1: \documentclass[aps,pre,showpacs,amsmath,amssymb,twocolumn]{revtex4}\usepackage{pslatex}
2: %\documentclass[preprint,aps,pre,showpacs,amsmath,amssymb,endfloats*]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{dcolumn}
5: \usepackage{natbib}
6:
7: \newcommand{\ST}{\mathcal S}
8:
9: \begin{document}
10:
11: \title{Logarithmic Relaxation in a Colloidal System}
12:
13: \author{M. Sperl}
14: \affiliation{Physik-Department, Technische Universit\"at M\"unchen,
15: 85747 Garching, Germany}
16:
17: \date{\today}
18:
19: \begin{abstract}
20: The slow dynamics for a colloidal suspension of particles interacting with
21: a hard-core repulsion complemented by a short-ranged attraction is
22: discussed within the frame of mode-coupling theory for ideal glass
23: transitions for parameter points near a higher-order glass-transition
24: singularity. The solutions of the equations of motion for the density
25: correlation functions are solved for the square-well system in
26: quantitative detail by asymptotic expansion using the distance of the
27: three control parameters packing fraction, attraction strength and
28: attraction range from their critical values as small parameters. For given
29: wave vectors, distinguished surfaces in parameter space are identified
30: where the next-to-leading order contributions for the expansion vanish so
31: that the decay functions exhibit a logarithmic decay over large time
32: intervals. For both coherent and tagged particle dynamics the
33: leading-order logarithmic decay is accessible in the liquid regime for
34: wave vectors of several times the principal peak in the structure factor.
35: The logarithmic decay in the correlation function is manifested in the
36: mean-squared displacement as a subdiffusive power law with an exponent
37: varying sensitively with the control parameters. Shifting parameters
38: through the distinguished surfaces, the correlation functions and the
39: logarithm of the mean-squared displacement considered as functions of the
40: logarithm of the time exhibit a crossover from concave to convex behavior,
41: and a similar scenario is obtained when varying the wave vector.
42: \end{abstract}
43:
44: \pacs{61.20.Lc, 82.70.Dd, 64.70.Pf}
45:
46: \maketitle
47:
48: %%%
49: \section{\label{sec:introduction}Introduction}
50:
51: The dynamics in an interacting many particle system is conveniently
52: described by density autocorrelation functions $\phi_q(t)$ for time $t$
53: and wave vector $q$. These correlation functions can be measured in both
54: experiment and computer simulation \cite{Boon1980}. Mode-coupling theory
55: for ideal glass transitions (MCT) discusses the transition from a liquid
56: to a glass as a bifurcation in the long-time limit of the correlator
57: $\phi_q(t)$ \cite{Goetze1991b}. In the liquid state, the correlation
58: function decays to zero. If a control parameter, say density, exceeds some
59: critical value, the long-time limit changes discontinuously from zero to a
60: finite value, a glass transition occurs \cite{Bengtzelius1984}. This
61: liquid-glass transition is identified with an $A_2$- or fold singularity
62: \cite{Arnold1986} in the equations of motion of MCT. The simplest example
63: for a liquid-glass transition is found in the hard-sphere system (HSS),
64: where the interaction potential among the particles is zero unless their
65: mutual separation becomes smaller than their diameter where the potential
66: becomes infinitely repulsive, thus preventing the particles from
67: overlapping. The HSS is the system MCT was applied to first
68: \cite{Bengtzelius1984}, and it is also the system for which the most
69: detailed predictions have been worked out \cite{Franosch1997,Fuchs1998}.
70: Close to the singularity, the equations of motion can be expanded in
71: asymptotic series. This yields a two-step decay with two related power
72: laws for the short-time and the long-time decay at a liquid-glass
73: transition \cite{Goetze1991b}. The HSS can be realized in colloidal
74: suspensions \cite{Pusey1986}. Experiments in these systems lead to the
75: conclusion that MCT is able to describe the main aspects of the glass
76: transition qualitatively and some aspects even quantitatively
77: \cite{Megen1995}.
78:
79: MCT can also exhibit other singularities than the fold \cite{Goetze1991b}.
80: These higher-order singularities were predicted recently to occur for
81: colloidal systems with short-ranged attraction where $A_3$- and
82: $A_4$-singularities were found that are also called cusp and swallowtail
83: \cite{Fabbian1999,Bergenholtz1999,Dawson2001}. In these systems, the
84: hard-core repulsion is supplemented by a short ranged attraction, e.g. in
85: the square-well system (SWS). A cusp singularity is the endpoint of a line
86: of glass-glass transitions that arises if two different mechanisms of
87: arrest are of the same importance. In the SWS the first mechanism is the
88: hard core repulsion that causes a transition as in the HSS via the well
89: known cage effect. The second mechanism leading to arrest is bond
90: formation introduced by the attractive part of the potential. This latter
91: transition was proposed as relevant for the transition to a gel
92: \cite{Bergenholtz1999}. If the difference in the two mechanisms is less
93: pronounced, the glass-glass transitions vanish and give rise to an
94: $A_4$-singularity. In the SWS this happens as the range of the attraction
95: is increased \cite{Dawson2001}. The range of attraction considered here is
96: of order less than $20\%$ of the particle diameter and the strength is
97: about several $k_\text{B}T$. This is accessible in colloid-polymer
98: mixtures with nonadsorbing polymer which is well under control
99: experimentally \cite{Poon2002}. Higher-order singularities have also been
100: identified for a number of short-ranged potentials with shapes differring from
101: the SWS yielding certain quantitative trends but no qualitative changes
102: \cite{Goetze2003b}.
103:
104: In addition to the success of MCT for the description of the HSS, two
105: findings support the use of this theory for the description of colloids
106: with attraction. First, a reentry phenomenon was predicted by the theory
107: where a glass state is melted upon increasing the attraction
108: \cite{Fabbian1999,Dawson2001}. This was subsequently found in several
109: experiments \cite{Eckert2002,Pham2002} and computer simulation studies
110: \cite{Pham2002,Foffi2002b,Zaccarelli2002b,Puertas2003}. Second, there are
111: indications of logarithmic decay \cite{Puertas2002} and related anomalous
112: decays \cite{Mallamace2000,Zaccarelli2002b} that are consistent with
113: scenarios found within MCT \cite{Fabbian1999,Dawson2001}. To investigate
114: the dynamics in such systems, apart from computer simulation dynamic light
115: scattering has already been used to determine correlation functions
116: \cite{Mallamace2000,Eckert2002,Pham2002}. Direct imaging techniques are
117: available to determine also the mean-squared displacement (MSD) with high
118: precision \cite{Kegel2000,Weeks2000,Weeks2002}. The purpose of the present
119: paper is the application of the general theory for higher-order
120: singularities \cite{Goetze2002} to the SWS and the derivation of testable
121: quantitative predictions for the correlation functions and the MSD.
122: Certain scenarios have been discussed before for schematic models
123: \cite{Goetze2002}. Some of these scenarios shall be identified also in the
124: microscopic model in the following.
125:
126: The paper is organized as follows. In Sec.~\ref{sec:asy} the equations of
127: motion and the asymptotic solution for the logarithmic decay are
128: summarized and the subdiffusive power law for the MSD is derived. The
129: theory is applied to the $A_4$-singularity in the SWS for the correlation
130: functions in Sec.~\ref{sec:corr} and for the MSD in Sec.~\ref{sec:MSD}.
131: Changes in the scenarios when moving from the $A_4$-singularity to
132: $A_3$-singularities are discussed in Sec.~\ref{sec:A3}, and
133: Sec.~\ref{sec:HCY} contains a comparison to results obtained for the
134: hard-core Yukawa system (HCY). Section\ref{sec:conclusion} presents a
135: conclusion.
136:
137: %%%
138: \section{\label{sec:asy}Asymptotic Solutions}
139:
140: We shall consider a system of $N$ particles with diameter $d$ in a volume
141: $V$ interacting with a spherical potential. When at time $t$ the $j$th
142: particle is located at $\vec{r}_j(t)$ the density variables are defined as
143: $\rho_q(t)=\sum_j\exp[i\vec{q}\vec{r}_j(t)]$.
144:
145: %%
146: \subsection{\label{subsec:eom}Equations of Motion}
147:
148: The equations of motion for the normalized density correlators $\phi_q(t)=
149: \langle\rho_{{\vec{q}}}^{*}(t)\rho_{\vec{q}}\rangle/\langle|\rho_{\vec{q}}|^2
150: \rangle$ within MCT, when Brownian dynamics for the motion in colloids is
151: assumed, are given by
152: \cite{Bengtzelius1984,Goetze1991b,Szamel1991,Fuchs1993a},
153: \begin{subequations}\label{eq:mct:MCT}
154: \begin{equation}\label{eq:mct:phi_col}
155: \tau_q\partial_t\phi_q(t)+\phi_q(t)+\int_0^t
156: m_q(t-t')\partial_{t'}\phi_q(t')\,dt'=0\,.
157: \end{equation}
158: The initial conditions are $\phi_q (0)=1$. The microscopic time scale
159: reads $\tau_q=S_q/(D_0q^2)$. It is given by the short-time diffusion
160: coefficient, $D_0$, characterizing the Brownian motion and the static
161: structure factor $S_q=\langle|\rho_{\vec{q}}|^2\rangle$. The mode-coupling
162: approximation results in expressing the kernels $m_q(t)$ in terms of the
163: correlators $\phi_q(t)$ \cite{Goetze1991b},
164: \begin{equation}\label{eq:mct:kernel}
165: m_q (t) = {\cal F}_q \left[\mathbf{V}, \phi_k (t) \right] \,\, .
166: \end{equation}
167: As a consequence of the factorization into pair modes for the structural
168: relaxation in simple liquids, ${\mathcal F}_q$ is a bilinear functional of
169: the density correlators \cite{Bengtzelius1984},
170: \begin{equation}\label{eq:mct:Fdef}
171: {\mathcal F}_q[\tilde{f}] = \frac{1}{2} \int \frac{{d}^3k}{(2 \pi )^3}
172: V_{\vec{q},\vec{k}} \tilde{f}_k \tilde{f}_{|\vec{q}-\vec{k}|}\,,
173: \end{equation}
174: and the vertex is determined completely by the static structure of the
175: liquid system \cite{Goetze1976b,Sjoegren1980b},
176: \begin{equation}\label{eq:mct:vertex}
177: V_{\vec{q},\vec{k}} = S_q S_k S_{|\vec{q}-\vec{k}|}\, \rho
178: \left[ {\vec{q}} \cdot
179: \vec{k}\,{c_k} +\vec{q} \cdot
180: (\vec{q}-\vec{k})\,{{c_{|\vec{q}-\vec{k}|}} }
181: \right]^2/q^4\,.
182: \end{equation}
183: \end{subequations}
184: The number density is given by $\rho=N/V$ and $c_q$ denotes the direct
185: correlation function which is related to the static structure factor $S_q$
186: in the Ornstein-Zernike relation, $S_q=1/[1-\rho\,c_q]$, both depend on
187: external control parameters like density or temperature \cite{Hansen1986}.
188: For the SWS with hard-core diameter $d$, depth of the potential $u_0$, and
189: range of the potential $\Delta$, we get three dimensionless control
190: parameters, the packing fraction $\varphi=d^3\rho\pi/6$, the attraction
191: strength $\Gamma=u_0/(k_\text{B}T)$ and the relative well width
192: $\delta=\Delta/d$. These can be combined to a control-parameter vector
193: $\mathbf{V}=(\varphi,\Gamma,\delta)$.
194:
195: It is the long-time limit of the correlation function,
196: $\lim_{t\rightarrow\infty}\phi_q(t) = f_q$, that determines whether a
197: system is in the liquid regime, where $f_q=0$, or in an arrested state,
198: where $0< f_q\leqslant 1$. In the latter case, the values $f_q$
199: characterize the arrested glassy state and the $f_q$ are called glass-form
200: factors or Debye-Waller factors. In the long-time limit, the equation of
201: motion Eq.\ (\ref{eq:mct:MCT}) reduces to an equation involving only the
202: mode-coupling functional and the glass-form factors \cite{Goetze1991b},
203: \begin{equation}\label{eq:Feq}
204: f_q/(1-f_q) = {\cal F}_q[f]\,.
205: \end{equation}
206:
207: Frequently studied is the dynamics of a single or tagged particle with the
208: single particle density $\rho_q^s(t)=\exp[i\vec{q}\vec{r}_s(t)]$. For the
209: correlation function of a tagged particle,
210: $\phi^s_q(t)=\langle\rho_{{\vec{q}}}^{s\,*}(t) \rho^s_{\vec{q}}\rangle$,
211: similar equations as Eqs.\ (\ref{eq:mct:MCT}) have been derived
212: \cite{Bengtzelius1984,Fuchs1998},
213: \begin{subequations}\label{eq:mct:tagged}
214: \begin{equation}\label{eq:mct:phis_col}
215: \tau^s_q\partial_t\phi^s_q(t)+\phi^s_q(t)+\int_0^t
216: m^s_q(t-t')\partial_{t'}\phi^s_q(t')\,dt'=0\,,
217: \end{equation}
218: with $\tau^s_q=1/(D^s_0q^2)$. The short-time diffusion coefficient for a
219: single particle, $D^s_0$, again specifies the Brownian dynamics. The
220: mode-coupling functional for the tagged particle motion,
221: \begin{equation}\label{eq:mct:Fsdef}
222: {\mathcal F}^s_q[f,f^s] = \int \frac{{d}^3k}{(2 \pi )^3} S_k
223: \frac{\rho}{q^4} {c^s_k}^2 (\vec{q}\vec{k})^2
224: f_k f^s_{|\vec{q}-\vec{k}|}\,,
225: \end{equation}
226: is also determined by the static structure of the liquid system where
227: $c^s_q$ is the single-particle direct correlation function
228: \cite{Hansen1986}.
229: \end{subequations}
230:
231: The dynamics of the tagged particle is coupled to the coherent density
232: correlator $\phi_q(t)$ and for that reason $\phi^s_q(t)$ also displays the
233: bifurcation dynamics that is driven by $\phi_q(t)$. The equation for the
234: long-time limits of the tagged particle correlations function,
235: $\phi^s_q(t\rightarrow\infty)=f_q^s$, reads
236: \begin{equation}\label{eq:mct:fqs}
237: f_q^s/(1-f_q^s)
238: ={\cal F}^s_q[f,f^s]\,.
239: \end{equation}
240: In the following, the tagged particle will be assumed as of the same sort
241: as the host fluid. If the host particles are in the liquid state, $f_q=0$,
242: a tagged particle cannot be arrested, and in that case Eq.\
243: (\ref{eq:mct:fqs}) implies $f^s_q=0$.
244:
245: The MSD is defined by
246: $\delta r^2(t) = \langle |\vec{r}_s(t)-\vec{r}_s(0)|^2 \rangle$ and
247: describes the average distance a particle has traveled within some time
248: $t$ \cite{Hansen1986}. It is obtained, e.g., as small wave-number limit of
249: the tagged-particle correlator in Eq.\ (\ref{eq:mct:tagged}),
250: $\phi^s_q(t)=1-q^2\delta r^2(t)/6+{\mathcal O}(q^4)$
251: \cite{Goetze1991b,Fuchs1998},
252: \begin{subequations}\label{eq:mct:MSD}
253: \begin{equation}\label{eq:mct:MSD_col}
254: \delta r^2(t) + D^s_0 \int_0^t m^{(0)}(t-t')\,\delta
255: r^2(t')\,dt'=6D^s_0t\,,
256: \end{equation}
257: $m^{(0)}(t)=\lim_{q\rightarrow 0}m^s_q(t)
258: ={\mathcal F}_{MSD}[\phi(t),\phi^s(t)]$. The mode-coupling functional for
259: the MSD reads
260: \begin{equation}\label{eq:mct:FMSDdef}
261: {\mathcal F}_{MSD}[f,f^s] = \int \frac{{d}k}{(6 \pi^2 )}\,\rho\, S_k
262: (c^s_k)^2 f_k f^s_k\,.
263: \end{equation}
264: \end{subequations}
265: A characteristic localization length $r_s$ is defined by the second moment
266: for the relaxation of the distribution of $\phi^s_q(t)$
267: \cite{Goetze1991b}, which can be identified with the functional in Eq.\
268: (\ref{eq:mct:FMSDdef}) $r_s^2=1/{\mathcal F}_{MSD}[f,f^s]$. It is the
269: long-time limit of the MSD. Its value at the critical point, $r_s^c$,
270: characterizes the arrested structure. The value $6r_s^{c\,2}$ represents
271: the plateau for the dynamics of $\delta r^2(t)$.
272:
273: Equations~(\ref{eq:mct:MCT}) to (\ref{eq:mct:MSD}) are solved numerically
274: using algorithms introduced in \cite{Fuchs1991b,Goetze1996}. Details of
275: the implementation are found in \cite{Franosch1997,Dawson2001}. We use $d$
276: as unit of length, $d=1$, and choose the unit of time so that
277: $1/D_0=1/D_0^s=160$. The structure factors for the SWS and the HCY are
278: calculated in mean-spherical approximation \cite{Dawson2001,Cummings1979}.
279: The wave numbers shall be discretized to a grid of $M$ points with a
280: spacing $\Delta q=0.4/d$. The cutoff in the calculations is ranging from
281: $M=300$ for $\delta>0.04$ up to $M=750$ for $\delta=0.02$.
282:
283: %%
284: \subsection{\label{subsec:log}Logarithmic Decay Laws}
285:
286: The asymptotic solution at higher-order singularities shall be quoted from
287: Ref.~\cite{Goetze2002} where also further details can be found. The
288: asymptotic expansion is performed in small deviations of the correlation
289: function from the critical long-time limit $f^c_q$ involving the
290: coefficients
291: \begin{equation}\label{eq:asy:Aqk_def}
292: A_{q k_1 \cdots k_n}^{(n)} (\mathbf{V}) = \frac{1}{n!} (1 - f_q^c)\,
293: \frac{\partial^n {\cal F}_q \left [\mathbf{V}, f_k^c \right ] }{ \partial
294: f_{k_1}^c \cdots\partial f_{k_n}^c } \,
295: (1 - f_{k_1}^c) \cdots (1 - f_{k_n}^c) \,,
296: \end{equation}
297:
298: which can be split into values at the singularity,
299: $A_{qk_1\cdots k_n}^{(n)c}$, and remainders,
300: $A_{q k_1 \cdots k_n}^{(n) } (\mathbf{V}) = A_{q k_1 \cdots k_n}^{(n)c}
301: + \hat{A}_{q k_1 \cdots k_n}^{(n)} (\mathbf{V})$.
302: The Jacobian matrix of Eq.\ (\ref{eq:Feq}) is singular at the critical
303: points and assumes the form $\bigl[\delta_{q k} - A_{q k}^{(1) c} \bigr]$.
304: The non-negative left and right eigenvectors of matrix $A_{q k}^{(1) c}$
305: shall be denoted by $a^*_q$ and $a_q$ and can be fixed uniquely by
306: requiring $\sum_q \, a^*_q \, a_q = 1$ and $\sum_q \, a^*_q \, a_q^2 = 1$.
307: The reduced resolvent $R_{q k}$ of $A_{q k}^{(1) c}$ maps vectors
308: orthogonal to $a^*_q$ to vectors orthogonal to $a_q$.
309: The leading-order result for Eq.\ (\ref{eq:mct:MCT}) is then given by
310: \begin{equation}\label{eq:asy:log_decay}
311: \phi_q (t) = f_q^c + h_q \left[ - B \ln (t / \tau) \right] \,,\;
312: B = \sqrt{\left[ - 6\varepsilon_1 (\mathbf{V}) / \pi^2 \right]}\,,
313: \end{equation}
314: with the critical amplitudes $h_q=(1-f^c_q)a_q$ and the separation
315: parameter $\varepsilon_1(\mathbf{V})=a^*_q \hat{A}_q^{(0)}(\mathbf{V})$,
316: which is restricted to negative values, $\varepsilon_1<0$. If
317: $\varepsilon$ indicates the distance of the control parameters
318: $\mathbf{V}$ from the critical point, the leading result is of order
319: $\sqrt{\varepsilon}$ and correct up to ${\cal O}(\varepsilon)$. The
320: next-to-leading-order approximation is
321: \begin{eqnarray}\label{eq:asy:G1G2q}
322: \phi_q (t) = (f_q^c + \hat{f}_q) &+& h_q \bigl [ (- B + B_1)
323: \ln (t / \tau) \nonumber
324: \\ && + (B_2 + K_q B^2) \ln^2 (t / \tau)\nonumber
325: \\ && + B_3 \ln^3 (t / \tau) + B_4 \ln^4 (t / \tau) \bigr ] \,,
326: \end{eqnarray}
327: what includes the terms of order $\varepsilon$ and neglects terms of order
328: $\varepsilon^{3/2}$. It involves corrections to the plateau values,
329: \begin{equation}\label{eq:asy:deltafq}
330: \hat{f}_q = (1 - f_q^c) R_{q k} \left[ \hat{A}_{k }^{(0)} (\mathbf{V}) -
331: \epsilon_1 (\mathbf{V}) a_k^2 \right]\,,
332: \end{equation}
333: correction amplitudes,
334: \begin{equation}\label{eq:asy:Kq}
335: K_q = R_{q k} \left[ A_{k k_1 k_2 }^{(2)c} a_{k_1} a_{k_2}
336: - a_{k}^2 \right] / a_q \,,
337: \end{equation}
338: and the prefactors,
339: \begin{subequations}\label{eq:asy:Bcoeffs}
340: \begin{equation}\label{eq:asy:Bcoeffs_B1}
341: B_1 = (0.44425\, \zeta - 0.065381\, \mu_3)\,\varepsilon_1 (\mathbf{V})
342: - 0.22213 \,\varepsilon_2 (\mathbf{V}) \,\, ,
343: \end{equation}
344: \begin{equation}\label{eq:asy:Bcoeffs_B2}
345: B_2 = (0.91189\,\zeta + 0.068713\,\mu_3)\,\varepsilon_1 (\mathbf{V})
346: - 0.15198\,\varepsilon_2 (\mathbf{V}) \,\, ,
347: \end{equation}
348: \begin{equation}\label{eq:asy:Bcoeffs_B3}
349: B_3 = - 0.13504 \,\mu_3 \,\varepsilon_1 (\mathbf{V}) \,\, ,
350: \quad B_4 = - 0.046197 \,\mu_3 \,\varepsilon_1 (\mathbf{V}) \,\, .
351: \end{equation}
352: \end{subequations}
353: The numbers characterizing the higher-order singularities are
354: \begin{equation}\label{eq:asy:zeta}
355: \zeta = \sum_q a^*_q \left[ a_q^2 K_q + a_q^3/2
356: \right] \,\, ,
357: \end{equation}
358: and
359: \begin{equation}\label{eq:asy:mu3}
360: \mu_3 = 2 \zeta - \sum_q a^*_q \left[ A_{q k_1 k_2 k_3}^{(3)c}
361: a_{k_1} a_{k_2} a_{k_3} + 2 A_{q k_1 k_2}^{(2)c} a_{k_1}a_{k_2}
362: K_{k_2}
363: \right] \,\, .
364: \end{equation}
365: For the leading correction also an additional separation parameter is
366: introduced,
367: \begin{equation}\label{eq:asy:epsilon2_A}\begin{split}
368: \varepsilon_2 (\mathbf{V}) = \sum_q a^*_q
369: \hat{A}_{q k}^{(1)} (\mathbf{V}) a_k
370: + 2 \varepsilon_1 (\mathbf{V}) \sum_q a^*_q a_q^2 K_q
371: \\ + 2 \sum_q a^*_q \left[ A_{q k_1 k_2}^{(2)c} a_{k_1}
372: \hat{f}_{k_2}/(1-f_{k_2}^c)
373: - a_q \hat{f}_q /(1-f_q^c) \right] \,.\end{split}
374: \end{equation}
375: The time scale $\tau$ is determined by matching asymptotic approximation
376: and numerical solution of $\phi_q(t)$ at the plateau $f_q^c$ or the
377: rescaled plateau $f_q^c+\hat{f}_q$ for Eq.\ (\ref{eq:asy:log_decay}) and
378: Eq.\ (\ref{eq:asy:G1G2q}), respectively.
379:
380: %%
381: \subsection{\label{subsec:couple}Coupled Variables}
382:
383: Inserting the asymptotic expansion of Eq.\ (\ref{eq:asy:G1G2q}) into the
384: long-time limit of Eq.\ (\ref{eq:mct:phis_col}), the approximation for the
385: tagged particle dynamics up to order $\varepsilon$ is obtained
386: \cite{Sperl2003},
387: \begin{eqnarray}\label{eq:asy:phis}
388: \phi^s_q (t) = (f_q^{s\,c} + \hat{f}^s_q) &+& h^s_q \bigl [ (- B + B_1)
389: \ln (t / \tau) \nonumber
390: \\ && + (B_2 + K^s_q B^2) \ln^2 (t / \tau)\nonumber
391: \\ && + B_3 \ln^3 (t / \tau) + B_4 \ln^4 (t / \tau) \bigr ]\,,
392: \end{eqnarray}
393: with the critical amplitudes $h^s_q=(1-f^{s\,c}_q)\,a_q^s$, the correction
394: amplitudes $K^s_q$ and the plateau corrections $\hat{f}^s_q$. The latter
395: are derived from the functional~(\ref{eq:mct:Fsdef}) and the related
396: coherent quantities by
397: \begin{subequations}\label{eq:asy:taggedTaylor}
398: \begin{equation}\label{eq:asy:taggedTaylor:asq}
399: \sum_k (\delta_{qk} - {A^{s\,c}_{q,k}})\, a_k^s = \sum_k
400: {A^{s\,c}_{qk}} a_k\,,
401: \end{equation}
402: \begin{equation}\label{eq:asy:taggedTaylor:Ksq}\begin{split}
403: \sum_k (\delta_{qk} - {A^{s\,c}_{q,k}})\,a_k^s K^s_k = -
404: {a^s_q}^2
405: +\sum_k {A^{s\,c}_{qk}} a_k K_k\\
406: +\sum_{k,p} [
407: {A^{s\,c}_{q,kp}} a^s_k a^s_p + {A^{s\,c}_{qkp}} a_k a_p +
408: {A^{s\,c}_{qk,p}} a_k a^s_p
409: ]\,,\end{split}
410: \end{equation}
411: \begin{equation}\label{eq:asy:taggedTaylor:fsq}
412: \sum_k (\delta_{qk} - {A^{s\,c}_{q,k}})\,a_k^s \hat{f}_k^s =
413: -\varepsilon_1(\mathbf{V})\,{a^s_q}^2
414: + \sum_k {A^{s\,c}_{qk}}\, a_k\hat{f}_k
415: +\hat{A}^{s}_q(\mathbf{V})
416: \,.
417: \end{equation}
418: \end{subequations}
419: The derivatives with respect to the coherent and tagged particle
420: glass-form factors are denoted before and after the comma, respectively.
421: The coefficients are
422: \begin{equation}\label{eq:asy:Asqk_def}\begin{split}
423: &A^s_{q k_1 \cdots k_n, p_1 \cdots p_m} (\mathbf{V}) =
424: \frac{1}{n!} \frac{1}{m!} (1 - {f^s_q}^c)\,
425: \frac{\partial^n \partial_m {\cal F}^s_q \left [\mathbf{V}, f_k^c
426: ,{f^s_q}^c\right ] }{ \partial f_{k_1} \cdots\partial f_{k_n}
427: \partial f^s_{p_1} \cdots\partial f^s_{p_n} } \,
428: \\&\qquad\times
429: (1 - f_{k_1}^c) \cdots (1 - f_{k_n}^c)
430: (1 - {f^s_{p_1}}^c) \cdots (1 - {f^s_{p_n}}^c) =\\
431: &\qquad={A^{sc}_{q k_1 \cdots k_n, p_1 \cdots p_m}} +
432: \hat{A}^{s}_{q k_1 \cdots k_n, p_1 \cdots p_m} (\mathbf{V})
433: \,\, .\end{split}
434: \end{equation}
435:
436: Similar arguments as above yield the asymptotic expansion for the MSD up
437: to order $\varepsilon$ ,
438: \begin{eqnarray}\label{eq:asy:MSD}
439: \frac{1}{6}\delta r^2 (t) = {r^c_{s}}\,^2 &-& {\hat{r}_{s}}^2 -
440: h_\text{MSD}\, \bigl [ (- B + B_1)
441: \ln (t / \tau) \nonumber
442: \\ && + (B_2 + K_\text{MSD} B^2) \ln^2 (t / \tau)\nonumber
443: \\ && + B_3 \ln^3 (t / \tau) + B_4 \ln^4 (t / \tau) \bigr ] \,,
444: \end{eqnarray}
445: with parameters
446: \begin{subequations}\label{eq:asy:MSDpar}
447: \begin{equation}\label{eq:asy:MSDpar:h}
448: h_\text{MSD} = {r_s^{c\,4}}\{{\cal F}^c_\text{MSD}[h_k,f_p^{s\,c}]+
449: {\cal F}^c_\text{MSD}[f_k^c,h_p^s]\}\,,
450: \end{equation}
451: \begin{equation}\label{eq:asy:MSDpar:K}\begin{split}
452: K_\text{MSD} = \,&r_s^{c\,4}\{{\cal F}^c_\text{MSD}[h_k,h_p^s]
453: +{\cal F}^c_\text{MSD}[h_k K_k,f_p^{s\,c}]
454: \\&\qquad+{\cal F}^c_\text{MSD}[f_k^c,h_p^s K_p^s]
455: \}/h_\text{MSD}
456: -
457: h_\text{MSD}/r_s^{c\,2}\,,
458: \end{split}\end{equation}
459: \begin{equation}\label{eq:asy:MSDpar:df}\begin{split}
460: \hat{r}^2_{s} =\,& {r_s^{c\,4}}\{
461: {\cal F}^c_\text{MSD}[h_k\hat{f}_k,f_p^{s\,c}]
462: +{\cal F}^c_\text{MSD}[f_k^c,h_p^s\hat{f}_p^s]\\&\qquad
463: +{\cal F}^c_\text{MSD}[f_k^c,f_p^{s\,c}](\mathbf{V})
464: -{\cal F}^c_\text{MSD}[f_k^c,f_p^{s\,c}](\mathbf{V}^c)
465: \}/h_\text{MSD}\\&
466: -\varepsilon_1(\mathbf{V})h^2_\text{MSD}/r_s^{c\,2}\,.
467: \end{split}\end{equation}
468: \end{subequations}
469: For the generic liquid-glass transition, the asymptotic expansion was
470: carried out with a different convention as in Eq.\ (\ref{eq:asy:Aqk_def}),
471: however, the quantities $h_q$, $K_q$, $h^s_q$, $K^s_q$, $h_\text{MSD}$,
472: $K_\text{MSD}$, and $\zeta$ are the same in both descriptions
473: \cite{Franosch1997,Fuchs1998}. The plateau corrections, $\hat{f}_q$,
474: $\hat{f}^s_q$, and $\hat{r}^2_{s}$ are different for liquid-glass
475: transitions and higher-order singularities. The expansions in Eqs.\
476: (\ref{eq:asy:G1G2q}), (\ref{eq:asy:phis}), and (\ref{eq:asy:MSD}) share
477: the coefficients $B$, $B_1$, $B_2$, $B_3$, and $B_4$. They differ in the
478: plateau and its correction, the critical amplitude and the correction
479: amplitude.
480:
481: %%
482: \subsection{\label{subsec:MSD}Subdiffusive Power Law in the MSD}
483:
484: The logarithmic decay laws shall be phrased for the MSD in a slightly
485: different form than in Eq.\ (\ref{eq:asy:MSD}). This is done in order to
486: account for the fact that the MSD is conveniently shown in a
487: double-logarithmic representation which is more sensitive to the detection
488: of power laws. The asymptotic approximation (\ref{eq:asy:MSD}) for the MSD
489: can be written as $z = a_0 + a_1\,y + a_2\,y^2+ a_3\,y^3 + a_4\, y^4$.
490: Here, $z=\delta r^2(t)/6$ and $y=\ln(t/\tau)$ The constant term represents
491: the square of the corrected localization length,
492: $a_0=r_s^{c\,2}-\hat{r}_s^2$, the coefficients $a_1=h_\text{MSD}(B-B_1)$,
493: $a_2=-h_\text{MSD}(B_2+K_\text{MSD}B^2)$ as well as $a_3=-h_\text{MSD}B_3$
494: and $a_4-h_\text{MSD}B_4$ are the separation dependent prefactors for the
495: leading and next-to-leading order terms. This yields the expansion
496: \begin{subequations}\label{eq:asy:expox}
497: \begin{equation}\label{eq:asy:expoxexp}
498: \ln z = \ln r_s^{c\,2} - \hat{r}_s^2/r_s^{c\,2}
499: + x' \,y + b_2\,y^2 + a_3/r_s^{c\,2}\, y^3 + a_4/r_s^{c\,2}\, y^4
500: + {\mathcal O}(\varepsilon^{3/2})\,,
501: \end{equation}
502: with
503: \begin{equation}\label{eq:asy:expoxpar}
504: x' = a_1/r_s^{c\,2}\,,\quad b_2=
505: \frac{2 r_s^{c\,2} a_2 - a_1^2}{2 r_s^{c\,4}}\,.
506: \end{equation}
507: \end{subequations}
508: In leading order, one gets a power law for the MSD,
509: \begin{subequations}\label{eq:asy:powerlead}
510: \begin{equation}\label{eq:asy:expoxlaw}
511: \delta r^2(t)/6= r_s^{c\,2}\,(t/\tau)^{x}\,,
512: \end{equation}
513: with an exponent
514: \begin{equation}\label{eq:asy:expoxcalc}
515: x = h_\text{MSD}B/{r_s^c}\,^2\,.
516: \end{equation}
517: \end{subequations}
518: Exponent $x$ varies with the square-root in the separation parameter
519: $\varepsilon_1$, cf. Eq.\ (\ref{eq:asy:log_decay}). Including the
520: corrections of order $\varepsilon$ rescales the exponent to
521: \begin{subequations}\label{eq:asy:powercorr}
522: \begin{equation}\label{eq:asy:b1dash}
523: x' = h_\text{MSD}(B-B_1)/r_s^{c\,2}\,.
524: \end{equation}
525: and the next-to-leading order result reads
526: \begin{equation}\label{eq:asy:powercorrb2}\begin{split}
527: \delta r^2(t)/6 = (t/\tau)^{x'}&\{r_s^{c\,2}\,
528: -\hat{r}_s^2+ b_2\,r_s^{c\,2} \ln(t/\tau)^2\\&
529: +a_3\ln(t/\tau)^3+a_4\ln(t/\tau)^4\}\,.\end{split}
530: \end{equation}
531: \end{subequations}
532:
533: %%%
534: \section{\label{sec:corr}Correlation Functions near an $A_4$-singularity}
535:
536: Before we can apply the asymptotic expansion of Eq.\ (\ref{eq:asy:G1G2q}),
537: we need to specify the values for $\mu_3$ and $\zeta$ appearing in the
538: prefactors of Eqs.~(\ref{eq:asy:Bcoeffs}). The $A_4$-singularity is
539: characterized by $\mu_3=0$. This condition has been used to locate the
540: $A_4$-singularity at $\mathbf{V}=\mathbf{V}^*$ for the SWS by:
541: \begin{equation}\label{eq:gtd:A4.SWS.MSA2nd}
542: \varphi^* = 0.52768\,,\quad\Gamma^* = 4.4759\,,\quad \delta^* = 0.04381\,.
543: \end{equation}
544: The vanishing parameter $\mu_3$ implies a considerable simplification in
545: the preceding formulas since $B_3=B_4=0$ \cite{Goetze2002}. The deviations
546: of the control parameter values specifying the $A_4$-singularity from the
547: ones reported in Ref.~\cite{Dawson2001} originate from refined numerical
548: procedures used here and they do not exceed $6\%$. The characteristic
549: parameter was $|\mu_3|<5\cdot 10^{-4}$ at the control-parameter values
550: specified above. The parameter $\zeta$ varies regularly and is
551: $\zeta=0.122$ at the $A_4$-singularity. This is smaller than the value in
552: the HSS, $\zeta_\text{HSS}=0.269$ \cite{Franosch1997}.
553:
554: \begin{figure}[htb]
555: \includegraphics[width=\columnwidth]{fqA4}
556: \caption{\label{fig:swsasy:fqA4}Wave-vector dependent amplitudes
557: characterizing the $A_4$-singularity, Eq.\ (\ref{eq:gtd:A4.SWS.MSA2nd}),
558: for coherent and tagged particle correlators of the square-well system
559: (SWS).
560: In the upper and middle panel the critical glass-form factors $f_q^*$,
561: Eq.\ (\ref{eq:Feq}), and the amplitudes $h_q^*$ are shown as full lines,
562: respectively. The dashed lines represent the values for $f_q^{s\,*}$, Eq.\
563: (\ref{eq:mct:fqs}), and $h_q^{s\,*}$, Eq.\
564: (\ref{eq:asy:taggedTaylor:asq}). For the hard-sphere system (HSS),
565: $f_q^{s\,c}$ and $h_q^{s}$ are shown dotted. The lower panel shows the
566: correction amplitudes $K_q^*$, Eq.\ (\ref{eq:asy:Kq}), and $K_q^{s\,*}$,
567: Eq.\ (\ref{eq:asy:taggedTaylor:Ksq}), as full and dashed lines,
568: respectively. A square at $q=24.2$ indicates the corrections for the path
569: calculated for Fig.\ \ref{fig:swsasy:A4quadlines} and the correlators
570: shown in Fig.\ \ref{fig:swsasy:A4log}. The correction amplitudes $K_q$
571: ($-\cdot-$) and $K_q^{s}$ ($\cdots$) for the HSS are shown for comparison.
572: The unit of length here and in the following figures is the hard-core
573: diameter $d$ of the particles.
574: }
575: \end{figure}
576:
577: The second prerequisite for the asymptotic description according to Eq.\
578: (\ref{eq:asy:G1G2q}) are the wave-vector dependent amplitudes $f_q^*$,
579: $h_q^*$ and $K_q^*$. These are shown for the $A_4$-singularity in Fig.\
580: \ref{fig:swsasy:fqA4} together with the related values for the
581: tagged-particle correlator, Eq.\ (\ref{eq:asy:phis}). The quantities for
582: the tagged particle motion are close to the ones for the coherent
583: correlator $\phi_q(t)$ except for values of $q$ smaller than, say, $q=10$.
584: This difference was observed already for the HSS \cite{Fuchs1998}. Since
585: we will not be concerned with small $q$ in the following, we restrict the
586: discussion to the coherent dynamics and imply that the same is applicable
587: also to the incoherent part with only minor changes. In comparison to the
588: HSS the $f_q^*$, $h_q^*$, $f^{s\,*}_q$ and $h^{s\,*}_q$ are extended over
589: a broader $q$-range. The maximum in $h^s_q$ is shifted from $q\approx 13$
590: to $q\approx 25$ reflecting the smaller localization length in the SWS as
591: noticed before, cf. \cite{Dawson2001}. We see in the lower panel of Fig.\
592: \ref{fig:swsasy:fqA4} that the distributions of the correction amplitudes
593: $K_q$ and $K^s_q$ share that trend of becoming broader from the HSS to the
594: $A_4$-singularity of the SWS. The zero in $K_q$ moves from around
595: $q\approx 14$ in the HSS to $q\approx 32$ in the SWS. In addition, the
596: amplitudes are shifted to lower values for small $q$.
597:
598: \begin{figure}[htb]
599: \includegraphics[width=\columnwidth]{A4quadlines}
600: \caption{\label{fig:swsasy:A4quadlines}Curves of vanishing quadratic
601: correction in Eq.\ (\ref{eq:asy:G1G2q}) at the $A_4$-singularity of the
602: SWS, $B_2(q)=0$ (dash-dotted), for $q=7.0$, $20.2$, $24.2$, $27.0$, and
603: for $K_q=0$ as labeled.
604: The full line shows a part of the glass-transition diagram for constant
605: $\delta=\delta^*$. The lines of vanishing separation parameters
606: $\varepsilon_1(\mathbf{V})$ and $\varepsilon_2(\mathbf{V})$ are shown by a
607: broken and a dotted line, respectively. For the wave vector $q=24.2$, a
608: path on the curve $B_2(24.2)=0$ is marked ($+$) and labeled by $n$, for
609: which the correlators are shown in Fig.\ \ref{fig:swsasy:A4log}. State
610: $n=2$ is analyzed also in Fig.\ \ref{fig:swsasy:A4logqvar}. For the points
611: ($\blacktriangle$) labeled a, b, and c the decay is shown in Figs.\
612: \ref{fig:swsasy:A4logVvar} and \ref{fig:conclog}.
613: }
614: \end{figure}
615:
616: Having specified the characteristic parameters for the $A_4$-singularity,
617: the solution at any point in the control-parameter space can be compared
618: to the asymptotic approximation as the control parameters are translated
619: into separation parameters $\varepsilon_1$ and $\varepsilon_2$. As done
620: for the schematic models in Ref.~\cite{Goetze2002}, we start by
621: determining the surfaces where the quadratic corrections in Eq.\
622: (\ref{eq:asy:G1G2q}) are zero, $B_2(q)=B_2+K_qB^2=0$. On these surfaces in
623: the control-parameter space the logarithmic decay is expected to show up
624: as straight line around the plateau $f_q^*$, as the cubic and quartic terms
625: in Eq.\ (\ref{eq:asy:G1G2q}) vanish because of $B_3=B_4=\mu_3=0$ at the
626: $A_4$-singularity, cf. Eq.\ (\ref{eq:asy:Bcoeffs_B3}).
627: We get a different surface for each wave vector $q$ and show typical examples
628: in Fig.\ \ref{fig:swsasy:A4quadlines} for a cut through the glass-transition
629: diagram for $\delta=\delta^*$. For $q=7.0$ one gets $K_q=-1.81$. The
630: solution of $B_2(\mathbf{V})=1.81\,B(\mathbf{V})^2$ yields the chain line
631: labeled $B_2(7.0)=0$ in Fig.\ \ref{fig:swsasy:A4quadlines} and is lying in
632: the arrested region close to the line of liquid-glass transitions. Since
633: the $K_q$ depend smoothly on $q$, the evolution of the curve where
634: $B_2(q)=0$, can be understood by inspecting the parameters $B$ and $B_2$.
635: The square $B^2$ is always positive and proportional to
636: $\varepsilon_1(\mathbf{V})$, cf. Eq.\ (\ref{eq:asy:log_decay}), therefore
637: $K_q\,B^2$ is proportional to $K_q\,|\varepsilon_1(\mathbf{V})|$ and
638: shares the sign of $K_q$. Inserting $\mu_3=0$ and $\zeta=0.1216$ into Eq.\
639: (\ref{eq:asy:Bcoeffs_B2}) yields
640: $B_2(\mathbf{V})=0.111\,\varepsilon_1(\mathbf{V})
641: -0.152\,\varepsilon_2(\mathbf{V})$, which has to be positive to comply
642: with $B_2(q)=0$. The second separation parameter is negative,
643: $\varepsilon_2(\mathbf{V})<0$, below the dotted curve for
644: $\varepsilon_2=0$ in Fig.\ \ref{fig:swsasy:A4quadlines}. In addition, the
645: value $|\varepsilon_2(\mathbf{V})|$ on the line $B_2(7.0)=0$ is larger
646: than $|\varepsilon_1(\mathbf{V})|$ which we can also infer from the fact
647: that the line $\varepsilon_1=0$ is closer than the line $\varepsilon_2=0$.
648: We now chose a point on the line $B_2(7.0)=0$, keep the first separation
649: parameter fixed, say $\varepsilon_1=\varepsilon_1'$, and move to higher
650: values for $K_q$, e.g., for $q=20.2$ where $K_q=-0.966$. $B^2$ stays the
651: same and the term $K_qB^2$ increases. To ensure that $B_2(20.2)=0$, the
652: value $B_2(\mathbf{V})$ has to decrease. We can achieve that by moving
653: closer to the line $\varepsilon_2=0$. For fixed $\varepsilon_1'$ this
654: implies a shift to lower $\varphi$ and higher $\Gamma$. Consequently the
655: entire line, where $B_2(q)=0$, is rotating clockwise around the
656: $A_4$-singularity as $K_q$ increases. This is seen for the chain line
657: $B_2(20.2)=0$ in Fig.\ \ref{fig:swsasy:A4quadlines}. Since $K^s_q$ is
658: monotonic increasing with $q$ and $K_q$ has the same trend when neglecting
659: the small oscillations, Fig.\ \ref{fig:swsasy:fqA4}, the line
660: $B^s_2(q)=B_2(\mathbf{V}) +B(\mathbf{V})^2K^s_q=0$ also rotates clockwise
661: with increasing wave-vector $q$.
662:
663: The variation of the lines $B_2(q)=0$ described above depends only on the
664: angle at which the lines $\varepsilon_1=0$ and $\varepsilon_2=0$ intersect
665: at the $A_4$-singularity. This intersection is in a sense generic that it
666: is shared by the close-by $A_3$-singularities of the SWS. It applies also
667: to the $A_4$-singularities of the other potentials which are similar to
668: the square well. This is so because the functionals determining the
669: separation parameters depend on quantities like the structure factors and
670: the glass-form factors which are similar for different potentials
671: \cite{Goetze2003b}. For a given wave vector $q$, the line $B_2(q)=0$ may
672: or may not lie in the liquid regime depending on $K_q$. For the SWS at
673: $\delta=\delta^*$ we get a range of $-1\lesssim K_q\lesssim 0.4$
674: corresponding to $20\lesssim q \lesssim 35$, where a line $B_2(q)=0$ is
675: found in the liquid regime. We illustrate this by adding lines for
676: $q=24.2$, $q=27.0$ and for $K_q=0$ to Fig.\ \ref{fig:swsasy:A4quadlines}.
677: The vanishing $K_q$ is corresponding to $q\approx 32.3$ yielding a line
678: $B_2(q)=0$ still in the liquid. For $q\gtrsim 35$, the latter line rotates
679: further around the $A_4$-singularity and into the arrested regime beyond the
680: almost horizontal line of liquid-glass transitions.
681:
682: \begin{figure}[htb]
683: \includegraphics[width=\columnwidth]{A4log}
684: \caption{\label{fig:swsasy:A4log}Logarithmic decay at the $A_4$-singularity
685: in the SWS for $q=24.2$ on the path indicated in
686: Fig.\ \ref{fig:swsasy:A4quadlines}.
687: The correlation functions are shown as full lines for the states
688: $n=1,\,2,\,3$ (see text) and at $\mathbf{V}=\mathbf{V}^*$ labeled
689: $\phi_q^*$. The horizontal line indicates the critical plateau value
690: $f_q^*$ for $q=24.2$, short lines the renormalized plateaus
691: $f_q^*+\hat{f}_q$. Broken lines show the approximation of Eq.\
692: (\ref{eq:asy:G1G2q}), $-(B-B_1)\,\ln(t/\tau)$, dotted lines the
693: approximation by Eq.\ (\ref{eq:asy:log_decay}). Filled and open symbols,
694: respectively, mark the points where the approximations deviate by $5\%$
695: from the solution. The cross indicates the time when the solution for
696: $n=3$ and the critical correlator $\phi_q^*$ differ by $5\%$. The unit of
697: time here and in the following figures is given by a short-time diffusion
698: coefficient of $D_0=1/160$.
699: }
700: \end{figure}
701:
702: We select a wave vector $q=24.2$ with $K_q=-0.596$ as indicated in Fig.\
703: \ref{fig:swsasy:fqA4} by a square and choose a path on the line
704: $B_2(24.2)=0$ marked by the plus symbols in Fig.\
705: \ref{fig:swsasy:A4quadlines}. For $n=1,\,2,\,3$, the control parameters
706: are $(\Gamma,\varphi)=(3.312,0.5125)$, $(4.271,0.5250)$, and
707: $(4.453,0.5274)$, respectively. The solutions are shown in Fig.\
708: \ref{fig:swsasy:A4log} together with the leading approximation, Eq.\
709: (\ref{eq:asy:log_decay}), (dotted) and the next-to-leading approximation,
710: Eq.\ (\ref{eq:asy:G1G2q}), (dashed). The time scales $\tau$ are matched at
711: the plateau $f_q^*$ for the leading approximation and at the renormalized
712: plateaus $f_q^*+\hat{f}_q$ for the first correction. We recognize that for
713: $n=3$, Eq.\ (\ref{eq:asy:G1G2q}) accounts for more than ten decades in
714: time with a relative accuracy better than $5\%$. The leading approximation is
715: acceptable on that level for nine decades. For $n=1$ two and more than one
716: decade are covered, respectively. Five and three orders of magnitude in
717: time are achieved for $n=2$. For $n=1,\,2,\,3$, the leading approximation
718: describes at least $30\%$ of the complete decay and when including the
719: correction, $65\%$ are covered on the chosen accuracy level. The distance
720: in the control parameter $\Gamma$ from the value at the $A_4$-singularity
721: is $25\%$ for $n=1$ and $4\%$ for $n=2$, so no fine-tuning was necessary
722: to obtain such large windows for the logarithmic decay. The curve $n=1$
723: requires about five decades for the complete decay which is well in the
724: reach of today's computer simulation techniques \cite{Kob2003pre}.
725:
726: It was possible to describe part of the critical decay at an
727: $A_3$-singularity in a one-component model by the expansion in polynomials
728: in $\ln t$ at a point away from the singularity \cite{Goetze2002}. We
729: therefore compare the critical decay $\phi_q^*(t)$ with the decay for
730: $n=3$ and indicate the point at $t\approx 5000$ where both differ by $5\%$
731: in Fig.\ \ref{fig:swsasy:A4log}. With only the leading correction at our
732: disposal, a $2\%$-criterion was not fulfilled as for the one-component
733: model, where also the next-to-leading correction could be used
734: \cite{Goetze2002}. The dashed line for $n=3$ does not come closer to the
735: critical decay than $4\%$. Allowing for $5\%$, the interval from $t\approx
736: 20$ to $t\approx 4000$ could be described. However, at the
737: $A_4$-singularity the approximation in Eq.\ (\ref{eq:asy:G1G2q}) always
738: yields a straight $\ln t$-decay as approximation on the chosen path with
739: $B_2(q)=0$. This disagrees qualitatively with the observed critical
740: decay.
741:
742: \begin{figure}[htb]
743: \includegraphics[width=\columnwidth]{A4logqvar}
744: \caption{\label{fig:swsasy:A4logqvar}Logarithmic decay at the
745: $A_4$-singularity for varying wave vector $q$.
746: The inset shows the correlation functions $\phi_q(t)$ at state $n=2$ from
747: Fig.\ \ref{fig:swsasy:A4quadlines} for wave vectors $q=4.2,\,24.2\,,$ and
748: $32.2$ from top to bottom and the short horizontal lines show the
749: corresponding critical plateau values $f_q^*$. The full panel shows the
750: same correlation functions rescaled according to $\hat{\phi}_q(t)=
751: (\phi_q(t)-f^*_q-\hat{f}_q)/h^*_q$ as full lines and labeled by the
752: respective wave vectors. Dashed lines show the asymptotic laws, Eq.\
753: (\ref{eq:asy:G1G2q}). The deviations of the approximations from the
754: solutions of $5\%$ are marked by the open symbols. Filled symbols for
755: $q=4.2$ ($\blacktriangledown$) and $q=32.2$ ($\blacksquare$) show the
756: $5\%$ deviation from the additional approximation of neglecting quadratic
757: terms in Eq.\ (\ref{eq:asy:G1G2q}) (see text).
758: }
759: \end{figure}
760:
761: To identify correctly some decay that is linear in the $\phi_q(t)$ versus
762: $\log t$ diagram with the logarithmic decay predicted by the asymptotic
763: laws, Eq.\ (\ref{eq:asy:G1G2q}), we check if a different correlator with a
764: different correction amplitude $K_q$ is \textit{not} linear in $\ln t$ at
765: the same point in the control-parameter space. For a two-component model a
766: characteristic alternation of concave, linear and convex decay in $\ln t$
767: was found \cite{Goetze2002,Sperl2003}. Not both correlators could be
768: linear in $\ln t$ at the same time. For the SWS this check is performed at
769: the point $n=2$ from Fig.\ \ref{fig:swsasy:A4quadlines} by variation of
770: the wave vector $q$. For the wave vectors $q=4.2$ and $32.2$ the
771: correction amplitudes are $K_q=-1.400$ and $-0.0413$, respectively.
772: Therefore $B_2(4.2)<0$ and $B_2(32.2)>0$. We expect $\phi_q(t)$ to be
773: concave or convex, accordingly, as is demonstrated by the inset of Fig.\
774: \ref{fig:swsasy:A4logqvar}. The rescaled correlators $\hat{\phi}_q(t)$
775: displayed in the full panel allow for a more detailed analysis. We see
776: that the solutions as well as the approximations clearly exhibit increased
777: curvature for larger $q$. Since the coefficient linear in $\ln t$ is not
778: depending on $q$, cf. Eq.\ (\ref{eq:asy:G1G2q}), the middle dashed line
779: represents the leading correction to all three correlators when the
780: quadratic terms are neglected. For $q=24.2$ we observe good agreement over
781: almost 5 decades as before, cf. Fig.\ \ref{fig:swsasy:A4log}. For $q=4.2$
782: and $32.2$, however, the additional approximation reduces the range of
783: applicability to less than one decade as marked by the filled symbols.
784: Including the quadratic terms from the approximation~(\ref{eq:asy:G1G2q})
785: extends this range by half a decade to later times and to earlier times by
786: one and almost two decades for $q=4.2$ and $32.2$, respectively. The time
787: window defined by a $5\%$ deviation from the
788: approximation~(\ref{eq:asy:G1G2q}) is larger by one and two orders of
789: magnitude for $q=24.2$ than for $q=32.2$ and $q=4.2$, respectively, what
790: indicates that $q$-dependent higher-order corrections significantly
791: influence the range of applicability for the leading
792: correction~(\ref{eq:asy:G1G2q}).
793:
794: The time scale $\tau$ in Fig.\ \ref{fig:swsasy:A4logqvar} was matched for
795: $q=24.2$, so the violation of scale universality inherent to an
796: approximation like in Eq.\ (\ref{eq:asy:G1G2q}) leads to different times
797: $\tau(q)$, where the correlators for different $q$ cross their respective
798: renormalized plateau $f_q^*+\hat{f}_q$ \cite{Goetze2002}. The
799: representation with the rescaled $\hat{\phi}_q(t)$ is particularly
800: sensitive to these deviations since the point where the plateau is crossed
801: is required to be zero, $\hat{\phi}_q(t/\tau)=0$. In Fig.\
802: \ref{fig:swsasy:A4logqvar} we see that the line crossing the zero is
803: slightly broader than a single curve. The deviations in $\tau(q)$ are
804: small enough to not exceed the numerical grid for the time axis which
805: around $\tau=2988$ is given by $\Delta t=172$. So we interpolate to get
806: for $q=4.2,\,24.2$, and $32.2$, $\tau(q)=2899$, $2988$, and $3017$,
807: respectively. These differences do not introduce larger errors in the
808: analysis carried out above.
809:
810: \begin{figure}[htb]
811: \includegraphics[width=\columnwidth]{A4logVvar}
812: \caption{\label{fig:swsasy:A4logVvar}Logarithmic decay at the
813: $A_4$-singularity for the three states marked by triangles in
814: Fig.\ \ref{fig:swsasy:A4quadlines}.
815: The inset shows the correlation functions $\phi_q(t)$ for $q=24.2$. The
816: plateau value $f_q^*$ is indicated by the short horizontal line. The full
817: panel shows $\hat{\phi}_q(t)=(\phi_q(t)-f_q^*-\hat{f}_q)/h_q$ divided by
818: the respective values for $(B-B_1)$ at the three states specified. The
819: dashed curves show the result from Eq.\ (\ref{eq:asy:G1G2q}). Filled
820: squares and circles mark the points where curve a and c deviate by $5\%$
821: from $-\ln(t/\tau)$, respectively. The deviation for curve~b
822: ($\triangle$) for short times is at $t/\tau\approx 10^{-4}$ and not
823: included in the figure.
824: }
825: \end{figure}
826:
827: In order to change from convex to concave behavior we can also change the
828: control parameters. For states above the line $B_2(24.2)=0$, we expect
829: concave behavior, $B_2(q)<0$, for states below, convex decay,
830: $B_2(24.2)>0$. For a demonstration of this result, the rescaled
831: correlators $\hat{\phi}_q(t/\tau)$ at the states labeled a, b, c in Fig.\
832: \ref{fig:swsasy:A4quadlines} are divided by the prefactor $(B-B_1)$ of
833: Eq.\ (\ref{eq:asy:G1G2q}). This way the part of the decay that is linear
834: in $\ln t$ shows up as straight line with slope $-\ln 10$ in Fig.\
835: \ref{fig:swsasy:A4logVvar}. The approximations~(\ref{eq:asy:G1G2q}) are
836: shown as dashed lines for each state representing
837: $-\ln(t/\tau)+[B_2(24.2)/(B-B_1)]\ln^2(t/\tau)$. For state~b the
838: approximation is identical to $-\ln(t/\tau)$ and the solution follows that
839: line over 5 decades before $5\%$ deviation is reached. The states a and c
840: are chosen to have the same value for $B-B_1\approx 0.015$ and
841: $B_2(24.2)=\mp0.0020$, respectively. The solutions at state~a and~c follow
842: the $-\ln t$-law closely within a $5\%$ margin for two decades or one
843: decade, respectively, which is significantly less than found for state~b.
844: We can infer from Fig.\ \ref{fig:swsasy:A4quadlines} that at state~a the
845: quadratic corrections would vanish again if we went from $q=24.2$ to the
846: higher wave vector $q=27.0$. A scenario similar to the one shown in Fig.\
847: \ref{fig:swsasy:A4log} can be found.
848:
849: The procedure outlined in Figs.~\ref{fig:swsasy:A4quadlines},
850: \ref{fig:swsasy:A4log}, \ref{fig:swsasy:A4logqvar}, and
851: \ref{fig:swsasy:A4logVvar} can be summarized as follows. From the
852: higher-order singularities there emanate surfaces in the control-parameter
853: space for a specific wave vector $\bar{q}$ where the quadratic term in
854: Eq.\ (\ref{eq:asy:G1G2q}) is zero, cf. Fig.\ \ref{fig:swsasy:A4quadlines},
855: and the decay is linear in $\ln t$. Moving closer to the singularity on
856: that surface, the window in time where the logarithmic decay is a valid
857: approximation increases, cf. Fig.\ \ref{fig:swsasy:A4log}. On a fixed
858: point on that surface the decay is concave for $q<\bar{q}$ and convex for
859: $q>\bar{q}$, cf. Fig.\ \ref{fig:swsasy:A4logqvar}. For fixed $\bar{q}$,
860: the change from concave to convex is achieved by crossing the mentioned
861: surface from above in the sense exemplified in Fig.\
862: \ref{fig:swsasy:A4logVvar}.
863:
864: The coupled quantities share the leading asymptotic behavior of the
865: density correlators. As a consequence of the factorization theorem of MCT,
866: only the glass-form factors and the critical amplitudes $h_q$ are
867: different for the coupled quantities \cite{Goetze1985}. The leading
868: corrections imply a violation of a generalized factorization theorem.
869: These are proportional to the correction amplitude $K_q$. Since for large
870: wave vectors, say $q>10$, the quantities $f^s_q$, $h^s_q$, and $K^s_q$ are
871: close to the ones for the coherent correlator, the approximation for the
872: tagged particle correlation functions $\phi^s_q(t)$ for these large $q$ is
873: the same as for $\phi_q(t)$. So the discussion for $\phi^s_q(t)$ is
874: already exhausted by Fig.\ \ref{fig:swsasy:fqA4}. Not much could be gained
875: from repeating the discussion of the previous section for $\phi^s_q(t)$.
876:
877:
878: %%%
879: \section{\label{sec:MSD}Mean Squared Displacement near an $A_4$-singularity}
880:
881: \begin{figure}[htb]
882: \includegraphics[width=\columnwidth]{MSDx}
883: \caption{\label{fig:swsasy:MSDx}Subdiffusive power law in the mean-squared
884: displacement (MSD).
885: The solutions for states 1, 2, and 3 in the inset are shown as full lines
886: in the full panel together with the leading (dotted) and next-to-leading
887: (dashed) approximation by Eq.~(\ref{eq:asy:MSD}). The long horizontal line
888: represents $6\,r_s^{*\,2}=0.01086$, the short horizontal lines the
889: corrections to the plateau, $6(r_s^{*\,2}-\hat{r}_s^2)$, cf. Eq.\
890: (\ref{eq:asy:MSDpar:df}). The straight full lines show the power law
891: $(t/\tau)^{x}$, Eq.\ (\ref{eq:asy:expoxlaw}), with exponents $x=0.365$,
892: $0.173$ and $0.0878$ for states $n=1,\,2,\,3$. The filled symbols show the
893: points where the solutions deviate by $5\%$ from the leading-order power
894: laws. The inset shows part of the glass-transition diagram for
895: $\delta=\delta^*$ and a chain line where $b_2=0$,
896: cf. Eq.\ (\ref{eq:asy:expoxpar}), (see text).
897: }
898: \end{figure}
899:
900: According to Eq.\ (\ref{eq:asy:expoxexp}), $\delta r^2(t)$ is expected to
901: exhibit power-law behavior around the plateau $6\,r_s^{c\,2}$ provided the
902: term $b_2$ vanishes. The power-law exponent $x$ is determined explicitly
903: in Eq.\ (\ref{eq:asy:expoxcalc}) by the localization length and the
904: critical amplitude, which are
905: \begin{equation}\label{eq:swsasy:A4rs}
906: r_s^{*}=0.04255\,,\quad h^*_\text{MSD}=0.004051 \,.
907: \end{equation}
908:
909: The inset of Fig.\ \ref{fig:swsasy:MSDx} shows the line where $b_2$ from
910: Eq.\ (\ref{eq:asy:expoxexp}) vanishes. This line is almost identical to
911: the one for $B_2(24.2)=0$ shown in Fig.\ \ref{fig:swsasy:A4quadlines} for
912: the correlators $\phi_q(t)$. The MSD for three states on that line is
913: shown in the full panel. It is described well by the approximation in Eq.\
914: (\ref{eq:asy:MSD}). For states $n=1,\,2,\,3$, one, three and six decades
915: are covered with deviations less than $5\%$, so the approximation yields a
916: description of similar accuracy as for the correlation functions in Fig.\
917: \ref{fig:swsasy:A4log}. The leading result from Eq.\
918: (\ref{eq:asy:log_decay}) describes the relaxation proportional to $\ln t$
919: (dotted) which always has negative curvature in the double-logarithmic
920: representation and does not provide a valid description for $n=1$ and 2.
921: The reason for the qualitative difference between the solution for the MSD
922: and the leading logarithmic law is that the corrections proportional to
923: $K_\text{MSD}=-1.708$ are large, Therefore, $K_\text{MSD}B^2+B_2$ is never
924: close to zero in the liquid regime except very close to the
925: $A_4$-singularity. This is seen for $n=3$ in Fig.\ \ref{fig:swsasy:MSDx}
926: where $\ln t$ develops a straightened decay around the plateau.
927:
928: The power law~(\ref{eq:asy:powerlead}) provides a different formulation of
929: a leading order approximation and is shown in Fig.\ \ref{fig:swsasy:MSDx}
930: as straight line for $n=1,\,2,\,3$. For $n=1$ this describes the MSD for
931: more than a decade as indicated by the squares. For $n=2$ three decades
932: are covered and six decades of power-law behavior are identified for curve
933: $n=3$. So the accuracy is similar to the one provided by the approximation
934: in next-to-leading order by Eq.\ (\ref{eq:asy:MSD}). Both asymptotic
935: descriptions fall on top of each other around the plateau and therefore
936: corroborate that the reformulation~(\ref{eq:asy:expoxexp}) is justified.
937: The interpretation of the behavior of the MSD is then much simpler when
938: considering the power laws instead of the logarithms of time. The
939: decreasing slope of the relaxation when approaching the $A_4$-singularity
940: as in Fig.\ \ref{fig:swsasy:MSDx} is just the exponent $x$ from Eq.\
941: (\ref{eq:asy:expoxcalc}) which decreases as $B$ with the square-root of
942: the separation parameter $\varepsilon_1$, cf. Eq.\
943: (\ref{eq:asy:log_decay}). The same parameter $B$ is the prefactor of the
944: leading-order logarithmic decay in Eq.\ (\ref{eq:asy:log_decay}). In that
945: sense Fig.\ \ref{fig:swsasy:MSDx} is the analog of Fig.\
946: \ref{fig:swsasy:A4log}.
947:
948: \begin{figure}[htb]
949: \includegraphics[width=\columnwidth]{MSDV}
950: \caption{\label{fig:swsasy:MSDV}Concave and convex deviations from the
951: power law, Eq.\ (\ref{eq:asy:powerlead}) in the MSD.
952: Solutions for the states a, b, and c are shown as full lines, the
953: approximation~(\ref{eq:asy:powerlead}) as straight full lines for
954: exponents $x=0.147$, $0.173$, and $0.285$, respectively. Filled symbols
955: denote the $5\%$ deviation of the solutions from the leading-order power
956: law. For state~b, the dashed line exhibits the corrected power law with
957: $x'=0.155$, Eq.\ (\ref{eq:asy:b1dash}), and the open triangle the $5\%$
958: deviations of the solution from it. Dashed lines show the approximation by
959: Eq.\ (\ref{eq:asy:powercorrb2}) for a and c with $b_2=0.00363$ and
960: $-0.00735$, respectively, and $x'=0.143$ and $0.214$. The open symbols
961: mark the $5\%$ deviations. The inset replots the one from Fig.\
962: \ref{fig:swsasy:MSDx} and shows by the crosses the state points a, b, and
963: c.
964: }
965: \end{figure}
966:
967: The term $b_2$ in Eq.\ (\ref{eq:asy:expoxexp}) varies regularly in the
968: separation parameters $\varepsilon_1$ and $\varepsilon_2$, and $b_2$ is
969: positive above the line $b_2=0$ and negative below. Therefore, similar to
970: the case for the correlators in the linear-$\log$ representation, in the
971: double-logarithmic representation, the behavior of the MSD can be changed
972: from convex to concave when crossing the line of vanishing $b_2$. This is
973: demonstrated for three states in Fig.\ \ref{fig:swsasy:MSDV}. State~b is
974: identical to the state $n=2$ in Fig.\ \ref{fig:swsasy:MSDx} and obeys
975: $b_2=0$. The power law $(t/\tau)^{x}$ is shown as straight full line. The
976: time scale $\tau$ is matched at the plateau $6\,r_s^{*\,2}$. Moving to
977: state~c below the chain line, $(\Gamma, \varphi)=(3.42,0.525)$, a
978: relaxation is obtained which clearly exhibits negative curvature and is
979: consistent with the calculated value $b_2=-0.00735$. The leading-order
980: power law with exponent $x=0.285$ fulfills a $5\%$-deviation criterion for
981: two decades which accidentally extends to short times as the approximation
982: crosses the solution twice. Reducing the allowed deviation to $4\%$ would
983: reduce that interval to less than a decade. If we include the term
984: proportional to $b_2$ from Eq.\ (\ref{eq:asy:powercorrb2}) and renormalize
985: the exponent to $x'$, Eq.\ (\ref{eq:asy:b1dash}), the approximation agrees
986: with the solution for three decades. It is obvious from a comparison with
987: curve~1 in Fig.\ \ref{fig:swsasy:MSDx}, that the leading-order power law
988: describes that solution better than it describes the solution at state~c
989: in Fig.\ \ref{fig:swsasy:MSDV} for comparable values for $\tau$ and the
990: plateau correction $\hat{r}^2_{s}$ . The deviation to convex behavior is
991: demonstrated by the dashed line at curve~a,
992: $(\Gamma,\varphi)=(4.57,0.523)$. Again the range of validity is extended
993: to earlier times but for later times no improvement can be found.
994:
995: In Fig.\ \ref{fig:swsasy:MSDx} the dashed line, which describes the
996: next-to-leading order approximation of Eq.\ (\ref{eq:asy:G1G2q}), deviates
997: from the leading order power law~(\ref{eq:asy:expoxlaw}) below the plateau
998: where the range of validity for the power law extends to much smaller
999: times than justified by its derivation. We also recognize that the
1000: exponent $x$ overestimates the slope of the relaxation in
1001: Figs.~\ref{fig:swsasy:MSDx} and~\ref{fig:swsasy:MSDV}. In Eq.\
1002: (\ref{eq:asy:expoxcalc}) only the term $B$ from the leading order
1003: approximation is present. Taking into account the renormalization of this
1004: prefactor to $B-B_1$ in Eq.\ (\ref{eq:asy:b1dash}) changes the exponent
1005: for state b from $x=0.173$ to $x'=0.155$. By comparing the full line for
1006: the leading result and the dashed line for the corrected one in Fig.\
1007: \ref{fig:swsasy:MSDV}, we find that the range of applicability is shifted
1008: to later times by one decade and extended by two decades. The corrected
1009: power law is valid from $t=10^2$ to $t=5\cdot10^6$ and comparison to Fig.\
1010: \ref{fig:swsasy:MSDx} shows that approximation~(\ref{eq:asy:G1G2q}) covers
1011: a similar range. The accidental extension to shorter times is removed. The
1012: approximation now covers the range also a naive power-law fit would yield.
1013:
1014: In summary, the correction amplitude $K_\text{MSD}$ for the MSD does not
1015: vanish within the liquid regime. Therefore a logarithmic relaxation law
1016: can be detected only for states very close to the singularity. However,
1017: there is a line of vanishing corrections for the logarithm of the MSD.
1018: Here a logarithmic relaxation can be observed and this describes a
1019: subdiffusive power law of the MSD. We can interpret Fig.\
1020: \ref{fig:swsasy:MSDV} as the analog of Fig.\ \ref{fig:swsasy:A4logVvar}.
1021: Some quadratic correction to a leading-order linear behavior can be set to
1022: zero on a surface in control-parameter space. Departing from that surface
1023: in opposite directions introduces either positive or negative corrections
1024: and the linear behavior is changed to convex or concave.
1025:
1026: %%%
1027: \section{\label{sec:A3}$A_3$-singularities}
1028:
1029: \begin{figure}[htb]
1030: \includegraphics[width=\columnwidth]{fqA3}
1031: \caption{\label{fig:swsasy:fqA3}Glass-form factors $f_q^\circ$ and
1032: $f_q^{s\,\circ}$, amplitudes $h_q^\circ$ and $h_q^{s\,\circ}$, and
1033: correction amplitudes $K_q^\circ$ and $K_q^{s\,\circ}$ at the
1034: $A_3$-singularity for $\delta=0.03$.
1035: Line styles are the same as in Fig.\ \ref{fig:swsasy:fqA4}. The values at
1036: the $A_4$-singularity, $f_q^*$ (dotted), $h_q^*$ (dash-dotted), and
1037: $K_q^*$ (dash-dotted), are shown for comparison. The values for $q=24.2$
1038: and $45.0$ are marked by diamonds. The inset shows $K_q$ for $4<q<11$ for
1039: $\delta=\delta^*$ (chain line), $0.03$ (full line) and $0.02$ (dotted
1040: line).
1041: }
1042: \end{figure}
1043:
1044: An $A_3$-singularity is not located on a liquid-glass-transition line but
1045: is the endpoint of a glass-glass-transition line \cite{Goetze1991b}. The
1046: parameter $\mu_3$ is no longer vanishing and for $\delta=0.03$ we get
1047: $\mu_3=0.109$ and $\zeta=0.157$. For this $A_3$-singularity the
1048: $q$-dependent amplitudes are shown in Fig.\ \ref{fig:swsasy:fqA3}. No
1049: qualitative changes are obvious compared to the results shown in Fig.\
1050: \ref{fig:swsasy:fqA4} for the $A_4$. The smaller length scale
1051: $\delta=0.03$ for the attractive well introduces a smaller localization
1052: length, and this implies the broader distributions in wave-vector space.
1053: So the trend seen when changing from the HSS to the $A_4$-singularity of
1054: the SWS is continued when approaching $A_3$-singularities at smaller
1055: $\delta$. There are only two notable exceptions at smaller $q$. First, the
1056: value for $K_q$ at the position of the structure factor peak is minimal
1057: for the $A_4$, $-1.81=K_q^*<K_q^\circ=-1.72$. The inset shows this region
1058: enlarged for $\delta=\delta^*$, $0.03$ and $0.02$, demonstrating that
1059: $K_q$ at the peak is again larger for the $A_3$-singularity with smaller
1060: well width $0.02$, where $K_q=-1.69$. Second, the zero-wave-vector limit
1061: of $K^s_q$ is also smallest at the $A_4$-singularity. The respective
1062: values for $\delta=\delta^*$, $0.03$ and $0.02$ are $-1.71$, $-1.64$, and
1063: $-1.62$. Therefore, one experiences the strongest $q$-dependent
1064: corrections at the $A_4$-singularity.
1065:
1066: \begin{figure}[htb]
1067: \includegraphics[width=\columnwidth]{A3quadlines}
1068: \caption{\label{fig:swsasy:A3quadlines}Curves of vanishing quadratic
1069: correction in Eq.\ (\ref{eq:asy:G1G2q}) for the $A_3$-singularity
1070: ($\bigcirc$) at $\delta=0.03$.
1071: The $\delta=0.03$ cut through the glass-transition diagram is displayed by
1072: full lines. The various lines are shown in the same style as in Fig.\
1073: \ref{fig:swsasy:A4quadlines} and labeled accordingly. The line $b_2=0$,
1074: cf. Eq.\ (\ref{eq:asy:expoxpar}), indicates the analogous line for the
1075: MSD, cf. inset of Fig.\ \ref{fig:swsasy:MSDx}.
1076: }
1077: \end{figure}
1078:
1079: Figure~\ref{fig:swsasy:A3quadlines} shows the analog of Fig.\
1080: \ref{fig:swsasy:A4quadlines} for a cut through the glass-transition
1081: diagram at $\delta=0.03$. The lines $\varepsilon_1=0$ and
1082: $\varepsilon_2=0$ for the $A_3$-singularity are obtained from a smooth
1083: transformation of the corresponding lines at the $A_4$-singularity, and
1084: they appear in similar locations in the diagram. The line
1085: $\varepsilon_2=0$ is again very close to the almost horizontal line of
1086: transitions. Just below, we find again the line where $B_2(q)=0$ when
1087: $K_q=0$. However, this now represents $q\approx 57.5$, cf. Fig.\
1088: \ref{fig:swsasy:fqA3}, which is a value almost twice as large as for the
1089: corresponding line in Fig.\ \ref{fig:swsasy:A4quadlines}. For the wave
1090: vector $q=24.2$ we find the line, where $B_2(24.2)=0$, completely in the
1091: glass state. Taking the same value for the correction amplitude as for
1092: $q=24.2$ at the $A_4$, $K^*_q\approx-0.6$, we obtain $q=45.0$, cf. Fig.\
1093: \ref{fig:swsasy:fqA3} and the line labeled accordingly in Fig.\
1094: \ref{fig:swsasy:A3quadlines}. Since the latter line comes close to the
1095: liquid-glass-transition line we take that as a reference and estimate the
1096: range of wave-vectors where the quadratic corrections can be put to zero
1097: in the liquid regime to $45\lesssim q \lesssim 70$. The lines where
1098: $B_2(q)=0$ can be rather sensitive to $q$-variation. This is demonstrated
1099: by the curve $B_2(46.2)=0$. Although the change in the wave vector is
1100: relatively small in comparison to $q=45.0$, the values for $K_q$ differ by
1101: more than $20\%$ for fixed $q$ and induce a rotation of the line
1102: $B_2(q)=0$ by quite a significant angle.
1103:
1104: Having in mind the drastic changes in the lines where $B_2(q)=0$, it may
1105: come with some surprise that the line for the MSD, where $b_2=0$, stays
1106: rather robust and accessible in the liquid regime as seen in Fig.\
1107: \ref{fig:swsasy:A3quadlines}. The variation in $q$ for the amplitudes is
1108: reflected in changes of the localization lengths. For the
1109: $A_3$-singularity at $\delta=0.03$ we get
1110: \begin{equation}\label{eq:swsasy:A3rs}
1111: r_s^\circ=0.0243\,,\quad h^\circ_\text{MSD}=0.00136\,.
1112: \end{equation}
1113: From Eq.\ (\ref{eq:swsasy:A4rs}) one gets $r_s^*/r_s^\circ= 1.75$ and the
1114: square of the latter ratio, $r_s^{*\,2}/r_s^{\circ\,2}\approx 3$, is the
1115: same as $h^*_\text{MSD}/h^\circ_\text{MSD}$. Since only the fraction
1116: $h_\text{MSD}/r_s^2$ could introduce larger modifications in Eq.\
1117: (\ref{eq:asy:expoxexp}), the changes in $b_2$ cancel approximately and the
1118: line specified by $b_2=0$ experiences only minor deformations when
1119: $\delta$ is varied. The wave vector for which the lines $B_2(q)=0$ and
1120: $b_2=0$ are closest to each other, is $q=45.8$ at the $A_3$-singularity
1121: for $\delta=0.03$.
1122:
1123: \begin{figure}[htb]
1124: \includegraphics[width=\columnwidth]{delta_var}
1125: \caption{\label{fig:swsasy:delta_var}Parameters for the asymptotic
1126: description at the $A_3$-singularities of the SWS for varying $\delta$.
1127: Panel~A displays $\mu_3$ ($\blacktriangle$), Eq.\ (\ref{eq:asy:mu3}), and
1128: $\zeta$ ($\Diamond$), Eq.\ (\ref{eq:asy:zeta}). The dashed curve shows the
1129: asymptotic $\sqrt{\delta^*-\delta}$-law for the $\mu_3$. The localization
1130: length $r_s^\circ$ is shown in panel~B. The correction amplitudes
1131: $K_\text{MSD}^\circ$ and the ratios
1132: $h_\text{MSD}^\circ/r_\text{MSD}^{\circ\,2}$ are shown in panels~C and~D.
1133: }
1134: \end{figure}
1135:
1136: To corroborate the finding for the MSD from the preceding paragraph, the
1137: parameters for the asymptotic description of the MSD at the
1138: $A_3$-singularities are shown in Fig.\ \ref{fig:swsasy:delta_var}. The
1139: $\mu_3$ vanish when we approach the $A_4$-singularity. The decrease close
1140: to $\delta^*$ is described asymptotically by a square-root variation,
1141: $\mu_3\propto \sqrt{\delta^*-\delta}$, shown by the dashed line
1142: \cite{Sperl2003}. The smallness of $\mu_3$ indicates that all the
1143: $A_3$-singularities are already influenced by the proximity of the
1144: close-by $A_4$-singularity. One can take advantage of this finding and
1145: conclude that the terms proportional to $\mu_3$ in Eq.\
1146: (\ref{eq:asy:G1G2q}) are small. Moreover, one may neglect $B_3$ and $B_4$
1147: in Eq.\ (\ref{eq:asy:Bcoeffs_B3}) entirely without introducing large
1148: additional errors. The leading correction to the logarithmic decay laws is
1149: then only quadratic also for the $A_3$-singularities. Parameter $\zeta$
1150: varies regularly around a finite value at $\delta^*$ but shares the
1151: variation of $\mu_3$ at $\delta^*$ due to Eq.\ (\ref{eq:asy:mu3}). Panel~B
1152: shows the decrease of the localization length at the $A_3$-singularity
1153: when $\delta$ is reduced. A change of $40\%$ in $r^\circ_s$ from
1154: $\delta=\delta^*$ to $\delta=0.03$, cf. Eqs.~(\ref{eq:swsasy:A4rs})
1155: and~(\ref{eq:swsasy:A3rs}), is reflected in the broadening of the
1156: distributions in $q$ seen in Figs.\ \ref{fig:swsasy:fqA4}
1157: and~\ref{fig:swsasy:fqA3}. This broadening is responsible for the large
1158: variation in $q$ when comparing Fig.\ \ref{fig:swsasy:A4quadlines} with
1159: Fig.\ \ref{fig:swsasy:A3quadlines}. It was noted in the discussion of the
1160: inset of Fig.\ \ref{fig:swsasy:fqA3} that $K_q$ introduces the strongest
1161: corrections for the correlation functions at the $A_4$-singularity. This
1162: is also true for the MSD as seen in panel~C for $K_\text{MSD}$ which is
1163: largest in absolute value at the $A_4$-singularity. The variation in
1164: $K_\text{MSD}$ with $\delta$ is however small and does not introduce
1165: significant changes to $a_2$ in Eq.\ (\ref{eq:asy:expoxpar}). The
1166: amplitude $h_\text{MSD}$ is the remaining parameter entering Eq.\
1167: (\ref{eq:asy:expoxpar}) that could alter the location of the line $b_2=0$
1168: in the glass-transition diagram. We noted above that only the ratio
1169: $h_\text{MSD}/r_s^{c\,2}$ needs to be considered which is shown in
1170: panel~D. From there one infers that the ratio varies only by less than
1171: $5\%$. We can conclude that the line of power law variation for the MSD
1172: stays in the liquid regime even when $\delta$ is changed significantly.
1173:
1174: %%%
1175: \section{\label{sec:HCY}Hard Core Yukawa System}
1176:
1177: The $A_l$-singularities occurring in MCT are topologically stable, smooth
1178: changes in the control parameters do not challenge their existence.
1179: Therefore, the results for the SWS can be applied also to other potentials
1180: with a short-ranged attraction. Nevertheless, the deformation of the
1181: potential might introduce changes large enough to be relevant for the
1182: detection of the higher-order singularities. Among several potentials the
1183: hard core Yukawa system (HCY) was found to differ by up to $20\%$ in
1184: certain properties at the $A_4$-singularity from the SWS
1185: \cite{Goetze2003b}. Since other potentials differ less we use that system
1186: as a second example for an $A_4$-singularity.
1187:
1188: \begin{figure}[htb]
1189: \includegraphics[width=\columnwidth]{HCYquadlines}
1190: \caption{\label{fig:HCYquadlines}Cut through the parameter space for the
1191: hard-core Yukawa system for $\delta=\delta^*$.
1192: Lines styles are the same as in Fig.\ \ref{fig:swsasy:A4quadlines}. The
1193: wave vectors $q=15.0$, $21.4$, $27.0$, and $28.2$ are approximately
1194: equivalent to $q=7.0$, $20.2$, $24.2$, and $27.0$ in
1195: Fig.\ \ref{fig:swsasy:A4quadlines}, respectively, after rescaling
1196: $\Gamma$ by a factor of 2.98 (see text).
1197: }
1198: \end{figure}
1199:
1200: Figure~\ref{fig:HCYquadlines} shows the analog of Fig.\
1201: \ref{fig:swsasy:A4quadlines} for the HCY. For a comparison, the
1202: $A_4$-singularity in the SWS was mapped on top of the $A_4$-singularity in
1203: the HCY by scaling in $\Gamma$ with a factor of $2.98$ and by a shift in
1204: $\varphi$ of $0.0065$.
1205: The same transformation was applied to the lines where $B_2(q)=0$ in the
1206: SWS. Fig.\ \ref{fig:HCYquadlines} displays the $B_2(q)=0$ lines for the
1207: HCY that come closest to the ones shown in Fig.\ \ref{fig:swsasy:A4quadlines}
1208: after the mapping. The correction amplitude $K_q$ for the HCY vanishes at
1209: $q\approx 34$, and the range in wave vector for which $B_2(q)=0$ is lying in
1210: the liquid regime is shifted to higher wave vectors, $21\lesssim q\lesssim 36$
1211: or $-0.9\lesssim K_q\lesssim 0.2$, in comparison to the SWS.
1212: For $q=27.0$ we get the line $B_2(q)=0$ for the HCY that is closest to the
1213: line $b_2=0$ for the MSD as compared to $B_2(24.2)=0$ in the SWS.
1214:
1215: %%%
1216: \section{\label{sec:conclusion}Conclusion}
1217:
1218: Logarithmic decay or, equivalently, $1/f$~noise in the fluctuation
1219: spectra, can arise in a number of situations and is explained by various
1220: approaches \cite{Weissman1988}. In the log-linear representation
1221: appropriate for the correlation functions, this decay exhibits a straight
1222: line. To discriminate the logarithmic decay laws originating from
1223: higher-order glass-transition singularities within MCT \cite{Goetze2002}
1224: from other possible scenarios one needs criteria to distinguish one from
1225: the other. The theory makes specific predictions where in the
1226: control-parameter space the logarithmic decay is expected and how the
1227: corrections introduce deviations from that behavior. In this paper, the
1228: scenarios are discussed in quantitative detail for an example relevant for
1229: studies of colloidal dynamics, the square-well system (SWS). To proceed,
1230: specific cuts through the three-dimensional parameter space are
1231: considered. Here, lines are identified where the corrections quadratic in
1232: the logarithm of time vanish for a chosen wave vector $q$, cf. Fig.\
1233: \ref{fig:swsasy:A4quadlines}. These lines emanate from the higher-order
1234: singularity and rotate clockwise around the higher-order singularity with
1235: increasing $q$. The correlation functions for states on these lines
1236: exhibit decays that are linear in the logarithm of time for several orders
1237: of magnitude in time, cf. Fig.\ \ref{fig:swsasy:A4log}. In leading order,
1238: the slope of the decay is given by the square-root of the distance from
1239: the higher-order singularity, Eq.\ (\ref{eq:asy:log_decay}). The
1240: mean-squared displacement MSD displays a power law, Eq.\
1241: (\ref{eq:asy:expoxlaw}), that is valid on a similar line in the
1242: control-parameter space, cf. Fig.\ \ref{fig:swsasy:MSDx}. The exponent $x$
1243: of this subdiffusive behavior is also decreasing with the square-root of
1244: the distance. Both the logarithmic decay and the power law are accessible
1245: in the liquid regime. The logarithmic decay is predicted for wave vectors
1246: $q$, which are equivalent to values of about three to four times the first
1247: peak of the static structure factor.
1248:
1249: In a semi-logarithmic representation for the correlation functions and a
1250: double-logarithmic plot of the MSD, characteristic convex and concave
1251: relaxation patterns are found when states are chosen that are off the
1252: specified lines, cf. Figs.~\ref{fig:swsasy:A4logVvar} and
1253: \ref{fig:swsasy:MSDV}. Due to the variation of the correction amplitude
1254: $K_q$ in Fig.\ \ref{fig:swsasy:fqA4}, a similar variation from convex to
1255: concave behavior is introduced by changes in the wave vector at a fixed
1256: point in control-parameter space, cf. Fig.\ \ref{fig:swsasy:A4logqvar}.
1257: These deviations from logarithmic behavior provide a test for the clear
1258: identification of dynamical scenarios that are consistent with Eq.\
1259: (\ref{eq:asy:G1G2q}) and hence originate from higher-order singularities.
1260:
1261: When the localization at the higher-order singularity is changed by either
1262: deforming the shape of the potential or by moving to $A_3$-singularities
1263: at smaller ranges of the attraction, the logarithmic decay of the
1264: correlation functions is shifted to higher wave vectors. Whereas the
1265: difference between the SWS and the hard-core Yukawa system at the
1266: $A_4$-singularity is modest, cf. Fig.\ \ref{fig:swsasy:A4quadlines} and
1267: Fig.\ \ref{fig:HCYquadlines}, the lines of vanishing quadratic correction
1268: change drastically at the $A_3$-singularity, cf. Fig.\
1269: \ref{fig:swsasy:A3quadlines}. In contrast, the line where the subdiffusive
1270: power law for the MSD is valid, is robust against changes of the well
1271: width and the potential shape, cf. Figs.\ \ref{fig:swsasy:MSDx},
1272: \ref{fig:swsasy:A3quadlines}, and \ref{fig:HCYquadlines}.
1273:
1274: \begin{figure}[htb]
1275: \includegraphics[width=\columnwidth]{conclog}
1276: \caption{\label{fig:conclog}Correlators for states a (panel a) and b
1277: (panel b) from Fig.\ \ref{fig:swsasy:A4quadlines} for wave vectors
1278: $q=4.2$, $20.2$, $24.2$, $27.0$, $32.2$, and $36.2$ from top to bottom.
1279: Full lines show the solutions of the MCT equations for the SWS, dashed
1280: lines the approximation by Eq.\ (\ref{eq:asy:G1G2q}). Triangles mark the
1281: $5\%$ deviation of the correlator from the approximation for $q=27.0$ and
1282: $24.2$, respectively. The dotted vertical lines indicate the time scales
1283: $\tau$, the short horizontal lines the corrected plateau value
1284: $f_q+\hat{f}_q$ for $q=27.0$ and $24.2$, respectively.
1285: }
1286: \end{figure}
1287:
1288: For comparing the solutions of the equations of motion, Eqs.\
1289: (\ref{eq:mct:MCT}), (\ref{eq:mct:tagged}), and (\ref{eq:mct:MSD}), with
1290: the asymptotic expansions, Eqs.\ (\ref{eq:asy:G1G2q}),
1291: (\ref{eq:asy:phis}), and (\ref{eq:asy:MSD}), all parameters are calculated
1292: explicitly except the time scale $\tau$ which is matched at the plateau.
1293: In an experiment or a computer simulation only the correlators are
1294: available directly. We show these in Fig.\ \ref{fig:conclog} for two
1295: states specified in Fig.\ \ref{fig:swsasy:A4quadlines} for different wave
1296: vectors. Since state b is closer to the $A_4$-singularity, the range of
1297: validity for the asymptotic approximation is larger than for state a.
1298: Especially the extension of the linear-$\log$ decay at some specific wave
1299: vector increases when moving closer to the singularity. As noted in
1300: connection with Fig.\ \ref{fig:swsasy:A4logqvar}, the range of validity
1301: for the approximation by Eq.\ (\ref{eq:asy:G1G2q}) may vary with $q$.
1302: Partly for that reason a larger absolute curvature is attributed to the
1303: correlators by the approximation than a fit would do. A free fit could
1304: identify logarithmic behavior at state~b for $q=20.2$ from $t\approx 5$ to
1305: $t\approx 5\cdot 10^7$ with a deviation of at most $5\%$. In addition,
1306: fitting the correlator for $q=24.2$ also for $t\geqslant 10^5$ would yield
1307: positive curvature. Therefore, a free fit in that region of the
1308: control-parameter space tends to find the logarithmic decay at a somewhat
1309: lower wave vector than predicted by Eq.\ (\ref{eq:asy:G1G2q}). However,
1310: with a choice of the time scale $\tau$ that is reasonably close to the
1311: theoretical value, the concave and convex decay patterns can still be
1312: identified unambiguously in the correlators without invoking additional
1313: assumptions.
1314:
1315: \begin{figure}[htb]
1316: \includegraphics[width=\columnwidth]{concfqA4}
1317: \caption{\label{fig:concfqA4}Comparison of $f_q$ and $h_q$ from the fit to
1318: the simulation of two different states \cite{Sciortino2003pre} with the
1319: values $f_q^*$ and $h_q^*$ for the SWS from Fig.\ \ref{fig:swsasy:fqA4}.
1320: For the comparison in the lower panel the theoretical values are
1321: multiplied by $0.14$.
1322: }
1323: \end{figure}
1324:
1325: A recent molecular dynamics study of a binary mixture of square-well
1326: particles identifies a power law with $x=0.28$ for the MSD over four
1327: decades and a related logarithmic decay of the correlation function at a
1328: wave vector $q=16.8$ \cite{Sciortino2003pre}. A scenario similar to Fig.\
1329: \ref{fig:conclog} was found for the correlation functions: Upon increasing
1330: $q$, a change from concave to convex decay is observed. For a second
1331: state, faster decay with larger prefactors for the logarithmic decay is
1332: reported together with a larger exponent, $x=0.44$, for the power law in
1333: the MSD. This finding is consistent with the assumption that this second
1334: state is further from the supposed higher-order singularity than the first
1335: state. Different from Fig.\ \ref{fig:conclog}, in the simulation $\delta$
1336: was changed to vary the distance while $\varphi$ and $\Gamma$ were kept
1337: fixed. The logarithmic decay was shifted to a higher wave vector for
1338: smaller $\delta$ \cite{Sciortino2003pre}. This is consistent with the
1339: expectation that can be inferred from Figs.\ \ref{fig:swsasy:A4quadlines}
1340: and \ref{fig:swsasy:A3quadlines} by observing, e.g., the rotation of the
1341: line $B_2(24.2)=0$. The analysis of the simulation data allowed for a fit
1342: of the values for $f_q^*$ and $h_q^*$ \cite{Sciortino2003pre}. These are
1343: shown in Fig.\ \ref{fig:concfqA4} together with the theoretical
1344: predictions for the SWS. The fitted parameters for both states almost fall
1345: on top of each other for $f_q^*$. The amplitude $h_q^*$ is deduced from
1346: the simulation data only up to some overall factor. It can be matched
1347: reasonably by a multiplication of the theoretical prediction for $h_q^*$.
1348: The extension in $q$ for the values obtained from MCT for the SWS are
1349: narrower, the width at half maximum for $f_q$ differs by $15\%$. A similar
1350: difference was observed for a binary mixture of hard spheres and agreement
1351: between theory and simulation could be improved by using the structure
1352: factor from the simulation as input to the MCT calculations
1353: \cite{Foffi2003pre}. For the amplitude $h_q^*$ the locations of the maxima
1354: disagree by $15\%$ and the width is different by $25\%$. The deviations
1355: for $q<7$ in both $f_q$ and $h_q$ can be attributed to the effects of
1356: mixing \cite{Foffi2003pre}.
1357:
1358: In summary, scenarios for logarithmic decay near higher-order
1359: glass-transition singularities are presented in this work. Some essential
1360: predictions are supported by the results of computer simulations. This
1361: should motivate further investigations in colloidal systems with
1362: short-ranged attraction. In particular the power-law behavior for the MSD
1363: including the deviations might be accessible to experiments.
1364:
1365: \acknowledgments
1366: I thank W.~G\"otze for valuable discussion. This work was supported by the
1367: Deutsche Forschungsgemeinschaft Grant Go154/13-1.
1368:
1369: \bibliographystyle{apsrev}
1370:
1371: \begin{thebibliography}{38}
1372: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1373: \expandafter\ifx\csname bibnamefont\endcsname\relax
1374: \def\bibnamefont#1{#1}\fi
1375: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1376: \def\bibfnamefont#1{#1}\fi
1377: \expandafter\ifx\csname citenamefont\endcsname\relax
1378: \def\citenamefont#1{#1}\fi
1379: \expandafter\ifx\csname url\endcsname\relax
1380: \def\url#1{\texttt{#1}}\fi
1381: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1382: \providecommand{\bibinfo}[2]{#2}
1383: \providecommand{\eprint}[2][]{\url{#2}}
1384:
1385: \bibitem[{\citenamefont{Boon and Yip}(1980)}]{Boon1980}
1386: \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Boon}} \bibnamefont{and}
1387: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Yip}},
1388: \emph{\bibinfo{title}{Molecular Hydrodynamics}}
1389: (\bibinfo{publisher}{McGraw-Hill}, \bibinfo{address}{New York},
1390: \bibinfo{year}{1980}).
1391:
1392: \bibitem[{\citenamefont{G\"otze}(1991)}]{Goetze1991b}
1393: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}}, in
1394: \emph{\bibinfo{booktitle}{Liquids, Freezing and Glass Transition}}, edited by
1395: \bibinfo{editor}{\bibfnamefont{J.~P.} \bibnamefont{Hansen}},
1396: \bibinfo{editor}{\bibfnamefont{D.}~\bibnamefont{Levesque}}, \bibnamefont{and}
1397: \bibinfo{editor}{\bibfnamefont{J.}~\bibnamefont{Zinn-Justin}}
1398: (\bibinfo{publisher}{North Holland}, \bibinfo{address}{Amsterdam},
1399: \bibinfo{year}{1991}), vol. \bibinfo{volume}{Session LI (1989)} of
1400: \emph{\bibinfo{series}{Les Houches Summer Schools of Theoretical Physics}},
1401: pp. \bibinfo{pages}{287--503}.
1402:
1403: \bibitem[{\citenamefont{Bengtzelius et~al.}(1984)\citenamefont{Bengtzelius,
1404: G\"otze, and Sj\"olander}}]{Bengtzelius1984}
1405: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Bengtzelius}},
1406: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}}, \bibnamefont{and}
1407: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Sj\"olander}},
1408: \bibinfo{journal}{J.~Phys.~C} \textbf{\bibinfo{volume}{17}},
1409: \bibinfo{pages}{5915} (\bibinfo{year}{1984}).
1410:
1411: \bibitem[{\citenamefont{Arnol'd}(1992)}]{Arnold1986}
1412: \bibinfo{author}{\bibfnamefont{V.~I.} \bibnamefont{Arnol'd}},
1413: \emph{\bibinfo{title}{Catastrophe Theory}} (\bibinfo{publisher}{Springer},
1414: \bibinfo{address}{Berlin}, \bibinfo{year}{1992}), \bibinfo{edition}{3rd} ed.
1415:
1416: \bibitem[{\citenamefont{Franosch et~al.}(1997)\citenamefont{Franosch, Fuchs,
1417: G\"otze, Mayr, and Singh}}]{Franosch1997}
1418: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Franosch}},
1419: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}},
1420: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1421: \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Mayr}}, \bibnamefont{and}
1422: \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Singh}},
1423: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{55}},
1424: \bibinfo{pages}{7153} (\bibinfo{year}{1997}).
1425:
1426: \bibitem[{\citenamefont{Fuchs et~al.}(1998)\citenamefont{Fuchs, G\"otze, and
1427: Mayr}}]{Fuchs1998}
1428: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}},
1429: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}}, \bibnamefont{and}
1430: \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Mayr}},
1431: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{58}},
1432: \bibinfo{pages}{3384} (\bibinfo{year}{1998}).
1433:
1434: \bibitem[{\citenamefont{Pusey and van Megen}(1986)}]{Pusey1986}
1435: \bibinfo{author}{\bibfnamefont{P.~N.} \bibnamefont{Pusey}} \bibnamefont{and}
1436: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{van Megen}},
1437: \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{320}},
1438: \bibinfo{pages}{340} (\bibinfo{year}{1986}).
1439:
1440: \bibitem[{\citenamefont{van Megen}(1995)}]{Megen1995}
1441: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{van Megen}},
1442: \bibinfo{journal}{Transp.~Theory~Stat.~Phys.} \textbf{\bibinfo{volume}{24}},
1443: \bibinfo{pages}{1017} (\bibinfo{year}{1995}).
1444:
1445: \bibitem[{\citenamefont{Fabbian et~al.}(1999)\citenamefont{Fabbian, G\"otze,
1446: Sciortino, Tartaglia, and Thiery}}]{Fabbian1999}
1447: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Fabbian}},
1448: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1449: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}},
1450: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1451: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Thiery}},
1452: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{59}},
1453: \bibinfo{pages}{R1347} (\bibinfo{year}{1999});
1454: \textbf{\bibinfo{volume}{60}},
1455: \bibinfo{pages}{2430(E)} (\bibinfo{year}{1999}).
1456:
1457:
1458: \bibitem[{\citenamefont{Bergenholtz and Fuchs}(1999)}]{Bergenholtz1999}
1459: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bergenholtz}} \bibnamefont{and}
1460: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}},
1461: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{59}},
1462: \bibinfo{pages}{5706} (\bibinfo{year}{1999}).
1463:
1464: \bibitem[{\citenamefont{Dawson et~al.}(2001)\citenamefont{Dawson, Foffi, Fuchs,
1465: G\"otze, Sciortino, Sperl, Tartaglia, Voigtmann, and
1466: Zaccarelli}}]{Dawson2001}
1467: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Dawson}},
1468: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Foffi}},
1469: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}},
1470: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1471: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}},
1472: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sperl}},
1473: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1474: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Voigtmann}},
1475: \bibnamefont{and}
1476: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Zaccarelli}},
1477: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{63}},
1478: \bibinfo{pages}{011401} (\bibinfo{year}{2001}).
1479:
1480: \bibitem[{\citenamefont{Poon}(2002)}]{Poon2002}
1481: \bibinfo{author}{\bibfnamefont{W.~C.~K.} \bibnamefont{Poon}},
1482: \bibinfo{journal}{J.~Phys.: Condens.~Matter} \textbf{\bibinfo{volume}{14}},
1483: \bibinfo{pages}{R859} (\bibinfo{year}{2002}).
1484:
1485: \bibitem[{\citenamefont{G\"otze and Sperl}(2003)}]{Goetze2003b}
1486: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}} \bibnamefont{and}
1487: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sperl}},
1488: \bibinfo{journal}{J.~Phys.: Condens.~Matter} \textbf{\bibinfo{volume}{15}},
1489: \bibinfo{pages}{S869} (\bibinfo{year}{2003}).
1490:
1491: \bibitem[{\citenamefont{Eckert and Bartsch}(2002)}]{Eckert2002}
1492: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Eckert}} \bibnamefont{and}
1493: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Bartsch}},
1494: \bibinfo{journal}{Phys.~Rev.~Lett.} \textbf{\bibinfo{volume}{89}},
1495: \bibinfo{pages}{125701} (\bibinfo{year}{2002}).
1496:
1497: \bibitem[{\citenamefont{Pham et~al.}(2002)\citenamefont{Pham, Puertas,
1498: Bergenholtz, Egelhaaf, Moussa\"{\i}d, Pusey, Schofield, Cates, Fuchs, and
1499: Poon}}]{Pham2002}
1500: \bibinfo{author}{\bibfnamefont{K.~N.} \bibnamefont{Pham}},
1501: \bibinfo{author}{\bibfnamefont{A.~M.} \bibnamefont{Puertas}},
1502: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bergenholtz}},
1503: \bibinfo{author}{\bibfnamefont{S.~U.} \bibnamefont{Egelhaaf}},
1504: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moussa\"{\i}d}},
1505: \bibinfo{author}{\bibfnamefont{P.~N.} \bibnamefont{Pusey}},
1506: \bibinfo{author}{\bibfnamefont{A.~B.} \bibnamefont{Schofield}},
1507: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Cates}},
1508: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}}, \bibnamefont{and}
1509: \bibinfo{author}{\bibfnamefont{W.~C.~K.} \bibnamefont{Poon}},
1510: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{296}},
1511: \bibinfo{pages}{104} (\bibinfo{year}{2002}).
1512:
1513: \bibitem[{\citenamefont{Foffi et~al.}(2002)\citenamefont{Foffi, Dawson,
1514: Buldyrev, Sciortino, Zaccarelli, and Tartaglia}}]{Foffi2002b}
1515: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Foffi}},
1516: \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Dawson}},
1517: \bibinfo{author}{\bibfnamefont{S.~V.} \bibnamefont{Buldyrev}},
1518: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}},
1519: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Zaccarelli}},
1520: \bibnamefont{and}
1521: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1522: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{65}},
1523: \bibinfo{pages}{050802} (\bibinfo{year}{2002}).
1524:
1525: \bibitem[{\citenamefont{Zaccarelli et~al.}(2002)\citenamefont{Zaccarelli,
1526: Foffi, Dawson, Buldyrev, Sciortino, and Tartaglia}}]{Zaccarelli2002b}
1527: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Zaccarelli}},
1528: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Foffi}},
1529: \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Dawson}},
1530: \bibinfo{author}{\bibfnamefont{S.~V.} \bibnamefont{Buldyrev}},
1531: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}},
1532: \bibnamefont{and}
1533: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1534: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{66}},
1535: \bibinfo{pages}{041402} (\bibinfo{year}{2002}).
1536:
1537: \bibitem[{\citenamefont{Puertas et~al.}(2003)\citenamefont{Puertas, Fuchs, and
1538: Cates}}]{Puertas2003}
1539: \bibinfo{author}{\bibfnamefont{A.~M.} \bibnamefont{Puertas}},
1540: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}}, \bibnamefont{and}
1541: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Cates}},
1542: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{67}},
1543: \bibinfo{pages}{031406} (\bibinfo{year}{2003}).
1544:
1545: \bibitem[{\citenamefont{Puertas et~al.}(2002)\citenamefont{Puertas, Fuchs, and
1546: Cates}}]{Puertas2002}
1547: \bibinfo{author}{\bibfnamefont{A.~M.} \bibnamefont{Puertas}},
1548: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}}, \bibnamefont{and}
1549: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Cates}},
1550: \bibinfo{journal}{Phys.~Rev.~Lett.} \textbf{\bibinfo{volume}{88}},
1551: \bibinfo{pages}{098301} (\bibinfo{year}{2002}).
1552:
1553: \bibitem[{\citenamefont{Mallamace et~al.}(2000)\citenamefont{Mallamace,
1554: Gambadauro, Micali, Tartaglia, Liao, and Chen}}]{Mallamace2000}
1555: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Mallamace}},
1556: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Gambadauro}},
1557: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Micali}},
1558: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1559: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Liao}}, \bibnamefont{and}
1560: \bibinfo{author}{\bibfnamefont{S.-H.} \bibnamefont{Chen}},
1561: \bibinfo{journal}{Phys.~Rev.~Lett.} \textbf{\bibinfo{volume}{84}},
1562: \bibinfo{pages}{5431} (\bibinfo{year}{2000}).
1563:
1564: \bibitem[{\citenamefont{Kegel and van Blaaderen}(2000)}]{Kegel2000}
1565: \bibinfo{author}{\bibfnamefont{W.~K.} \bibnamefont{Kegel}} \bibnamefont{and}
1566: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{van Blaaderen}},
1567: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{287}},
1568: \bibinfo{pages}{290} (\bibinfo{year}{2000}).
1569:
1570: \bibitem[{\citenamefont{Weeks et~al.}(2000)\citenamefont{Weeks, Crocker,
1571: Levitt, Schofield, and Weitz}}]{Weeks2000}
1572: \bibinfo{author}{\bibfnamefont{E.~R.} \bibnamefont{Weeks}},
1573: \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Crocker}},
1574: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Levitt}},
1575: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Schofield}},
1576: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Weitz}},
1577: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{287}},
1578: \bibinfo{pages}{627} (\bibinfo{year}{2000}).
1579:
1580: \bibitem[{\citenamefont{Weeks and Weitz}(2002)}]{Weeks2002}
1581: \bibinfo{author}{\bibfnamefont{E.~R.} \bibnamefont{Weeks}} \bibnamefont{and}
1582: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Weitz}},
1583: \bibinfo{journal}{Phys.~Rev.~Lett.} \textbf{\bibinfo{volume}{89}},
1584: \bibinfo{pages}{095704} (\bibinfo{year}{2002}).
1585:
1586: \bibitem[{\citenamefont{G\"otze and Sperl}(2002)}]{Goetze2002}
1587: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}} \bibnamefont{and}
1588: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sperl}},
1589: \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{66}},
1590: \bibinfo{pages}{011405} (\bibinfo{year}{2002}).
1591:
1592: \bibitem[{\citenamefont{Szamel and L\"owen}(1991)}]{Szamel1991}
1593: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Szamel}} \bibnamefont{and}
1594: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{L\"owen}},
1595: \bibinfo{journal}{Phys.~Rev.~A} \textbf{\bibinfo{volume}{44}},
1596: \bibinfo{pages}{8215} (\bibinfo{year}{1991}).
1597:
1598: \bibitem[{\citenamefont{Fuchs}(1993)}]{Fuchs1993a}
1599: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}}, Ph.D. thesis,
1600: \bibinfo{school}{TU M\"unchen} (\bibinfo{year}{1993}).
1601:
1602: \bibitem[{\citenamefont{G\"otze and L\"ucke}(1976)}]{Goetze1976b}
1603: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}} \bibnamefont{and}
1604: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{L\"ucke}},
1605: \bibinfo{journal}{Phys.~Rev.~B} \textbf{\bibinfo{volume}{13}},
1606: \bibinfo{pages}{3825} (\bibinfo{year}{1976}).
1607:
1608: \bibitem[{\citenamefont{Sj\"ogren}(1980)}]{Sjoegren1980b}
1609: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Sj\"ogren}},
1610: \bibinfo{journal}{Phys.~Rev.~A} \textbf{\bibinfo{volume}{22}},
1611: \bibinfo{pages}{2883} (\bibinfo{year}{1980}).
1612:
1613: \bibitem[{\citenamefont{Hansen and McDonald}(1986)}]{Hansen1986}
1614: \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}} \bibnamefont{and}
1615: \bibinfo{author}{\bibfnamefont{I.~R.} \bibnamefont{McDonald}},
1616: \emph{\bibinfo{title}{Theory of Simple Liquids}}
1617: (\bibinfo{publisher}{Academic Press}, \bibinfo{address}{London},
1618: \bibinfo{year}{1986}), \bibinfo{edition}{2nd} ed.
1619:
1620: \bibitem[{\citenamefont{Fuchs et~al.}(1991)\citenamefont{Fuchs, G\"otze,
1621: Hofacker, and Latz}}]{Fuchs1991b}
1622: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fuchs}},
1623: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1624: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Hofacker}}, \bibnamefont{and}
1625: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Latz}},
1626: \bibinfo{journal}{J.~Phys.: Condens.~Matter} \textbf{\bibinfo{volume}{3}},
1627: \bibinfo{pages}{5047} (\bibinfo{year}{1991}).
1628:
1629: \bibitem[{\citenamefont{G\"otze}(1996)}]{Goetze1996}
1630: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1631: \bibinfo{journal}{J.~Stat.~Phys.} \textbf{\bibinfo{volume}{83}},
1632: \bibinfo{pages}{1183} (\bibinfo{year}{1996}).
1633:
1634: \bibitem[{\citenamefont{Cummings and Smith}(1979)}]{Cummings1979}
1635: \bibinfo{author}{\bibfnamefont{P.~T.} \bibnamefont{Cummings}} \bibnamefont{and}
1636: \bibinfo{author}{\bibfnamefont{E.~R.} \bibnamefont{Smith}},
1637: \bibinfo{journal}{Chem.~Phys.} \textbf{\bibinfo{volume}{42}},
1638: \bibinfo{pages}{241} (\bibinfo{year}{1979}).
1639:
1640: \bibitem[{\citenamefont{Sperl}(2003)}]{Sperl2003}
1641: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sperl}}, Ph.D. thesis,
1642: \bibinfo{school}{TU M\"unchen} (\bibinfo{year}{2003}).
1643:
1644: \bibitem[{\citenamefont{Kob}(2003)}]{Kob2003pre}
1645: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Kob}}, in
1646: \emph{\bibinfo{booktitle}{Slow Relaxations and Nonequilibrium Dynamics in
1647: Condensed Matter}}, edited by \bibinfo{editor}{\bibfnamefont{J.-L.}
1648: \bibnamefont{Barrat}},
1649: \bibinfo{editor}{\bibfnamefont{M.}~\bibnamefont{Feigelman}},
1650: \bibnamefont{and} \bibinfo{editor}{\bibfnamefont{J.}~\bibnamefont{Kurchan}}
1651: (\bibinfo{publisher}{Springer}, \bibinfo{address}{Berlin},
1652: \bibinfo{year}{2003}), vol. \bibinfo{volume}{Session LXXVII (2002)} of
1653: \emph{\bibinfo{series}{Les Houches Summer Schools of Theoretical Physics}}.
1654:
1655: \bibitem[{\citenamefont{G\"otze}(1985)}]{Goetze1985}
1656: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1657: \bibinfo{journal}{Z.~Phys.~B} \textbf{\bibinfo{volume}{60}},
1658: \bibinfo{pages}{195} (\bibinfo{year}{1985}).
1659:
1660: \bibitem[{\citenamefont{Weissman}(1988)}]{Weissman1988}
1661: \bibinfo{author}{\bibfnamefont{M.~B.} \bibnamefont{Weissman}},
1662: \bibinfo{journal}{Rev.~Mod.~Phys.} \textbf{\bibinfo{volume}{60}},
1663: \bibinfo{pages}{537} (\bibinfo{year}{1988}).
1664:
1665: \bibitem[{\citenamefont{Sciortino et~al.}(2003)\citenamefont{Sciortino,
1666: Tartaglia, and Zaccarelli}}]{Sciortino2003pre}
1667: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}},
1668: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1669: \bibnamefont{and}
1670: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Zaccarelli}}
1671: (\bibinfo{year}{2003}), \eprint{cond-mat/0304192}.
1672:
1673: \bibitem[{\citenamefont{Foffi et~al.}(2003)\citenamefont{Foffi, G\"otze,
1674: Sciortino, Tartaglia, and Voigtmann}}]{Foffi2003pre}
1675: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Foffi}},
1676: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{G\"otze}},
1677: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}},
1678: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}},
1679: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Voigtmann}}
1680: (\bibinfo{year}{2003}), \eprint{preprint}.
1681:
1682: \end{thebibliography}
1683:
1684: \end{document}
1685:
1686: