1: \documentclass[twocolumn, nofootinbib]{revtex4}
2: %\documentclass[preprint ,nofootinbib]{revtex4}
3: \usepackage{amsmath} \usepackage{graphicx} \usepackage{dcolumn}
4: \usepackage{bm} \usepackage{epsf} \usepackage{amssymb}
5:
6:
7: \begin{document}
8:
9: \sloppy
10:
11: \newcommand{\prtl}{\partial}
12: \newcommand{\la}{\left\langle}
13: \newcommand{\ra}{\right\rangle}
14: \newcommand{\dla}{\la \! \! \! \la}
15: \newcommand{\dra}{\ra \! \! \! \ra}
16: \newcommand{\we}{\widetilde}
17: %
18: \newcommand{\ws}{\we{s}}
19: \newcommand{\wF}{\we{F}}
20: \newcommand{\sins}{{\mbox{\scriptsize ins}}}
21: \newcommand{\smfp}{{\mbox{\scriptsize mfp}}}
22: \newcommand{\sfull}{{\mbox{\scriptsize full}}}
23: \newcommand{\smp}{{\mbox{\scriptsize mp}}}
24: \newcommand{\sLZ}{{\mbox{\scriptsize LZ}}}
25: \newcommand{\sph}{{\mbox{\scriptsize ph}}}
26: \newcommand{\sinhom}{{\mbox{\scriptsize inhom}}}
27: \newcommand{\sneigh}{{\mbox{\scriptsize neigh}}}
28: \newcommand{\srlxn}{{\mbox{\scriptsize rlxn}}}
29: \newcommand{\svibr}{{\mbox{\scriptsize vibr}}}
30: \newcommand{\smicro}{{\mbox{\scriptsize micro}}}
31: \newcommand{\smax}{{\mbox{\scriptsize max}}}
32: \newcommand{\seq}{{\mbox{\scriptsize eq}}}
33: \newcommand{\sstr}{{\mbox{\scriptsize str}}}
34: \newcommand{\teq}{{\mbox{\tiny eq}}}
35: \newcommand{\sinn}{{\mbox{\scriptsize in}}}
36: \newcommand{\tin}{{\mbox{\tiny in}}}
37: \newcommand{\scr}{{\mbox{\scriptsize cr}}}
38: \newcommand{\sscr}{{\mbox{\scriptsize scr}}}
39: \newcommand{\sL}{{\mbox{\scriptsize L}}}
40: \newcommand{\sTS}{{\mbox{\scriptsize TS}}}
41: \newcommand{\stheor}{{\mbox{\scriptsize theor}}}
42: \newcommand{\sGS}{{\mbox{\scriptsize GS}}}
43: \newcommand{\sNMT}{{\mbox{\scriptsize NMT}}}
44: \newcommand{\sRFOT}{{\mbox{\scriptsize RFOT}}}
45: \newcommand{\tRFOT}{{\mbox{\tiny RFOT}}}
46: \newcommand{\sbulk}{{\mbox{\scriptsize bulk}}}
47: \newcommand{\tbulk}{{\mbox{\tiny bulk}}}
48: \newcommand{\tDC}{{\mbox{\tiny DC}}}
49: \newcommand{\sauto}{{\mbox{\scriptsize auto}}}
50: \newcommand{\tauto}{{\mbox{\tiny auto}}}
51: \newcommand{\sescape}{{\mbox{\scriptsize escape}}}
52: \newcommand{\sth}{{\mbox{\scriptsize th}}}
53: \newcommand{\svib}{{\mbox{\scriptsize vib}}}
54: %
55: \newcommand{\sT}{{\mbox{\scriptsize T}}}
56: \newcommand{\sTLS}{{\mbox{\scriptsize TLS}}}
57: \newcommand{\sd}{{\mbox{\scriptsize d}}}
58: \newcommand{\sext}{{\mbox{\scriptsize ext}}}
59: \newcommand{\scav}{{\mbox{\scriptsize cav}}}
60: \newcommand{\bmu}{\bm \mu}
61: \newcommand{\bE}{\bm E}
62: \newcommand{\bD}{\bm D}
63: \newcommand{\bd}{\bm d}
64: \newcommand{\br}{\bm r}
65: \newcommand{\bj}{\bm j}
66: \newcommand{\bi}{\bm i}
67: \newcommand{\bnabla}{\bm \nabla}
68: \newcommand{\smol}{{\mbox{\scriptsize mol}}}
69: %
70: \def\Xint#1{\mathchoice
71: {\XXint\displaystyle\textstyle{#1}}%
72: {\XXint\textstyle\scriptstyle{#1}}%
73: {\XXint\scriptstyle\scriptscriptstyle{#1}}%
74: {\XXint\scriptscriptstyle\scriptscriptstyle{#1}}%
75: \!\int}
76: \def\XXint#1#2#3{{\setbox0=\hbox{$#1{#2#3}{\int}$}
77: \vcenter{\hbox{$#2#3$}}\kern-.5\wd0}}
78: \def\ddashint{\Xint=}
79: \def\dashint{\Xint-}
80: %
81: %
82: %
83: \title{Charge and momentum transfer in supercooled melts: \\ Why
84: should their relaxation times differ?}
85:
86: \author{Vassiliy Lubchenko} \affiliation{Department of Chemistry,
87: University of Houston, Houston, TX 77204-5003}
88:
89: %\date{\today}
90:
91: \begin{abstract}
92:
93: The steady state values of the viscosity and the intrinsic
94: ionic-conductivity of quenched melts are computed, in terms of
95: independently measurable quantities. The frequency dependence of the
96: ac dielectric response is estimated. The discrepancy between the
97: corresponding characteristic relaxation times is only apparent; it
98: does not imply distinct mechanisms, but stems from the intrinsic
99: barrier distribution for $\alpha$-relaxation in supercooled fluids
100: and glasses. This type of intrinsic ``decoupling'' is argued not to
101: exceed four orders in magnitude, for known glassformers. We explain
102: the origin of the discrepancy between the stretching exponent
103: $\beta$, as extracted from $\epsilon(\omega)$ and the dielectric
104: modulus data. The actual width of the barrier distribution always
105: grows with lowering the temperature. The contrary is an artifact of
106: the large contribution of the dc-conductivity component to the
107: modulus data. The methodology allows one to single out other
108: contributions to the conductivity, as in ``superionic'' liquids or
109: when charge carriers are delocalized, implying that in those
110: systems, charge transfer does not require structural
111: reconfiguration.
112:
113:
114: \end{abstract}
115:
116: \date{\today}
117:
118: \maketitle
119:
120: \section{Introduction}
121:
122:
123: Molecular motions in deeply supercooled melts and glasses are
124: cooperative so that transporting a single molecule requires concurrent
125: rearrangement of up to several hundreds of surrounding molecules. Such
126: high degree of cooperativity results in high barriers even for the
127: smallest scale molecular translations. These high barriers underly the
128: slow, activated dynamics in deeply supercooled melts and the emergence
129: of a mechanically stable aperiodic lattice, if a melt is quenched
130: sufficiently rapidly. The Random First Order Transition (RFOT)
131: methodology, developed by Wolynes and coworkers, provides a
132: constructive microscopic picture of the structural rearrangements in
133: supercooled melts and quenched glasses. The RFOT has quantitatively
134: explained or predicted the signature phenomena accompanying the glass
135: transition, including the connection between the thermodynamic and
136: kinetic anomalies \cite{KTW, XW, XWbeta}, the length scale of the
137: cooperative rearrangements \cite{XW}, deviations from Stokes-Einstein
138: hydrodynamics \cite{XWhydro}, aging \cite{LW_aging}, the low
139: temperature anomalies \cite{LW, LW_BP, LW_RMP}, and more. (See
140: \cite{LW_ARPC} for a recent review.)
141:
142: Perhaps the most dramatic experimental signature of the glass
143: transition is the rapid super-Arrhenius growth of the relaxation times
144: with lowering the temperature, from about a picosecond, near the
145: melting point $T_m$, to as long as hours, at the glass transition
146: temperature $T_g$. These relaxation times are deduced via several
147: distinct experimental methodologies and all display an extraordinarily
148: broad dynamical range. Nevertheless, making detailed comparisons
149: between those distinct methodologies has required additional
150: phenomenological assumptions. Mysteriously, these comparisons show a
151: significant degree of mismatch, sometimes by several orders of
152: magnitude. For example, the phenomenological ``conductivity relaxation
153: time'' $\tau_\sigma$ \cite{Macedo1972}, is consistently shorter than
154: the mechanical relaxation time $\tau_s$, especially at lower
155: temperatures. The apparent time scale separation varies wildly from
156: system to system: For instance in molten nitrates, it is about four
157: orders of magnitude at $T_g$ \cite{Howell1974}, while in silver
158: containing superionic melts, the $\tau_s/\tau_\sigma$ ratio becomes as
159: large as $10^{11}$ \cite{McLinAngell}, i.e. almost as much as the
160: whole dynamical range accessible to the melt! This disparity
161: suggested that the mechanical relaxation and the electrical
162: conductivity in these systems were in fact due to distinct mechanisms:
163: At higher temperatures, the time scale separation is small so that the
164: two processes strongly affect each other, or ``mix'', while at lower
165: temperatures, the processes become increasingly ``decoupled''
166: \cite{Angell}. At such low temperatures, the mechanical relaxation
167: occurs via the aforementioned, activated concerted events, also called
168: the primary, or $\alpha$-relaxation. Other processes that seem to
169: decouple from the mechanical relaxation include nuclear spin
170: relaxation, rotational diffusion, and the diffusion of small
171: probes. (For reviews, see \cite{Angell, Ngai, CiceroneEdiger})
172:
173: Here we focus on two specific transport phenomena: low-frequency
174: momentum transfer, i.e. the viscous response, and the ionic conduction
175: in supercooled melts. Notwithstanding the complications needed to
176: analyze the electrical modulus data \cite{Elliott1994, Roling1998,
177: ColeTombari, Sidebottom1995, Moynihan1994, Dyre1991, Doi1988}, the
178: mismatch between the typical relaxation times, corresponding to the
179: two types of transport, is clearly present. Furthermore, in the case
180: of superionic compounds, one may show (see below) that conduction
181: occurs without distorting the liquid's structure beyond the typical
182: vibrational displacements. This is much less obvious for compounds
183: where the ionic motions are ``decoupled'' from the bulk structural
184: relaxation by four orders of magnitude or less, the latter dynamic
185: range comparable to the breadth of the $\alpha$-peaks in dielectric
186: dispersion in insulating melts near $T_g$ (see
187: e.g. \cite{LSBL}). Accounting for the distribution width is essential
188: here because the viscosity and conductivity are distinct, in fact
189: exactly reciprocal types of response: In momentum transfer, the
190: velocity gradient is the source, and the passed-on rate-of-force is
191: the response. In charge transfer, the force on the ion is the source,
192: while the arising velocity field is the response. Consistent with this
193: general notion, in computing the viscosity, we will average the
194: relaxation time, with respect to local inhomogeneities, while the
195: intrinsic ionic conductivity will be determined by the average {\em
196: rate} of $\alpha$-relaxation. Because of the mentioned, extremely
197: broad distribution of structural relaxation times $\tau$, the quantity
198: $ \la \tau \ra \la \tau^{-1} \ra $ may reach several orders of
199: magnitude, and so an apparent decoupling is indeed expected; no
200: additional microscopic mechanisms need to be invoked.
201:
202: The microscopic calculation and comparison, of the viscosity and the
203: ionic conductivity, are thus the main focus of this article. The two
204: quantities are computed, in terms of the barrier distribution and
205: other measurable material properties, in Sections \ref{visc} and
206: \ref{cond} respectively. To perform comparisons with experiment and
207: assess the upper limit on the ``inherent decoupling'' between the two
208: phenomena, we will discuss the barrier distribution in some detail, in
209: Section \ref{Barrier}. We will find that indeed, the degree of
210: decoupling should increase with the width of the barrier distribution,
211: and hence at lower temperatures, as demonstrated by the RFOT
212: methodology \cite{XWbeta}. We will assess the deviation between
213: $\alpha$-relaxation times, as deduced from viscosity, ionic
214: conductance, and the maximum in $\epsilon''(\omega)$. Further, we
215: will exemplify potential ambiguities in using the dielectric modulus
216: formalism in estimating the relaxation time distribution. The latter
217: techique has suggested that for some substances, the distribution
218: width in fact decreases with lowering the temperature, in conflict
219: with the present results and the correlation between the stretching
220: exponent $\beta$ and temperature, predicted earlier by the RFOT theory
221: \cite{XWbeta}. We will see that the conflict is artificial and results
222: from the large contribution of the ac-component to the modulus data,
223: consistent with earlier, phenomenological arguments \cite{Johari1988,
224: Roling1998}.
225:
226: \section{Viscosity}
227: \label{visc}
228:
229:
230: The key microscopic notion behind the RFOT methodology is that, {\em
231: regardless of the detailed interparticle potentials}, local
232: aperiodic arrangements in classical condensates become meta{\em
233: stable} below a certain temperature $T_A$ (or above a certain
234: density) \cite{dens_F1, dens_F2}. Chemical detail and molecular
235: structure affect the value of $T_A$, and the viscosity of the
236: fluid. If the viscosity is high enough, one may cool the liquid so
237: that it becomes locally trapped in metastable minima, while avoiding
238: the nucleation of a periodic crystal, which would have been the lowest
239: free energy state. Another system-dependent quantity is the size $a$
240: of the elemental structural unit in the metastable liquid, or
241: ``bead'': the length $a$ plays the role of the lattice spacing in the
242: aperiodic structure, and is indeed quite analogous to the size of the
243: unit cell in an oxide crystal, or it may correspond to the size of a
244: rigid monomer or side chain in a polymer. The size $a$ characterizes
245: the range of the local chemical order that sets in during a crossover,
246: at a temperature $T_\scr$, from collision dominated transport to
247: activated dynamics \cite{LW_soft}. The temperature $T_\scr$ is related
248: to the meanfield temperature $T_A$ but is always smaller. The bead
249: size $a$ may be unambiguously determined from the fusion entropy of
250: the corresponding crystal, when the latter entropy is known
251: \cite{LW_soft}, or else can be computed from the fragility $D$ using
252: the universal relationship between the latter and the heat capacity
253: jump at $T_g$: $D = 32/\Delta c_p$, as derived in RFOT \cite{XW,
254: StevensonW}. Alternatively, if the configuration entropy can be
255: reliably estimated, one may use the RFOT-derived relation for the
256: configurational entropy per bead $s_c \sim .8 k_B$ \cite{XW}, which is
257: somewhat sensitive to the barrier-softening effects though
258: \cite{LW_soft}.
259:
260: \begin{figure}[t]
261: \includegraphics[width= .95 \columnwidth]{xiFN.eps}
262: \caption{\label{xiFN} {\bf (a):} Typical nucleation profile of one
263: aperiodic lattice, within another, in deeply supercooled liquids.
264: $N \equiv (4\pi/3) (r/a)^3$, $N^\ddagger$ is the typical
265: transition state size: $(dF/dN)_{N^\ddagger} = 0$. $\xi$ is the
266: volumetrically defined cooperative length: $N^* \equiv (\xi/a)^3$,
267: where $F(N^*) = 0$. {\bf (b):} Cartoon of a structural
268: rearrangement. The shown magnitude of $\xi$ corresponds to a
269: temperature near $T_g$ on 1 hour scale. The two sets of circles -
270: solid and dashed ones - denote two alternative structural
271: states. $d_L \simeq a/10$, or the ``Lindemann length'', is the
272: typical bead displacement during a transition.}
273: \end{figure}
274:
275: Once locally metastable, the liquid may reconfigure but in an
276: activated fashion, i.e. by nucleating a new aperiodic structure within
277: the present one. Such activated events occur, on average, once per
278: typical $\alpha$-relaxation time $\tau$, per region of size $\xi$.
279: The nucleus grows in a sequence of individual, nearly-barrierless bead
280: moves of length $d_L \simeq a/10$ \cite{dens_F1, L_Lindemann} and time
281: $\tau_\smicro \simeq 1$ ps \cite{L_Lindemann}. The overall sequence of
282: elemental moves typically corresponds to the following activation
283: profile, see Fig.\ref{xiFN}(a):
284: \begin{equation} \label{F(N)} F(N) = \gamma \sqrt{N} - T s_c N,
285: \end{equation}
286: where $N$ is the size of a reconfigured region. The ``surface term''
287: $\gamma \sqrt{N}$ is the mismatch penalty for creating one aperiodic
288: structure within another. $s_c$ is the excess, ``configurational''
289: entropy of the liquid per bead, hence the entropic, bulk term $(-T s_c
290: N)$, which drives the transition and reflects the multiplicity of
291: possible aperiodic arrangements in a region of size $N$. The maximum
292: of the profile:
293: \begin{equation} \label{Fsc} F^\ddagger = \mbox{max}\{F(N)\} =
294: \frac{\gamma^2}{4 s_c T}
295: \end{equation}
296: is achieved at $N^\ddagger= N^*/4$, where $F(N^*)=0$, so that the
297: typical relaxation time is
298: \begin{equation} \label{tau} \tau = \tau_\smicro e^{F^\ddagger/k_B T}
299: \equiv \tau_\smicro e^{D T_K/(T - T_K)}.
300: \end{equation}
301: This formula works well at $\tau > 1$ nsec or so. The form on the
302: r.h.s. is the Vogel-Fulcher law, derived in the RFOT.
303:
304: The end result of a cooperative, activated event is a reconfigured
305: region of size $\xi$, where each of the $N^* \equiv (\xi/a)^3$ beads
306: has moved the Lindemann length $d_L$, or so, see Fig.\ref{xiFN}(b).
307: Both $\xi$ and the nucleation critical size, $r^\ddagger =
308: \xi/4^{1/3}$, increase with lowering the temperature, roughly as
309: $r^\ddagger \propto \xi \propto 1/(T - T_K)^{2/3}$ \cite{KTW, XW,
310: LW_soft}. Here, $T_K$ is the so called ideal glass transition
311: temperature, where the excess liquid entropy $s_c$, extrapolated below
312: $T_g$, would vanish \cite{Kauzmann}. At $T_g$, $\xi$ is still quite
313: modest, only about six beads across \cite{XW, LW}. Activated
314: transport becomes dominant below the temperature $T_\scr$, such that
315: $r^\ddagger(T_\scr) = a$, which corresponds, apparently universally,
316: to $\tau/\tau_\smicro \simeq 10^3$, or viscosities on the order of 10
317: Ps \cite{LW_soft}. At times shorter than $10^3 \tau_\smicro$, one may
318: then speak of a local aperiodic lattice on length scales of the
319: cooperativity size $\xi$, since the slow structural reconfigurations
320: have now time-scale separated from the vibrations
321: \cite{LW_soft}. Because of the local nature of structural relaxations,
322: one speaks of dynamic heterogeneity, or a ``mosaic'' of cooperative
323: rearrangments \cite{XW}. The heterogeneity is two-fold: On the one
324: hand, a {\em local} rearrangment implies that the surrounding
325: structure is static during the transition, up to vibrations. On the
326: other hand, because of the spatial and temporal variation in the local
327: density of states (and hence variations in $s_c$), local
328: reconfigurations are generally subject to somewhat different barriers
329: in different regions \cite{XW} (see Eq.(\ref{Fsc}) and also Section
330: \ref{Barrier}).
331:
332: Computation of the viscosity in such a dynamically heterogeneous
333: environment may be done in two steps: First compute the viscosity in a
334: medium with a homogeneous relaxation time, call it $\tau'$, and then
335: average out with respect to the true distribution of the relaxation
336: times. This procedure is valid in view of the equivalence of time and
337: ensemble average. When the relaxation rate is strictly spatially
338: homogeneous, one may formally define a diffusion constant for an
339: individual bead: $D' = d_L^2/6 \tau'$, since a bead moves the
340: Lindemann length, once per time $\tau'$, on average. Note that since a
341: bead's movements, as embodied in $d_L$ and $\tau'$, are dictated by
342: its cage, this is an example of ``slaved'' motion, to borrow
343: Frauenfelder's adjective for conformational changes of a protein
344: encased in a stiff solvent \cite{FFMP, LWF}. One may associate, by
345: detailed balance, a low-frequency drag coefficient to that diffusion
346: constant: $\zeta' = k_B T/D'$. Such dissipative response implies
347: irreversible momentum exchange between a chosen particle and its
348: homogeneous (!) surrounding, hence a Stokes' viscosity $\eta' =
349: \zeta'/(6 \pi a/2)$, where $a/2$ is used for the radius of the region
350: carved out in the liquid by a single bead. Averaging with respect to
351: $\tau'$ yields for the steady state viscosity of the actual
352: heterogeneous liquid:
353: \begin{equation} \label{eta1} \eta = \frac{2 k_B T}{\pi a d_L^2} \la
354: \tau \ra,
355: \end{equation}
356: where we have removed the prime at $\tau$, implying averaging with
357: respect to the actual barrier distribution. The equation above can be
358: rewritten as
359: \begin{equation} \eta = (a/d_L)^2 \frac{2 k_B T}{\pi a^3} \la \tau \ra
360: \simeq 60 \: \frac{k_B T_g}{a^3} \la \tau \ra,
361: \end{equation}
362: since the Lindemann ratio, $d_L/a \simeq 0.1$ has been argued to
363: change at most by 10\% between $T_m$ and $T_g$ \cite{L_Lindemann}.
364:
365: Another instructive way to present Eq.(\ref{eta1}) is to note that the
366: Lindemann length is nearly equal to the typical amplitude of
367: high-frequency vibrations: $d_L \simeq d_\svibr$, within 5\% or so,
368: see Fig.3 of Ref.\cite{L_Lindemann}. The vibrational amplitude is
369: fixed by the equipartition theorem, since per bead: $K_\infty a^3
370: (d_\svibr/a)^2 = k_B T$, where $K_\infty$ is the high-frequency
371: elastic constant of the aperiodic lattice. One gets, as a result, a
372: Maxwell-type expression:
373: \begin{equation} \label{eta3} \eta \sim K_\infty \la \tau \ra.
374: \end{equation}
375: The last equation provides an easy way to see that the estimates in
376: Eqs.(\ref{eta1})-(\ref{eta3}) agree well with the experiment: Judging
377: from the sound speed in glasses \cite{FreemanAnderson,
378: BerretMeissner}, the typical high frequency elastic modulus is about
379: $10^{9} - 10^{10}$ Pa, i.e. comparable but somewhat less than those of
380: crystals. The range of relaxation times $10^{-12} - 10^{4}$ sec,
381: implies $10^{-3} - 10^{13}$ Pa$\cdot$sec for the viscosity, as is
382: indeed observed. Alternatively, one may obtain these figures by
383: substituting a typical $a \sim 3$\AA ~(see \cite{LW_soft, StevensonW}
384: for specific estimates of bead sizes/densities).
385:
386: Finally, the exploited equivalence between the time and ensemble
387: averages implies that crystallization has not begun during the
388: experiment, of course. The latter possibility adds uncertainty into
389: viscosity measurements, as the presence of crystallites would greatly
390: broaden the dynamic range of local relaxations owing to relatively
391: slow crystal nucleation events and the slow hydrodynamics near the
392: crystallites. Similarly, long chain motions in polymeric melts would
393: also introduce additional long time scales into the problem. Our
394: derivation does not apply to those situations. We note that optical
395: transparency, which is often used as an indicator of no-crystallinity,
396: does not ensure that crystallites - hundreds of nanometers across or
397: smaller - are absent. Therefore a rigorous experimental study should,
398: in the least, check whether performing viscoelastic measurements has
399: enhanced the crystallization of the sample. Ideally, X-ray diffraction
400: should be monitored in the course of viscosity measurements.
401:
402:
403: \section{Ionic Conductivity}
404: \label{cond}
405:
406: In any supercooled melt, whether regarded ionic or not, the beads
407: carry an additional charge, relative to the corresponding crystal,
408: because of the lack of crystalline symmetry. As a result, each
409: structural reconfiguration is characterized by a transition-induced
410: electric dipole, see Fig.\ref{domain_dipole}.
411: \begin{figure}[t]
412: \includegraphics[width=.95\columnwidth]{domain_dipole.eps}
413: \caption{\label{domain_dipole} Shown on the left is a fragment of the
414: mosaic of cooperatively reconfiguring regions in the supercooled
415: liquid. Expanded portion shows how rotation of a bond leads to
416: generating an elemental dipole change during a transition, where the
417: partial charges on the two beads are $\pm \zeta q$.}
418: \end{figure}
419:
420: The latter was estimated to be about a Debye or so, for most molecular
421: substances \cite{LSWdipole}. This value comes about as we may break up
422: the whole domain into $N^*/2$ pairs, where is pair has an elemental
423: dipole $\bmu_i$, $i = 1 \ldots N^*/2$: $\la |\bmu_i| \ra = \zeta e
424: q$. Here $q$ is the elementary charge: $q = 10^0 e$, and $\zeta < 1$
425: characterizes the excess charge. This quantity $\zeta$ is usually
426: small reflecting small deviations from the crystalline symmetry, in
427: the case of molecular crystals, or reflecting the weak interaction in
428: Van der Waals systems. Alternatively, the overall density of
429: charged/polar beads may be low. Ionic melts, by the very meaning of
430: the term, are distinct from molecular/Van der Waals systems in that
431: nearly {\em all} beads are strongly charged, implying $\zeta \sim 1$.
432: To be more specific, the conclusions of this article will be
433: exemplified with an often studied mixture of 40\%
434: Ca(NO$_3$)$_2$-60\%KNO$_3$ (``CKN''), $T_g \simeq 330$K.
435:
436: During a transition, each dipole turns by an angle $(d_L/a)$. The
437: total transition dipole:
438: \begin{equation} \bmu_T = \sum_{j}^{N^*/2} \left[ \bmu_j^{(f)} -
439: \bmu_j^{(in)} \right]
440: \end{equation}
441: scales as $\sqrt{N^*}$ because of the random orientation of the
442: elemental dipoles \cite{LSWdipole}:
443: \begin{equation}
444: \mu_T \simeq \zeta (q a) [(\xi/a)^3/2]^{1/2} (d_L/a).
445: \label{mu_Tqual}
446: \end{equation}
447: When the dipole density is uniform, every transition results in a
448: local arrangement equally representative of the liquid structure. In
449: other words, structural transitions do not modify the overall pattern
450: of the immediate coordination shell. Transitions lead to a local
451: ionic currents: $\bi' = \bmu_T'/\tau'$, per region of volume $\xi^3$.
452: In the presence of an electric field, the net current density:
453: \begin{equation}
454: \bj = \la \bj' \ra = \frac{1}{\xi^3} \la \frac{\bmu_T}{\tau} \ra,
455: \end{equation}
456: is non-zero because the dipole moment at the transition state is
457: correlated with the overall transition dipole moment. The latter can
458: be shown using Wolynes' library construction of liquid states
459: \cite{LW_aging}. Repeating that argument, but in the presence of
460: electric field $\bE$, yields for the typical free energy profile for
461: structural reconfiguration in steady state:
462: \begin{equation} \label{F(N)E} F(N) = \gamma \sqrt{N} - \bmu_T(N)
463: \bE_c - T s_c N,
464: \end{equation}
465: where $N$ is the size of the rearranged region and
466: \begin{equation}
467: \bmu_T(N) = \sum_j^{N/2} \left[ \mu_j^{(f)} - \mu_j^{(in)} \right]
468: \end{equation}
469: is the overall dipole change in that region. The subscript ``$c$'' in
470: $\bE_c$ signifies that the latter is a cavity field. The field
471: dependent term $|\bmu_T(N) \bE_c| \ll k_B T$ is overwhelmingly smaller
472: than the other terms in Eq.(\ref{F(N)E}). As a result, the transition
473: state dipole moment is not field-induced, but, again, is ``slaved'' to
474: the lattice. More formally, one may use the argument from
475: Ref.\cite{LW} showing that the density of structural states at the
476: reconfiguration bottle-neck is of the order $1/T_g$, implying the
477: field will not affect the specific sequence of elemental moves, but
478: will affect the dynamics merely by shifting the energies along
479: structurally dictated sequences of moves. Thus in the lowest order in
480: $\bE_c$, $\tau^{-1}(\bE_c) = \tau^{-1}(\bE_c=0)(1+ \bmu_T^\ddagger
481: \bE_c/k_B T)$, yielding
482: \begin{equation} \bj = \la \frac{\bmu_T (\bmu_T^\ddagger \bE_c)}{\tau}
483: \ra \frac{1}{k_B T},
484: \end{equation}
485: where $\bmu_T^\ddagger \equiv \bmu_T(N^\ddagger)$ is the transition
486: dipole moment at the zero-field transition state. The cavity field,
487: see e.g. \cite{TitulaerDeutch}, is related to the external field
488: $E^{(\infty)}$ by
489: \begin{equation} E_c(\omega) = E^{(\infty)}(\omega) \frac{3
490: \epsilon_b(\omega)}{2 \epsilon_b(\omega)+1},
491: \end{equation}
492: where $\epsilon_b(\omega)$ is the dielectric constant of the
493: surrounding bulk. Since a steady current is implied in the derivation,
494: (the imaginary part of) $\epsilon_b(\omega)$ diverges at zero
495: frequency, implying $E_c = E^{(\infty)} (3/2)$. One thus obtains for
496: the ionic conductivity tensor:
497: \begin{equation} \sigma_{ij} = \la \frac{(\bmu_T)_i
498: (\bmu_T^\ddagger)_j}{\tau} \ra \frac{3}{2 k_B T \xi^3}.
499: \end{equation}
500: Bearing in mind that $N^\ddagger = N^*/4 = (\xi/a)^3/4$, and that the
501: liquid is isotropic, on average: $\la (\Delta \bmu)_i (\Delta \bmu)_j
502: \ra = \delta_{ij} (\Delta \mu)^2/3$, one finally has:
503: \begin{equation} \label{sij} \sigma_{ij} = \delta_{ij} \, \frac{\Delta
504: \mu_\smol^2}{8 \, a^3 k_B T} \la \frac{1}{\tau} \ra,
505: \end{equation}
506: where
507: \begin{equation} \Delta \mu_\smol^2 \equiv \la \left[ \mu_j^{(f)} -
508: \mu_j^{(in)} \right]^2 \ra
509: \end{equation}
510: is the average elemental dipole change squared. Finally note that in
511: covalently networked materials, where dipole assignment may be
512: ambiguous, one may still estimate local dipole changes using the known
513: piezoelectric properties of the corresponding crystal, see
514: \cite{LSWdipole} for details.
515:
516: The derivation above does not apply to systems where the dipole
517: density is significatly non-uniform. For instance, glycerol has one
518: polar, OH group per non-polar, aliphatic group, implying the liquid is
519: non-homogeneous, dipole moment wise, on the $\alpha$-relaxation time
520: scale. In such systems, the premise that structural rearrangements
521: result in equally representative configurations of the liquid does not
522: hold. In the glycerol example, ionic conduction would imply breaking
523: OH or CH bonds. In CKN, on the other hand, the overall bond pattern,
524: around any atom, does not change significantly during a transition,
525: even though individual bonds distort by the Lindemann length, as
526: mentioned. Eq.(\ref{sij}) thus places the absolute upper limit on the
527: {\em intrinsic} ionic conductivity of a melt. By ``intrinsic'' we mean
528: that the computed currents are always present in the fluid and result
529: from the intrinsic activated transport: Local bond pattern does not
530: change significantly in the course of an individual activated event,
531: but only in the course of many consecutive events, since during an
532: individual event, the molecular displacements barely exceed typical
533: vibrational displacements. Conversely, if a system displays a higher
534: conductivity than prescribed by Eq.(\ref{sij}), one may conclude that
535: the ion motion does not require structural reconfiguration. Here, the
536: bond pattern actually changes, however these are not scaffold bonds of
537: the aperiodic lattice comprising the fluid (or glass). (More on this
538: below.)
539:
540:
541: To simplify comparison of Eq.(\ref{sij}) with experiment, let us
542: express the combination of the bead charge $\zeta q$ and size $a$, in
543: Eq.(\ref{sij}), through the {\em finite} frequency dielectric
544: response, a measurable quantity in principle (see below). The latter
545: is the response of a rearranging region in the absence of bulk
546: current, i.e. with a fixed environment, up to vibrations. It is
547: convenient to choose such regions at volume $\xi^3$, so that each
548: region has two structural states available, within thermal reach from
549: each other, separated by a barrier sampled from the actual barrier
550: distribution in the liquid. If the two states, ``1'' and ``2'' are
551: characterized by dipole moments $\bmu_1$ and $\bmu_2$ respectively,
552: the expectation value of the dipole moment of the region is $\mu =
553: (\bmu_1 + \bmu_2)/2 + \Delta \bmu (p_2 - p_1)/2$, where $\Delta \bmu
554: \equiv (\bmu_2 - \bmu_1)$; $p_1$ and $p_2 = (1 - p_1)$ are the
555: probabities to occupy state 1 and 2 respectively. The relative
556: population $(p_2 - p_1)$ depends on the field via $\delta \ln(p_2/p_1)
557: = \Delta \bmu \bE_c/k_B T$. At realistic field strengths, i.e.
558: $|\Delta \bmu \bE_c/k_B T| \ll 1$, one has for the field-induced shift
559: of the relative population: $\delta (p_2 - p_1) \simeq 2 p_1 p_2
560: (\Delta \bmu \bE_c)$. Similarly to the preceding argument, $\la
561: (\Delta \bmu)_i (\Delta \bmu)_j \ra = (N^*/2) \delta_{ij} (\Delta
562: \mu_\smol)^2/3$. Further, since we have frozen the structural
563: transitions in the surrounding region, in estimating the cavity field,
564: one must use $\epsilon(\omega)$ with the $\alpha$-relaxation
565: contribution subtracted. This does not introduce much ambiguity
566: because in most ionic substances, the dielectric constant even at very
567: high frequences is significantly larger than unity. In CKN, for
568: instance, $\epsilon'(\infty) \simeq 7$ \cite{Howell1974}, allowing us
569: to write as before: $E_c \simeq E^{(\infty)}(3/2)$. One thus obtains,
570: in a standard fashion, for the frequency dependent dielectric constant
571: in the absence of macroscopic current:
572: \begin{equation} \label{epsmu} \epsilon_\sins(\omega)-\epsilon_\infty
573: = 4 \pi \la p_1 p_2 \ra \frac{\Delta \mu_\smol^2}{4 \, a^3 k_B T}
574: \la \frac{1}{1 - i \omega \tau} \ra,
575: \end{equation}
576: where the label ``ins'' signifies the absence of dc conductivity.
577:
578:
579: In the presence of steady current, the full response per domain is the
580: sum of the steady current from Eq.(\ref{sij}) and the ac current from
581: Eq.(\ref{epsmu}). The addition of the dc contribution, $i
582: \sigma/\omega$, to the full dielectric response will increase the
583: absolute value of $\epsilon_b(\omega)$. This means that
584: Eq.(\ref{epsmu}) should work even better. One thus gets for the full
585: dielectric response of a conducting substance:
586: \begin{equation} \label{epsfull} \epsilon(\omega) - \epsilon_\infty =
587: 4 \pi \la p_1 p_2 \ra \frac{\Delta \mu_\smol^2}{4 \, a^3 k_B T} \la
588: \frac{1}{1 - i \omega \tau} \ra+ i \frac{\sigma}{\omega},
589: \end{equation}
590: where $\sigma$ is the dc conductivity from Eq.(\ref{sij}).
591:
592: One needs to know the distribution of the transition energies $E$ to
593: estimate the quantity $\la p_1 p_2 \ra \equiv \la 1/4 \cosh^2(E/2k_B
594: T) \ra$. Since $\xi$ is the smallest possible size of a rearranging
595: unit, these rearrangments correspond to the elementary excitations in
596: the system. We thus conclude, based on equipartition, that the typical
597: value of $E$ is roughly $k_B T$, implying that $\la p_1 p_2 \ra$ is
598: close to its maximum value of one quarter but is likely smaller by
599: another factor of two or so. Assuming then, for the sake of argument
600: that $\la p_1 p_2 \ra = 1/8$, one gets $\frac{\Delta \mu_\smol^2}{ a^3
601: k_B T} \simeq 2 (\epsilon'_0-\epsilon_\infty)$, within an order of
602: magnitude. By Eq.(\ref{sij}), this implies a Maxwell-like relation
603: between the dc conductivity and the real part of the dielectric
604: response:
605: \begin{equation} \label{Maxwell_cond} \sigma \sim
606: (\epsilon'_0-\epsilon_\infty) \la \frac{1}{\tau} \ra,
607: \end{equation}
608: with an important distinction, though, that here one averages the {\em
609: inverse} relaxation time. The $(\epsilon'_0-\epsilon_\infty)$
610: difference in CKN, to be concrete, is about $(20 - 7) = 13$
611: \cite{Pimenov, Howell1974}. This implies, by Eq.(\ref{epsmu}) and
612: CKN's $T_g \simeq 330$K \cite{Pimenov}, that $\zeta q \simeq 3e$, at
613: $a = 3$\AA, a reasonable value for the bead charge. Note that $a =
614: 3$\AA ~is consistent with CKN's $\Delta c_p \simeq .12$ cal/g K
615: \cite{AngellTorell} and the mentioned $D = 32/\Delta c_p$ \cite{XW}.
616:
617: Now, substituting CKN's $(\epsilon'_0-\epsilon_\infty)$ into
618: Eq.(\ref{sij}) yields for the conductivity $\sigma \simeq 3 \cdot
619: 10^{-10} \la \tau^{-1} \ra$ (Ohm m)$^{-1}$sec. Naively replacing $\la
620: \tau^{-1} \ra$ with $1/\la \tau \ra$ would imply, at the glass
621: transition, where $\la \tau \ra \sim 10^{2} - 10^{3}$ sec, a
622: conductivity of the order $10^{-13} - 10^{-14}$ (Ohm cm)$^{-1}$, which
623: is three to four orders of magnitude below the observed value
624: \cite{Pimenov, Howell1974}. Note that the value $\la \tau^{-1} \ra
625: \la \tau \ra = 10^3 - 10^4$ is just the magnitude of decoupling
626: observed in CKN near $T_g$ \cite{Howell1974, Angell}, and is in fact
627: expected for a fragile substance such as CKN is, as we will argue in
628: the following.
629:
630:
631:
632: \section{Barrier distribution and the Decoupling}
633: \label{Barrier}
634:
635: Relaxation barriers in supercooled liquids are distributed because the
636: local density of states is non-uniform, leading to variations in the
637: local value of the configurational entropy and hence the RFOT-derived
638: barrier from Eq.(\ref{Fsc}). In the simplest argument, the gaussian
639: fluctuations of the entropy translate into gaussian fluctuations in
640: the barrier, where the relative deviations of the two quantities, from
641: the most probable value, are given by \cite{XWbeta}:
642: \begin{equation} \label{relf} \delta \wF \equiv \frac{\delta
643: F}{F_\smp} = \frac{\delta s_c}{s_c} = \frac{1}{2 \sqrt{D}},
644: \end{equation}
645: where $D$ is the liquid's fragility from Eq.(\ref{tau}). The quantity
646: $1/2\sqrt{D}$ varies between 0.05 and 0.25 or so, for known
647: glassformers, the low and high limits corresponding to strong and
648: fragile substances respectively.
649:
650: Xia and Wolynes (XW) further argued that the real barrier distribution
651: should be cut-off at the most probable value because a liquid region
652: with relatively low density of states is likely neighbors with a
653: relatively fast region \cite{XWbeta}. In addition we may recall that
654: in the library construction, the most likely liquid state is the one
655: where the liquid is {\em guaranteed} to have an escape trajectory
656: \cite{LW_aging}. This means that the most probable barrier is also the
657: maximum barrier. One may conclude then that the naive Gaussian
658: distribution is adequate at small barriers, but significantly
659: overestimates the probability of barriers larger than the typical
660: barrier. Put another way, the trajectories corresponding to higher
661: than most probable barrier in the naive Gaussian, all contribute to
662: the $F \le F_\smp$ range. XW have implemented this notion by replacing
663: the r.h.s. of the simplest Gaussian distribution by a delta-function
664: centered at $F_\smp$ \cite{XWbeta}:
665: \begin{equation} \label{pF} p_1(\wF) = \frac{e^{-(1/\wF-1)^2/2\delta
666: \wF^2}}{\sqrt{2 \pi (\delta \wF)^2} \wF^2} +
667: \frac{1}{2}\delta(\wF-1),
668: \end{equation}
669: where $\wF \equiv F/F_\smp < 1^+$, and we took advantage of the
670: temperature-independence of the relative width in Eq.(\ref{relf}).
671: This approximate form does not use adjustable parameters and
672: quantitatively accounts for the correlation between the fragility and
673: the stretching exponent $\beta$ \cite{XWbeta}, and the deviations from
674: the Stokes-Einstein relation. The distribution in Eq.(\ref{pF}) is
675: shown in Fig.\ref{pFfig}. The only difference of Eq.(\ref{pF}) with
676: the XW's form is that they used a purely gaussian form for $\wF < 1$,
677: whereas we follow their own suggestion and employ the more accurate $F
678: \propto 1/s_c$ (where $s_c$ is gaussianly distsributed of course). The
679: accurate evaluation of the left wing of the distribution is imperative
680: in estimating the average rate $\tau_\smicro^{-1} e^{-F/k_B T}$,
681: because the latter is a rapidly varying function of $F$. ($k_B T$ is
682: significantly less that $\delta F$ at low temperatures.) Note that
683: because of the rapid decay of the exponential at small $\wF$ in
684: Eq.(\ref{pF}), accounting for the lowest order, quadratic fluctuations
685: of entropy suffices. The quantity $\la \tau \ra \la \tau^{-1} \ra$,
686: that characterizes the apparent decoupling, computed with the XW's
687: distribution, is shown with the dashed line in Fig.\ref{t_tt}, at
688: $T_g$, as a function of the relative distribution width $\delta \wF$.
689:
690: How robust is the prediction based on the simple functional form for
691: the barrier distribution from Eq.(\ref{pF})? In spite of its
692: quantitative successes, one may argue that the true barrier
693: distribution should be a smoother function, near $\wF =1$. One way to
694: see this is to computing $\epsilon''(\omega)$ from Eq.(\ref{epsmu})
695: via the distribution in Eq.(\ref{pF}): The obtained curves are a sum
696: of two peaks, one of which is broader, one the other is narrower than
697: the experimental $\epsilon''(\omega)$. The two peaks correspond to
698: the half-Gaussian and the delta-function in Eq.(\ref{pF})
699: respectively. Let us see that knowing the precise form of the barrier
700: distribution however is not essential in quantitative estimates of the
701: decoupling so long as we account correctly for the overall width of
702: the distribution and its decay at the low barrier side.
703:
704: \begin{figure}[t]
705: \includegraphics[width=.6\columnwidth]{pFfig.eps}
706: \caption{\label{pFfig} The barrier distributions from Eqs.(\ref{pF})
707: and (\ref{pFm}). The $\delta$-function portion of the distribution
708: from Eq.(\ref{pF}) is not shown. $\delta \wF = 0.25$.}
709: \end{figure}
710:
711: It is straightforward to show that there exists a distribution that
712: (a) satisfies these requirements without introducing adjustable
713: constants, (b) reproduces the experimental $\epsilon(\omega)$ and does
714: as well as the XW form for the $\beta$ vs. $D$ correlation. As we
715: have already discussed, the low barrier wing of the distribution in
716: Eq.(\ref{pF}) is adequate. On the other hand, the high barrer wing
717: should include the contributions from both sides of the original
718: Gaussian peak, which are both of width $\delta F/2$. ``Stacking''
719: these two on top of each other, to the left of $F_\smp$, results in a
720: distribution of width $\delta F/4$ (see aslo Appendix). Further,
721: based on the known $\epsilon(\omega)$ data, the barrier distribution
722: should be well approximated by an exponential, suggesting we use $p(F)
723: \propto e^{F/(\delta F/4)}$ near $F_\smp$. Indeed, this implies
724: $p(\tau) \propto \tau^{(4 k_B T/\delta F)-1}$. At frequencies not too
725: close to the maximum of $\epsilon''(\omega)$ and the rapid drop-off at
726: small $\wF$, one has an approximate power law:
727: \begin{equation} \label{appr} \epsilon''(\omega) \simeq \int_0^\infty
728: d\tau \, \tau^{(4 k_B T/\delta F)-1} \frac{\omega \tau}{1+(\omega
729: \tau)^2} \propto \omega^{-4 k_B T/\delta F}.
730: \end{equation}
731: We thus arrive at the following form:
732: \begin{equation} \label{pFm} p(\wF) = \left\{ \begin{array}{ll}
733: \frac{c_1}{\wF^2} e^{-(1/\wF-1)^2/2\delta \wF^2}, & \wF \le \wF_e\\
734: \frac{c_2}{\wF^2} e^{\wF/(\delta \wF/4)}, & \wF_e < \wF \le 1,
735: \end{array}
736: \right.
737: \end{equation}
738: where $F_e$ and the normalization constants $c_1$ and $c_2$ are chosen
739: so that the distribution is normalized, continuous, and its first
740: derivative is continuous too. The distribution from Eq.(\ref{pFm}) is
741: plotted in Fig.\ref{pFfig}. The decoupling strength $\la \tau \ra \la
742: \tau^{-1} \ra$, computed for the composite distribution from
743: Eq.(\ref{pFm}), is shown in Fig.\ref{t_tt}, as a function of the
744: relative distribution width $\delta \wF$, at $T_g$. We therefore
745: observe that in fragile liquids, the apparent time-scale separation
746: may reach as much as four orders of magnitude near the glass
747: transition - even though only one process is present! - because the
748: inrinsic ionic conductivity is dominated by the fastest relaxing
749: regions.
750:
751:
752: \begin{figure}[t]
753: \includegraphics[width=.85\columnwidth]{t_tt.eps}
754: \caption{\label{t_tt} The decoupling between the viscosity and the
755: intrinsic ionic conductivity, as a function of the relative barrier
756: width $\delta \wF$ from Eq.(\ref{relf}), at $T_g$. The dashed and
757: solid line pertain to the specific barrier distributions from
758: Eq.(\ref{pF}) and (\ref{pFm}) respectively. }
759: \end{figure}
760:
761: Conversely, when the apparent decoupling exceeds the intrinsic value
762: prescribed by Fig.\ref{t_tt}, we may conclude that ionic conduction
763: does not in fact require structural relaxation. This notion is of
764: significance for the mechanisms of electrical conductance in glasses
765: and will be discussed in detail in the Conclusions.
766:
767:
768:
769: \begin{figure}[t]
770: \includegraphics[width=.85\columnwidth]{t_FT.eps}
771: \caption{\label{t_FT} Different relaxation times derived from the
772: barrier distribution in Eq.(\ref{pFm}), as functions of the most
773: probable barrier (left) and the corresponding temperature
774: (right). $\delta \wF = 0.25$, corresponding to fragility $D=4$.}
775: \end{figure}
776:
777:
778: One may also illustrate the effects of apparent decoupling for a
779: specific value of fragility, by plotting several varieties of
780: relaxation times, as functions of the most probable barrier, or the
781: corresponding temperature, see Fig.\ref{t_FT}. (Given the time scale
782: at the glass transition, say $D T_K/(T_g - T_K) = \ln(10^{16}) \simeq
783: 37$ (see Eq.(\ref{tau})), there is a one-to-one correspondence between
784: the fragility $D$ and the $T_g/T_K$ ratio.) We observe that the
785: average relaxation time and the one derived from the inverse of the
786: maximum position of $\epsilon''(\omega)$ are close, and are near the
787: most probable value of the relaxation time. (The $\epsilon''(\omega)$
788: was computed using Eqs.(\ref{epsfull}) and (\ref{pFm}), see below.)
789: The apparent conductivity relaxation time is strongly decoupled,
790: consistent with data of Howell at el. \cite{Howell1974} for CKN. Note
791: that the value of fragility used in Fig.\ref{t_FT}, $D=2$, is probably
792: smaller than in CKN. In addition, we have ignored here, for clarity,
793: the effects of barrier softening \cite{LW_soft}, that would require
794: introducing a system-specific adjustable constant $T_A$. The latter
795: effects would change the slopes of the curves somewhat, without
796: affecting their vertical separations.
797:
798:
799:
800:
801: To test the predictions from Figs.\ref{t_tt} and \ref{t_FT}, one needs
802: to know the width of the barrier distribution for $\alpha$-relaxation.
803: As already mentioned, the gross features of this distribution have
804: been predicted by the RFOT theory, and have lead to quantitative
805: predictions of the correlation between the stretching exponent $\beta$
806: and the fragility $D$, and the deviations from the Stokes-Einstein
807: relation. The corresponding trends are as follows: more fragile
808: liquids are predicted to have broader barrier distribution leading to
809: a smaller value of $\beta$, and vice versa for stronger substances
810: \cite{XWbeta}. A correlation with the fragility comes about by virtue
811: of Eq.(\ref{relf}). Several {\em \`{a} priori} ways to determine
812: $\beta$ and $D$ have been employed, by experimenters, that sometimes
813: produce conflicting results. For example, the fragility extracted
814: from $\tau_\sigma$ will be consistently lower than that extracted from
815: the mechanical relaxation time $\tau_s$, because $\tau_\sigma <
816: \tau_s$. The exponent $\beta$ from the stretched exponential is
817: extracted from fits of various relaxation processes to a stretched
818: exponential profile $e^{-(t/\tau)^\beta}$. Alternatively, one may
819: choose to fit the Fourier transform of the stretch exponential, or the
820: Cole-Davidson form, to the imaginary part of $\epsilon(\omega)$ in
821: insulators \cite{LindseyPatterson}. These usually produce comparable
822: results for the corresponding exponent $\beta$, with a notable
823: exception of ionic conductors, which happen to be the main focus of
824: this paper. In ionically conducting systems, the dc component of the
825: full dielectric response from Eq.(\ref{epsfull}) largely ``swamps''
826: the ac part so that reliable determinations of the latter are
827: complicated. The reader is reminded that dielectric measurements on
828: ion melts are difficult because electrodes generally block ionic
829: current. The effects of build-up charge are often treated
830: phenomenologically, by means of equivalent circuits \cite{Macdonald,
831: Pimenov}. Given these complications, many have chosen to plot the
832: reciprocal of $\epsilon(\omega)$, i.e. the dielectric modulus
833: \cite{Howell1974, Macdonald}:
834: \begin{equation} \label{Mo} M(\omega) \equiv 1/\epsilon(\omega).
835: \end{equation}
836: $M(\omega)$ is well behaved and even shows a peak in the imaginary
837: component, similarly to $\epsilon(\omega)$ of a near insulator. In
838: the absence of an {\em \`{a} priori} microscopic picture and by
839: analogy with $\epsilon(\omega)$, one might interpret this peak as as
840: the response of the electric field $\bE$ to the dielectric
841: displacement $\bD$. This in fact would be appropriate in a layered
842: dielectric \cite{Howell1974}. See also the discussions in
843: Refs.\cite{Elliott1994, Roling1998, ColeTombari, Sidebottom1995,
844: Moynihan1994, Dyre1991, Doi1988}. Yet the resulting values of the
845: most probable relaxation time and the stretching exponent deviate from
846: those obtained with other methods \cite{Angell, Sidebottom1995,
847: Pimenov}. In fact, the modulus-derived $\beta_M$ increases, while
848: the width of the $M''(\omega)$ peak decreases with lowering the
849: temperature, in conflict with the general trends for poor conductors,
850: and the conclusions of the RFOT theory.
851:
852:
853: \begin{figure}[t]
854: \includegraphics[width=.9\columnwidth]{epsM.eps}
855: \caption{\label{epsM} The top panel shows the imaginary part of the
856: non-conductive part of the dielectric response, $\epsilon(\omega)$,
857: Eq.(\ref{epsmu}), for four values of the most probable
858: $\alpha$-relaxation barrier: $F_\smp/k_B T =
859: \ln(\tau/\tau_\smicro)$, as indicated on the graph. ($\delta \wF =
860: 0.25$, $\tau_\smicro = 1$ ps.) These barrier values were chosen
861: because for $\tau/\tau_\smicro > e^{7} \simeq 10^{3}$, the RFOT
862: approach is quantitatively accurate, while $\tau/\tau_\smicro \simeq
863: e^{37} \simeq 10^{16}$ is at the upper limit of the dynamical range
864: routinely accessible in the lab. The bottom panel show the
865: corresponding modulus, from Eq.(\ref{Mo}), that includs the dc-part,
866: due to the intrinsic ionic current. The same four values of the
867: barrier are used. We have used the CKN's values for $\epsilon_0$ and
868: $\epsilon_\infty$, and assumed that $\frac{\Delta \mu_\smol^2}{ a^3
869: k_B T} \simeq 2 (\epsilon'_0-\epsilon_\infty)$, see the discussion
870: preceeding Eq.(\ref{Maxwell_cond}).}
871: \end{figure}
872:
873:
874: The RFOT theory and the present results allow one to address these
875: difficulties, to which we devote the rest of this Section. One first
876: notes that structural reconfigurations are {\em compact}, and so the
877: layered-dielectric view of supercoold melts is not microscopically
878: justified. We next plot, in the top panel of Fig.\ref{epsM}, the
879: non-conductive $\epsilon_\sins''(\omega)$, from Eq.(\ref{epsmu})
880: averaged with respect to the barrier distribution from Eq.(\ref{pFm}).
881: We have used CKN's values for $\epsilon_0$ and $\epsilon_\infty$, as
882: before. For the sake of argument, we use $\delta \wF = 0.25$,
883: corresponding to $\beta \simeq 0.4$ at $T_g$. In Fig.\ref{epsM},
884: bottom, we show the imaginary part $M''(\omega)$, of the full modulus.
885: Clearly the two functions exhibit qualitatively different behaviors.
886: Note that the effect of the dc component on the apparent relaxation
887: profile has been discussed previously \cite{Johari1988, Roling1998},
888: including the possibility of a double peak \cite{Johari1988}. The
889: latter has been observed by Funke at el. \cite{Funke}, but has not
890: been reproduced by Pimenov at el. \cite{Pimenov}. Nevertheless, the
891: dielectric modulus obtained here is qualitatively consistent with
892: CKN's data from Ref.\cite{Pimenov}. Finally note that for smaller
893: dc-conductivities, the modulus data would become more similar to
894: $\epsilon''(\omega)$.
895:
896:
897: \begin{figure}[t]
898: \includegraphics[width=.7\columnwidth]{w_T.eps}
899: \caption{\label{w_T} The widths of the peaks in the imaginary parts
900: of the dielectric constant $\epsilon(\omega)$ and modulus
901: $M(\omega)$, as functions of temperature
902: (c.f. Fig.\ref{epsM}). $\delta \wF = 0.25$.}
903: \end{figure}
904:
905:
906: We conclude from the above analysis that if one were to use the
907: modulus data to extract the characteristics of the barrier
908: distribution, one must measure first the dc-current, add it to the
909: $\epsilon_\sins(\omega)$ from a microscopic theory, and then compare
910: the result to the measured $M(\omega)$ data. But again, because of the
911: large contribution of the dc component, the corresponding fits would
912: not discriminate well between different forms of
913: $\epsilon_\sins(\omega)$. On the other hand, treating the electric
914: field as a response to the displacement may lead to erroneous
915: conclusions on the temperature dependence of the barrier width. In
916: fact, the barrier widths derived from $\epsilon_\sins''(\omega)$ or
917: $M''(\omega)$ show the opposite trends, as we have seen already in
918: Fig.\ref{epsM}. One may further quantify this observation: In the
919: absence of a microscopic theory, one often characterizes the width of
920: the $\epsilon_\sins''(\omega)$ peak by a stretching exponent $\beta$,
921: as derived e.g. from Davidson-Cole fits. The distribution from
922: Eq.(\ref{pFm}) indeed gives rise a power law behavior, consistent with
923: Eq.(\ref{appr}), see Appendix. In contrast, the corresponding
924: $M''(\omega)$ curves do not exhibit a similar power-law behavior. I
925: have chosen to illustrate the opposite temperature trends in the
926: widths of $\epsilon_\sins''[\ln(\omega)]$ and $M''[\ln(\omega)]$
927: peaks, by measuring the latter widths at one-third-height and plotting
928: them as functions of temperature, see Fig.\ref{w_T}. Similar opposite
929: trends, too, would be observed for the corresponding {\em apparent}
930: barrier widths or effective $\beta$'s. Clearly, interpreting the
931: dielectric modulus of an ionic conductor as a response function may
932: lead to a significant underestimation of the actual barrier width at
933: low temperatures, and qualitatively incorrect conclusions on the
934: temperature dependence of the width.
935:
936:
937:
938:
939:
940: \section{Conclusions}
941:
942: We have computed, from the first principles, the viscosity and the
943: intrinsic ionic conductivity of supercooled liquids. The viscosity is
944: determined by four microscopically defined quantities: the length
945: scale of the local chemical order that sets in at temperature
946: $T_\scr$, where liquid dynamics become activated; the Lindemann
947: length, characterizing molecular displacements at the mechanical
948: stability edge; the temperature; and the average relaxation time
949: $\tau$ of the activated reconfigurations that dominate the liquid
950: dynamics below $T_\scr$. The extraordinarily long $\tau$ range is what
951: gives rise to the high viscosity of the liquid when it approaches the
952: glass transition. When the local chemically stable units (or
953: ``beads'') are charged, the fluid will also exhibit an ionic
954: conductivity, which we have called the ``intrinsic'' conductivity, to
955: constrast it with electric conduction via delocalized electronic
956: carriers or via mobile ions that are not bonded to the metastable
957: aperiodic lattice forming the supercooled liquid. Computing the
958: conductivity requires an additional microscopic characteristic, the
959: electric charge on a ``bead''. Fortunately, this additional parameter
960: may be deduced from the ac dielectric response, which we have also
961: estimated. Perhaps the main finding of this work is that in contrast
962: with the viscosity, the ionic conductivity is dominated by the fastest
963: relaxing regions in the liquid, as reflected in Eq.(\ref{sij}).
964:
965: We have discussed ways to test the above predictions, the most
966: important aspect of which is the large separation, or ``decoupling'',
967: between the apparent time scales, suggested previously by viscosity
968: and ionic conductivity data on purely phenomenological grounds. We
969: have shown that such apparent time-scale separation is indeed expected
970: because of the very broad barrier distribution for
971: $\alpha$-relaxation, derived earlier in the Random First Order
972: Transition (RFOT) methodology. The decoupling thus stems essentially
973: from the same cause as the violation of Stokes-Einstein relation in
974: supercooled liquids \cite{XWhydro}. Now, we have seen that the value
975: of the decoupling is not very sensitive to the precise form of the
976: barrier distribution so long as one acounts for the RFOT-derived gross
977: characteristics of this. We have thus quantified the degree of
978: ``decoupling'': The intrinsic ionic conductivity was argued to
979: decouple at most by four orders of magnitude from the low-frequency
980: momentum transport. Conversely, any conductivity exceeding this limit
981: must be due to other charge carriers that do not disturb the liquid
982: structure beyond typical vibrational displacements. Indeed, suppose
983: the apparent decoupling exceeds the value prescribed by the width of
984: the barrier distribution. This means that there will be ions that
985: travel a distance exceeding the Lindemann length in a time it takes
986: the local environment to relax. Therefore, local relaxation is not a
987: necessary condition for a non-zero current of these ions. Some
988: interaction with relaxation may still be present, however at large
989: enough decouplings, we may say that the ion (or any other carrier)
990: interacts with the liquid as if the latter were a perfectly stable,
991: albeit disordered lattice. In such cases, one may think of the ionic
992: current in superionic conductors in terms of regular, not slaved
993: diffusion. In regular diffusion, the total travel time is dominated
994: by the slowest step, in contrast to Eq.(\ref{sij}).
995:
996:
997:
998: The intrinsic difficulty in experimental assessment of the barrier
999: distribution in moderately conductive melts is that the dc current
1000: dominates the overall dielectric response. This gives rise to
1001: ambiguities as to what the actual width of the barrier distribution
1002: is, since mechanical relaxation and dielectric {\em modulus} data
1003: disagree. We have shown that this is expected, and argued that the
1004: mechanical relaxation offers the preferred method of estimating the
1005: actual barrier distribution.
1006:
1007:
1008:
1009:
1010:
1011: {\em Acknowledgments:} The author thanks Peter G. Wolynes for critical
1012: comments and useful suggestions. He gratefully acknowledges the GEAR,
1013: the New Faculty Grant, and the Small Grant Programs at the University
1014: of Houston.
1015:
1016:
1017:
1018:
1019:
1020: \section*{Appendix}
1021:
1022:
1023:
1024: \begin{figure}[t]
1025: \includegraphics[width=.95\columnwidth]{epsom.eps} \vspace{3mm}
1026: \caption{\label{epsom} The four thick lines are the same as in
1027: Fig.\ref{epsM}, but in the double-log scale. The thin lines are
1028: the Davidson-Cole forms, following from a simple approximation,
1029: see Appendix. The dash-dotted line illustrates how the stretching
1030: exponent $\beta$ was extracted from the curves.}
1031: \end{figure}
1032:
1033:
1034: \begin{figure}[t]
1035: \includegraphics[width=.85\columnwidth]{bD.eps}
1036: \caption{\label{bD} Two approximate relations of the stretching
1037: exponent $\beta$ to the fragility $D$ from the Vogel-Fulcher form,
1038: from Eq.(\ref{betaD}) and as derived from the slopes of the
1039: high-frequency wing of the $\epsilon''(\omega)$ peaks, such as in
1040: Fig.\ref{epsom}.}
1041: \end{figure}
1042:
1043: Let us see that the distribution in Eq.(\ref{pFm}) is qualitatively
1044: consistent with experimental $\epsilon_\sins (\omega)$ and the
1045: empirical correlation of $\beta$ and $D$. For this, we replot the top
1046: panel of Fig.\ref{epsM} in the double-log format, in
1047: Fig.\ref{epsom}. We note the general adequacy of the barrier
1048: distribution from Eq.(\ref{pFm}): Similarly to the experimental
1049: $\epsilon(\omega)$ in poor conductors, the resulting high-frequency
1050: wing is significantly broader than the low-frequency one. Note that
1051: the actual data would also often display an {\em additional}
1052: high-frequency wing, which is ascribed to the secondary,
1053: $\beta$-processes, also called Johari-Goldstein relaxation
1054: \cite{JohariGoldstein}. (See \cite{LunkLoidl} for a review). The
1055: present results suggest that $\beta$-relaxation does not contribute to
1056: the intrinsic ionic conductivity. At any rate, the derived
1057: $\epsilon(\omega)$ show several decades of nearly power-law decay,
1058: allowing one to extract the corresponding exponent: $\epsilon(\omega)
1059: \propto \omega^{-\beta}$. The effective $\beta$'s were deduced from
1060: the slopes of the curves at the points of maximum second derivative,
1061: as exemplified by the dash-dotted line in Fig.\ref{epsom}. The
1062: dependence of the thus obtained exponent $\beta$ on the fragility $D$,
1063: at a fixed $F_\smp/k_B T = 37$, is shown by the dashed-dotted line in
1064: Fig.\ref{bD}. This $\beta$ is, again, qualitatively consistent with
1065: experiment. Greater accuracy should not be expected here, as we have
1066: not treated the higher-frequency range associated with
1067: $\beta$-relaxation, which would affect the experimentally determined
1068: stretching exponents.
1069:
1070:
1071:
1072: In addition, we verify that the informal argument in the main text
1073: that the width of the barrier distribution should be about $\delta F/4
1074: $, at the half-height or so. Indeed, for a gaussian barrier
1075: distribution with width $\delta F/4 = F/8\sqrt{D}$ implies the
1076: following approximate expression for the stretching exponent $\beta$
1077: at $T_g$ (c.f. Eq.(9) of Ref.\cite{XWbeta}):
1078: \begin{equation} \label{betaD} \beta = \left[ 1 + \left(
1079: \frac{F/k_BT}{8 \sqrt{D}} \right)^2 \right]^{-1/2},
1080: \end{equation}
1081: shown as the solid line in Fig.\ref{bD}. This expression is in very
1082: good agreement with experiment, see Fig.2 from Ref.\cite{XWbeta}. (At
1083: $T_g$ on scale $\tau/\tau_\smicro = 10^{16}$, $F/k_B T \simeq 37$.)
1084: Note also Eq.(\ref{betaD}) is consistent with Eq.(\ref{relf}),
1085: assuming the Davidson-Cole \cite{CD} and William-Watts \cite{KWW}
1086: stretching exponents $\beta$ are close \cite{LindseyPatterson}. That
1087: the latter is the case indeed I demonstrate by graphing the
1088: Davidson-Cole (DC) form $(\epsilon_{\tDC}(\omega) - \epsilon_\infty) =
1089: (\epsilon_0 - \epsilon_\infty) (1 - i \omega \tau)^{-\beta}$ \cite{CD}
1090: with $\tau = \tau_\smicro e^{F_\smp/k_B T}$ and $\beta$ from
1091: Eq.(\ref{betaD}). These are shown in Fig.\ref{epsom} as thin dashed
1092: lines.
1093:
1094:
1095:
1096:
1097:
1098:
1099: \bibliography{/Users/vas/Documents/tex/ACP/lowT}
1100:
1101:
1102: \end{document}
1103: