailed
1256: experiments. The results we have obtained should hold to considerably
1257: lower temperatures at least before much aging has actually occured.
1258: Essentially the average rate will become Arrhenius below $T_K$ while
1259: the breadth of relaxation times will continue to increase as $T$
1260: approaches the absolute zero. Quantitatively several effects may
1261: intervene that are worthy of further study, however. These effects
1262: are 1) secondary relaxations, 2) fluctuations in mismatch energies and
1263: changes in the ``wetting'' mechanism, and 3) quantum effects. We
1264: comment on these in turn.
1265:
1266: 1) As a practical matter, less and less of the range of relaxation
1267: times can be accessed on deep cooling during a typical laboratory
1268: experiment - at the lowest temperatures only the fast end of the
1269: relaxation time distribution can be accessed. This part of the
1270: distribution is caused by regions of high energy that correspond to a
1271: small domain size. At the shorter length scale several effects that
1272: are system specific, will occur.
1273:
1274: This region of the relaxation time spectrum is often called the
1275: $\beta$-relaxation. The term $\beta$-relaxation tends to bring to mind
1276: universal characteristics and indeed the RFOT theory does for example
1277: suggests a scale for the maximum rate corresponding to the cost of
1278: overturning a single molecular unit; essentially this is
1279: $\gamma_0$. On the other hand most glass formers have internal
1280: structure and these inner parts or side-chains of the molecule will
1281: have multiple conformations that can relax, in a specific ways as well
1282: those predictable from RFOT alone \cite{slaving}. These ``rapid''
1283: relaxations, also will slow with temperature, and might come into the
1284: measurement window. Dyre has argued that just such a ``contamination''
1285: of the main $\alpha$ aging process by a $\beta$ relaxation \cite{Dyre}
1286: may explain some of the Leheny-Nagel measurements on glycerol
1287: \cite{Leheny}.
1288:
1289: 2) Well below $T_K$ the mismatch energies may change their scaling
1290: with size. The ``wetting'' mechanism relies on there being a
1291: multiplicity of states to interpolate through the interface. Such a
1292: multiplicity exists at $T_K$ but becomes less important with
1293: decreasing $T$. This leads to a ``hardening'' of the interface
1294: (increase in $\gamma$) or perhaps a crossover from $N^{1/2}$ scaling
1295: to $N^{2/3}$ scaling of the mismatch energy. At the same time, rare
1296: fluctuations of the barriers may allow some regions to relax more
1297: rapidly than expected. Plotkin and Wolynes have analyzed just this
1298: sort of effect in the context of ``buffing'' of energy landscapes in
1299: protein \cite{buffing} and Lubchenko and Wolynes have studied a
1300: similar effect in the quantum regime which allows a tail of resonant
1301: tunneling states to appear \cite{LW}. Also well below $T_K$ - small
1302: regions may reconfigure by crystallizing - an effect we ignore in the
1303: present paper.
1304:
1305: 3) Eventually below $T_K$ classical barrier crossing will be
1306: supplanted by quantum tunneling. This gives rise eventually to two
1307: level systems \cite{LW} and the Boson peak \cite{LW_BP} which we have
1308: discussed in detail elsewhere.
1309:
1310: \vspace{-5mm}
1311:
1312: \section{Aging and History Dependence}
1313:
1314: The local energy landscape theory put forward in this paper predicts
1315: the relaxation of a system where the statistics of the energies in the
1316: local energy landscape libraries of the sample is assumed to be known
1317: at any instant. The explicit formula for the typical relaxation rate
1318: depends on the mean bulk energy of the cooperative regions and the
1319: explicit formula for the stretching exponent $\beta$ also contains the
1320: variance of these energies. In the quenched sample we have assumed
1321: these statistics are characteristic of an equilibrium system at the
1322: temperature $T_g$ where the system ``fell out of equilibrium''. Since
1323: the cooperative regions are large it is natural to assume these
1324: statistics are Gaussian as we have done. In simple quench histories
1325: $T_g$ can be estimated by the temperature where the apparent heat
1326: capacity most rapidly falls during cooling. But more complicated
1327: thermal histories are possible and even in a simple quench $T_g$
1328: really must be determined self-consistently by the dynamics of the
1329: system, itself. In general the statistics of the local landscape
1330: libraries for a nonequilibrium system will be determined by the
1331: system's dynamics and its detailed past thermal history.
1332:
1333: The most general description of the nonequilibrium statistics is quite
1334: complex since the bulk energies of any region can be considered
1335: functions over the shape and size of domains. These functions might be
1336: described as a set of fluctuating bulk energy fields, but the
1337: resulting construction is complex. When the system is cooling, high
1338: energy regions of size $N^*$ will be replaced by regions equilibrated
1339: to the ambient temperature $T$. Thus there will be a patchwork of
1340: equilibrated and nonequilibrated mosaic cells (see Figure
1341: \ref{inhomT}).
1342: \begin{figure}[tb]
1343: \includegraphics[width=.55\columnwidth]{inhomT.eps}
1344: \caption{\label{inhomT} After a considerable period of aging well
1345: below $T_g$ a patchwork of equilibrated and nonequilibrated mosaic
1346: cells will be found. If the equilibrium energy at $T$ is more than a
1347: standard deviation of the configurational energies at $T_g$, the
1348: distribution of energies will be noticeably bimodal and the idea of a
1349: single fictive temperature will break down. The unimodal distribution
1350: with a single fictive temperature should be quite safe if $\Delta T =
1351: T_g -T < \sqrt{k_B T^2/\Delta c_p N^*} = \delta T^*$. For $T_g$
1352: relevant to 1 hr. quenches this gives $\delta T^*/T_g \simeq
1353: 0.07$. Most of the Alegria et al. \cite{Alegria} data lie in this
1354: modest quenching range, while ``hyperquenched'' samples (with $\Delta
1355: T \gg \delta T^*$) will often fall outside the allowed range of using
1356: a single fictive temperature. When a sample has a two peaked
1357: distribution of local energies, an ultra-slow component of relaxation
1358: will arise. Notice that an equilibrated region at the temperature
1359: $T=T_g - \delta T^*$ will relax on the tens to hundreds of hours scale
1360: (using the relation that $\Delta E_{n.e.L.T.}^\ddagger \simeq \Delta
1361: F_g^\ddagger$), if $\tau_g$ is taken to be one hour. Percolation of
1362: the ultra slow regions will lead to the possibility of observing
1363: multiple length scales. }
1364: \end{figure}
1365: The nature of the new statistics, after some transitions occur and
1366: substantial aging has progressed, depends on how big is the difference
1367: between the typical current energy of a region and the target
1368: equilibrium energy. If the gap is big, a two peaked structure in the
1369: distribution of local bulk energies will develop and the statistics
1370: will be far from Gaussian: some regions that are newly equilibrated
1371: will relax further at a (slower) rate characteristic of equilibrium at
1372: $T$, while the other, not yet transformed regions will still relax at
1373: the faster nonequilibrium rate discussed already. Such a situation, if
1374: it arises, might account for ``ultra-slow'' relaxations which have
1375: occasionally been reported in aging studies \cite{MillerMcPhail} (see
1376: the Figure \ref{inhomT} caption for a criterion for such
1377: behavior). The statistics in this case of very deep quenches studied
1378: for very long times is very complex kinematically. Two peaked
1379: distributions of local energies may also arise if the system is
1380: abruptly but briefly heated from a low temperature state to a much
1381: higher one. Fortunately in the more usual situation of modest
1382: monotonic quenches we can expect the distribution of local energies to
1383: remain unimodal. In such cases, to first order then the distribution
1384: will be characterized by a mean bulk energy per particle
1385: $\bar{\epsilon}$! If the ambient temperature were to remain fixed,
1386: $\bar{\epsilon}$ should relax to $\epsilon_{eq}(T)$ since at that
1387: point detailed balance applied to the microscopic rates will replenish
1388: any states from which the system locally escapes. (We remind the
1389: reader that $\epsilon$ denotes a (library) free energy $\phi$ without
1390: the vibrational entropy contribution.) The typical escape rate at any
1391: value of $\bar{\epsilon}$ will be given by Eq.(\ref{k(x)g}) at ambient
1392: temperature $T$. Thus the equation of motion for will $\bar{\epsilon}$
1393: have the form
1394: \begin{widetext}
1395: \begin{equation}
1396: \bar{\epsilon}-\epsilon_{eq}(T) = - \int_{-\infty}^t d t_1
1397: \dot{\rho}(t-t_1; T(t), \tau[\bar{\epsilon}(t_1), T(t_1)]) \:
1398: \{\bar{\epsilon}(t_1) - \epsilon_{eq}[T(t_1)]\}.
1399: \end{equation}
1400: \end{widetext}
1401: where $\rho$ is the relaxation function when the statistics of
1402: libraries are known and fixed by $\bar{\epsilon}(t_1)$, $T(t_1)$ is
1403: the ambient temperature at time $t_1$ - we assume vibrational energies
1404: equilibrate quickly by thermal conduction. $\rho$ not only depends on
1405: $T(t_1)$ and $\bar{\epsilon}$ but contains the parameter $\beta$ which
1406: in turn depends on the variance of energies of the local region. As a
1407: first approximation the variance can be taken as that characteristic
1408: of an equilibrium system at a temperature $T^*$ such that
1409: \begin{equation}
1410: \epsilon_{eq}(T^*) = \bar{\epsilon}.
1411: \end{equation}
1412: In this approximation, we may use Eq.(\ref{beta_ratio}) for
1413: \begin{equation}
1414: \beta \equiv \beta_{in}[T(t),T(\bar{\epsilon})].
1415: \end{equation}
1416: We see that if the shape of the distribution of local libraries does
1417: not change much, the RFOT picture leads to a situation where a single
1418: parameter $\bar{\epsilon}$ suffices to characterize the
1419: system. $T^*(\bar{\epsilon})$ thus essentially fixes the ``fictive
1420: temperature'' $T_f$ in the NTM phenomenology, albeit with different
1421: expressions for $\tau(T, T_f)$ and $\beta=\beta(T, T_f)$. The
1422: non-Gaussian statistics alluded to earlier, however suggests the use
1423: of a single fictive temperature is only approximate and that a more
1424: complete characterization of the statistics may be needed. Multiple
1425: fictive temperatures defining the higher moments could in principle be
1426: defined. In addition spatial fluctuations of fictive temperatures are
1427: needed to capture the co-existence of equilibrated and nonequilibrated
1428: domains in the mosaic structure. At least for moderate quenches it may
1429: be possible to ignore the spatial inhomogenity and merely monitor the
1430: fluctuations in energies of domains as a secondary variable.
1431:
1432: \vspace{-5mm}
1433:
1434: \section{Summary}
1435:
1436: We described a local energy landscape theory of the dynamics of
1437: supercooled liquids and glasses. In the equilibrated supercooled
1438: regime this theory is just a microcanonical ensemble reformulation of
1439: the random first order transition theory and its notion of entropic
1440: droplets. New results are obtained in the aging regime of
1441: nonequilibrium quenched glasses. The key equation is (\ref{k(x)g})
1442: which shows it is the difference between the equilibrated free energy
1443: at the quench temperature and the initial free energy of the
1444: particular frozen state that drives motions. The theory approximately
1445: reproduces the phenomenology of Narayanaswamy and Moynihan and
1446: Tool. Thus the nonlinearity parameter in NMT theory can be
1447: calculated. This parameter is shown to be correlated with the
1448: supercooled liquid's fragility in agreement with experiment. This
1449: correlation is quantitatively very similar to that obtained by the
1450: Hodge-Scherer-Adam-Gibbs extrapolation that assumes configurational
1451: entropy is fixed at $T_g$. Deviations from that extrapolation which
1452: assumes Arrhenius behavior in the glassy state are predicted
1453: however. The previously puzzling, modest non-Arrhenius temperature
1454: dependence of relaxation observed within the glassy state is explained
1455: by the RFOT theory, although it is a small effect. The variations of
1456: nonexponentiality of relaxation in the glassy state are predicted but
1457: are also rather small in the moderately quenched regime. The
1458: comparison of nonexponentiality with experiment is less conclusive
1459: than the comparison of mean relaxation rates however, owing to the
1460: difficulty of accessing the complete relaxation behavior during the
1461: quench.
1462:
1463:
1464: One advantage of the aging theory based on RFOT theory is that in
1465: principle the behavior upon very deep quenches is predicted. Most
1466: importantly $\beta$ is predicted to continue to decrease with quench
1467: temperature. We hope that more experiments in this regime will be
1468: done. We have noted however that some non-universal effects may enter
1469: for such quenches. Also the kinematics of deep quenches may be complex
1470: owing to spatial fluctuations of fictive temperature predicted by our
1471: theory.
1472:
1473: Although the present approach justifies to some extent the use of a
1474: single fictive temperature to characterize the glassy state the
1475: limitations of this idea have been made clear. A straightforward
1476: extension of the usual formulation to include fluctuations in the
1477: moderately cooled regime more fully was proposed. In the strongly
1478: cooled regime, the theory predicts patches of especially slowly
1479: relaxing regions will appear. This prediction may be tested by single
1480: molecule imaging approaches \cite{RussellIsraeloff,VandenBout}.
1481:
1482:
1483: Acknowledgement: Peter G. Wolynes would like to thank Xiaoyu Xia for
1484: early discussions on this topic. The work was supported by the
1485: National Science Foundation grant CHE0317017. Vassiliy Lubchenko is
1486: grateful for the kind support and indulgence of Robert Silbey at
1487: M.I.T.
1488:
1489: \vspace{-5mm}
1490:
1491: \appendix
1492:
1493: \section*{Appendix}
1494:
1495: \setcounter{equation}{0}
1496: \renewcommand{\theequation}{A.\arabic{equation}}
1497:
1498: Several auxiliary results are derived in this Appendix (here,
1499: $k_B=1$). First, we derive formula (\ref{nonArrh_rat}) from the main
1500: text. Using $s = -\partial f/\partial T$, one has (neglecting
1501: differences in vibrational entropy)
1502: \begin{eqnarray}
1503: %\begin{equation}
1504: f_{eq}(T) &=& \phi_K - \int_{T_K}^{T} s dT \nonumber \\ &=& \phi_K -
1505: \Delta c_p(T_g) T_g \left(\frac{T-T_K}{T_K} - \ln \frac{T}{T_K}
1506: \right), \hspace{7mm}
1507: \label{-10}
1508: \end{eqnarray}
1509: %\end{equation}
1510: where we have used Angell's empirical form \cite{XW} for the
1511: configurational entropy, also used in the main text: $s_c(T) = \Delta
1512: c_p(T_g) T_g(1/T_K-1/T), \mbox{ } \Delta c_p(T) = \Delta c_p(T_g)
1513: (T_g/T)$. Noting that $f_{eq}(T_g) = \phi_g - T_g s_c(T_g)$ fixes the
1514: ideal glass state energy $\phi_K$:
1515: \begin{equation}
1516: \phi_K = \phi_g - \Delta c_p(T_g) T_g \ln(T_g/T_K).
1517: \label{-9}
1518: \end{equation}
1519: Eqs.(\ref{-10}) and (\ref{-9}) immediately yield
1520: Eq.(\ref{nonArrh_rat}).
1521:
1522: Next, we derive the more accurate expression for the
1523: $\beta_{n.e.}(T)/\beta(T_g)$, as follows from Eq.(\ref{beta}). We also
1524: provide the equilibrium value $\beta_{eq}$ of the non-exponentiality
1525: parameter $\beta$, previously obtained by Xia and Wolynes \cite{XW},
1526: in terms of experimental parameters $T_K/T_g$ and
1527: $\ln(\tau_Q/\tau_{micro})$, used throughout this paper.
1528:
1529: First we derive $\beta_{eq}$. According to Ref.\cite{XW}, the barrier
1530: fluctuations are directly related to the local fluctuations in the
1531: configurationlal entropy at the scale $N^*$ corresponding to a local
1532: equilibrium unit, of which the global composite landscape is
1533: comprised: $\delta \Delta F^\ddagger/\Delta F^\ddagger = \delta
1534: S_c/S_c = \sqrt{\Delta c_P N^*}/s_c N^*$. This yields:
1535: \begin{equation}
1536: \frac{\delta \Delta F^\ddagger}{T} = \frac{\Delta F^\ddagger}{T}
1537: \frac{\sqrt{\Delta c_P}}{s_c \sqrt{N^*}}.
1538: \end{equation}
1539:
1540: It directly follows from $F^\ddagger(N) = \gamma \sqrt{N} - Ts_cN$,
1541: $(dF/dN)_{N=N^\ddagger} = 0$, and $F(N^*) = 0$ that
1542: \begin{equation}
1543: \frac{\Delta F^\ddagger}{T} = \frac{(\gamma/T)^2}{4s_c}
1544: \label{1}
1545: \end{equation}
1546: and
1547: \begin{equation}
1548: \sqrt{N^*} = \frac{\gamma/T}{s_c}.
1549: \end{equation}
1550: Using Eq.(\ref{tQ}), Eq.(\ref{1}) and the temperature independence of
1551: the $\gamma/T$ ratio in equilibrium \cite{XW}, one gets
1552: \begin{equation}
1553: (\gamma/T) = (\gamma/T)_{T=T_g} = 2 \sqrt{s_c(T_g)
1554: \ln(\tau_Q/\tau_{micro})}.
1555: \end{equation}
1556: With the help of the equations above and Angell's empirical form for
1557: the configurational entropy one easily obtains that
1558: \begin{equation}
1559: \frac{\delta \Delta F^\ddagger}{T} =
1560: \frac{\sqrt{\ln(\tau_Q/\tau_{micro})}}{2}
1561: \frac{\sqrt{1/T_K-1/T_g}}{\sqrt{T}(1/T_K-1/T)}.
1562: \label{2}
1563: \end{equation}
1564: This and Eq.(\ref{beta}) can be used to compute the
1565: $\beta_{eq}(T)/\beta(T_g)$ ratio.
1566:
1567: The non-equilibrium $\beta$ requires even less effort. Using
1568: Eq.(\ref{dFdFratio}), one gets
1569: \begin{equation}
1570: \frac{\delta \Delta F^\ddagger_{n.e.}}{\delta \Delta F^\ddagger_g} =
1571: \left[ \frac{F_{n.e.}^\ddagger(T)}{F_g^\ddagger} \right]^2 \left[
1572: \frac{\gamma(T_g)}{\gamma(T)} \right]^2.
1573: \end{equation}
1574: Hence,
1575: %\vspace{-10mm}
1576: \begin{equation}
1577: \beta_{n.e.}(T) = \left[ 1+ \left(\frac{\delta \Delta
1578: F^\ddagger_g}{T}\right)^2
1579: \left[\frac{F_{n.e.}^\ddagger(T)}{F_g^\ddagger} \right]^4 \left[
1580: \frac{\gamma(T_g)}{\gamma(T)} \right]^4 \right]^{-1/2}.
1581: \end{equation}
1582: The ratio $F_{n.e.}^\ddagger(T)/F_g^\ddagger$ was computed in the
1583: beginning of the Appendix and given in the main text as
1584: Eq.(\ref{nonArrh_rat}); $\delta \Delta F^\ddagger(T_g)$ is obtained
1585: from Eq.(\ref{2}) at $T=T_g$.
1586:
1587: %\vspace{-5mm}
1588:
1589: \begin{thebibliography}{99}
1590:
1591: %\vspace{-5mm}
1592:
1593: \bibitem{OLSW} J.N. Onuchic, Z. Luthey-Schulten and P.G. Wolynes,
1594: Ann. Rev. Phys. Chem {\bf 48}, 545-600 (1997).
1595:
1596: \bibitem{Wales} D.J. Wales, Energy Landscapes with application to
1597: Clusters, Biomolecules and Glasses (Cambridge U. Press, Cambridge,
1598: 2003)
1599:
1600: \bibitem{KW_PRA} T.R.Kirkpatrick and P.G.Wolynes, Phys. Rev. A {\bf
1601: 31}, 939 (1985).
1602:
1603: \bibitem{KW_PRB} T.R.Kirkpatrick and P.G.Wolynes, Phys. Rev. B {\bf
1604: 36}, 8552 (1987).
1605:
1606: \bibitem{KTW} T.R.Kirkpatrick, D.Thirumalai and P.G.Wolynes,
1607: Phys. Rev. A {\bf 40}, 1045 (1989).
1608:
1609: \bibitem{XW} X.Xia and P.G.Wolynes, Proc. Nat. Acad. Sci. {\bf 97},
1610: 2990 (2000).
1611:
1612: \bibitem{XWbeta} X.Xia and P.G.Wolynes, Phys. Rev. Lett. {\bf 86}, 5526 (2001).
1613:
1614: \bibitem{LW_soft} V.Lubchenko and P.G.Wolynes, J. Chem. Phys. {\bf
1615: 119}, 9088 (2003).
1616:
1617: \bibitem{LW} V.Lubchenko and P.G.Wolynes, Phys. Rev. Lett. {\bf 87},
1618: 195901 (2001).
1619:
1620: \bibitem{LW_BP} V.Lubchenko and P.G.Wolynes,
1621: Proc. Natl. Acad. Sci. USA {\bf 100}, 1515 (2003).
1622:
1623: \bibitem{Bouchaud} J.P. Bouchaud, et al. in ``Spin Glasses and Random
1624: Fields,'' A.P. Young, ed. World Scientific, Singapore 1998.
1625:
1626: \bibitem{Struik} L.C. Struik, Physical Aging in Amorphous Polymers and
1627: other materials (Elsevier, Houston, 1978).
1628:
1629: \bibitem{Tool} A.Q.Tool, J. Am. Ceram. Soc. {\bf 29}, 240 (1946).
1630:
1631: \bibitem{Narayanaswamy} O.S.Narayanaswamy, J. Am. Ceram. Soc. {\bf
1632: 54}, 491 (1971).
1633:
1634: \bibitem{Moynihan} C.T.Moynihan, A.J.Easteal, M.A.Debolt, and
1635: J.Tucker, J. Am. Ceram. Soc. {\bf 59}, 12 (1976).
1636:
1637: \bibitem{Scherer} G.W. Scherer, J. Am. Chem. Soc. {\bf 67}, 504 (1984)
1638:
1639: \bibitem{Hodge} I.M.Hodge, Macromol. {\bf 20}, 2897 (1987);
1640: J. Noncryst Solids {\bf 169}, 211 (1994).
1641:
1642: \bibitem{AdamGibbs} G.Adam and J.H.Gibbs, J. Chem. Phys. {\bf 43}, 139 (1965).
1643:
1644: \bibitem{Alegria} A.Alegria, E.Guerrica-Echevarr\'{i}a, L.Goitiand\'{i}a,
1645: I.Teller\'{i}a, and J.Colmenero, Macromol. {\bf 28}, 1516 (1995).
1646:
1647:
1648: \bibitem{MPV} M. Mezard, G.Parisi and M.Virasoro, Spin Glass Theory
1649: and Beyond (World Scientific, Singapore, 1985)
1650:
1651: \bibitem{Coluzzi} B. Coluzzi, G. Parisi and P. Verrocchio,
1652: J. Chem. Phys. {\bf 112}, 2933 (2000)
1653:
1654: \bibitem{inherent} F.H.Stillinger and T.A.Weber, Phys. Rev. A {\bf
1655: 25}, 978 (1982).
1656:
1657: \bibitem{Mezei} F.Mezei, in ``Liquids, Freezing and the Glass
1658: Transition'', ed. J.P.Hansma, D.Levesque and J.Zinn-Justin,
1659: North-Holland, Amsterdam, p.629, 1991.
1660:
1661: \bibitem{SW} J.P.Stoessel and P.G.Wolynes,
1662: J. Chem. Phys. {\bf 80}, 4502 (1984).
1663:
1664: \bibitem{SSW} Y.Singh, J.P.Stoessel, and P.G.Wolynes,
1665: Phys. Rev. Lett. {\bf 54}, 1059 (1985).
1666:
1667: \bibitem{DasguptaValls} C.Dasgupta and O.T.Valls, Phys. Rev. E {\bf
1668: 59}, 3123 (1999).
1669:
1670: \bibitem{HW} R.W.Hall and P.G.Wolynes, Phys. Rev. Lett. {\bf 90},
1671: 085505 (2003).
1672:
1673: \bibitem{Parisi} G.Parisi, Nuovo Cimento {\bf 16D}, 939 (1994).
1674:
1675: \bibitem{Villain} J.Villain, J. Physique {\bf 46}, 1843 (1985).
1676:
1677: \bibitem{Ediger} M. Ediger, Ann. Rev. Phys. Chem {\bf 51}, 99 (2000)
1678:
1679: \bibitem{Nelson} D.R.Nelson, ``Defects and Geometry in Condensed
1680: Matter Physics'', Cambridge UniversityPress, 2002.
1681:
1682: \bibitem{Kivelson} D.Kivelson, S.A.Kivelson, X.L.Zhao, Z.Nussinov, and
1683: G.Tarjus, Physica A {\bf 219}, 27 (1995).
1684:
1685: \bibitem{Andersen} G.H.Fredrickson and H.C.Andersen,
1686: J. Chem. Phys. {\bf 83}, 5822 (1985).
1687:
1688: \bibitem{Eastwood} M.P.Eastwood and P.G.Wolynes, Europhys. Lett. {\bf
1689: 60}, 587 (2002).
1690:
1691: \bibitem{Nussinov} Z.Nussinov, Phys. Rev. B {\bf 64}, 014208 (2004)
1692:
1693: \bibitem{Angell_cp1} C.Alba, L.E.Busse, D.J.List, and C.Angell,
1694: J. Chem. Phys. {\bf 92}, 617 (1990)
1695:
1696: \bibitem{Angell_cp2} R.Richert and C.A. Angell, J. Chem. Phys. {\bf
1697: 108}, 9016 (1998)
1698:
1699: \bibitem{McKenna} G.B.McKenna and C.A.Angell, J. Noncryst. Solids {\bf
1700: 131-133}, 528 (1991)
1701:
1702: \bibitem{app_phys_rev} C.A.Angell, K.L.Ngai, G.B.McKenna, P.F.Millan, and
1703: S.W.Martin, Appl. Phys. {\bf 88}, 3113 (2000)
1704:
1705: \bibitem{Castaing} B.Castaing and J.Souletie, J de Phys I {\bf 1}, 403
1706: (1991); C. Monthus and J. Bouchaud, J. Phys. A {\bf 29}, 3847 (1990)
1707:
1708: \bibitem{Leheny} R.L.Leheny and S.R.Nagel, Phys. Rev. B {\bf 57}, 5154
1709: (1998).
1710:
1711: \bibitem{slaving} H. Frauenfelder, V. Lubchenko and P.G. Wolynes,
1712: unpublished.
1713:
1714: \bibitem{Dyre} N.B. Olsen, T. Christensen and J. Dyre,
1715: Phys. Rev. Lett. {\bf 86}, 1271 (2001).
1716:
1717: \bibitem{buffing} S.S.Plotkin and P.G.Wolynes,
1718: Proc. Natl. Acad. Sci. {\bf 100}, 4417 (2003).
1719:
1720: \bibitem{MillerMcPhail} R.S. Miller and R.A. McPhail,
1721: J. Chem. Phys. {\bf 106}, 3393 (1997)
1722:
1723: \bibitem{RussellIsraeloff} E.V.Russell and N. Israeloff, Nature {\bf
1724: 408}, 695 (2000).
1725:
1726: \bibitem{VandenBout} L.A.Deschenes, D.A.Vanden Bout, Science {\bf 292}
1727: 255 (2001).
1728:
1729: \end{thebibliography} \end{document}
1730: