1: \documentclass[aps,pre,amsmath,amssymb,twocolumn,nofootinbib]{revtex4}
2: %\documentclass[preprint,aps,pre,amsmath,amssymb,nofootinbib,endfloats]{revtex4}
3: \usepackage{pslatex}
4: \usepackage{graphicx}
5: \usepackage{dcolumn}
6: \usepackage{bm}
7: \usepackage{natbib}
8: %\usepackage{showkeys}
9:
10: \begin{document}
11: \title{The Cole-Cole Law for Critical Dynamics in Glass-Forming Liquids}
12:
13: \author{M. Sperl}
14: \affiliation{Duke University, Department of Physics, Box 90305,
15: Durham, NC 27708, USA}
16:
17: \date{\today}
18: \begin{abstract}
19:
20: Within the mode-coupling theory (MCT) for glassy dynamics, the asymptotic
21: low-frequency expansions for the dynamical susceptibilities at critical
22: points are compared to the expansions for the dynamic moduli; this shows
23: that the convergence properties of the two expansions can be quite
24: different. In some parameter regions, the leading-order expansion formula
25: for the modulus describes the solutions of the MCT equations of motion
26: outside the transient regime successfully; at the same time, the leading-
27: and next-to-leading order expansion formulas for the susceptibility fail.
28: In these cases, one can derive a Cole-Cole law for the susceptibilities;
29: and this law accounts for the dynamics for frequencies below the band of
30: microscopic excitations and above the high-frequency part of the
31: $\alpha$-peak. It is shown that this scenario explains the
32: optical-Kerr-effect data measured for salol and benzophenone (BZP). For
33: BZP it is inferred that the depolarized light-scattering spectra exhibit a
34: wing for the $\alpha$-peak within the Gigahertz band. This wing results
35: from the crossover of the von~Schweidler-law part of the $\alpha$-peak to
36: the high-frequency part of the Cole-Cole peak; and this crossover can be
37: described quantitatively by the leading-order formulas of MCT for the
38: modulus.
39:
40: \end{abstract}
41:
42: \pacs{61.20.Lc, 82.70.Dd, 64.70.Pf}
43:
44: \maketitle
45:
46: %%%
47: \section{Introduction\label{sec:intro}}
48:
49: During the past 15~years, several new spectrometers have been introduced
50: for the study of the glassy dynamics of liquids. The evolution of this
51: complex slow dynamics upon decreasing the temperature $T$ or increasing
52: the density $\rho$ has been documented for many systems for times $t$,
53: which exceed the natural time scale $t_\text{mic}$ for condensed-matter
54: motions by three or more orders of magnitude. Similar progress has been
55: made for molecular-dynamics simulations of liquid models. In parallel to
56: these experimental activities, a theory for the evolution of glassy
57: dynamics has been developed which is referred to as mode-coupling theory
58: (MCT). This theory is based on regular equations of motion for a set of
59: auto-correlation functions. The MCT equations lead to fold bifurcations
60: for the correlators' long-time limits if some control parameter like $T$
61: reaches a critical value $T_c$; this bifurcation describes a transition
62: from a liquid to an amorphous solid. The distance of the control parameter
63: from the critical value, say $\varepsilon = (T_c-T)/T_c$, can be used for
64: the discussion of the bifurcation dynamics as a small parameter. For
65: $\varepsilon$ tending to zero and times increasing to infinity, it is
66: possible to calculate asymptotic solutions of the MCT equations. The
67: leading-order results provide a set of general formulas, which explain the
68: qualitative features of the bifurcation scenario. Many fits of data with
69: these general formulas have been studied in order to test the relevance of
70: the MCT for the explanation of the experimental facts \cite{Goetze1999}.
71:
72: A set of general MCT results, which is of main interest in this paper,
73: concerns the critical dynamics. This dynamics is described by
74: auto-correlation functions $\phi(t)$ for control parameters at the
75: bifurcation point, say $T=T_c$. Equivalently, one can consider the
76: corresponding loss spectra $\chi''(\omega)$. These are the products of the
77: frequency $\omega$ and the Fourier-cosine transform of $\phi(t)$. The
78: central asymptotic formula is specified by (1) a positive number $f^c$,
79: which is called the plateau, (2) a positive amplitude, say $A$, and (3)
80: the critical exponent $a$, obeying $0<a<0.396$, $\lim_{t\rightarrow\infty}
81: t^a[\phi(t)-f^c] = A$. This formula is equivalent to
82: $\lim_{\omega\rightarrow 0} \chi''(\omega)/\omega^a = \sin(\pi
83: a/2)\Gamma(1-a)A$, with $\Gamma$ denoting the gamma function. For a given
84: transition point, all correlators are specified by the same exponent $a$,
85: but different critical points can differ in their value for $a$. The
86: leading-order long-time result for the correlator describes a power-law
87: decay: $\phi(t)-f^c \propto 1/t^a$. Equivalently, the leading-order result
88: for the low-frequency critical loss spectrum is given by a power law
89: variation $\chi''(\omega) \propto \omega^a$. For states near the
90: transition point, $\phi(t)$ and $\chi''(\omega)$ can be replaced by their
91: respective critical functions for short times, $t \ll t_\sigma$, or large
92: frequencies, $\omega t_\sigma \gg 1$. The time scale $t_\sigma$ is the
93: same for all correlators and diverges for states approaching the
94: transition point. For $t \geqslant t_\sigma$ and $\omega t_\sigma
95: \leqslant 1$, the correlators and spectra depend sensitively on
96: $\varepsilon$; for shorter times, $t \ll t_\sigma$ and $\omega t_\sigma
97: \gg 1$, the correlators and spectra depend on $\varepsilon$ smoothly. An
98: $\omega^a$ spectrum was identified first for the glass-forming molten salt
99: $0.4 Ca(NO_3)_2 0.6 K(NO_3)$ (CKN) in data obtained by neutron-scattering
100: spectroscopy \cite{Knaak1988}. This system was also used to document for
101: the first time the evolution of glassy dynamics within the full Giga-Hertz
102: band \cite{Li1992}: Using depolarized light-scattering spectroscopy, a
103: spectrum compatible with the $\omega^a$ law was found extending from 1GHz
104: to 400GHz. The $t^{-a}$ decay in the time domain was measured first for
105: density correlators by photon-correlation spectroscopy for a colloidal
106: suspension of hard spheres \cite{Megen1993b} with the density as a control
107: parameter.
108:
109: Other glass-forming systems that can be studied experimentally are found
110: in many van-der~Waals liquids. The natural time scale for inter-molecular
111: vibrations is a picosecond. For normal liquid behavior, one expects
112: correlations to decay to zero for times around some picoseconds. The
113: normal-liquid excitation spectra extend from, say, 0.5THz to, say, 5THz.
114: Glassy-dynamics spectra for several systems have been measured by
115: depolarized light-scattering spectroscopy, and the data were shown to be
116: consistent with the MCT bifurcation scenario. For example, spectra for
117: toluene have been fitted successfully with leading-order asymptotic
118: results for frequencies between 0.5GHz and 1THz and for temperatures
119: decreasing from $T=T_c+140$K to $T_c+10$K; $T_c \approx 150$K
120: \cite{Wiedersich2000}. For all these systems, the weight of the
121: glassy-dynamics part of the spectra is large compared to the part of
122: normal-liquid dynamics, i.e., the so-called $\alpha$-peak is large. In
123: agreement with the MCT prediction for such situations, the amplitude $A$
124: for the $\omega^a$ spectrum is small, and the low-frequency contributions
125: of the normal-liquid dynamics affect the region of the expected $\omega^a$
126: behavior of the spectra. As a result, no frequency interval has ever been
127: identified for a van-der-Waals liquid, where an $\omega^a$ spectrum can be
128: identified explicitly.
129:
130: Torre \textit{et. al} \cite{Torre1998} have introduced optical-Kerr-effect
131: (OKE) spectroscopy as a technique for the study of glassy dynamics. The
132: measurement provides the response function $\chi(t) \propto
133: -\partial_t\phi(t)$ for the same probing variable, which is studied in
134: depolarized light-scattering experiments. The Fourier-sine transforms of
135: $\chi(t)$ are proportional to the loss spectra $\chi''(\omega)$ mentioned
136: in the preceding paragraph. The evolution of the glassy dynamics of
137: m-toluidine was measured for temperatures decreasing from 295K to 250K.
138: The response functions could be fitted well by the scaling-law results
139: predicted by the leading-order asymptotic formulas for the MCT bifurcation
140: \cite{Torre2000}. The analysis implies a critical temperature $T_c$ near
141: 220K and a critical exponent $a$ near 0.3. However, lowering the
142: temperature to 225K, the critical power-law decay $\chi(t) \propto
143: 1/t^{1+a}$ was not observed \cite{Ricci2002}. The negative slope of the
144: measured $\log\chi$-versus-$\log t$ curve is not $1+a$, rather it is a
145: number smaller than unity, say $1-b'$. Decay laws $\chi(t) \propto
146: 1/t^{1-b'}$ with exponents $b'$ around 0.2 have been identified by Cang
147: \textit{et. al} \cite{Cang2003c} for a number of other van-der-Waals
148: liquids. The OKE response of salol was studied by Hinze \textit{et. al},
149: and the data were shown to be consistent with the known MCT scaling-law
150: formulas for the temperature decreasing from 340K to 266K
151: \cite{Hinze2000b}. The glassy response for $T = 257$K was measured with an
152: impressive accuracy for times increasing up to 0.5~$\mu$s; and the
153: dynamics for $t >$ 100ps displays the behavior expected from the
154: leading-order asymptotic results of MCT. However, the glassy dynamics for
155: 2ps $<t<$ 20ps manifests itself by a $\chi(t) \propto 1/t$ decay. The
156: corresponding correlator shows a logarithmic time dependence. All OKE
157: response functions measured so far demonstrate a glassy dynamics that
158: cannot be described by the general leading-order asymptotic formulas for
159: the MCT bifurcation in the regime $t \lesssim 30$ps. In a recent
160: alternative approach \cite{Berthier2005} is was possible to fit some of
161: the OKE data for $t > 30$ps.
162:
163: It was argued recently that the new facets of glassy dynamics discovered
164: by OKE spectroscopy \cite{Ricci2002,Cang2003c,Hinze2000b} can be
165: understood as generic implications of the Cole-Cole law for the critical
166: dynamics \cite{Goetze2004}. In the following, this statement shall be
167: explained in detail. Section~\ref{sec:general_MCT} summarizes the
168: equations of motion and the known scaling-law results of MCT. It is
169: explained in Sec.~\ref{sec:crit} that an asymptotic expansion of the
170: modulus can have a much larger range of validity than the expansion for
171: the susceptibility; the Cole-Cole law -- introduced in 1941 as empirical
172: law \cite{Cole1941} -- is derived in full generality from the microscopic
173: equations of motion. The relevance of these results is demonstrated for a
174: schematic model in Sec.~\ref{sec:appl} using parameter values that
175: describe the mentioned OKE data. The interplay between the Cole-Cole peak
176: and the $\alpha$-peak is investigated in Sec.\ref{sec:alphabeta} where for
177: specific parameter values the $\alpha$-peak displays a wing.
178: Section~\ref{sec:sum} presents a conclusion.
179:
180: %%%
181: \section{Essential MCT Formulas\label{sec:general_MCT}}
182: \subsection{Equations of Motion\label{subsec:basic_eqs}}
183:
184: Within the basic version of MCT, the dynamics of the system is described
185: by $M$ correlators $\phi_q(t)$, $q=1,\dots, M$. These are real and even
186: functions of the time $t$, which obey the initial conditions $\phi_q(t=0)
187: = 1$, $\partial_t\phi_q(t=0)=0$. The corresponding set of normalized
188: response functions is given by
189:
190: \begin{subequations}\label{eq:susc}
191: \begin{equation}\label{eq:susc:der}
192: \chi_q(t) = -\partial_t \phi_q(t)\,.
193: \end{equation}
194: Laplace transforms map functions from the time domain, say $F(t)$, in the
195: frequency domain. They shall be used with the convention
196: $\mathrm{LT}[F(t)](z) = i \int_0^\infty\,dt\,\exp[izt] F(t)$, $z = \omega
197: + i 0$. One gets $\mathrm{LT}[\chi_q(t)](z) = [1+\omega \phi_q(\omega)]$,
198: where $\phi_q(\omega) = \mathrm{LT}[\phi_q(t)](z)$. The normalized
199: dynamical susceptibilities are given by
200:
201: \begin{equation}\label{eq:susc:FDT}
202: \chi_q(\omega) = 1+\omega\phi_q(\omega)\,.
203: \end{equation}
204: \end{subequations}
205: The loss spectra $\chi''_q(\omega) = \mathrm{Im}\, \chi_q(\omega)$ are
206: related trivially to the fluctuation spectra $\phi_q''(\omega) =
207: \mathrm{Im}\, \phi_q(\omega)$: $\chi_q''(\omega) = \omega
208: \phi_q''(\omega)$.
209:
210: The Zwanzig-Mori formalism provides a fraction representation of
211: $\phi_q(\omega)$ in terms of a fluctuating-force correlator $M_q(\omega)$:
212: $\phi_q(\omega) = -1/\{ \omega - \Omega_q^2 / [\omega + M_q(\omega)] \}$.
213: The positive frequency $\Omega_q$ quantifies the initial decay of the
214: correlator $\phi_q(t) = 1 - [\Omega_q t]^2/2 + {\cal O}(t^3)$. Within MCT,
215: a white-noise term $\nu_q$ is split off from the kernel $M_q(\omega)$; the
216: remainder is represented in terms of a dimensionless function $m_q(t)$:
217: $M_q(\omega) = i\nu_q + \Omega_q^2 m_q(\omega)$; $\nu_q\geqslant 0$. Here,
218: $m_q(t)$ and $m_q(\omega)$ are related by Laplace-transformation. The
219: fraction representation is equivalent to the equations of motion
220:
221: \begin{subequations}\label{eq:EOM}
222: \begin{equation}\label{eq:EOM:int}
223: \begin{split}
224: \partial_t^2 \phi_q(t) +& \nu_q\partial_t\phi_q(t)\hfill \\
225: +& \Omega_q^2 \left[
226: \phi_q(t)+\int_0^t\,dt'\,m_q(t-t')\partial_{t'}\phi_q(t')
227: \right] = 0\,.
228: \end{split}
229: \end{equation}
230: The essential approximation in MCT is the expression of $m_q(t)$ as a
231: polynomial
232: ${\cal F}_q$ of the correlators:
233: \begin{equation}\label{eq:EOM:kernel}
234: m_q(t) = {\cal F}_q[\phi_1(t), \dots, \phi_M(t)]\,.
235: \end{equation}
236: \end{subequations}
237: There is no monomial contribution of order zero. The coefficients of the
238: polynomial are called mode-coupling coefficients. They are the coupling
239: constants of the theory and must not be negative. Within the microscopic
240: theory, the polynomials are of second order and the coefficients are given
241: by the equilibrium structure functions, which in turn are smooth functions
242: of control parameters like the temperature T for the states considered.
243:
244: At the generic transition mentioned in the preceding section, the
245: correlator's long-time limits depend singularly on the distance parameter
246: $\varepsilon$. For $\varepsilon < 0$, fluctuations disappear for long
247: times, $\phi_q(t\rightarrow\infty) = 0$. For $\varepsilon \geqslant 0$,
248: $\phi_q(t \rightarrow \infty) = f_q$, $0<f_q<1$. $q = 1,\dots, M$; the
249: fluctuations arrest. Within the microscopic version of MCT, the arrested
250: part $f_q$ has the meaning of the Debye-Waller factor of the solid
251: amorphous state. The leading-order variation with changes of $\varepsilon$
252: for the arrested part is given by $f_q - f^c_q \propto
253: \sqrt{\varepsilon}$, $\varepsilon \rightarrow 0^+$, $f_q^c > 0$,
254: $q=1,\dots, M$. The regularity of the MCT equations implies the following.
255: For $\varepsilon$ tending to arbitrarily small values, there appears an
256: arbitrarily large time interval, where $\phi_q(t)$ is arbitrarily close to
257: $f^c_q$. This critical arrested part $f^c_q$ has the meaning of a plateau
258: for the $\phi_q(t)$-versus-$\log t$ curves for states near the transition.
259: The corresponding plateau for the force correlators is $m_q(t \rightarrow
260: \infty) = f^{m\,c}_q = {\cal F}^c_q[f^c_1, \dots, f^c_M]$. At the
261: transition, $\phi_q(\omega)$ and $m_q(\omega)$ exhibit poles
262: $-f^c_q/\omega$ and $-f^{m\,c}_q/\omega$, respectively. Hence, there is a
263: region of small $|\varepsilon|$ and small $\omega$, where $\omega+i\nu_q$
264: can be neglected compared to $\Omega_q^2 m_q(\omega)$. This region is the
265: one for the MCT glassy dynamics. The fraction representation of the
266: correlators simplifies to $\phi_q(\omega) = -1/[\omega - 1/m_q(\omega)]$.
267: This formula can also be noted as
268:
269: \begin{subequations}\label{eq:MCTlong}
270: \begin{equation}\label{eq:MCTlong:m}
271: \omega m_q(\omega) = \omega \phi_q(\omega)/\left[ 1 + \omega\phi_q(\omega)
272: \right]\,.
273: \end{equation}
274: Equation~(\ref{eq:susc:FDT}) yields the equivalent expression for the
275: dynamical susceptibility:
276:
277: \begin{equation}\label{eq:MCTlong:chi}
278: \chi_q(\omega) = 1/\left[ 1-\omega m_q(\omega) \right]\,.
279: \end{equation} \end{subequations}
280: Within the regime of glassy dynamics, $\left[1-\omega m_q(\omega)\right]$
281: has the meaning of a modulus for the response described by $\chi_q(t)$.
282: The pair of Eqs.~(\ref{eq:EOM:kernel}, \ref{eq:MCTlong:m}) can fix the
283: solution only up to some overall time scale $t_0$. The latter is
284: determined by matching of the transient dynamics with the glassy dynamics.
285: For further details and for a list of original papers, the reader can
286: consult Ref.~\cite{Franosch1997}.
287:
288:
289: %%
290: \subsection{Scaling Laws\label{subsec:scaling_laws}}
291:
292: There is a straight-forward recipe to calculate from the coupling
293: coefficients at the transition point and from the critical arrested parts
294: $f^c_q$ a number $\lambda$, $1/2 \leqslant \lambda < 1$. It is called the
295: exponent parameter for the chosen transition point of the model under
296: discussion. It fixes the critical exponent $a$, mentioned in
297: Sec.~\ref{sec:intro}. It fixes a further exponent $b$, $0<b\leqslant 1$,
298: which is called the von~Schweidler exponent. The equation for the two
299: exponents reads $\Gamma(1-a)^2/\Gamma(1-2a) = \lambda =
300: \Gamma(1+b)^2/\Gamma(1+2b)$. Parameter $\lambda$ also specifies a pair of
301: equations for a pair of functions $g_\pm(\hat{t})$, which are defined for
302: $\hat{t}>0$. The equations read $\pm 1 + \lambda g_\pm(\hat{t})^2 =
303: (d/d\hat{t}) \int_0^{\hat{t}}\,d\hat{t}\, g_\pm(\hat{t}-\hat{t}')
304: g_\pm(\hat{t}')$, and they have to be solved with the initial condition
305: $\lim_{\hat{t}\rightarrow 0} \hat{t}^a g_\pm(\hat{t}) = 1$. Up to
306: corrections of order $\hat{t}^a$, one gets for small $\hat{t}$:
307:
308: \begin{subequations}\label{eq:gpm}
309: \begin{equation}\label{eq:gpm:ta}
310: g_\pm(\hat{t}\ll 1) = 1/\hat{t}^a\,.
311: \end{equation}
312: The function $g_+(\hat{t})$ approaches its long-time limit exponentially,
313: $g_+(\hat{t}\gg 1) = 1/\sqrt{1-\lambda}$.
314: The function $g_-(\hat{t})$ exhibits a power-law divergence for large
315: $\hat{t}$. Up to corrections of order $1/\hat{t}^b$, one gets
316:
317: \begin{equation}\label{eq:gpm:B}
318: g_-(\hat{t}\gg 1) = -B\hat{t}^b\,.
319: \end{equation}
320: \end{subequations}
321: There are tables allowing the determination of $a$, $b$ and $B$ from a
322: given $\lambda$. There are also tables to determine $g_\pm(\hat{t})$ with
323: an accuracy sufficient for all practical purposes \cite{Goetze1990}.
324:
325: The functions $g_\pm(\hat{t})$ are the shape functions for the first
326: scaling law of MCT. For $\varepsilon$ tending to zero, there appears a
327: time interval of diverging length, within which $\Delta_q(t) = |\phi_q(t)
328: - f^c_q|$ is arbitrary small. The leading-order solution for the small
329: parameter $\Delta_q(t)$ yields the first scaling-law. The solution
330: assumes the form
331:
332: \begin{subequations}\label{eq:phi_}
333: \begin{equation}\label{eq:phi_gpm}
334: \phi_q(t) = f^c_q + h_q \sqrt{|\sigma|} g_\pm(t/t_\sigma)\,,\quad
335: \varepsilon \gtrless 0\,.
336: \end{equation}
337: The amplitudes $h_q > 0$, $q = 1, \dots, M$ are calculated from the
338: mode-coupling coefficients at the critical point. The separation parameter
339: $\sigma$ is defined similarly: it is a smooth function of the control
340: parameters and can be linearized close to the transition point, $\sigma =
341: C\varepsilon + {\cal O}(\varepsilon^2)$, with a positive coefficient $C$
342: that depends on the chosen control parameter. The first critical time
343: scale $t_\sigma$ is given by the critical exponent $a$, the separation
344: parameter $\sigma$, and the time scale $t_0$, which is defined by the
345: short-time dynamics:
346:
347: \begin{equation}\label{eq:sigma:tsigma}
348: t_\sigma = t_0/|\sigma|^\delta\,,\quad \delta = 1/(2a)\,.
349: \end{equation} \end{subequations}
350: The strong control-parameter dependence of the correlators near the
351: plateau is described solely by that of the correlation scale
352: $\sqrt{|\sigma|}$ and of the time scale $t_\sigma$.
353:
354: From Eqs.~(\ref{eq:gpm:ta},\ref{eq:phi_gpm},\ref{eq:sigma:tsigma}), one
355: gets the leading-order asymptotic law for the decay of the critical
356: correlator discussed in Sec.~\ref{sec:intro}: $\phi_q(t) - f^c_q = h_q
357: (t_0/t)^a$. From Eqs.~(\ref{eq:gpm:B}, \ref{eq:phi_gpm},
358: \ref{eq:sigma:tsigma}) one gets the von~Schweidler law for the decay of
359: the liquid correlators below the plateau:
360:
361: \begin{equation}\label{eq:vS:phi}
362: \phi_q(t) = f^c_q + h_q (t/t_{\sigma'})^b\,,\quad t_\sigma \ll t \ll
363: t_{\sigma'}\,,\quad \varepsilon < 0 \,.
364: \end{equation}
365: Here, the second critical time scale of MCT reads $t_{\sigma'} = t_0/[
366: B^{1/b} |\sigma|^\gamma]$, with $\gamma = 1/(2a) + 1/(2b) \,. $ Up to
367: errors of order $|\varepsilon|$, the decay of the liquid correlators below
368: the plateau is described by the second scaling law of MCT
369:
370: \begin{equation}\label{eq:alpha}
371: \phi_q(t) = \tilde{\phi}_q(t/t'_\sigma) \,,\quad
372: t'_\sigma \ll t\,,\quad \varepsilon <0\,.
373: \end{equation}
374: Here, $\tilde{\phi}_q(\tilde{t})$ is an $\varepsilon$-independent shape
375: function. Its initial part is given by von~Schweidler's law:
376: $\tilde{\phi}_q(\tilde{t}\ll 1) = f^c_q - h_q \tilde{t}^b$. Corrections of
377: order $\sqrt{|\varepsilon|}$ modify the formula (\ref{eq:alpha}) for the
378: below-plateau decay in a regime where $\phi_q(t)$ is close to the plateau.
379: These corrections are described by Eq.~(\ref{eq:phi_gpm}) for $t \geqslant
380: t_\sigma$. The below-plateau decay is referred to as $\alpha$-process, and
381: Eq.~(\ref{eq:alpha}) formulates the superposition principle. Details of
382: the derivation of the cited results, references to the original work, and
383: a comprehensive demonstration for the hard-sphere system can be found in
384: Refs.~\cite{Franosch1997,Fuchs1998}.
385:
386:
387: %%%
388: \section{Critical Dynamics and Cole-Cole law\label{sec:crit}}
389:
390: In this section, correlators and susceptibilities shall be discussed for
391: states at the transition point, say $T=T_c$. Focusing on the range of
392: validity, the essential formulas are analyzed for the correlators in
393: \ref{sec:crit:corr}, and for the susceptibilities in \ref{sec:crit:susc}.
394: The Cole-Cole law is derived in \ref{sec:crit:cole}.
395:
396: %%
397: \subsection{Power-Law Solution for Correlation
398: Functions\label{sec:crit:corr}}
399:
400: The leading-order scaling-law formulas~(\ref{eq:gpm:ta},\ref{eq:phi_gpm})
401: yield for the critical correlators
402:
403: \begin{equation}\label{eq:crit:corr}
404: \phi_q(t) = f^c_q + h_q (t_0/t)^a\,.
405: \end{equation}
406: A first question, whose answer is not implied by the results cited in
407: Sec.~\ref{subsec:scaling_laws}, concerns the range of validity of
408: Eq.~(\ref{eq:crit:corr}) for short times. Let us define an onset time
409: $t^*_q$ for the power law by the request, that Eq.~(\ref{eq:crit:corr})
410: describes $\phi_q(t)- f^c_q$ within a relative error of, say, 10\%, for $t
411: \geqslant t^*_q$. For normal-liquid dynamics, one would expect the
412: correlators to decay to zero for times around $t_\text{mic}$. A second
413: question to be discussed is: How can one describe the critical correlators
414: in cases where $t^*_q$ exceeds $t_\text{mic}$, i.e., when there is a gap
415: between the end of the transient regime and the onset of the critical
416: power law?
417:
418: The asymptotic solution~(\ref{eq:crit:corr}) can be extended to an
419: asymptotic series expansion in powers of $t^{-a}$. The result up to the
420: next-to-leading term shall be noted as
421:
422: \begin{subequations}\label{eq:critcorr}
423: \begin{equation}\label{eq:critcorr:phi}
424: \phi_q(t) = f^c_q + h_q (t_0/t)^a \left[ 1+ \hat{K}_q (t_0/t)^a \right]\,.
425: \end{equation}
426: Here, remainders which are given by $(t_0/t)^{3a}$ times some power of
427: $\ln(t/t_0)$ are dropped. The Tauberian theorem yields an equivalent
428: formula in the frequency domain:
429:
430: \begin{equation}\label{eq:critcorr:chi}
431: \begin{split}
432: \omega\phi_q(\omega) = -f^c_q - &\Gamma(1-a) h_q (-i\omega t_0)^a\\
433: &- \Gamma(1-2 a) h_q \hat{K}_q (-i\omega t_0)^{2a}\,.
434: \end{split}\end{equation}
435: \end{subequations}
436: In this formula, terms are dropped, which are proportional to
437: $\omega^{3a}$ times some power of $\ln\omega$. The preceding two formulas
438: are the starting point for the following derivations. It is a
439: straight-forward procedure to calculate the correction amplitude
440: $\hat{K}_q$ from the mode-coupling coefficients at the critical point
441: $T=T_c$ \cite{Franosch1997}. From Eq.~(\ref{eq:critcorr:phi}), one can
442: estimate an onset time of $t^*_q/t_0 = [10|\hat{K}_q|]^\frac{1}{a}$. This
443: is an estimate based on the assumption that higher-order expansion terms
444: in Eq.~(\ref{eq:critcorr:phi}) do not influence the results seriously for
445: $t \geqslant t^*_q$. Typically for many systems, the critical exponent $a$
446: is around 0.3, and $1/a \gtrsim 3$. Hence, $t^*_q$ depends sensitively on
447: the correction amplitude $\hat{K}_q$. As a result, $t^*_q$ can vary
448: considerably for different $q$. As a relevant example, let us cite the
449: results for the density-fluctuation correlators of a system of hard
450: spheres of diameter $d$ \cite{Franosch1997}: The time $t_1^*$ referring to
451: a wave-number $q_1d = 7.0$ exceeds the time $t_2^*$ for $q_2d = 10.6$ by a
452: factor of around 100.
453:
454: %%
455: \subsection{Power-Law Solution for the
456: Susceptibilities\label{sec:crit:susc}}
457:
458: From Eqs.~(\ref{eq:susc:der}, \ref{eq:crit:corr}), one obtains the
459: long-time result for the response functions up to leading-order
460: corrections,
461: \begin{equation}\label{eq:chit0}
462: \chi(t) t_0 = a h_q (t_0/t)^{1+a}\,.
463: \end{equation}
464: According to Eq.~(\ref{eq:critcorr:phi}), the leading order result in
465: Eq.~(\ref{eq:crit:corr}) is valid for $t \geqslant t_q^*$, but the onset
466: time for Eq.~(\ref{eq:chit0}) is later, $t \geqslant t^*_q 2^{1/a}$, where
467: typically $2^{1/a} \approx 10$. Hence, the detection of the critical
468: dynamics in its leading asymptotic form is more difficult for the response
469: functions than for the correlators.
470:
471: From Eq.~(\ref{eq:critcorr:chi}), one arrives at the expansion formula for
472: the absorptive part of the dynamical susceptibility
473:
474: \begin{subequations}\label{eq:chi_power}
475: \begin{equation}\label{eq:chi_power:chi}\begin{split}
476: \chi_q''(\omega) = \left[ \Gamma(1-a) \sin\left( \pi a/2 \right) \right]
477: &h_q (\omega t_0)^a \\&\times
478: \left[ 1+ k_a \hat{K}_q (\omega t_0)^a \right]\,,
479: \end{split}\end{equation}
480:
481: \begin{equation}\label{eq:chi_power:ka}
482: k_a = 2 \Gamma(1-a) \cos\left( \pi a/2 \right)/\lambda\,.
483: \end{equation}
484: The leading-order power-law result reads
485:
486: \begin{equation}\label{eq:chi_power:leading}
487: \chi_q''(\omega) = \left[ \Gamma(1-a) \sin\left( \pi a/2 \right) \right]
488: h_q (\omega t_0)^a\,,
489: \end{equation}
490: \end{subequations}
491: which has an onset frequency of $\omega_q^* = 1/(t^*_q k_a^{1/a})$; for
492: $\omega \leqslant \omega_q^*$, the leading-order result in
493: Eq.~(\ref{eq:chi_power:leading}) describes the critical spectrum with an
494: error smaller than 10\%. If $\lambda$ decreases from 1 to $1/2$, $k_a$
495: increases from 2 to near 5. For $\lambda$ near a typical value of 0.7,
496: $k_a$ is above 3, and the onset frequency for the $\omega^a$-law is about
497: 30 times smaller than $1/t^*_q$. Therefore, the detection of the critical
498: power-law in the loss spectra is even more difficult then in the response
499: function.
500:
501: Including the leading-correction terms in the asymptotic formulas, as
502: noted in Eqs.~(\ref{eq:critcorr:phi}, \ref{eq:chi_power:chi}), is an
503: obvious manner to extend the range of applicability of the analytic
504: description of the dynamics. This is demonstrated comprehensively for the
505: MCT for the hard-sphere system in Refs.~\cite{Franosch1997, Fuchs1998}.
506: But there are cases, where this procedure does not lead to satisfactory
507: results. The mean-squared displacement is an example, where the
508: description of the increase towards the plateau by the analog of
509: Eq.~(\ref{eq:critcorr:phi}) cannot account for the glassy dynamics
510: \cite{Fuchs1998}.
511:
512: %%
513: \subsection{The Cole-Cole Law\label{sec:crit:cole}}
514:
515: Equation~(\ref{eq:critcorr:chi}) is an asymptotic expansion in terms of
516: powers of the small quantity $\xi = (-i\omega t_0)^a$ that holds up to
517: errors of $\xi^3$. Substitution of this expansion into
518: Eq.~(\ref{eq:MCTlong:m}) provides an analogous expansion for $\omega
519: m_q(\omega)$. Let us indicate the non-trivial parts of the coefficients by
520: a superscript $m$:
521:
522: \begin{equation}\label{eq:mqw}\begin{split}
523: -\omega m_q(\omega) = f_q^{m\,c} +& \Gamma(1-a) h^m_q (-i \omega t_0)^a \\
524: &+ \Gamma(1-2a) h^m_q \hat{K}^m_q (-i \omega t_0)^{2a}\,.
525: \end{split}\end{equation}
526: Comparing coefficients of equal powers of $\xi$, one finds:
527:
528: \begin{subequations}\label{eq:cmq}
529: \begin{equation}\label{eq:fqmc}
530: f_q^{m\,c} = f_q^c/(1-f_q^c)\,,\quad f_q^c = f_q^{m\,c} /(1 +
531: f_q^{m\,c} ) \,.
532: \end{equation}
533: \begin{equation}\label{eq:hmq}
534: h_q^m = h_q /(1-f_q^c)^2\,,\quad h_q = h_q^m/(1+ f_q^{m\,c} )^2\,.
535: \end{equation}
536: \begin{equation}\label{eq:cmq:cmq}\begin{split}
537: \hat{K}^m_q = \hat{K}_q + \lambda [h_q/(1-f_q^c)]\,,\quad\\
538: \hat{K}_q = \hat{K}^m_q -\lambda [h^m_q / (1 + f_q^{m\,c} )]\,.
539: \end{split}
540: \end{equation}
541: \end{subequations}
542:
543: Since the auto-correlation functions are normalized, $\phi_q(t=0) = 1$,
544: one gets $f_q^c < 1$. The kernel $m_q(t=0)$ can have any positive value.
545: Therefore, in principle, $f_q^{m\,c}$ can be any positive number. The
546: different normalizations can be eliminated in the critical amplitudes by
547: comparing the ratios $h_q/f_q^c$ and $h^m_q/f_q^{m\,c}$:
548:
549: \begin{equation}\label{eq:hqfq}
550: [h^m_q/f^{m\,c}_q] = [h_q/f_q^c]/(1-f^c_q)\,.
551: \end{equation}
552: These ratios determine the relative amplitude of the dynamics around the
553: plateau within the first scaling-law regime. Equation (\ref{eq:phi_gpm})
554: yields $(\phi_q(t)-f^c_q)/f^c_q = [h_q/f_q^c] \sqrt{|\sigma|}
555: g_\pm(t/t_\sigma)$, and a corresponding identity holds for $m_q(t)$. From
556: Eq.~(\ref{eq:hqfq}) one concludes that the relative amplitude of the
557: first-scaling-law contribution is larger by $1/(1-f_q^c) > 1$ for the
558: modulus than for the susceptibility; and $1/(1-f_q^c)$ increases with
559: $f^c_q$, the $\alpha$-peak contribution of the loss spectrum.
560: Consequently, within the range of validity of the first scaling law, it is
561: easier to measure the first-scaling-law contribution for the modulus than
562: that for the susceptibility.
563:
564: The expansion~(\ref{eq:mqw}) can be substituted into
565: Eq.~(\ref{eq:MCTlong:chi}) in order to obtain an asymptotic expansion for
566: the inverse of the critical susceptibility, i.e., for the modulus. The
567: result shall be noted in the form
568:
569: \begin{equation}\label{eq:ccc}
570: \chi_q(\omega) = \chi_{0\,q}^{cc} /\left[ 1 + \left(
571: -i\omega/\omega_q^{c} \right)^a + \hat{K}^{cc}_q \left(
572: -i\omega/\omega_q^{c} \right)^{2a}\right] \,.
573: \end{equation}
574: The expression $\chi_q(\omega)^{-1}$ is correct up to errors of the order
575: $\omega^{3a}$. The three parameters specifying this formula can be
576: expressed in terms of the coefficients in Eq.~(\ref{eq:mqw}) and, via
577: Eqs.~(\ref{eq:cmq}), in terms of the coefficients
578: specifying the correlators in Eq.~(\ref{eq:critcorr:phi}). One gets
579: $\chi_{0\,q}^{cc} = 1/(1+f_q^{m\,c})$. This amplitude is the complement of
580: the $\alpha$-peak strength $f^c_q$ of the normalized loss spectrum,
581:
582: \begin{subequations}\label{eq:cccpar}
583: \begin{equation}\label{eq:cccpar:chi0}
584: \chi_{0\,q}^{cc} = 1- f^c_q\,.
585: \end{equation}
586: The characteristic frequency entering the new formula for the
587: susceptibility, is given by $(\omega_q^{c}t_0)^a = (1 + f_q^{m\,c}) /
588: \left( h^m_q \Gamma(1-a) \right)$ or by
589:
590: \begin{equation}\label{eq:cccpar:wcc}
591: \omega_q^{c}t_0 = \left[ (1-f^c_q) / \left(h_q \Gamma(1-a)\right)
592: \right]^{1/a}
593: \,.
594: \end{equation}
595: The correction amplitude reads
596: $\hat{K}_q^{cc} = \hat{K}^m_q (1+f^{m\,c}_q)/(\lambda h^m_q)$,
597: which is equivalent to
598: \begin{equation}\label{eq:cccpar:cqcc}
599: \hat{K}_q^{cc} = 1 + \left[(1-f^c_q)/h_q\right] \hat{K}_q / \lambda\,.
600: \end{equation}
601: \end{subequations}
602:
603: If the frequencies are so small that $|\hat{K}_q^{cc}
604: (-i\omega/\omega_q^{c})| \ll 1$, Eq.~(\ref{eq:ccc}) simplifies to the
605: transparent expression of the Cole-Cole law,
606:
607: \begin{equation}\label{eq:ccleft:chi}
608: \chi_q(\omega) = \chi_{0\,q}^{cc}/\left[ 1 + (-i\omega/\omega_q^{c})^a
609: \right]\,.
610: \end{equation}
611: The condition of validity for this formula means that the critical modulus
612: can be described by the simple power law in the first line of
613: Eq.~(\ref{eq:mqw}). The modulus spectrum obeys a formula analog to
614: Eq.~(\ref{eq:chi_power:leading}): $\omega m_q''(\omega) = \Gamma(1-a)
615: \sin\left(\pi a/2\right) h^m_q (\omega t_0)^a$. Using the 10\% criterion
616: from above, the last term in Eq.~(\ref{eq:mqw}) can be dropped for
617: frequencies $\omega t_0 \leqslant \left[ \lambda/\left( 10 \Gamma(1-a)
618: |\hat{K}^m_q| \right) \right]^{1/a}$. The correlator corresponding to the
619: loss spectrum in Eq.~(\ref{eq:ccleft:chi}) is given by the Mittag-Leffler
620: function of index $a$: $M_a(x) = \sum_{n=0}^\infty x^n/\Gamma(1+n a)$,
621:
622: \begin{equation}\label{eq:ML}
623: \phi_q(t) = f_q^c + \chi_{0\,q}^{cc} M_a\left[-(t\omega_q^{c})^a\right]
624: \,.
625: \end{equation}
626: The Mittag-Leffler function can be calculated efficiently by Fourier-back
627: transformation of $\chi''_q(\omega) / \omega$.
628:
629: Equation~(\ref{eq:ccleft:chi}) describes a very broad peak for the loss
630: spectrum $\chi_q''(\omega)$ with a maximum located at $\omega =
631: \omega_q^{c}$. The spectrum is invariant under the interchange of
632: $(\omega/\omega_q^{c})$ to $(\omega_q^{c}/\omega)$: the
633: $\chi_q''(\omega)$-versus-$\log\omega$ curve is symmetric. The width of
634: the peak increases strongly with decreasing exponent $a$. For $a \approx
635: 0.3$, $\omega$ has to increase by more than a factor of $10^4$ in order to
636: scan the interval of frequency where $\chi_q''(\omega) \geqslant
637: \chi_q''(\omega_q^{c}) / 2$. This susceptibility formula known as
638: Cole-Cole law, Eq.~(\ref{eq:ccleft:chi}), was introduced as an empirical
639: formula for dielectric loss spectra in glassy systems \cite{Cole1941}. In
640: particular, broad peaks located above the $\alpha$ peaks -- often called
641: $\beta$-peaks in this context -- have been fitted by it.
642:
643: The Cole-Cole formula, Eq.~(\ref{eq:ccleft:chi}), exhibits simple limits
644: for small and large frequencies. For $(\omega/\omega_q^{c})^a\ll 1$, the
645: Cole-Cole susceptibility reproduces the general power-law for the loss
646: spectrum, Eq.~(\ref{eq:chi_power:leading}), $\chi_q''(\omega) \propto
647: \omega^a$. Similarly, the corresponding Mittag-Leffler correlator
648: reproduces the general long-time asymptote for the response,
649: Eq.~(\ref{eq:chit0}). For $(\omega/\omega_q^{c})^a \gg 1$, the Cole-Cole
650: spectrum describes a critical spectrum which decreases with increasing
651: frequency: $\chi_q''(\omega) \propto 1/\omega^a$. This is similar to a
652: von~Schweidler-law spectrum. The corresponding Mittag-Leffler correlator
653: decreases according to the law $\phi(t) -\text{const.} \propto
654: - t^a$. This yields a response function $\chi(t) \propto 1/t^x$ with $x =
655: 1-a <1$. Such behavior is consistent with the one detected by the OKE
656: results for van-der Waals liquids \cite{Cang2003c}.
657:
658: The crossover of the critical spectrum from the low-frequency wing to the
659: high-frequency wing of the Cole-Cole loss peak has a counterpart for the
660: Mittag-Leffler correlator that is most easily seen in a semilogarithmic
661: plot. For $(t\omega_q^{c})^a \ll 1$, the $\left[ \phi(t) - \text{const.}
662: \right]$-versus-$\log t$ curve is bent downward, as known for the
663: von~Schweidler curves. For $(t\omega_q^{c})^a \gg 1$, there is the
664: critical power-law decay which shows up as an upward-bent curve. Hence,
665: the $\phi(t)$-versus-$\log t$ curve exhibits an inflection point. For an
666: exponent $a$ around 0.3, the curve is nearly straight for an increase of
667: $\log_{10} t$ by a factor of about 3. In this case, there is
668: nearly-logarithmic decay of the critical correlator for a time variation
669: over three orders of magnitude.
670:
671: It depends on the size of $1/\omega_q^{c}$ relative to the microscopic
672: time scale $t_\text{mic}$ which part of the Mittag-Leffler correlator
673: dominates the critical dynamics in the time range of interest. The crucial
674: question concerns the range of validity of the leading-order result for
675: the susceptibility and for the modulus. These ranges are determined by the
676: correction amplitudes. If $|\hat{K}_q|$ is large and $|\hat{K}^m_q|$
677: small, the Cole-Cole formula has a larger range of applicability than the
678: general simple power-law formula~(\ref{eq:chi_power:leading}). But, if
679: $|\hat{K}^m_q|$ is larger than $|\hat{K}_q|$,
680: Eq.~(\ref{eq:chi_power:leading}) is a better approximation for the
681: critical decay than the Cole-Cole spectrum. Since the brackets in
682: Eqs.~(\ref{eq:cmq}) are positive, there are two obvious results. If
683: $\hat{K}_q$ is positive, there holds $\hat{K}^m_q > \hat{K}_q$. In this
684: case, the leading-order result for the susceptibility is a better
685: description of the critical spectrum than the Cole-Cole spectrum. If
686: $\hat{K}^m_q$ is negative, there holds $|\hat{K}^m_q| < |\hat{K}_q|$ and
687: the application of the Cole-Cole law is superior to the one of the general
688: leading-order result for the loss spectrum. For negative correction
689: amplitudes $\hat{K}_q$, there is a trend for cancellation of the two
690: contributions to $\hat{K}_q^m$ in Eq.~(\ref{eq:cmq:cmq}). It is a generic
691: situation, that $|\hat{K}^m_q|$ is smaller than $|\hat{K}_q|$. Hence, for
692: $\hat{K}_q < 0$, it is expected that the Cole-Cole law is a good
693: description of the critical dynamics.
694:
695: A detailed discussion of the MCT equations has shown that a large arrested
696: part $f^c_q$ implies a large negative correction amplitude $\hat{K}_q$ for
697: the critical decay \cite{Franosch1997}. In this case, the onset time
698: $t^*_q$ may exceed the microscopic time scale $t_\text{mic}$ by several
699: orders of magnitude. The time range for the applicability of the
700: leading-order asymptotic description of the critical decay may be outside
701: the accessible range of the existing spectrometers. But, this is the
702: situation, where $|\hat{K}^m_q|$ may be so small that the Cole-Cole law or
703: the Mittag-Leffler correlator can provide an adequate description of the
704: critical dynamics.
705:
706:
707: %%%
708: \section{Application to Schematic Models\label{sec:appl}}
709:
710: The preceding results shall be demonstrated in a simple schematic model
711: for two different cases where the model parameters are adjusted to
712: reproduce the data measured for benzophenone (BZP) and salol
713: \cite{Hinze2000b,Cang2003c,Goetze2004}
714:
715: The simplest MCT models deal with a single correlation function only. For
716: these models, the formulas in the preceding sections simplify since the
717: label $q$ can be dropped. The correlator, the critical arrested part, and
718: the critical amplitude shall be denoted by $\phi(t)$, $f^c$, and $h$,
719: respectively. The model is specified by the frequency $\Omega$ and the
720: friction coefficient $\nu$ in the equation of motion~(\ref{eq:EOM:int}).
721: Furthermore, there are the non-negative coefficients $v_l,\, l \geqslant
722: 1$, which specify the mode-coupling monomial of order $l$ for ${\cal F}$
723: in Eq.~(\ref{eq:EOM:kernel}). In applications for data descriptions, the
724: specified numbers are considered as smooth functions of the physical
725: control parameters.
726:
727: The simplest model for the polynomial ${\cal F}$, which can reproduce all
728: possible values for the exponent parameter $\lambda$, is given by the
729: following formula for the fluctuating-force kernel
730:
731: \begin{equation}\label{eq:mF12}
732: m(t) = v_1\phi(t) + v_2 \phi(t)^2\,.
733: \end{equation}
734: The two coupling constants $v_1$ and $v_2$ specify the state of the system
735: by a point in the first quadrant of the $v_1$-$v_2$ plane, cf. lower inset
736: in Figs.~\ref{fig:BZP} and \ref{fig:salol}. The points $(v_1^c, v_2^c)$
737: for generic fold bifurcations are located on a piece of a parabola. The
738: position of the specific transition point can be characterized by the
739: value for $\lambda$ \cite{Goetze1984}:
740:
741: \begin{subequations}\label{eq:vc}
742: \begin{equation}\label{eq:vc:vc}
743: v_1^c = (2\lambda-1)/\lambda^2 \,,\; v_2^c = 1/\lambda^2 \,,\; f^c =
744: 1-\lambda \,,\; h=\lambda\,.
745: \end{equation}
746: The separation of some point $(v_1, v_2)$ from the transition point
747: $(v_1^c, v_2^c)$, see upper insets Figs.~\ref{fig:BZP} and
748: \ref{fig:salol}, is given by
749: \begin{equation}\label{eq:vc:sigma}
750: \sigma = \left[ \hat{v}_1 + \hat{v}_2 (1-\lambda) \right] \lambda
751: (1-\lambda) \,,\quad \hat{v}_{1,\,2} = v_{1,\,2} - v^c_{1,\,2}
752: \,.
753: \end{equation}
754: \end{subequations}
755: The arrested part of the correlator for the glass state reads $\phi(t
756: \rightarrow \infty) = h[\sigma/(1-\lambda)]^{1/2}+{\cal O}(\sigma^{3/2})$.
757: The splitting of this value in an amplitude $h$ and a remainder is not
758: unique. The cited value for the critical amplitude $h$ follows the
759: conventions made in the preceding literature \cite{Franosch1997}. The
760: general expression for the correction amplitude $\hat{K}$ for $M=1$ models
761: \cite{Goetze1989c} is specialized easily to $\hat{K} = \kappa(x)$, where
762: the functions $\kappa(x)$ is defined by
763:
764: \begin{equation}\label{eq:kappa}
765: \kappa (x) = \frac{1}{2} \Gamma(1-x)^3/ \left[
766: \lambda \Gamma(1-3x) - \Gamma(1-x) \Gamma(1-2x)
767: \right] \,.
768: \end{equation}
769:
770: Function $\kappa(x)$ increases monotonically with increasing $\lambda$: it
771: decreases with increasing $x$. For $\lambda = 1/2$, i.e., for $a =
772: 0.395\dots$, one gets $\kappa(a) = -0.169\dots$ . For $\lambda$
773: approaching 1, $a$ tends to zero and $\kappa(a)$ diverges. For $a = 1/3$,
774: $\lambda = 0.684\dots$, the correction amplitude vanishes. All
775: higher-order corrections for the critical decay outside the transient
776: regime vanish as well for this special value of $\lambda$
777: \cite{Fuchs1999b}. Therefore, for $\lambda$ near 0.7, the simple power-law
778: formulas of the leading-order asymptotic-expansion theory like
779: Eqs.~(\ref{eq:crit:corr}, \ref{eq:chit0}, \ref{eq:chi_power:leading})
780: describe the critical dynamics very well.
781:
782: For the specified model, there holds $f^c\leqslant 1/2$. Hence, this model
783: cannot be used to describe glassy-dynamics data for systems where the
784: $\alpha$-peak loss spectra have a weight $f^c$ which exceeds 50\% of the
785: total weight. Moreover, the relation between the exponent parameter
786: $\lambda$ and the critical arrested part $f^c$, as formulated by
787: Eq.~(\ref{eq:vc:vc}), is an artifact of the model. The minimum requirement
788: for a schematic model for data analysis is the freedom to adjust $\lambda$
789: and $f^c$ independently. This goal can be achieved by introducing a second
790: correlator. The first correlator $\phi_{q=1}(t) = \phi(t)$ is used as a
791: caricature of the density-fluctuation dynamics. It provides the exponent
792: parameter. The second one describes the dynamics of some probing variable,
793: say, $A$. The correlator shall be denoted by $\phi_{q=2}(t) = \phi_A(t)$.
794: All quantities referring to this correlator shall be indicated by an index
795: $A$. The equation of motion~(\ref{eq:EOM:int}) is specified by $\Omega_A >
796: 0$, $\nu_A \geqslant 0$, and a kernel $m_A(t)$. The two frequencies
797: quantify the transient dynamics. The kernel is a polynomial
798: (\ref{eq:EOM:kernel}) of the two correlators involved. The simplest model
799: describing the coupling of the probing variable to the density-fluctuation
800: is given by \cite{Sjoegren1986}
801:
802: \begin{equation}\label{eq:mA}
803: m_A(t) = v_A \phi(t) \phi_A(t)\,.
804: \end{equation}
805: The coupling of the probing variable to the density fluctuations is
806: quantified by $v_A >0$. Since there is no influence of the dynamics of the
807: probing variable on the density dynamics, the same scaling-law function
808: $g_\pm(t/t_\sigma)$, Eq.~(\ref{eq:phi_gpm}), describes the leading-order
809: near-plateau dynamics of $\phi_A(t)$ as it does for that of $\phi(t)$.
810: Notice in particular that the parameters $\Omega_A$, $\nu_A$, and $v_A$ do
811: not modify the time scale $t_0$. In applications for data descriptions,
812: the model parameters $\Omega_A$, $\nu_A$, and $v_A$ are to be considered
813: as smooth functions of the physical control parameters, cf. lower insets
814: in Figs.~\ref{fig:BZP} and \ref{fig:salol}.
815:
816: The loss spectrum $\chi_A''(\omega)$ develops a $\beta$-peak upon
817: increasing the mode-coupling coefficient $v_A$ \cite{Buchalla1988}. For
818: this special case and in the limit of infinite $v_A$, the Cole-Cole
819: susceptibility and the Mittag-Leffler correlator have been derived for the
820: critical dynamics of the probing variable in Ref. \cite{Goetze1989c}. The
821: model specified by Eqs.~(\ref{eq:mF12}, \ref{eq:mA}) was used repeatedly
822: for the description of experimental data
823: \cite{Singh1998,Ruffle1999,Brodin2002,Wiebel2002,Goetze2004,Cang2005}.
824: Large data sets for the evolution of the glassy dynamics of propylene
825: carbonate have been analyzed in Ref.~\cite{Goetze2000b}; results measured
826: for different probing variables $A$ obtained by neutron-scattering,
827: depolarized-light-scattering, and dielectric-loss spectroscopy have been
828: fitted by a common first correlator $\phi(t)$. The different probes have
829: been characterized by adjusting the parameters for the second correlator
830: $\phi_A(t)$ only.
831:
832: Up to order $(t_0/t)^{2a}$, the kernel $m_A(t)$ can be calculated from
833: Eq.~(\ref{eq:mA}) substituting the expression~(\ref{eq:critcorr:phi}) for
834: $\phi(t)$ and the analog expression for the second correlator,
835:
836: \begin{equation}\label{eq:phiA}
837: \phi_A(t) = f_A^c + h_a (t_0/t)^a \left[ 1 + \hat{K}_A (t_0/t)^a \right]
838: \,.
839: \end{equation}
840: Laplace transform yields $\omega m_A(\omega)$ as an asymptotic power
841: series in the small parameter $\xi = (-i\omega t_0)^a$. The coefficients
842: are linear functions of $f^c_A$, $h_A$, and $\hat{K}_A$. This is
843: substituted on the left-hand side of Eq.~(\ref{eq:MCTlong:m}) for $q=2$.
844: Expanding the right-hand side up to orders $\xi^2$, one can compare the
845: coefficients in order to arrive at:
846:
847: \begin{equation}\label{eq:fAhA}
848: f_A^c = 1- \left[ 1/ (f^c v_A) \right] \,,\quad
849: h_A = \lambda / \left( f^{c\,2} v_A\right)\,.
850: \end{equation}
851:
852: The correction amplitude shall be noted in the convention of the general
853: theory \cite{Franosch1997}:
854:
855: \begin{subequations}\label{eq:cAKA}
856: \begin{equation}\label{eq:cAKA:cA}
857: \hat{K}_A = \kappa(a) + K_A\,,
858: \end{equation}
859:
860: \begin{equation}\label{eq:cAKA:KA}
861: K_A = \lambda v_a \left[ \frac{1}{v_A(1-\lambda)} -\lambda \right] /
862: \left[ v_A(1-\lambda) - 1 \right]\,.
863: \end{equation}
864: \end{subequations}
865: This leads to the parameters for the Cole-Cole susceptibility:
866:
867: \begin{equation}\label{eq:chi0ccA}
868: \chi^{cc}_{0\,A} = 1/\left(f^c v_A\right) \,,\quad
869: \omega^{c}_A t_0 = \left[\left(1-\lambda\right)/\left(\lambda\Gamma(1-a)
870: \right)\right]^{1/a} \,.
871: \end{equation}
872: Remarkably, the comparison of Eq.~(\ref{eq:chi0ccA}) with the general
873: expression in Eq.~(\ref{eq:cccpar:wcc}) shows, that the Cole-Cole
874: frequency $\omega_A^{c}$ does not depend on the coupling coefficient
875: $v_A$. $\omega_A^{c}t_0$ decreases with increasing $\lambda$ from
876: 0.37\dots for $\lambda=1/2$ to about $10^{-4}$ for $\lambda$ near 0.88.
877: For $\lambda$ around 0.7, $\omega_A^{c}t_0$ is about 0.02.
878:
879: %%
880: \subsection{Critical Relaxation for a Small Cole-Cole
881: Frequency\label{subsec:BZP}}
882:
883: %%Fig. 1
884: \begin{figure}
885: \includegraphics[width=\columnwidth]{BZP}
886: \caption{\label{fig:BZP}OKE response functions measured for BZP for $T/K =
887: 251,\,260,\,290,\,320$ \cite{Cang2003c} (full lines from bottom to top)
888: and fits by the schematic-model functions $\chi_A(t)$ \cite{Goetze2004}
889: (dotted lines). The straight dashed line has slope $-0.80$.
890: The dash-dotted lines labeled \textit{sc} are the
891: approximations by the scaling functions Eq.~(\ref{eq:chiA_beta}), the thin
892: dashed lines are fits with Eq.~(\ref{eq:chiA_beta}) for freely adjusted
893: $s_\sigma$ and $t_\sigma$. The insets show the separation parameter
894: $\sigma$ (right) and the fitting parameters (left); two additional state
895: points ($+$ and $\times$) indicate extrapolations for which solutions are
896: shown in Fig.~\ref{fig:BZPwing}.
897: }
898: \end{figure}
899:
900: Figure~\ref{fig:BZP} reproduces OKE-response functions measured for
901: benzophenone (BZP) \cite{Cang2003c} and fits to these data by the response
902: $\chi_A(t)$ calculated for the model defined in the preceding section. The
903: fit parameters for the mode-coupling coefficients $v_1$, $v_2$, and $v_A$
904: are specified in the lower inset. The analysis shall be done by
905: anticipating a transition point for $\lambda=0.70$, which implies the
906: exponents $a=0.33$ and $b=0.64$. The cross in the inset is close to this
907: bifurcation point. The $\sigma$-versus-$T$ diagram is shown in the upper
908: inset; and extrapolation to $\sigma=0$ suggests the critical temperature
909: $T_c=235\text{K}$ with an estimated uncertainty of $\pm 5\text{K}$. The
910: $251\text{K}$ result exhibits the expected von~Schweidler decay,
911: $\log\chi(t) = \text{const.}-(1-b)\log t$ for times between about 0.3ns
912: and about 6ns. Further details can be inferred from
913: Ref.~\cite{Goetze2004}.
914:
915:
916: %%Fig. 2
917: \begin{figure}
918: \includegraphics[width=\columnwidth]{BZPAt}
919: \caption{\label{fig:BZPphiAt} Correlation functions for BZP.
920: The lower panel exhibits the correlators
921: $\phi_A(t)$ underlying the fits in Fig.~\ref{fig:BZP} (heavy full lines
922: from left to right for $T/K = 320,\,290,\,260,\,$ and $251$,
923: respectively). The light full line with label \textit{c} shows the
924: critical correlator calculated for $\lambda = 0.7$ and $v_A = 30$. The
925: dotted line is the leading asymptotic law, Eq.~(\ref{eq:crit:corr}), with
926: the time scale $t_0 = 0.3775$ps. The dashed line marked \textit{vS} shows
927: a von~Schweidler law, Eq.~(\ref{eq:vS:phi}), with $b=0.64$ and a time
928: scale adjusted to match the $T=251$K curve. The dashed line labeled
929: \textit{K} is a fit by the Kohlrausch law, $\phi_A(t) = f_A^c
930: \exp[-(t/\tau)^\beta],\,\beta=0.91$, with $\tau$ adjusted to match the
931: $T=260$K curve. The times $t_0$ and $1/\omega^c_A = 15$ps are indicated by
932: arrows. The upper panel shows the critical decay (\textit{c}) together
933: with the leading-order ($t^{-a}$) and next-to-leading-order ($t^{-2a}$)
934: asymptotic power-law solution, Eq.~(\ref{eq:critcorr:phi}). The dotted
935: curve labeled \textit{cc} displays the leading-order Cole-Cole solution,
936: Eq.~(\ref{eq:ML}), with $\chi_0^{cc}=1/9$. The points of 10\% deviation of
937: the approximations $t^{-a}$, $t^{-2a}$, and \textit{cc} from the critical
938: decay are indicated by the diamond, circle, and triangle, respectively.
939: }
940: \end{figure}
941:
942: %%Fig. 3
943: \begin{figure}
944: \includegraphics[width=\columnwidth]{BZPchi}
945: \caption{\label{fig:BZPchi} Spectra for BZP.
946: The lower panel shows the susceptibility spectra for the correlators from
947: the lower panel of Fig.~\ref{fig:BZPphiAt}. The leading-order critical
948: spectrum, Eq.~(\ref{eq:chi_power:leading}), is the dotted straight
949: line marked by
950: $\omega^a$. A dashed straight line with slope $-0.2$ is fitted to the
951: $T=251$K spectrum to indicate an $\omega^{-b'}$ law for $b' = -0.2$.
952: In the upper panel, the dashed curve marked $\omega^{2a}$ shows the
953: approximation by Eq.~(\ref{eq:chi_power:chi}). The approximation by the
954: Cole-Cole function (\textit{cc}), Eq.~(\ref{eq:ccleft:chi}), is shown
955: dotted, the inclusion of the correction, Eq.~(\ref{eq:ccc}) yields the
956: dashed curve labeled \textit{ccc} for $\omega_A^c = 67$ns$^{-1}$.
957: The points of 10\% deviation of the approximations $\omega^{-a}$,
958: $\omega^{-2a}$, and \textit{cc} from the critical spectrum are indicated
959: by the diamond, circle, and triangle, respectively. The frequencies
960: $\omega_A^c$ and $\nu = \omega/(2\pi) = 1$THz are marked by arrows.
961: }
962: \end{figure}
963:
964: The MCT results describe the measured evolution of the glassy dynamics
965: adequately for times exceeding $t_\text{mic} = 1\text{ps}$ with two
966: exceptions. The $T = 251\text{K}$ data exhibit a pronounced oscillation
967: near $1.1\text{ps}$, while the calculated curve shows this oscillation
968: near $0.9\text{ps}$. Furthermore, there are some deviations between the
969: theoretical results and the data for the $T = 260\text{K}$ results for
970: times around $10\text{ns}$. The four correlators $\phi_A(t)$ and the four
971: loss spectra $\chi''_A(t)$, which are shown in the lower panels of
972: Figs.~\ref{fig:BZPphiAt} and~\ref{fig:BZPchi}, respectively, are the MCT
973: results corresponding to the MCT responses $\chi_A(t)$ shown in
974: Fig.~\ref{fig:BZP}. The data fits have been calculated for fixed relative
975: values of the four frequencies, which specify the transient dynamics:
976: $\Omega_A = \Omega$, $\nu_A = \nu = 5\Omega$. The scale $\Omega$ for the
977: fits is chosen different for different temperatures $T$ \cite{Goetze2004}.
978: The curves in Fig.~\ref{fig:BZPphiAt} and~\ref{fig:BZPchi} are presented
979: with a $T$-independent scale corresponding to that one used for the fit
980: for $T = 251$K: $\Omega = 1.67$ps$^{-1}$. With the mentioned reservations,
981: these curves can be considered as the measured quantities for the
982: OKE-probing variable $A$. The full lines marked $c$ in
983: Figs.~\ref{fig:BZPphiAt} and~\ref{fig:BZPchi} show the critical correlator
984: $\phi_A^c(t)$ and the critical loss spectrum ${\chi''_A}^c(\omega)$,
985: respectively, for the transition point specified by $\lambda = 0.70$ and
986: $v_A = 30$. The correlator relaxes to the plateau $f_A^c = 8/9$ with a
987: critical amplitude $h_A = 7/27$.
988:
989: The second scaling law for the decay below the plateau of the correlators,
990: Eq.~(\ref{eq:alpha}), implies a corresponding scaling law for the
991: response:
992:
993: \begin{equation}\label{eq:alpha_chi}
994: \chi_A(t) = \tilde{\chi}_A(t/t'_\sigma)/t'_\sigma\,.
995: \end{equation}
996: The control-parameter independent shape function reads
997: $\tilde{\chi}_A(\tilde{t}) = -\partial\tilde{\phi}_A(\tilde{t}) /
998: \partial\tilde{t}$. The reader can check this superposition principle for
999: the results in Fig.~\ref{fig:BZP} as follows. The $\log
1000: \chi_A(t)$-versus-$\log t$ curve for $T = 320\text{K}$ and $t \geqslant
1001: 2\text{ps}$ can be translated so that it collapses with the curves for
1002: $T/\text{K} = 290,\,260,\,$ and $251$ for times exceeding $4\text{ps}$,
1003: $90\text{ps}$, and $290\text{ps}$, respectively. Equivalently, the four
1004: loss spectra in Fig.~\ref{fig:BZPchi} for $\chi''_A(\omega) \geqslant 0.1$
1005: are connected by the superposition principle: $\chi''_A(\omega) =
1006: \tilde{\chi}''_A(\omega t'_\sigma)$. The von~Schweidler law describes a
1007: part of the high-frequency wing of the loss peak as shown by the straight
1008: dashed line with label \textit{vS} for the $T = 251\text{K}$ curve. The
1009: Kohlrausch law for a stretching exponent $\beta = 0.91$ describes the
1010: upper part of the $\alpha$-peak as shown by the dashed line with label $K$
1011: for the $T = 260\text{K}$ result.
1012:
1013: There is an interval of times between the end of the transient dynamics
1014: and the start of the von~Schweidler decays that deals with the relaxation
1015: towards and through the plateau $f_A^c$. The dynamics causes the loss
1016: spectra for $\chi''_A(\omega) \leqslant 0.1$ and $\omega/(2\pi) < 0.1$THz
1017: shown in Fig.~\ref{fig:BZPchi}. The leading-order asymptotic solution of
1018: the MCT equations of motion deals with this part of the dynamics together
1019: with the von~Schweidler-law part by Eq.~(\ref{eq:phi_gpm}). This formula
1020: yields the first scaling law for the response
1021:
1022: \begin{equation}\label{eq:chiA_beta}
1023: \chi_A(t) = h_A s_\sigma \hat{\chi}(t/t_\sigma)\,;\quad
1024: s_\sigma = \sqrt{|\sigma|}/t_\sigma\,.
1025: \end{equation}
1026:
1027: The shape function reads $\hat{\chi}(\hat{t}) = -\partial g_\pm(\hat{t}) /
1028: \partial\hat{t}$, $T \gtrless T_c$. For $T > T_c$, Eq.~(\ref{eq:gpm:B})
1029: yields the von~Schweidler response for large rescaled times $\hat{t} = t /
1030: t_\sigma$: $\hat{\chi}(\hat{t} \gg 1) =B\cdot b / \hat{t}^x\,,\quad x = 1-
1031: b < 1$. For small rescaled times, Eq.~(\ref{eq:gpm:ta}) yields
1032: $\hat{\chi}(\hat{t} \ll 1) = a/ \hat{t}^x\,,\quad x = 1 + a > 1$. The
1033: crossover from one asymptote to the other occurs for times near
1034: $t_\sigma$. The light dashed lines in Fig.~\ref{fig:BZP} exhibit fits of
1035: the data by Eq.~(\ref{eq:chiA_beta}) for $\lambda = 0.70$. However, both
1036: scales $s_\sigma$ and $t_\sigma$ are adjusted with the aim to achieve a
1037: good match in the von~Schweidler-law regime. The $1/t^{1+a}$-law is not
1038: exhibited by the results in Fig.~\ref{fig:BZP}. For $T = 320\text{K}$, the
1039: crossover time $t_\sigma$ is located near $1\text{ps}$, i.e., it is within
1040: or close to the transient regime. For the other three temperatures and
1041: times within the interval $t_\text{mic} < t < t_\sigma$, the measured
1042: response and the calculated functions $\chi_A(t)$ are below the values of
1043: $h_A s_\sigma\hat{\chi}(t / t_\sigma)$.
1044:
1045: The straight dashed line in Fig.~\ref{fig:BZP} has a slope of $-0.80$. It
1046: demonstrates a pseudo-von~Schweidler decay of the $T = 251\text{K}$
1047: results for $2\text{ps} < t < 20\text{ps}$: $\phi_A(t) - \text{const.}
1048: \propto -t^{b'}\,,\quad b' = 0.20$. The $\phi_A(t)$-versus-$\log t$ curves
1049: in Fig.~\ref{fig:BZPphiAt} for $t > 2\text{ps}$ approach and cross the
1050: plateau as downward-bent curves before entering the von~Schweidler decay
1051: regime. There are no inflection points of the curves for some time near
1052: $t_\sigma$, as implied by Eq.~(\ref{eq:phi_gpm}).
1053:
1054: The first scaling law describes a loss minimum of some value
1055: $\chi_\text{min}$ at some position $\omega_\text{min}$. The shape of the
1056: $\log \chi_A''(\omega)$-versus-$\log\omega$ curves is independent of the
1057: separation parameter $\sigma$ and the scales fix the position of the
1058: curve: $\chi_\text{min} \propto \sqrt{|\sigma|}$, $\omega_\text{min}
1059: \propto 1/t_\sigma$. The minimum is due to the crossover from the
1060: von~Schweidler wing of the $\alpha$-peak, $\chi''(\omega) \propto 1 /
1061: \omega^b$, to the power-law asymptote for the critical dynamics,
1062: $\chi''(\omega) \propto \omega^a$. For most practical purposes, the
1063: minimum can be approximated by the interpolation formula
1064: \cite{Sjoegren1990b}:
1065:
1066: \begin{equation}\label{eq:interpol}
1067: \chi''_A(\omega) / \chi_\text{min} = \left[
1068: b \left(\omega/\omega_\text{min}\right)^a
1069: + a \left(\omega_\text{min}/\omega\right)^b
1070: \right]/(a+b)\,.
1071: \end{equation}
1072: However, the loss spectra in the lower panel of Fig,~\ref{fig:BZPchi} do
1073: not exhibit loss minima, which can be described by the interpolation
1074: formula for the leading-order asymptotic result. The minima for
1075: frequencies $\omega$ near $1\text{ps}^{-1}$ do not depend sensitively on
1076: $\sigma$. With decreasing $|\sigma|$, the minimum position even shifts
1077: slightly upwards rather than downwards. With increasing frequency, the
1078: high-frequency part of the $\alpha$-peak exhibits a crossover from the
1079: von~Schweidler decay, $\chi''(\omega) \propto 1/\omega^b$, to a pseudo
1080: von~Schweidler decay, $\chi''(\omega) \propto 1/\omega^{b'}$. For $T =
1081: 251$K, this high-frequency wing extends from $\omega \approx
1082: 0.02\text{ps}^{-1}$ to about 0.4ps$^{-1}$ with $b' = 0.2$, as is
1083: demonstrated by the dashed straight line. The observed minimum is due to
1084: the crossover from the $1/\omega^{b'}$ wing to the spectral peak for the
1085: normal-liquid dynamics. The latter is located near $\nu= 1$THz.
1086:
1087: The scaling-law fits, which are shown in Fig.~\ref{fig:BZP} by the light
1088: dashed lines, are misleading because the scales $s_\sigma$ and $t_\sigma$
1089: have been adjusted freely. Calculating these scales from
1090: Eqs.~(\ref{eq:sigma:tsigma}, \ref{eq:vc:sigma}, \ref{eq:fAhA}), one
1091: obtains for $T = 251\text{K}$ and 260K the dash-dotted lines marked by
1092: $sc$. The two distance parameters $\varepsilon = (T_c -T) / T_c$ are -0.07
1093: and -0.11, respectively. These values are too large for a leading-order
1094: asymptotic formula to be applicable.
1095:
1096: The correlator $\phi_A(t)$ for the lowest temperature under discussion is
1097: close to the critical one for $t$ up to about $30$ps, as is shown in the
1098: lower panel of Fig.~\ref{fig:BZPphiAt}. Hence, the $T = 251$K curves
1099: exhibit critical glassy dynamics for the times between 1ps and 30ps. From
1100: Eqs.~(\ref{eq:cAKA}) one obtains the correction amplitude $\hat{K}_A =
1101: -1.51$. This large value yields an onset time $t_A^*$ for the $t^{-a}$-law
1102: which is beyond $t = 10^3$ps as is marked by the diamond in the upper
1103: panel of the figure. Including the leading correction term yields the
1104: approximation by the dashed line denoted $t^{-2a}$. It improves the
1105: description of the critical decay so that it can be understood for times
1106: larger than about $t = 30$ps. But the expansion formula~(\ref{eq:phiA})
1107: cannot be used to describe that part of the critical decay, which is
1108: measured for BZP and described by the solution of the two-component
1109: schematic MCT model.
1110:
1111: Substituting the cited values for $\lambda$, $f_A^c$, and $h_A$ into
1112: Eqs.~(\ref{eq:fqmc}, \ref{eq:hmq}), one gets the plateau for the modulus,
1113: $f_A^{m\,c} = 8$, and the critical amplitude $h_A^m = 21$. Notice that the
1114: relative strength of the first scaling law amplitude for the modulus is
1115: much larger than that for the correlator: $h_A^m/f_A^{m\,c} = 2.6$ versus
1116: $h_A/f_A^c = 0.29$. From Eq.~(\ref{eq:cmq:cmq}), one obtains the
1117: correction amplitude $\hat{K}_A^m = 0.124$. As expected from the
1118: discussions of
1119: Sec.~\ref{sec:crit}, the onset time $t_A^{m\,*}$ for the leading-order
1120: formula for the modulus,
1121:
1122: \begin{subequations}\label{eq:CC}
1123: \begin{equation}\label{eq:CC:crit_modulus}
1124: m_A(t) = f_A^{m\,c} + h_A \left(t_0 /t \right)^a\,,
1125: \end{equation}
1126: is smaller than $t_A^*$ by more than a factor $1000$. As a result, see the
1127: upper panel of Fig.~\ref{fig:BZPphiAt}, the equivalent
1128: formula~(\ref{eq:ML}) for the correlator,
1129:
1130: \begin{equation}\label{eq:CC:ML}
1131: \phi_A(t) = f_A^c + (1-f_A^c)
1132: \text{M}_a\left[-\left(t\omega_A^c\right)^a\right]\,,
1133: \end{equation}\end{subequations}
1134: describes the probing variable $A$ for all times exceeding $t = 1$ps. The
1135: leading-order asymptotic result for the modulus explains the response for
1136: $T = 251$K quantitatively within the interval $1$ps$\leqslant t \leqslant
1137: 30$ps.
1138:
1139: Figure~\ref{fig:BZPchi} shows that the leading-asymptotic description of
1140: the critical loss spectrum, Eq.~(\ref{eq:chi_power:leading}), would be
1141: relevant for the explanation of the loss minimum only in cases with
1142: $\omega_\text{min} < 2\cdot10^{-5}$ps$^{-1}$. The dashed line labeled
1143: $\omega^{2a}$ exhibits the asymptotic expansion for the critical loss up
1144: to the leading correction, Eq.~(\ref{eq:chi_power:chi}). This formula is
1145: relevant for $\omega < 10^{-3}$ps$^{-1}$. Even this expression is
1146: unsatisfactory for the discussion of the BZP results because of the large
1147: correction amplitude. However, the Cole-Cole spectrum describes the
1148: critical loss reasonably for $\omega < 0.2$ps$^{-1}$. The correction
1149: amplitude for the Cole-Cole law, Eq.~(\ref{eq:cccpar:cqcc}) is very small,
1150: $\hat{K}_A^{cc} = 0.076$. The Cole-Cole law with leading correction,
1151: Eq.~(\ref{eq:ccc}), is shown in the upper panel of Fig.~\ref{fig:BZPchi}
1152: by the curve marked $ccc$, and it slightly above the leading-order result
1153: in the regime of large frequencies.
1154:
1155: For the system under study, the Cole-Cole frequency reads $\omega_A^c =
1156: 0.067$ps$^{-1}$. This frequency $\omega_A^c/(2\pi) \approx 10$GHz is small
1157: compared to the $1$THz scale for the normal-liquid dynamics. The value
1158: $-\log\omega_A^c$ is in the center of the $\log t$ interval studied by the
1159: results of Fig.~\ref{fig:BZP}. Therefore, the critical spectrum relevant
1160: for the understanding of the data differs qualitatively from the
1161: leading-order power-law formula. As a result, the loss minima in
1162: Fig.~\ref{fig:BZPchi}, which are caused by the superposition of the
1163: critical spectra for $\omega$ near and above $\omega_A^c$ and the
1164: von~Schweidler-law wing of the $\alpha$-peak, differ drastically from the
1165: general low-frequency shape described by Eq.~(\ref{eq:interpol}). For $T =
1166: 251$K, the superposition yields the pseudo-von~Schweidler law manifested
1167: as a wing specified by the exponent $b' = 0.20$. This wing will be
1168: analyzed further in Sec.~\ref{sec:alphabeta}.
1169:
1170:
1171: %%
1172: \subsection{Critical Relaxation for an Intermediate Cole-Cole
1173: Frequency\label{subsec:salol}}
1174:
1175: Five OKE response functions measured for salol \cite{Hinze2000b} and fits
1176: by the functions $\chi_A(t)$ of the above explained schematic model are
1177: reproduced in Fig.~\ref{fig:salol}. The filled dots in the insets specify
1178: the mode-coupling coefficients $v_1$, $v_2$, and $v_A$ used for the
1179: calculations. Further details can be found in Ref.~\cite{Goetze2004},
1180: where, however, the value for $\Omega$ was misprinted. The series of
1181: states $(v_1, v_2)$ extrapolates to a transition point with $\lambda =
1182: 0.73$, which implies a critical exponent $a=0.31$ and a von~Schweidler
1183: exponent $b=0.59$. The extrapolation of the $\sigma$-versus-$T$ parameters
1184: to $\sigma = 0$ suggests a critical temperature $T_c = 245$K with an
1185: estimated uncertainty of $\pm 3$K. The MCT correlators $\phi_A(t)$ and
1186: loss spectra $\chi''_A(\omega)$, which are equivalent to the response
1187: $\chi_A(t)$ in Fig.~\ref{fig:salol}, are shown in the lower panels of
1188: Figs.~\ref{fig:salolphiAt} and~\ref{fig:salolchiAw}, respectively. They
1189: can be considered as the measured results for the glassy dynamics of
1190: salol, since the calculated and measured curves in Fig.~\ref{fig:salol}
1191: agree outside the transient regime. This holds with the reservation, that
1192: fits and measurements for the $T = 270$K results exhibit some
1193: discrepancies for times exceeding 20ns. The lines with label $c$ in
1194: Figs.~\ref{fig:salolphiAt} and~\ref{fig:salolchiAw} show critical
1195: correlators and critical loss spectra, respectively, calculated for
1196: $\lambda = 0.73$ and $v_A = 55$. The correlators relax to the plateau
1197: $f_A^c = 0.93$ with a critical amplitude $h_A = 0.18$.
1198:
1199: %%Fig 4
1200: \begin{figure}
1201: \includegraphics[width=\columnwidth]{salol}
1202: \caption{\label{fig:salol}
1203: OKE-response functions measured for salol for $T/K = 247,\, 257,\, 270,\,
1204: 300,\, 340$ (full lines from bottom to top) \cite{Hinze2000b}. The dotted
1205: lines are fits by the schematic-model response $\chi_A(t)$ calculated with
1206: $\Omega = 2\Omega_A = 10\nu_A = 15.9$ps$^{-1}$ \cite{Goetze2004}. Symbols
1207: and curve styles are the same as in Fig.~\ref{fig:BZP}. The slope of the
1208: straight dashed line is $-1.15$.
1209: }
1210: \end{figure}
1211:
1212:
1213: %%Fig 5
1214: \begin{figure}
1215: \includegraphics[width=\columnwidth]{phiAt}
1216: \caption{\label{fig:salolphiAt} Correlation functions for salol.
1217: The lower panel shows the correlators
1218: $\phi_A(t)$ used for the response fits in Fig.~\ref{fig:salol} for $T/K =
1219: 247,\,257,\,270,\,300,\,$ and $340$ (heavy full lines from right to left).
1220: The light full line shows the critical correlator calculated for $\lambda
1221: = 0.73$, and $v_A = 55$. The dashed line marked \textit{vS} exhibits a
1222: von~Schweidler law, Eq.~(\ref{eq:vS:phi}), for a time scale chosen to
1223: match the $T=247$K correlator. The dashed line marked \textit{K} is a
1224: Kohlrausch-law fit for the $T=257$K curve with exponent $\beta = 0.95$.
1225: The dotted line exhibits the formula~(\ref{eq:crit:corr}) with $t_0 =
1226: 0.00246$ps. The upper panel shows the critical decay (\textit{c}) together
1227: with the leading- ($t^{-a}$) and next-to-leading-order asymptotic
1228: solution~(\ref{eq:critcorr:phi}) ($t^{-2a}$). The dotted curve labeled
1229: \textit{cc} displays the leading-order solution from Eq.~(\ref{eq:ML})
1230: with $\chi_0^{cc}=0.067$ and $\omega_A^c = 7.25$ps. The points of 10\%
1231: deviation of the approximations $t^{-a}$, $t^{-2a}$, and \textit{cc} from
1232: the critical decay are indicated by the diamond, circle, and triangle,
1233: respectively. The times $t_0$ and $1/\omega_A^c$ are marked by arrows.
1234: }
1235: \end{figure}
1236:
1237: The test of the superposition laws for the long-time relaxation parts of
1238: the correlators is left to the reader. This second scaling law of MCT
1239: relates the $\alpha$-relaxation peaks for the loss spectra for
1240: $\chi''_A(\omega) \geqslant 0.1$. Kohlrausch-law fits and von~Schweidler
1241: asymptotes have been added to the data for $T = 257$K and $T = 247$K,
1242: respectively, in order to emphasize that the results for the evolution of
1243: the below-plateau-decay process follows the familiar pattern. Notice that
1244: the response for $T = 257$K exhibits the von~Schweidler-law decay,
1245: $\log\chi_A(t) = \text{const.} -(1-b) \log t$ for the large time interval
1246: 0.1ns $\leqslant t \leqslant $ 10ns. The measurement for this temperature
1247: permits a rather precise determination of the exponent $b$ and, thereby,
1248: of $\lambda$.
1249:
1250: %%Fig 6
1251: \begin{figure}
1252: \includegraphics[width=\columnwidth]{chiA}
1253: \caption{\label{fig:salolchiAw} Spectra for salol.
1254: The lower panel shows the susceptibility spectra $\chi_A''(\omega)$ for
1255: the correlators shown in the lower panel of Fig.~\ref{fig:salolphiAt}. The
1256: leading-order approximation to the critical spectrum,
1257: Eq.~(\ref{eq:chi_power:chi}), is shown as dotted straight line marked
1258: $\omega^a$. The dashed line labeled \textit{sc'} exhibits the
1259: interpolation formula (\ref{eq:interpol}) with the von~Schweidler exponent
1260: $b=0.59$ and the critical exponent $a$ replaced by $a'' = 0.24$.
1261: The upper panel exhibits the critical spectrum as full line with label
1262: \textit{c}, the leading-order asymptotic approximation,
1263: Eq.~(\ref{eq:chi_power:leading}), (dotted, labeled $\omega^a$) and
1264: Eq.~(\ref{eq:chi_power:chi}), (dashed, labeled $\omega^{2a}$). The
1265: approximation by the Cole-Cole function for $\omega_A^c = 7.25$ps$^{-1}$
1266: (\textit{cc}), Eq.~(\ref{eq:ccleft:chi}), is shown dotted, the inclusion
1267: of the correction, Eq.~(\ref{eq:ccc}), yields the dashed curve labeled
1268: \textit{ccc}. The points of 10\% deviation of the approximations
1269: $\omega^{-a}$, $\omega^{-2a}$, and \textit{cc} from the critical spectrum
1270: are indicated by the diamond, circle, and triangle, respectively. The
1271: frequencies $\omega_A^c$ and $\omega/(2\pi) = \nu = 1$THz are marked by
1272: arrows.
1273: }
1274: \end{figure}
1275:
1276: The light-dashed lines in Fig.~\ref{fig:salol} are fits of the data by the
1277: first-scaling-law expression~(\ref{eq:chiA_beta}) with freely adjusted
1278: scales $s_\sigma$ and $t_\sigma$. For $T \geqslant 270$K, these fits
1279: describe the data outside the transient regime up to times within the
1280: von~Schweidler-law region. A more detailed documentation of this result
1281: can be found in Ref.~\cite{Hinze2000b}, where also consistency of the
1282: fitted scales $s_\sigma$, $t_\sigma$, and $t'_\sigma$ with the
1283: MCT-power-law formulas was demonstrated. The scaling-law results
1284: calculated with the MCT values for the scales $h_A$, $s_\sigma$, and
1285: $t_\sigma$ in Eq.~(\ref{eq:chiA_beta}) are shown in Fig.~\ref{fig:salol}
1286: as dash-dotted lines marked $sc$ for the $T = 257$K and $T = 247$K data.
1287: For $T = 257$K, $\varepsilon = -0.05$ is so large, that
1288: Eq.~(\ref{eq:chiA_beta}) cannot account quantitatively for the scales
1289: ruling the von~Schweidler-law decay. For $T \geqslant 270$K the distance
1290: parameter $\varepsilon$ exceeds 10\%; the leading asymptotic expansion
1291: formula for the plateau-crossing process is not applicable for such large
1292: distances of the control parameter from the critical value. In contrast,
1293: the scaling law~(\ref{eq:chiA_beta}) describes the $T = 247$K result well
1294: for the large time interval 0.04ns $<t<$ 10ns. In this case, the distance
1295: parameter $\varepsilon = -0.008$ is so small that the leading-order
1296: expansion result for the plateau crossing, Eq.~(\ref{eq:phi_gpm}),
1297: accounts for a major part of the long-time response -- readjusting the
1298: scales $(h_As_\sigma)$, and $t_\sigma$ would not improve the fit.
1299:
1300: Figure~\ref{fig:salolphiAt} shows that the $T = 247$K response is close to
1301: the critical one for $t \leqslant 1$ns. The straight dashed line with
1302: slope $-1.15$, which is shown in Fig.~\ref{fig:salol}, demonstrates that
1303: the critical response exhibits a power-law decay $\chi_A(t) \propto
1304: 1/t^{1+a'}$, $a' = 0.15$, for 2ps $<t<$ 100ps. This is equivalent to a
1305: power-law decay of the correlator: $\phi_A(t) - f_a^c \propto 1/t^{a'}$.
1306: In qualitative agreement with the prediction by Eq.~(\ref{eq:phi_gpm}),
1307: the crossover from the $t^{-a}$ decay to the $-t^b$ decay causes an
1308: inflection point for the $\phi_A(t)$-versus-$\log t$ curve in
1309: Fig.~\ref{fig:salolphiAt}. The corresponding crossover from the
1310: $\omega^{-b}$ wing of the $\alpha$-peak to some $\omega^{a''}$ spectrum
1311: causes the loss minimum located near $\log\omega = -2.5$ in
1312: Fig.~\ref{fig:salolchiAw}. The dashed curve marked $sc'$ exhibits a
1313: description of this minimum by the function
1314: $\chi''_A(\omega)/\chi_\text{min} = \left[ b \left(\omega /
1315: \omega_\text{min}\right)^{a''} + a'' \left(\omega_\text{min} /
1316: \omega\right)^b \right]/(a''+b)$ which is suggested by
1317: Eq.~(\ref{eq:interpol}). For salol, as opposed to BZP, the minimum
1318: position decreases strongly with decreasing $|\varepsilon|$, in
1319: qualitative agreement with the first-scaling-law results.
1320:
1321: The inconsistent exponents for $T = 247$K in the regime 2ps $<t<$100ps
1322: indicate that the leading-order asymptotic description is still not
1323: applicable; the exponents $a' = 0.15$ obtained from the fit of the OKE
1324: data and $a'' = 0.24$ obtained from the fit of the corresponding minimum
1325: are different and both are smaller than the correct value of $a = 0.31$.
1326: The deviation of the critical correlator from
1327: the leading-order power-law result~(\ref{eq:crit:corr})
1328: and~(\ref{eq:chit0}) within the specified time interval is caused by the
1329: large value for the correction amplitude: $\hat{K}_A = -1.78$. This value
1330: implies an onset time for the $t^{-a}$ decay of beyond 20ps as shown by
1331: the diamond in Fig.~\ref{fig:salolphiAt}. The high-frequency part of the
1332: loss minimum is located in the region 10ns$^{-1}<\omega<$ 1ps$^{-1}$.
1333: Figure~\ref{fig:salolchiAw} demonstrates that the $\omega^a$ law does not
1334: describe the critical loss there. Inclusion of the leading correction
1335: terms for the analytic description of the critical dynamics, i.e., using
1336: Eq.~(\ref{eq:critcorr:phi}) for the correlator and
1337: Eq.~(\ref{eq:chi_power:chi}) for the loss spectrum, explains the result
1338: for $t \geqslant 0.6$ps and $\omega \leqslant 25$ns$^{-1}$, as
1339: demonstrated in Figs.~\ref{fig:salolphiAt} and~\ref{fig:salolchiAw},
1340: respectively.
1341:
1342: From Eq.~(\ref{eq:cmq:cmq}), one derives the correction amplitude for the
1343: modulus $\hat{K}_A^m = 0.177$. The leading-order formula
1344: (\ref{eq:CC:crit_modulus}) for the relaxation kernel describes the kernel
1345: $m_A$ outside the transient regime. The corresponding expression
1346: (\ref{eq:CC:ML}) for the correlator and the equivalent Cole-Cole formula
1347: for the loss spectrum account for the critical glassy dynamics, as shown
1348: in Figs.~\ref{fig:salolphiAt} and~\ref{fig:salolchiAw}. The Cole-Cole
1349: formula with correction term of amplitude $\hat{K}_A^c = 0.09$ yields a
1350: slight improvement compared to the equation based on $\hat{K}_A^c = 0$, as
1351: shown in Fig.~\ref{fig:salolchiAw} by the curve with label $ccc$.
1352: Equation~(\ref{eq:cccpar:wcc}) leads to the Cole-Cole frequency for salol:
1353: $\omega_A^c = 7.25$ps$^{-1}$. This value is close to the loss peak for the
1354: normal-liquid dynamics: $\omega_A^c t_\text{mic} \approx 1$. The part of
1355: the glassy critical loss spectrum which is relevant for the explanation of
1356: the data in Fig.~\ref{fig:salol} deals with the regime $10^{-4} \leqslant
1357: \omega/\omega_A^c \leqslant 10^{-1}$. Within this frequency interval, the
1358: Cole-Cole spectrum increases smoothly with $\omega$. Therefore, the
1359: leading-order formulas (\ref{eq:chit0}) or (\ref{eq:chi_power:leading})
1360: describe the dynamics qualitatively. However, the true critical spectrum
1361: differs from its low-frequency asymptote,
1362: Eq.~(\ref{eq:chi_power:leading}). Hence, the description of the dynamics
1363: by the scaling law is not correct quantitatively. Within the specified
1364: interval, the $\log\chi''_A(\omega)$-versus-$\log\omega$ spectrum can be
1365: approximated reasonably by a straight line, $\log\chi''_A(\omega) \approx
1366: \text{const.} + a'' \log\omega$, $a'' = 0.24$. As a result, the critical
1367: dynamics is approximated well by a power law decay specified by an
1368: exponent $a''$ smaller than the critical exponent $a$.
1369:
1370: In the preceding section, the critical spectra have been discussed for
1371: both small and intermediate Cole-Cole frequencies $\omega_q^c$; the case
1372: of a large Cole-Cole frequency is found in the
1373: depolarized-light-scattering spectra for CKN \cite{Li1992,Sperl2005bpre}.
1374: In the fits with a schematic model by Alba-Simionesco and coworkers
1375: \cite{Krakoviack1997,Krakoviack2002} one can identify a Cole-Cole peak for
1376: the critical spectra, but the description with the power-law solution is
1377: superior if terms up to order $\omega^{2a}$ are considered.
1378:
1379:
1380: %%%
1381: \section{Cole-Cole wing\label{sec:alphabeta}}
1382:
1383: Figures~\ref{fig:BZP}--\ref{fig:BZPchi} and
1384: \ref{fig:salol}--\ref{fig:salolchiAw} demonstrate the scenarios for the
1385: evolution of the glassy dynamics for $\omega^c_q \ll t^{-1}_\text{mic}$
1386: and $\omega^c_q \approx t^{-1}_\text{mic}$, respectively, for the probing
1387: variable described by the second correlator of a two-component schematic
1388: MCT model. Solutions for this model, which exemplify a scenario as shown
1389: in Fig.~\ref{fig:BZPchi}, have been discussed before by Cummins
1390: \cite{Cummins2005}. He pointed out that the evolution of the wing
1391: phenomenon obtained from the model is similar to the one known for
1392: dielectric-loss spectroscopy for glycerol and several van-der-Waals
1393: liquids \cite{Dixon1990}. He emphasizes also that the cited measurements
1394: refer to temperatures $T$ below $T_c$, while the calculations are done for
1395: $T > T_c$. Furthermore, the experimental data demonstrate the wing only
1396: for frequencies below 10GHz$\approx 10^{-2}t^{-1}_\text{mic}$. So far, no
1397: wing has been reported which occurs in the two-decade frequency window
1398: adjacent to the microscopic excitation region.
1399:
1400: In order to describe the wing phenomenon quantitatively, one can extend
1401: the Cole-Cole formula so that the spectrum can be described also for small
1402: but non-vanishing separation parameters $\sigma = C \varepsilon \propto
1403: (T_c-T)/T_c$. The leading-order expression for the plateau-crossing
1404: process of the fluctuating-force correlators reads in analogy to
1405: Eq.~(\ref{eq:phi_gpm}): $m_q(t) = f_q^{m\,c} + h_q^m
1406: g_\pm(t/t_\sigma),\,\varepsilon\gtrless 0$. Laplace transformation yields
1407: $\omega m_q(\omega) = -f_q^{m\,c} + h_q^m C(\omega)$. Function $C(\omega)$
1408: obeys a scaling law,
1409:
1410: \begin{equation}\label{eq:bet_C}
1411: C(\omega) = \sqrt{|\sigma|} c_\pm(\omega t_\sigma)\,,\quad
1412: \varepsilon \gtrless 0\,,
1413: \end{equation}
1414: with the control-parameter independent shape functions $c_\pm(\hat{\omega})$
1415: given by the Laplace transforms $g_\pm(\hat{\omega})$ of the shape
1416: functions $g_\pm(\hat{t})$:
1417: $c_\pm (\hat{\omega}) = \hat{\omega} g_\pm (\hat{\omega})$.
1418: Substitution of $\omega m_q(\omega)$ into Eq.~(\ref{eq:MCTlong:chi})
1419: yields the desired result. Expansion of the right-hand side in terms of
1420: the small parameter $C(\omega)$ and comparison with the result for
1421: $\chi_q(\omega)$ following from Eq.~(\ref{eq:phi_gpm}) reproduces the
1422: relations (\ref{eq:fqmc},\ref{eq:hmq}) for the critical arrested parts and
1423: critical amplitudes. One gets
1424:
1425: \begin{subequations}\label{eq:ccbeta}
1426: \begin{equation}\label{eq:ccbeta_lead}
1427: \chi_q(\omega) = \left( 1- f_q^c \right) /
1428: \left\{ 1 - \left[h_q/(1-f_q^c)\right] C(\omega)\right\}\,.
1429: \end{equation}
1430: Equation~(\ref{eq:ccbeta_lead}) was found for the schematic model in
1431: Ref.~\cite{Goetze1989c} for the special case of the second correlator
1432: $\phi_A(t)$ and in the limit of infinite coupling strength $v_A$. It is
1433: shown here, that Eq.~(\ref{eq:ccbeta_lead}) can be derived without such
1434: restrictions. The leading correction to Eq.~(\ref{eq:ccbeta_lead}) can be
1435: calculated from the asymptotic expansions in \cite{Franosch1997,Fuchs1998}
1436: in a straightforward manner; however, the evaluation of the result for
1437: fitting data is rather involved. Focusing on the critical decay, one can
1438: extend the preceding formula by adding the large-frequency part of the
1439: correction:
1440:
1441: \begin{equation}\label{eq:ccbeta_corr}\begin{split}
1442: \chi_q(\omega) = \left( 1- f_q^c \right) /
1443: \left\{ 1 - \left[h_q/(1-f_q^c)\right] C(\omega)\right.\\\left.
1444: + \hat{K}_q^\text{cc}\left(-i\omega/\omega_q^c\right)^{2a}\right\}\,.
1445: \end{split}\end{equation}
1446: \end{subequations}
1447: This expression describes the small-$\varepsilon$ dynamics for $\omega
1448: t_\text{mic}<1$ and frequencies extending down to the beginning of the
1449: von~Schweidler-law decay.
1450:
1451: %% Fig 7
1452: \begin{figure}[ht]
1453: \includegraphics[width=\columnwidth]{BZPwing}
1454: \caption{\label{fig:BZPwing}Asymptotic description of the wing in BZP.
1455: The heavy full line reproduces the BZP loss spectrum for $T=251$K from
1456: Fig.~\ref{fig:BZPchi} and the dashed one exhibits the spectrum of the
1457: extended Cole-Cole susceptibility, Eq.~(\ref{eq:ccbeta_lead}). The
1458: triangles mark the points of 10\% deviation between these spectra.
1459: The dotted line \textit{sc} is the first-scaling-law result for the
1460: susceptibility, and the line \textit{cc} shows the Cole-Cole spectrum for
1461: the critical point. The dash-dotted line represents the
1462: second-scaling-law result for the loss peak. The thin full lines are
1463: solutions corresponding to the state points indicated by $+$ and $\times$
1464: in Fig.~\ref{fig:BZP}.
1465: }
1466: \end{figure}
1467:
1468: BZP exemplifies the case of such a small $\hat{K}^{cc}_q$ that the
1469: leading-order result for the modulus, Eq.~(\ref{eq:ccbeta_lead}), can be
1470: used. Figure~\ref{fig:BZPwing} demonstrates that this result accounts for
1471: the $T=251$K spectrum for the large dynamical range $10^{-3}\text{ps}^{-1}
1472: < \omega < 1\text{ps}^{-1}$. The scaling-law result for the loss spectrum,
1473: $\chi_A''(\omega) = h_A C_A''(\omega)$, is shown as dotted line
1474: \textit{sc}; it can describe only a part of the von~Schweidler-law
1475: spectrum for $\omega < 10^{-3}\text{ps}^{-1}$, but is inadequate for
1476: larger frequencies. This observation for the dynamics in the frequency
1477: domain is equivalent to the one demonstrated in Fig.~\ref{fig:BZP} for the
1478: results in the time domain. The second-scaling-law result for the loss
1479: spectrum is exhibited as dash-dotted line. It follows from
1480: Eq.~(\ref{eq:alpha}), $\chi_A''(\omega) = (\omega t_\sigma')
1481: \tilde{\phi}_A''(\omega t_\sigma')$, and accounts for the loss peak for
1482: $\omega < 10^{-3}\text{ps}^{-1}$. Combining the result for the second
1483: scaling law for the modulus with the result for the first scaling law for
1484: the loss spectrum explains the BZP spectrum for $T = 251$K for $\omega <
1485: 1\text{ps}^{-1}$ including in particular the $\alpha$-peak wing. The small
1486: but systematic discrepancies between the loss spectrum in the
1487: structural-relaxation regime and the asymptotic MCT expressions result
1488: from the fact that the distance parameter $\varepsilon = (T-T_c)/T_c
1489: \approx 0.07$ is so large, that the asymptotic results for the scales
1490: still have noticeable errors while the shape functions already agree well
1491: with the nontrivial spectra.
1492:
1493: Starting from the state that fits the 251K data for BZP, a number of
1494: extrapolations are possible within the schematic model. Keeping all
1495: parameters but $(v_1, v_2)$ fixed, two additional states shall be
1496: considered; they are indicated by $\times$ and $+$ in the left inset of
1497: Fig.~\ref{fig:BZP}; and their spectra are marked accordingly in
1498: Fig.~\ref{fig:BZPwing}. The first state point ($\times$) is close to the
1499: transition point and is characterized by $\sigma = -0.003$ ($\varepsilon
1500: =-0.01$). Solutions at this point are shown in Ref.~\cite{Goetze2004} for
1501: $\chi(t)$ and $\chi_A(t)$. The respective spectrum $\chi_A''(\omega)$ in
1502: Fig.~\ref{fig:BZPwing} exhibits nearly-constant loss for $0.4\,
1503: 10^{-3}$ps$^{-1} < \omega < 0.4$ps$^{-1}$ within a 10\% margin. In this
1504: frequency window the maximum around $\omega^c_A$ indicates an emerging
1505: Cole-Cole peak; the minimum at $\omega = 10^{-3}$ps$^{-1}$ is caused by
1506: the crossover of the high-frequency wing of the $\alpha$-peak and the
1507: low-frequency wing of the Cole-Cole peak. This minimum is ruled by the
1508: first scaling law and the divergent time scale $t_\sigma$. The second
1509: state ($+$) is an interpolation between $\times$ and the solution for BZP;
1510: the separation parameter is $\sigma = -0.01$ ($\varepsilon = -0.03$), and
1511: the point is chosen similar to the ones used in Fig.~4 of
1512: Ref.~\cite{Cummins2005}. The spectrum in Fig.~\ref{fig:BZPwing} shows a
1513: wing, $\chi_A''(\omega) \propto \omega^{-0.1}$, for almost three orders of
1514: magnitude in frequency. Hence, the crossover between $\alpha$- and
1515: Cole-Cole peaks can be interpreted as a power-law wing $\omega^{-b'}$ for
1516: some range in frequencies and control parameters without fine-tuning. The
1517: evolution of the crossover progresses from a wing with gradually lower
1518: exponents $b'$ to nearly-constant loss and eventually the emergence of a
1519: Cole-Cole peak. While the extrapolations presented in this paragraph are
1520: not at all implied by the fit of the data, the scenarios seem nevertheless
1521: possible for realistic parameter values.
1522:
1523: %% Fig 8
1524: \begin{figure}[ht]
1525: \includegraphics[width=\columnwidth]{BZPOKE}
1526: \caption{\label{fig:BZPOKE}Asymptotic description of the OKE data for BZP.
1527: The full line reproduces the BZP data for $T=251$K from \cite{Cang2003c};
1528: the dotted line reproduces the fit from Fig.~\ref{fig:BZP}. The dashed
1529: curve, the dash-dotted curve \textit{sc}, and the dotted curve \textit{cc}
1530: respectively show the Fourier backtransforms of the extended Cole-Cole
1531: susceptibility in Eq.~(\ref{eq:ccbeta_lead}), of the first scaling law,
1532: and of the Cole-Cole law in Eq.~(\ref{eq:ccleft:chi}), cf.
1533: Fig.~\ref{fig:BZPwing}.
1534: }
1535: \end{figure}
1536:
1537: To conclude the discussion of BZP, let us return to the OKE data for
1538: $T=251$K as shown in Fig.~\ref{fig:BZPOKE}. In addition to data and fit
1539: for this state, the Fourier backtransforms of the asymptotic curves from
1540: Fig.~\ref{fig:BZPwing} are displayed. While the asymptotic solution at the
1541: critical point (labeled \textit{cc}) covers only less than a decade, the
1542: extended Cole-Cole susceptibility in Eq.~(\ref{eq:ccbeta_lead}) describes
1543: the data over three orders of magnitude in time and thereby explains the
1544: $t^{-0.8}$-law found empirically. The remaining discrepancies are again
1545: due to the relatively large distance parameter for this state; Fig.~3 of
1546: Ref.~\cite{Goetze2004} shows prefect agreement with the Cole-Cole law for
1547: states closer to the transition including state $\times$ from
1548: Fig.~\ref{fig:BZP}.
1549:
1550: %Fig 9
1551: \begin{figure}[ht]
1552: \includegraphics[width=\columnwidth]{Salolwing}
1553: \caption{\label{fig:Salolwing}Asymptotic description of the minimum in
1554: salol. The full line reproduces the salol loss spectrum for $T=247$K from
1555: Fig.~\ref{fig:salolchiAw} and the dashed one exhibits the spectrum of the
1556: extended Cole-Cole susceptibility, Eq.~(\ref{eq:ccbeta_lead}). The
1557: triangles mark the points of 10\% deviation between these spectra.
1558: The dotted line \textit{cc} is the Cole-Cole spectrum for the critical
1559: point, and the line \textit{sc} shows the first-scaling-law result for the
1560: susceptibility; the diamonds mark the points where \textit{sc} deviates by
1561: 10\% from the loss spectrum. The dash-dotted line represents the
1562: second-scaling-law result for the loss peak.
1563: }
1564: \end{figure}
1565:
1566: Different from the case for BZP, the lowest temperature available for
1567: salol, $T = 247$K, is described by a much smaller distance parameter,
1568: $\varepsilon \approx 0.01$, and therefore shares a larger frequency
1569: regime, say $\omega\lesssim 0.1$ps$^{-1}$, with the critical loss
1570: spectrum, see the lower panel of Fig.~\ref{fig:salolchiAw}. Being so close
1571: to the critical point, the second scaling law in Fig.~\ref{fig:Salolwing}
1572: -- labeled $\alpha$ -- is indistinguishable from the loss spectrum for
1573: $\omega \leqslant 10^{-4}$ps$^{-1}$. The first scaling law \textit{sc}
1574: describes the solution up to around $\omega\lesssim 0.005$ps$^{-1}$ which
1575: is the same as for the solution at the critical point, cf. upper panel of
1576: Fig.~\ref{fig:salolchiAw}. As for the critical spectrum, the leading-order
1577: approximation in Eq.~(\ref{eq:ccbeta_lead}) improves the description to
1578: larger frequencies by two decades up to $\omega\approx 0.1$ps$^{-1}$
1579: before crossing over to the Cole-Cole peak which in this case is hidden
1580: under the microscopic excitations.
1581:
1582:
1583: %%%
1584: \section{Conclusion\label{sec:sum}}
1585:
1586: Within the regime of glassy dynamics, susceptibilities $\chi_q(\omega)$
1587: and moduli $\omega m_q(\omega)$ are related by Eq.~(\ref{eq:MCTlong:chi});
1588: and at the critical point, susceptibilities and moduli are both described
1589: by the universal power law $\omega^a$. The range of validity of this power
1590: law is given by amplitudes for the correction $\omega^{2a}$ -- $\hat{K}_q$
1591: for the susceptibilities and $\hat{K}_q^m$ for the moduli, which are
1592: related by Eq.~(\ref{eq:cmq:cmq}). For some parameter regions a large
1593: value of $|\hat{K}_q|$ can render the power-law expansion for the
1594: susceptibilities irrelevant, while the correction amplitude
1595: $|\hat{K}_q^m|$ is so small that the universal power-law can be used for
1596: the description of the modulus successfully. Typically, this occurs if
1597: $\hat{K}_q$ is large and negative, and this can be expected if the plateau
1598: value $f^c_q$ is high, cf. Eq.~(\ref{eq:cmq:cmq}). In this case, the
1599: susceptibility can be described well by the expansion of its inverse,
1600: Eq.~(\ref{eq:ccc}). For vanishing correction $\hat{K}_q^m = K_q^{cc} = 0$,
1601: formula, the susceptibility at the critical point is given by the
1602: Cole-Cole law~(\ref{eq:ccleft:chi}), a formula first introduced in 1941
1603: \cite{Cole1941}.
1604:
1605: The Cole-Cole frequency $\omega_q^c$ in Eq.~(\ref{eq:ccleft:chi})
1606: introduces a characteristic scale for assorting the critical dynamics into
1607: three categories. (1) If $\omega_q^c$ is large compared to the scale
1608: $t^{-1}_\text{mic}$ for the band of normal liquid excitations, the
1609: Cole-Cole susceptibility reduces to the leading-order scaling-law formula
1610: which is implied by Eq.~(\ref{eq:phi_gpm}). There is a control-parameter
1611: sensitive minimum as discussed for Eq.~(\ref{eq:interpol}) that
1612: interpolates between the von~Schweidler-law tail of the $\alpha$-peak,
1613: $\chi_q''(\omega) \propto 1/\omega^b$, and the critical spectrum
1614: $\chi_q''(\omega) \propto \omega^a$, where $a$ and $b$ are related via the
1615: exponent parameter $\lambda$. As an example for such a scenario one can
1616: cite the molten salt CKN. While the leading-order result alone is not
1617: sufficient to describe the measured data quantitatively, the power-law
1618: expansion in Eq.~(\ref{eq:chi_power:chi}) yields a satisfactory
1619: description that is superior to the Cole-Cole solution
1620: \cite{Sperl2005bpre}. (2) If $\omega^c_q$ is close to $t^{-1}_\text{mic}$,
1621: one encounters a scenario demonstrated in Fig.~\ref{fig:Salolwing} for
1622: salol. There is a control-parameter sensitive loss minimum. It originates
1623: from the crossover between the von~Schweidler-law tail and the critical
1624: spectrum; and the critical spectrum approaches the maximum of the
1625: underlying Cole-Cole peak. A description by Eq.~(\ref{eq:interpol}) is
1626: possible only, if the exponent $a$ is replaced by some effective one
1627: $a''$, which is smaller than $a$, cf. Fig.~\ref{fig:salolchiAw}. (3) If
1628: $\omega_q^c$ is smaller than the microscopic time scale, say $\omega_q^c
1629: t_\text{mic} \approx 0.05$, we obtain the scenario seen in
1630: Fig.~\ref{fig:BZPwing} for benzophenone (BZP). In this case, the
1631: von~Schweidler-law part of the $\alpha$-peak, $\chi_q'' \propto
1632: 1/\omega^b$, crosses over to some flatter wing, $\chi_q'' \propto
1633: 1/\omega^{b'}$, $a < b' < b$. This wing is caused by approaching the
1634: Cole-Cole spectrum for frequencies around the maximum at $\omega^c_q$. For
1635: yet higher frequencies, one encounters a crossover -- from the
1636: high-frequency part of the Cole-Cole spectrum for structural relaxation to
1637: the spectrum due to normal-liquid excitations -- producing a different
1638: kind of a minimum. The position of this minimum is control-parameter
1639: insensitive. For the understanding of this minimum, both the scaling-law
1640: (\ref{eq:chiA_beta}) and the interpolation formula (\ref{eq:interpol}) are
1641: irrelevant. On the other hand, as shown in Sec.~\ref{sec:alphabeta} and
1642: especially in Fig.~\ref{fig:BZPwing}, it is possible to describe the wing
1643: rather accurately by the leading-order formula~(\ref{eq:ccbeta_lead}). If
1644: the $\alpha$-peak is shifted to yet lower frequencies, the wing can give
1645: way to a separate Cole-Cole peak.
1646:
1647: All figures in this work have been prepared with model parameters that
1648: reproduce the OKE-response functions of benzophenone (BZP) and salol.
1649: Hence, Fig.~\ref{fig:BZPwing} implies that the BZP spectra for
1650: temperatures near 250K, when measured by depolarized light-scattering in
1651: backward direction, should exhibit an $\alpha$-peak wing for frequencies
1652: $\nu$ between about 1GHz and about 100GHz. The crossover from the
1653: von~Schweidler-law wing to the wing induced by the Cole-Cole peak is
1654: expected to occur around 1GHz. This crossover position depends sensitively
1655: on the temperature, because the von~Schweidler-law relaxation depends
1656: sensitively on $T$.
1657:
1658: The Cole-Cole law with correction, Eq.~(\ref{eq:ccc}), has been derived in
1659: its fully general microscopic version in Sec.~\ref{sec:crit}; only in
1660: Secs.~\ref{sec:appl} and~\ref{sec:alphabeta} the theory was specialized to
1661: schematic models. While for the schematic models a number of parameters
1662: can be fixed to describe experimental data, for microscopic models the
1663: static structure of the model system determines these parameters uniquely.
1664: A first example for Cole-Cole dynamics in a microscopic model has been
1665: discussed recently for the mean-squared displacement $\delta r^2(t)$ of
1666: the hard-sphere system \cite{Sperl2005}; including the Mittag-Leffler
1667: function for the critical relaxation allows for an analytic description of
1668: the mean-squared displacement for the full range of the dynamics similar
1669: to BZP in Fig.~\ref{fig:BZPwing}. In addition, the Cole-Cole dynamics was
1670: identified in the data measured by van~Megen \textit{et. al}
1671: \cite{Megen1998} where Eq.~(\ref{eq:ccc}) accounts for the data for an
1672: interval in time of three orders of magnitude adjacent to the transient
1673: dynamics. Hence, the results discussed above seem relevant for both
1674: molecular and colloidal glasses and can be expected to facilitate more
1675: detailed investigations.
1676:
1677: I want to thank W.~G\"otze for repeated help during the preparation of
1678: this work; I am also grateful to L. Berthier, H.~Z.~Cummins, E.~R\"o\ss
1679: ler, and Th.~Voigtmann for enlightening discussions. Support was provided
1680: by DFG grant No. SP~714/3-1 and NSF grants DMR0137119 and DMS0244492.
1681:
1682: %%% bib
1683: \bibliographystyle{apsrev}
1684: \bibliography{lit,add}
1685:
1686: \end{document}
1687:
1688: