cond-mat0010159/PRE.tex
1: \documentstyle[aps,pre,twocolumn,graphicx,psfig]{revtex}
2: 
3: \def\lessgtr{\raise2.5pt\hbox{$<$}\llap{\lower2.5pt\hbox{$>$}}}
4: \def\gtrless{\raise2.5pt\hbox{$>$}\llap{\lower2.5pt\hbox{$<$}}}
5: 
6: \newcommand{\be}{\begin{equation}}
7: \newcommand{\ee}{\end{equation}}
8: \newcommand{\bea}{\begin{eqnarray}}
9: \newcommand{\eea}{\end{eqnarray}}
10: 
11: \begin{document}
12: 
13: \draft
14: 
15: \title{A mode-coupling theory for the glassy dynamics of 
16: a diatomic probe molecule immersed in a simple liquid}
17: \author{S.-H. Chong, W.~G{\"o}tze, and A.P. Singh}
18: \address{Physik-Department, Technische Universit{\"a}t M{\"u}nchen,
19: 85747 Garching, Germany}
20: \date{Phys. Rev. E, in print}
21: \maketitle
22: 
23: \begin{abstract}
24: 
25: Generalizing the mode-coupling theory for ideal
26: liquid-glass transitions, equations of motion are derived for the
27: correlation functions describing the glassy dynamics of a diatomic
28: probe molecule immersed in a simple glass-forming system. The
29: molecule is described in the interaction-site representation and
30: the equations are solved for a dumbbell molecule
31: consisting of two fused hard spheres in a hard-sphere system. The
32: results for the molecule's arrested position in the glass state
33: and the reorientational correlators for angular-momentum index
34: $\ell = 1$ and $\ell = 2$ near the glass transition are compared
35: with those obtained previously within a
36: theory based on a tensor-density description of the molecule in
37: order to demonstrate that the two approaches yield
38: equivalent results. 
39: For strongly hindered reorientational motion, the dipole-relaxation
40: spectra for the $\alpha$-process can be mapped on the dielectric-loss
41: spectra of glycerol if a rescaling is performed according to 
42: a suggestion by Dixon {\it et al.} 
43: [Phys.~Rev.~Lett. {\bf 65}, 1108 (1990)].
44: It is demonstrated that the glassy dynamics is
45: independent of the molecule's inertia parameters.
46: 
47: \bigskip
48: 
49: \noindent PACS numbers: 64.70.Pf, 61.25.Em, 61.20.Lc
50: \end{abstract}
51: 
52: \section{INTRODUCTION}
53: 
54: The mode-coupling theory (MCT) for the evolution of structural
55: relaxation in glass-forming liquids was originally developed
56: for atomic systems and for mixtures of atoms or ions. Detailed
57: tests of the theory have been provided by comparisons of the 
58: predictions for the hard-sphere system (HSS) with 
59: dynamic-light-scattering data for
60: hard-sphere colloids, as can be inferred from Ref.~\cite{Megen95}
61: and the papers quoted there. Quantitative tests have also been
62: made by comparing molecular-dynamics-simulation data for a
63: binary mixture with the MCT results for the model
64: \cite{Nauroth97,Gleim98,Gleim00}. A series of general 
65: implications of the MCT equations were derived like
66: scaling laws and relations between anomalous exponents describing
67: power-law spectra and relaxation-time scales which
68: establish some universal features of the dynamics~\cite{Goetze92}.
69: It was conjectured that these 
70: results also apply to molecular liquids. Indeed, there is a large
71: body of literature, which is reviewed in Ref.~\cite{Goetze99},
72: dealing with the analysis of data from experiments or
73: from molecular-dynamics simulations for complicated systems in
74: terms of the universal MCT formulas. These studies
75: suggest that MCT describes some essential features of the glassy
76: dynamics for molecular liquids. Therefore, it seems desirable
77: to develop a detailed microscopic theory also for systems of
78: non-spherical constituents.
79: 
80: A mode-coupling theory for molecular systems has been studied in
81: Refs.~\cite{Schilling97,Theis98,Fabbian99b,Winkler00,Theis00,Letz00} where
82: the structure is described by tensor-density
83: fluctuations. The basic concepts of the MCT for simple systems
84: like density correlators and relaxation kernels have been
85: generalized to infinite matrices. The equations for the non-ergodicity
86: parameters and critical amplitudes were solved. These
87: quantities generalize the Debye-Waller factors of the arrested
88: glass structure and characterize its changes with temperature.
89: Comparison of the theoretical findings with
90: molecular-dynamics-simulation data for 
91: water~\cite{Fabbian99b,Theis00}
92: and for a system of linear molecules \cite{Winkler00} demonstrates
93: that the theory can cope with microscopic details.
94: However, the derived equations 
95: are so involved that further simplifications would be
96: required before correlators or spectra
97: could actually be calculated.
98: 
99: The simplest question of glassy dynamics of the rotational
100: degrees of freedom concerns the motion of a single linear molecule in a
101: simple liquid. This problem is equivalent to the study of a dilute
102: solution of linear molecules in an atomic liquid as solvent. For
103: this system, a MCT has been developed, generalizing the
104: equation for a tagged particle in a simple liquid to an 
105: infinite-matrix equation for a tagged molecule \cite{Franosch97c}. 
106: The equations were solved for a molecule
107: consisting of two fused hard spheres immersed in a HSS
108: \cite{Franosch97c,Goetze00c}. The validity of the
109: universal laws for the reorientational dynamics was
110: demonstrated. Characteristic differences for the
111: $\alpha$-process of the relaxation for angular-momentum index
112: $\ell = 1$ and $\ell = 2$ were identified 
113: which explain the differences of
114: spectra measured for $\ell = 1$ by dielectric-loss spectroscopy
115: and for $\ell = 2$ by depolarized-light-scattering spectroscopy.
116: The experimentally established large ratio of the
117: $\alpha$-relaxation times for the $\ell = 2$--reorientational
118: process and the 
119: longitudinal elastic modulus was also obtained \cite{Goetze00c}.
120: These examples show that MCT can provide general insight into the
121: glassy dynamics of rotational degrees of freedom which goes beyond
122: the contents of the universal formulas. 
123: 
124: Within the basic version of MCT, 
125: the tagged-particle-density-fluctuation correlator
126: for wave number $q$ considered as function of time $t$, 
127: $\phi_q^s (t)$, or the dynamical structure
128: factor for frequency $\omega$, $S_{q}(\omega)$, can be written as:
129: $\phi_q^s (t) = \phi_{q}^{s \, *} (t/t_0)$ and 
130: $S_{q}(\omega) = S_{q}^{*} (\omega t_0)$. 
131: Here the functions $\phi_{q}^{s \, *} (\tilde t)$ and 
132: $S_{q}^{*} (\tilde \omega)$ are completely
133: determined by the equilibrium structure. 
134: This holds for times outside the transient regime, $t /
135: t_0 \gg 1$, or for frequencies below the band of microscopic
136: excitations, $\omega t_0 \ll 1$. The subtleties of the transient
137: dynamics like the dependence of oscillation frequencies on mass
138: ratios enter the long-time dynamics and the low-frequency
139: spectra via a common time scale $t_0$ only. This means that the
140: statistical information on the long-time dynamics is 
141: determined up to a scale $t_0$ by the statistics of the system's
142: orbits in configuration space rather than 
143: by the orbits in phase space. The glassy
144: dynamics as described by functions like $\phi_q^{s} (t)$ or 
145: $S_{q}(\omega)$ deals with the probabilities of paths through the
146: high-dimensional potential-energy landscape. The complicated dynamics on
147: microscopic time scales is irrelevant in the
148: long-time regime; it merely determines the scale $t_0$ for the
149: exploration of the configuration space. 
150: The cited results of the MCT for simple systems and mixtures
151: \cite{Goetze92,Franosch98,Fuchs99b} are not valid for the
152: mentioned theories for molecular systems
153: \cite{Schilling97,Theis98,Fabbian99b,Winkler00,Theis00,Franosch97c,Goetze00c},
154: which imply isotope effects for the glassy dynamics.
155: A change of the mass ratio of the molecule's constituents shifts
156: the center of gravity, and thereby the mode-coupling coefficients
157: are altered. This leads to shifts of the glass-transition
158: temperature, the particle's localization lengths, and the like. 
159: In this respect a system of A-B
160: molecules would behave qualitatively different than an A-B
161: mixture. There are no experimental observations
162: demanding that the long-time dynamics is independent of the
163: inertia parameters of the molecules. But we
164: consider the specified isotope effects as artifacts of the
165: approximations underlying the so far studied extensions of MCT.
166: This critique and the formidable complexity of the theories based on the
167: tensor-density descriptions appear as a motivation to search for
168: an alternative approach towards the glassy dynamics of molecular
169: systems. An alternative MCT was proposed by 
170: Kawasaki \cite{Kawasaki97}. But so far, nothing is known about the
171: solutions of his equations nor on the results concerning
172: the inertia-parameter issue. In this paper the
173: suggestion of Chong and Hirata \cite{Chong98b} will be
174: followed, and the MCT will be based on the interaction-site
175: representation of the system \cite{Chandler72,Hansen86}.
176: 
177: The description of a molecular liquid by interaction-site
178: densities is inferior to the one by tensor densities. The
179: correlators of tensor densities can be used to express
180: the ones of interaction-site densities but not vice
181: versa. Interaction-site theories have also difficulties to handle
182: reorientational correlators. Therefore, it is a major issue of
183: this paper to show, that the indicated ad hoc objections against a
184: MCT based on an interaction-site representation do not fully
185: apply if the theory is restricted to a parameter regime where the
186: cage effect is the dominant mechanism for the dynamics. To proceed,
187: the same dumbbell-molecule problem shall be studied, which was
188: analyzed previously \cite{Franosch97c,Goetze00c}. 
189: 
190: The paper is organized as follows. 
191: The basic equations for the model are introduced in Sec.~\ref{sec:2}.
192: Then, the MCT for a diatomic molecule
193: in a simple liquid is formulated in Sec.~\ref{sec:3}. The major problem is
194: the derivation of formulas for the mode-coupling coefficients.
195: This will be done within the Mori-Fujisaka formalism, and the
196: details are presented in Appendix~\ref{appen:B}. 
197: In Sec.~\ref{sec:4}, the results of
198: the theory for the dumbbell in a HSS are discussed.
199: The findings are summarized in Sec.~\ref{sec:5}.
200: 
201: \section{THE MODEL}
202: \label{sec:2}
203: 
204: A system of $N$ identical atoms distributed with density $\rho$ at
205: positions $\vec r_\kappa, \, \kappa = 1, \cdots , N$, is
206: considered as solvent. The structure can be described by the
207: density fluctuations for wave vectors $\vec q$: 
208: $\rho_ {\vec q} = \sum_\kappa \exp (i {\vec q} \cdot {\vec r}_\kappa)$. 
209: The structure factor 
210: $S_q = \langle \mid \rho_{\vec q} \mid^2 \rangle / N$
211: provides the simplest information on the equilibrium distribution
212: of these particles. Here $\langle \cdots \rangle$ denotes canonical
213: averaging for temperature $T$. Because of isotropy, $S_q$ only
214: depends on the wave-number $q = \mid \vec q \mid$. The
215: Ornstein-Zernike equation, $S_q = 1 / [1 - \rho c_q]$, relates
216: $S_q$ to the direct correlation function $c_q$. 
217: The structural dynamics is described in a statistical
218: manner by the normalized density correlators 
219: $\phi_q (t) = \langle \rho_{\vec q} (t)^*  \rho_{\vec q} \rangle / N S_q$. 
220: They are real even functions of time $t$ and exhibit the initial
221: behavior: $\phi_q (t) = 1 - \frac{1}{2} (\Omega_q t)^2 + O (|t|^3)$.
222: Here $\Omega_q = q v / \sqrt{S_q}$ is the bare phonon dispersion;
223: $v = \sqrt{k_B T/m}$ denotes the thermal velocity of the particles
224: with mass $m$ \cite{Hansen86}.
225: 
226: A rigid molecule of two atoms $A$ and $B$ shall be considered as
227: solute. Let $\vec r_a, \, a = A$ or $B$, denote the position vectors
228: of the atoms, so that $L = | \vec r_A - \vec r_B |$ denotes
229: the distance between the two interaction sites. Vector $\vec e =
230: (\vec r_A - \vec r_B) / L$ abbreviates the axis of the 
231: molecule. If $m_a$ denotes the mass of atom $a$, the total mass $M
232: = m_A + m_B$ and the moment of inertia $I = m_A m_B L^2 / M$
233: determine the thermal velocities $v_T = \sqrt{k_B T / M}$ and $v_R
234: = \sqrt{k_B T / I}$ for the molecule's translation and rotation,
235: respectively. Let us introduce also the center-of-mass position
236: $\vec r_C = (m_A \vec r_A + m_B \vec r_B) / M$ and the coordinates
237: $z_{a}$ of the atoms along the molecule axis: $z_A = L (m_B /
238: M), \, z_B = - L (m_A / M)$. The position of the molecule shall be
239: characterized by the two interaction-site-density fluctuations
240: \begin{equation}
241: \rho^{a}_{\vec q} = \exp ( i {\vec q} \cdot {\vec r_{a}} ),
242: \quad a = A \mbox{ or } B.
243: \label{eq:rho-def}
244: \end{equation}
245: The two-by-two matrix ${\bf w}_q$ of static fluctuation correlations
246: $w_q^{ab} = \langle \rho_{\vec q}^{a*} \rho_{\vec q}^b \rangle$ is
247: given by
248: \begin{equation}
249: w^{ab}_{q} = \delta^{ab} + (1 - \delta^{ab}) \, j_{0}(qL),
250: \label{eq:w-def}
251: \end{equation}
252: where here and in the following $j_\ell (x)$ denotes the spherical
253: Bessel function of index $\ell$. 
254: The solute-solvent interaction is described 
255: by the pair-correlation function 
256: $h_q^a = \langle \rho_{\vec q}^{*} \, \rho_{\vec q}^a \rangle / \rho$, 
257: which is expressed by a
258: direct correlation function $c_q^a$ \cite{Chandler72}
259: \begin{equation}
260: h^{a}_{q} = S_{q} \sum_{b} w^{ab}_{q} \, c^{b}_{q}.
261: \label{eq:huv-def}
262: \end{equation}
263: 
264: The dynamics of the molecule shall be characterized by the 
265: interaction-site-density correlators 
266: \begin{equation}
267: F^{ab}_{q}(t) = \langle \rho^{a}_{\vec q}(t)^{*} \rho^{b}_{\vec q} \rangle.
268: \label{eq:Fab-def}
269: \end{equation}
270: These are real even functions of time obeying $F_q^{ab} (t)
271: = F_q^{ba} (t)$. They shall be combined to a two-by-two-matrix
272: correlator ${\bf F}_q (t)$. Its short-time expansion can be noted
273: as
274: \begin{equation}
275: {\bf F}_{q}(t) = {\bf w}_{q} - 
276: {\textstyle \frac{1}{2}} \, q^{2} \, {\bf J}_{q} \, t^{2} +
277: {\bf O}(|t|^{3}).
278: \label{eq:Fab-short-time}
279: \end{equation}
280: The continuity equation reads 
281: $\dot \rho_{\vec q}^a = i {\vec q} \cdot {\vec j}_{\vec q}^a$, 
282: where the current fluctuation is 
283: $\vec j_{\vec q}^a = \vec v^a \rho_{\vec q}^a$ with $\vec v^a$ denoting the
284: velocity of atom $a$. Therefore, one gets 
285: $J_q^{ab} = 
286: \langle 
287: ({\vec q} \cdot {\vec j}_{\vec q}^a)^* ({\vec q} \cdot {\vec j}_{\vec q}^b) 
288: \rangle / q^2$. 
289: The result splits in a translational and rotational part, 
290: ${\bf J}_q = {\bf J}^{\rm T}_{q} + {\bf J}^{\rm R}_{q}$, 
291: where~\cite{Chong98}:
292: \begin{mathletters}
293: \label{eq:J-TR}
294: \bea
295: J^{{\rm T} \, ab}_{q} &=& v_{\rm T}^{2} \, w^{ab}_{q},
296: \label{eq:J-TR-a}
297: \\
298: J^{{\rm R} \, ab}_{q} &=& v_{\rm R}^{2} \, 
299: ( {\textstyle \frac{2}{3}} z_{a} z_{b} ) \,
300: [ \delta^{ab} + 
301: \nonumber \\
302: & & \qquad \qquad \qquad
303: (1-\delta^{ab}) \, (j_{0}(qL) + j_{2}(qL)) ].
304: \label{eq:J-TR-b}
305: \eea
306: 
307: Let us note the small-$q$ expansion of the density correlators in
308: the form
309: \end{mathletters}
310: \begin{equation}
311: F^{ab}_{q}(t) = 1 - {\textstyle \frac{1}{6}} \, q^{2} \, C^{ab}(t) + O(q^{4}). 
312: \label{eq:Fab-small-q}
313: \end{equation}
314: The diagonal elements of the symmetric matrix ${\bf C}(t)$ are the
315: mean-squared displacements
316: \begin{mathletters}
317: \label{eq:Cab-mat}
318: \begin{equation}
319: \delta r_{a}^{2}(t) = 
320: \langle ( {\vec r}_{a}(t) - {\vec r}_{a}(0) )^{2} \rangle =
321: C^{aa}(t),
322: \label{eq:Cab-mat-a}
323: \end{equation}
324: while the off-diagonal elements can be related to the
325: dipole correlator
326: \bea
327: C_{1}(t) &=& \langle {\vec e}(t) \cdot {\vec e} \, \rangle
328: \nonumber \\
329: &=&
330: [ C^{AB}(t) - {\textstyle \frac{1}{2}} ( C^{AA}(t) + C^{BB}(t) ) ] \, / \, L^{2}.
331: \label{eq:Cab-mat-b}
332: \eea
333: The mean-squared displacement of the center of mass can be
334: expressed as 
335: \bea
336: \delta r_{C}^{2}(t) &=& 
337: \langle ( {\vec r}_{C}(t) - {\vec r}_{C}(0) )^{2} \rangle
338: \nonumber \\
339: &=&
340: [ m_{A} \delta r_{A}^{2}(t) + m_{B} \delta r_{B}^{2}(t)] / M 
341: \nonumber \\
342: & &
343: \qquad \qquad \qquad 
344: + \, 
345: (2 I / M) \, [ C_{1}(t) - 1].
346: \label{eq:Cab-mat-c}
347: \eea
348: Expanding Eq.~(\ref{eq:Fab-short-time}) in $q$ yields the initial decay
349: \end{mathletters}
350: \begin{equation}
351: {\bf C}(t) = {\bf C}_{0} + 3 \, {\bf J}_{0} \, t^{2} + {\bf O}(|t|^{3}).
352: \label{eq:Cab-short-time}
353: \end{equation}
354: Here the initial value ${\bf C}_{0}$ is due to the expansion of 
355: Eq.~(\ref{eq:w-def}), 
356: while the prefactor of the $t^2$-term is due to the
357: zero-wave-number limit of Eqs.~(\ref{eq:J-TR}):
358: \begin{equation}
359: C^{ab}_{0} = L^{2} \, (1 - \delta^{ab}), \quad
360: J^{ab}_{0} = v_{\rm T}^{2} + {\textstyle \frac{2}{3}} \, 
361: v_{\rm R}^{2} \, z_{a} \, z_{b}.
362: \label{eq:Cab0-Jab0}
363: \end{equation}
364: 
365: For symmetric molecules one gets $m_{A} = m_{B} = M/2$,
366: $I = ML^{2}/4$, and $z_{A} = - z_{B} = L/2$.
367: In this case, there are only two independent density correlators,
368: since $F^{AA}_{q}(t) = F^{BB}_{q}(t)$.
369: It is convenient to perform an orthogonal transformation to fluctuations
370: of total number-densities $\rho_{N}({\vec q} \,)$ and 
371: ``charge'' densities $\rho_{Z}({\vec q} \,)$:
372: \begin{mathletters}
373: \label{eq:rho-NZ}
374: \begin{equation}
375: \rho_{x}({\vec q} \,) = 
376: ( \rho^{A}_{\vec q} \pm \rho^{B}_{\vec q} ) \, / \, \sqrt{2}, \quad
377: x = N \mbox{ or } Z.
378: \label{eq:rho-NZ-a}
379: \end{equation}
380: The transformation matrix ${\bf P} = {\bf P}^{-1}$ reads
381: \begin{equation}
382: {\bf P} = \frac{1}{\sqrt{2}}
383: \left(
384: \begin{array}{rr}
385: 1 &  1 \\
386: 1 & -1 
387: \end{array}
388: \right).
389: \label{eq:rho-NZ-b}
390: \end{equation}
391: It diagonalizes the matrices ${\bf w}_{q}$ from Eq.~(\ref{eq:w-def})
392: and ${\bf J}_{q}$ from Eqs.~(\ref{eq:J-TR}):
393: \bea
394: ( {\bf P} \, {\bf w}_{q} \, {\bf P} )^{xy} &=& 
395: \delta^{xy} \, w_{x}(q), \quad
396: w_{x}(q) = 1 \pm j_{0}(qL),
397: \label{eq:rho-NZ-c}
398: \\
399: ( {\bf P} \, {\bf J}^{\rm T}_{q} \, {\bf P} )^{xy} &=& 
400: \delta^{xy} \, v_{\rm T}^{2} \, w_{x}(q),
401: \label{eq:rho-NZ-d}
402: \\
403: ( {\bf P} \, {\bf J}^{\rm R}_{q} \, {\bf P} )^{xy} &=& 
404: \delta^{xy} \, {\textstyle \frac{1}{6}} \, v_{\rm R}^{2} \, L^{2} \,
405: [1 \mp (j_{0}(qL) + j_{2}(qL))],
406: \label{eq:rho-NZ-e}
407: \eea
408: where $x,y = N$ or $Z$.
409: \end{mathletters}
410: Also the matrix of density correlators is diagonalized.
411: Introducing the normalized correlators
412: $\phi_{q}^{x}(t)$, one gets
413: \begin{mathletters}
414: \label{eq:phi-NZ}
415: \bea
416: & &
417: \phi^{x}_{q}(t) =
418: \langle \rho_{x}({\vec q},t)^{*} \rho_{x}({\vec q}) \rangle \, / \, w_{x}(q), 
419: \\
420: & &
421: ( {\bf P} \, {\bf F}_{q}(t) \, {\bf P} )^{xy} =
422: \delta^{xy} \, \phi^{x}_{q}(t) \, w_{x}(q).
423: \eea
424: The mean-squared displacements are equal and shall be denoted by
425: $\delta r^{2}(t) = \delta r_{A}^{2}(t) = \delta r_{B}^{2}(t)$,
426: so that Eq.~(\ref{eq:Cab-mat-c}) reads
427: $\delta r^{2}(t) = \delta r_{C}^{2}(t) + 
428: (1/2) L^{2} [1 - C_{1}(t)]$.
429: \end{mathletters} 
430: The matrix ${\bf C}(t)$ is diagonalized 
431: \begin{mathletters}
432: \label{eq:Cmat-symm}
433: \bea
434: ( {\bf P} \, {\bf C}(t) \, {\bf P} )^{NN} &=&
435: 2 \delta r_{C}^{2}(t) + L^{2}, 
436: \\
437: ( {\bf P} \, {\bf C}(t) \, {\bf P} )^{ZZ} &=&
438: - L^{2} C_{1}(t). 
439: \eea
440: 
441: In Appendix A it is shown how the correlation functions in the
442: interaction-site representation can be expressed in terms of the
443: ones in the tensor-density representation.
444: \end{mathletters}
445: 
446: \section{APPROXIMATIONS}
447: \label{sec:3}
448: 
449: \subsection{The solvent-density correlator}
450: \label{subsec:MCT-v}
451: 
452: The density correlator of the solvent is needed to formulate the
453: equations for the probe molecule. This quantity is discussed
454: comprehensively in the preceding literature on the MCT for simple
455: systems~\cite{Franosch97}. 
456: Let us note here only those equations which have to
457: be solved in order to obtain the input information for
458: the calculations of the present paper. First, there is the exact
459: Zwanzig-Mori equation of motion \cite{Hansen86} relating the
460: correlator for density fluctuations $\phi_q (t)$ to the correlator
461: $m_q (t)$ for the force fluctuations:
462: \begin{equation}
463: \partial_{t}^{2} \phi_{q}(t) + \Omega_{q}^{2} \, \phi_{q}(t) +
464: \Omega_{q}^{2} \int_{0}^{t} dt' \, m_{q}(t-t') \, \partial_{t'} \phi_{q}(t') = 0.
465: \label{eq:GLE-v}
466: \end{equation}
467: Second, there is the approximate expression for kernel $m_q (t)$
468: as mode-coupling functional
469: \begin{mathletters}
470: \label{eq:MCT-v}
471: \begin{equation}
472: m_{q}(t) = {\cal F}_{q}[\phi(t)].
473: \label{eq:MCT-v-a}
474: \end{equation}
475: The functional ${\cal F}_q$ is rederived as Eq.~(\ref{eq:BXX-1}) in 
476: Appendix~\ref{appen:B}. 
477: The wave numbers are discretized to $M$ values with spacing
478: $h$: $q/h = 1/2, \, 3/2, \cdots, M-1/2$. Then $\phi (t)$
479: and similar quantities are to be viewed as vectors of $M$
480: components $\phi_q (t)$, $q=1,\cdots, M$, and the functional is 
481: \begin{equation}
482: {\cal F}_{q}[\tilde{f}] = \sum_{kp} V_{q,kp} \, \tilde{f}_{k} \, \tilde{f}_{p}.
483: \label{eq:MCT-v-b}
484: \end{equation}
485: Third, Eqs.~(\ref{eq:GLE-v}) and (\ref{eq:MCT-v-a}) imply the equation for
486: the long-time limit ${f_q = \phi_q (t \to \infty)}$:
487: \end{mathletters}
488: \begin{equation}
489: f_{q} = {\cal F}_{q}[f] \, / \, \{ 1 + {\cal F}_{q}[f] \}.
490: \label{eq:DW-v}
491: \end{equation}
492: For the liquid state, there is only the trivial solution $f_q =
493: 0$. The glass is characterized by a non-ergodicity
494: parameter $0 < f_q < 1$, which has the meaning of the Debye-Waller factor of the
495: arrested structure. At the liquid-glass transition, the long-time
496: limit of the correlator changes discontinuously from zero to the
497: critical value $f_q^c > 0$.
498: 
499: \subsection{The solute-interaction-site-density correlators}
500: \label{subsec:MCT-u}
501: 
502: For matrices of correlation functions as defined in 
503: Eq.~(\ref{eq:Fab-def}),
504: the Zwanzig-Mori formalism also leads to an exact equation of motion
505: \cite{Hansen86}:
506: \begin{mathletters}
507: \label{eq:GLE-u}
508: \begin{equation}
509: \partial_{t}^{2} {\bf F}_{q}(t) + {\bf \Omega}_{q}^{2} \, {\bf F}_{q}(t) +
510: {\bf \Omega}_{q}^{2} 
511: \int_{0}^{t} dt' \, {\bf m}_{q}(t-t') \, \partial_{t'} {\bf F}_{q}(t') = {\bf 0}.
512: \label{eq:GLE-u-a}
513: \end{equation}
514: From the short-time expansion together with Eq.~(\ref{eq:Fab-short-time}), 
515: one gets
516: \begin{equation}
517: {\bf \Omega}_{q}^{2} = q^{2} \, {\bf J}_{q} \, {\bf w}^{-1}_{q}.
518: \label{eq:GLE-u-b}
519: \end{equation}
520: The r.h.s. of this equation is a product of two symmetric positive
521: definite matrices. 
522: \end{mathletters}
523: Hence it can be written as square of a matrix
524: ${\bf \Omega}_q$.
525: Splitting off this matrix in front of the
526: convolution integral is done for later convenience.
527: 
528: The difficult problem is the derivation of an approximation for
529: the matrix ${\bf m}_{q}(t)$ of fluctuating-force correlations, so that the
530: cage effect is treated reasonably. 
531: The result, Eq.~(\ref{eq:BXX-2}) from 
532: Appendix~\ref{appen:B}, can be formulated as mode-coupling
533: functional 
534: $\mbox{\boldmath ${\cal F}$}_{q}$:
535: \begin{mathletters}
536: \label{eq:MCT-u}
537: \begin{equation}
538: m^{ab}_{q}(t) = {\cal F}^{ab}_{q}[{\bf F}(t), \phi(t)]. 
539: \label{eq:MCT-u-a}
540: \end{equation}
541: After the discretization of the wave numbers as explained above,
542: $\mbox{\boldmath ${\cal F}$}_{q}$ reads
543: \begin{equation}
544: {\cal F}^{ab}_{q}[\tilde{\textit{\textbf f}}, \tilde{f}] = q^{-2} 
545: \sum_{c} w^{ac}_{q} \, \sum_{kp} V^{cb}_{q,kp} \, 
546: \tilde{f}^{cb}_{k} \, \tilde{f}_{p}. 
547: \label{eq:MCT-u-b}
548: \end{equation}
549: The preceding equations are matrix generalizations of the MCT
550: equations for the tagged-particle-density correlator $\phi_q^s
551: (t)$ in a simple liquid \cite{Fuchs98}. 
552: \end{mathletters}
553: 
554: The equation for the non-ergodicity parameters of the molecule,
555: $F_q^{ab \, \infty} = F_q^{ab} (t \to \infty)$, can be obtained from
556: Eqs.~(\ref{eq:GLE-u-a}) and (\ref{eq:MCT-u-a}). 
557: It is a matrix generalization of Eq.~(\ref{eq:DW-v}):
558: \begin{equation}
559: {\bf F}^{\infty}_{q} = 
560: \mbox{\boldmath ${\cal F}$}_{q}[{\bf F}^{\infty},f] \,
561: \{ {\bf 1} + \mbox{\boldmath ${\cal F}$}_{q}[{\bf F}^{\infty},f] \}^{-1} \, 
562: {\bf w}_{q}.
563: \label{eq:DW-u}
564: \end{equation}
565: If the solvent is a liquid, i.e., if $f_q = 0$, one gets
566: ${\bf F}_q^\infty = {\bf 0}$. If the solvent is a glass, 
567: the long-time limits $F_q^{ab \, \infty}$ can be non trivial. In
568: this case, the solvent properties enter via the Debye-Waller
569: factors $f_q$, which renormalize the coupling coefficients 
570: $V_{q,kp}^{cb}$ in Eq.~(\ref{eq:MCT-u-b}).
571: 
572: Let us specialize to symmetric molecules. 
573: Multiplying Eqs.~(\ref{eq:GLE-u}) to (\ref{eq:DW-u}) from left and right with 
574: $\bf P$ given by Eq.~(\ref{eq:rho-NZ-b}) and inserting ${\bf 1} = {\bf P P}$ 
575: between every pair of matrices, all equations are
576: transformed to diagonal ones. Thus, there are two equations of
577: motion
578: \bea
579: & &
580: \partial_{t}^{2} \phi^{x}_{q}(t) + \Omega^{x \, 2}_{q} \, \phi^{x}_{q}(t) 
581: \nonumber \\
582: & & 
583: \qquad \qquad \qquad
584: + \,
585: \Omega^{x \, 2}_{q} 
586: \int_{0}^{t} dt' \, m^{x}_{q}(t-t') \, \partial_{t'} \phi^{x}_{q}(t') = 0, 
587: \label{eq:GLE-NZ}
588: \eea
589: for $x=N$ or $Z$. 
590: The two characteristic frequencies $\Omega_q^x$, which specify the
591: initial decay of the correlators by $\phi_q^x (t) = 1 -
592: \frac{1}{2} (\Omega_q^x t)^2 + O (|t|^3)$, read:
593: \begin{mathletters}
594: \label{eq:Omega-NZ}
595: \bea
596: \Omega^{N \, 2}_{q} &=& (v_{\rm T} q)^{2} + 
597: {\textstyle \frac{1}{6}} 
598: (v_{\rm R} L q)^{2} [1 - j_{0}(qL) - j_{2}(qL)] 
599: \nonumber \\
600: & &
601: \qquad \qquad \qquad \qquad \qquad
602: / \, 
603: [1 + j_{0}(qL)],
604: \label{eq:Omega-NZ-a}
605: \\
606: \Omega^{Z \, 2}_{q} &=& (v_{\rm T} q)^{2} +
607: {\textstyle \frac{1}{6}} 
608: v_{\rm R}^{2} [1 + j_{0}(qL) + j_{2}(qL)] (qL)^{2} 
609: \nonumber \\
610: & &
611: \qquad \qquad \qquad \qquad \qquad
612: / \, 
613: [1 - j_{0}(qL)].
614: \label{eq:Omega-NZ-b}
615: \eea
616: The relaxation kernels can be written as $m_q^x (t) = {\cal F}_q^x
617: [\phi^x (t), \phi (t)]$, where Eq.~(\ref{eq:BXX-2}) gives 
618: \end{mathletters}
619: \begin{mathletters}
620: \label{eq:MCT-NZ}
621: \bea
622: {\cal F}^{x}_{q}[\tilde{f}^{x}, \tilde{f}] &=& [w_{x}(q)/q^{2}] \,
623: \int \frac{d{\vec k}}{2 (2\pi)^{3}} \,
624: ({\vec q} \cdot {\vec p} / q)^{2} 
625: \nonumber \\
626: & &
627: \qquad \qquad
628: \times \,
629: w_{x}(k) \, \rho S_{p} \, c^{N}(p)^{2} \,
630: \tilde{f}^{x}_{k} \, {\tilde f}_{p}. 
631: \label{eq:MCT-NZ-a}
632: \eea
633: Here ${\vec p} = {\vec q} - {\vec k}$, and 
634: $c^N (p) = \sqrt{2} c^A_p = \sqrt{2} c^B_p$.
635: The above specified discretization of the wave numbers yields
636: ${\cal F}_q^x$ as polynomial
637: \begin{equation}
638: {\cal F}^{x}_{q}[\tilde{f}^{x}, \tilde{f}] = 
639: [w_{x}(q)/q^{2}] \,
640: \sum_{kp} V^{x}_{q, kp} \, \tilde{f}^{x}_{k} \, \tilde{f}_{p}.
641: \label{eq:MCT-NZ-b}
642: \end{equation}
643: One gets for the non-ergodicity parameters $f_q^x = \phi_q^x
644: (t \to \infty) = (F_q^{AA \, \infty} \pm F_q^{AB \, \infty}) / w_x (q)$
645: \end{mathletters}
646: \begin{equation}
647: f^{x}_{q} = {\cal F}^{x}_{q}[f^x, f] \, / \, 
648: \{ 1 + {\cal F}^{x}_{q}[f^x, f] \}.
649: \label{eq:DW-NZ}
650: \end{equation}
651: 
652: There is no coupling between the fluctuations of the total density
653: and the ones of the ``charge'' density. The mathematical structure
654: of the two sets of equations for $x = N$ and $x = Z$, respectively,
655: is the same as the one studied previously for the density correlator
656: $\phi_q^s (t)$ of a tagged particle in a simple liquid
657: \cite{Fuchs98}. For the density dynamics one also finds the
658: small-$q$ asymptote for the frequency
659: $\Omega_q^{N \, 2} = (v_{\rm T} q)^2 + O (q^4)$, 
660: reflecting free translation of the probe molecule.
661: There also is the $q^{-2}$-divergency of the mode-coupling
662: coefficients in ${\cal F}_q^N$, which implies the approach towards
663: unity of the Lamb-M\"ossbauer factor for vanishing wave number:
664: $f_{q \to 0}^N = 1$. For the ``charge'' dynamics, one gets 
665: a non-zero small-$q$ limit for the frequency 
666: characterizing free rotation 
667: $\Omega_{q \to 0}^{Z \, 2} = 2 v_{\rm R}^2 + O(q^2)$. 
668: The mode-coupling coefficients do not diverge for $q \to 0$, 
669: since $6 w_Z (q) / (L q )^2 \to 1$. 
670: Therefore, the non-ergodicity parameter for the variable 
671: $\rho_Z ({\vec q},t)$ approaches a limit smaller than unity:
672: $f^{Z}_{q \to 0} < 1$.
673: 
674: \subsection{The dipole correlator and the mean-squared displacements}
675: \label{subsec:MCT-C1-MSD}
676: 
677: According to Eqs.~(\ref{eq:Cab-mat}), the knowledge of the dipole correlator 
678: $C_1(t)$ and two of the mean-squared displacements $\delta r_a^2 (t)$ for $a
679: = A, B$, or $C$ is equivalent to the knowledge of the three
680: independent elements of the symmetric matrix ${\bf C} (t)$. Using
681: Eq.~(\ref{eq:Fab-small-q}) and expanding 
682: Eq.~(\ref{eq:GLE-u-a}) for small wave numbers one gets
683: \begin{equation}
684: \partial_{t}^{2} {\bf C}(t) - {\bf D} + {\bf \Omega}^{2}_{0} {\bf C}(t) +
685: {\bf J}_{0} 
686: \int_{0}^{t} dt' {\bf m}(t-t') \partial_{t'} {\bf C}(t') = {\bf 0}.
687: \label{eq:GLE-Cmat}
688: \end{equation}
689: This exact equation of motion for ${\bf C} (t)$ has to be solved
690: with the initial condition from Eq.~(\ref{eq:Cab-short-time}). 
691: The frequency matrix is obtained as zero-wave-number limit from 
692: Eq.~(\ref{eq:GLE-u-b}):
693: \begin{equation}
694: {\bf \Omega}^{2}_{0} = (2 v_{\rm R}^{2} / L) \, 
695: \left(
696: \begin{array}{rr}
697: z_{A} & - z_{A} \\
698: z_{B} & - z_{B} 
699: \end{array}
700: \right).
701: \label{eq:Omega-Cmat}
702: \end{equation}
703: Equation~(\ref{eq:GLE-Cmat}) implies 
704: $\ddot{\bf C}(0) - {\bf D} + {\bf \Omega}_0^2 {\bf C} (0) = {\bf 0}$. 
705: Thus, one gets from Eq.~(\ref{eq:Cab-short-time}): 
706: ${\bf D} = 6 {\bf J}_0 + {\bf \Omega}_0^2 {\bf C}_0$, i.e., 
707: $D^{ab} = 6 \, v_{\rm T}^{2} + 2 \, v_{\rm R}^{2} \, (z_{A} + z_{B}) \, z_{a}$.
708: 
709: The MCT approximation for the kernel ${\bf m}(t)$ is obtained by combining
710: Eqs.~(\ref{eq:GLE-u-b}) and (\ref{eq:MCT-u-b}) and 
711: taking the zero-wave-vector limit. 
712: With Eq.~(\ref{eq:BXX-2}), one finds
713: \begin{mathletters}
714: \label{eq:MCT-Cmat}
715: \bea
716: {\bf m}(t) &=& \mbox{\boldmath ${\cal F}$}[{\bf F}(t), \phi(t)],
717: \label{eq:MCT-Cmat-a}
718: \\
719: {\cal F}^{ab}[\tilde{\textit{\textbf f}}, \tilde{f}] &=& \frac{1}{6 \pi^{2}}
720: \int_{0}^{\infty} dk \, k^{4} \, \rho S_{k} \, c^{a}_{k} \, c^{b}_k \, 
721: \tilde{f}^{ab}_{k} \, \tilde{f}_{k}.
722: \label{eq:MCT-Cmat-b}
723: \eea
724: 
725: Again, the theory simplifies considerably for symmetric molecules.
726: In this case, one can transform Eq.~(\ref{eq:GLE-Cmat}) as explained in
727: connection with the derivation of Eq.~(\ref{eq:GLE-NZ}). 
728: \end{mathletters}
729: Using Eqs.~(\ref{eq:Cmat-symm}) one
730: gets the exact equation of motion for the mean-squared
731: displacement 
732: \begin{equation}
733: \partial_{t}^{2} \delta r^{2}_{C}(t) - 6 v_{\rm T}^{2} + v_{\rm T}^{2}
734: \int_{0}^{t} dt' \, m_{N}(t-t') \, 
735: \partial_{t'} \delta r^{2}_{C}(t') = 0,
736: \label{eq:GLE-MSD}
737: \end{equation}
738: to be solved with the initial behavior 
739: $\delta r_{C}^{2} (t) = 3 (v_{\rm T} t)^2 + O (|t|^3)$. 
740: Similarly, one obtains for the dipole correlator
741: \bea
742: & &
743: \partial_{t}^{2} C_{1}(t) + 2 v_{\rm R}^{2} \, C_{1}(t) 
744: \nonumber \\
745: & & 
746: \qquad \qquad
747: + \, 
748: 2 v_{\rm R}^{2}
749: \int_{0}^{t} dt' \, m_{Z}(t-t') \, \partial_{t'} C_{1}(t') = 0,
750: \label{eq:GLE-C1}
751: \eea
752: to be solved with the initial decay 
753: $C_1 (t) = 1 - (v_{\rm R} t)^2 + O(|t|^3)$. 
754: The MCT approximation for the kernels is obtained from 
755: Eq.~(\ref{eq:MCT-Cmat-b}):
756: \begin{mathletters}
757: \label{eq:MCT-MSD-C1}
758: \bea
759: m_{x}(t) &=& {\cal F}_{x}[\phi^{x}(t),\phi(t)], \quad
760: x = N \mbox{ or } Z,
761: \label{eq:MCT-MSD-C1-a}
762: \\
763: {\cal F}_{x}[\tilde{f}^{x}, \tilde{f}] &=& \alpha_{x} 
764: \int_{0}^{\infty} dk \,
765: k^{4} \, \rho S_{k} \, c^{N}(k)^{2} \, w_{x}(k) \, \tilde{f}^{x}_{k} \, \tilde{f}_{k},
766: \label{eq:MCT-MSD-C1-b}
767: \eea
768: where $\alpha_N = 1 / (6 \pi^2)$ and $\alpha_Z = L^2 / (72\pi^2)$. 
769: \end{mathletters}
770: Equations~(\ref{eq:GLE-C1}) and (\ref{eq:MCT-MSD-C1}) for the dipole correlator
771: have the standard form of the MCT equation. If $\phi_q^Z (t)$
772: approaches zero for large times, the same approach towards
773: equilibrium is exhibited by $C_1 (t)$. If the solvent is a glass,
774: $f_q > 0$, and if the ``charge''-density fluctuations
775: $\phi_q^Z (t)$ exhibit non-ergodic behavior, $f_q^Z > 0$, also the
776: $\ell = 1$--reorientational correlator exhibits non-ergodic
777: dynamics:
778: \bea
779: C_{1}(t \to \infty) = f_{1} = 
780: {\cal F}_{Z}[f^{Z}, f] \, / \, \{ 1 + {\cal F}_{Z}[f^{Z}, f] \}.
781: \label{eq:f1-symm}
782: \eea
783: Parameter $f_{1}$ is the long-wave-length limit of $f_{q}^{Z}$
784: discussed in Eq.~(\ref{eq:DW-NZ}):
785: $f^{Z}_{q \to 0} = f_{1}$.
786: 
787: \subsection{The quadrupole correlator}
788: \label{subsec:MCT-C2}
789: 
790: The quadrupole correlator 
791: $C_2 (t) = \langle 3 [{\vec e}(t) \cdot {\vec e}\,]^{2} - 1 \rangle / 2$ 
792: cannot be extracted from the correlators $F_q^{a b} (t)$ with $a, b
793: = A$ or $B$. But let us consider a linear symmetric triatomic
794: molecule. 
795: The third atom, labeled $C$, has its position in the
796: center $\vec r_C$. The preceding theory can be extended by
797: adding as third variable the fluctuations $\rho_{\vec q}^C = \exp
798: (i \vec q \cdot \vec r_C)$. The basic quantities are now the 
799: elements of the 3-by-3 matrix correlator, defined as in 
800: Eq.~(\ref{eq:Fab-def}) with $a,b = A, B$ or $C$. 
801: The correlator formed with
802: $\rho_{Q}({\vec q}) = 
803: \rho_{\vec q}^A + \rho_{\vec q}^B - 2 \rho_{\vec q}^C$ 
804: is a linear combination of the nine functions $F_q^{ab} (t)$. 
805: An expansion for small $q$ yields
806: \begin{equation}
807: \langle \rho_{Q}({\vec q},t)^{*} \rho_{Q}({\vec q}) \rangle =
808: {\textstyle \frac{1}{180}} \, (qL)^{4} \, 
809: [C_{2}(t) + {\textstyle \frac{5}{4}}] + O(q^{6}).
810: \label{eq:C2-ABC}
811: \end{equation}
812: In this case, $C_2 (t)$ can be obtained in a similar manner as
813: discussed above for $C_1 (t)$. A diatomic molecule can be
814: considered as a special mathematical limit of a triatomic one.
815: Hence, there is no problem in principle to obtain $C_2 (t)$
816: within a theory based on an interaction-site description.
817: Motivated by  this observation, an auxiliary site $C$ shall be
818: introduced \cite{Hoye77,Sullivan81} 
819: and $\rho_{\vec q}^C$ will be used as third
820: basic variable. 
821: However, a complete theory with 3-by-3 matrices shall not
822: be developed. Rather some additional approximations will be
823: introduced so that $C_2 (t)$ is obtained
824: as corollary of the above formulated closed theory.
825: 
826: The quadrupole correlator can be written as small-$q$ limit of a
827: correlation function formed with tensor-density fluctuations
828: defined in Eq.~(\ref{eq:rho-tensor-def}) for 
829: ${\vec q}_{0}=(0,0,q)$:
830: $C_2 (t) = \lim_{q \to 0} 
831: \langle \rho_2^0 ({\vec q}_0, t)^* \rho_2^0 ({\vec q}_0) \rangle$.
832: Therefore, an exact Zwanzig-Mori equation can be derived as usual:
833: \bea
834: & &
835: \partial_{t}^{2} C_{2}(t) + 6 v_{\rm R}^{2} \, C_{2}(t) 
836: \nonumber \\
837: & & 
838: \qquad \qquad
839: + \, 
840: 6 v_{\rm R}^{2}
841: \int_{0}^{t} dt' \, m_{2}^{\rm R}(t-t') \, \partial_{t'} C_{2}(t') = 0.
842: \label{eq:GLE-C2}
843: \eea
844: The relaxation kernel $m_2^{\rm R} (t)$ is a correlator for fluctuating
845: forces $F_{{\rm R} 2} (q 0, t)$ referring to angular-momentum index
846: $\ell = 2$ and helicity $m = 0$:
847: \begin{equation}
848: m_{2}^{\rm R}(t) = \lim_{q \to 0} 
849: \langle F_{{\rm R} 2}(q0,t)^{*} F_{{\rm R} 2}(q0) \rangle.
850: \label{eq:m2-def}
851: \end{equation}
852: The time evolution of the fluctuating force is generated by the
853: reduced Liouvillian ${\cal L}^\prime = {\cal Q L Q}$, where $\cal
854: Q$ projects perpendicular to $\rho_2^0 (\vec q_0)$ and ${\cal L}
855: \rho_2^0 (\vec q_0)$, and the Liouvillian ${\cal L}$ is
856: defined by $i {\cal L} A (t) = \partial_t A (t)$.
857: The involved notation has been chosen in order to get the formulas
858: in agreement with the ones of the more general theory in 
859: Ref.~\cite{Franosch97c}. The procedure used for the theory of simple
860: liquids \cite{Goetze91b} shall be applied to derive an approximation for
861: the kernel. First, the forces will be approximated by the
862: projection onto the space of the simplest modes contributing 
863: $F_{{\rm R} 2} \to {\cal P}' F_{{\rm R} 2}$. 
864: Here ${\cal P}'$ projects onto the space
865: spanned by the pair modes
866: \begin{mathletters}
867: \label{eq:m2-MCT}
868: \begin{equation}
869: A^{a}({\vec k}, {\vec p}) = \rho^{a}_{\vec k} \, \rho_{\vec p} \, / 
870: \sqrt{N S_{p}}, \quad
871: a = A, B \mbox{ or } C.
872: \label{eq:m2-MCT-a}
873: \end{equation}
874: The essential step is the second one, where correlations of the
875: pairs are replaced by products of correlations: 
876: $\langle A^{a} ({\vec k}, {\vec p}, t)^* A^{a'} ({\vec k}', {\vec p} \, ') \rangle \to 
877: \delta_{{\vec k} {\vec k}'} \delta_{{\vec p} {\vec p} \, '}
878: \langle \rho_{\vec k}^a (t)^{*} \rho_{\vec k}^{a'} \rangle \, 
879: \langle \rho_{\vec p}(t)^{*}    \rho_{\vec p} \rangle / NS_{p}$. 
880: This approximation is
881: done in particular for $t = 0$, thereby getting for the
882: normalization matrix for the pair modes: 
883: $\langle A^{a} ({\vec k}, {\vec p})^* A^{b} ({\vec k}', {\vec p} \, ') \rangle \to 
884: \delta_{{\vec k} {\vec k}'} \, \delta_{{\vec p} {\vec p} \, '} \, 
885: w_k^{ab}$. 
886: Here $w^{ab} (k) = j_0 (k (z_a - z_b))$ 
887: generalizes Eq.~(\ref{eq:w-def}) to a 3-by-3 matrix.
888: As a result, one gets
889: \bea
890: m^{\rm R}_{2}(t) &=& \lim_{q \to 0} \sum_{{\vec k} {\vec p}}
891: \sum_{abcd} \langle F_{{\rm R} 2}(q0)^{*} A^{a}({\vec k},{\vec p} \, ) \rangle \,
892: ({\bf w}_{k}^{-1})^{ab} 
893: \nonumber \\
894: &\times&
895: F_{k}^{bc}(t) \, \phi_{p}(t) \,
896: ({\bf w}_{k}^{-1})^{cd} \, 
897: \langle A^{d}({\vec k},{\vec p} \, )^{*} F_{{\rm R} 2}(q0) \rangle.
898: \label{eq:m2-MCT-b}
899: \eea
900: The $q \to 0$ limit is carried out easily, reducing the sum over
901: $\vec k$ and $\vec p$ to the one over $\vec k$ with ${\vec k} = - {\vec p}$.
902: \end{mathletters}
903: One obtains the kernel as mode-coupling functional
904: \begin{equation}
905: m^{\rm R}_{2}(t) = \int_{0}^{\infty} dk 
906: \sum_{ab \in \{A, B, C\}} 
907: V_{2}^{ab}(k) \, F^{ab}_{k}(t) \, \phi_{k}(t). 
908: \label{eq:m2-final}
909: \end{equation}
910: Let us restrict the discussion to symmetric molecules.
911: For this case, an explicit expression for $V_2^{a b} (k)$ is noted as 
912: Eqs.~(\ref{eq:m2-exp}) in Appendix~\ref{appen:A}. 
913: 
914: The correlators $\phi_k (t)$ and $F_k^{a b} (t)$ with $a, b = A$
915: or $B$ are taken from Sec.~\ref{subsec:MCT-v} and Sec.~\ref{subsec:MCT-u}, 
916: respectively.
917: The theory of Sec.~\ref{subsec:MCT-C1-MSD} provides the
918: results for the mean-squared displacement $\delta r_C^2 (t)$. The
919: Gaussian approximation shall be used to evaluate 
920: $F_k^{C C} (t) \approx \exp [- \frac{1}{6} k^2 \delta r_C^2 (t)]$. 
921: The two remaining functions can be expressed in terms of tensor-density
922: fluctuations according to Eq.~({\ref{eq:Fab-site-tensor}): 
923: $F_k^{a C} (t) = \sum_\ell
924: \sqrt{(2 \ell + 1)} j_\ell (k z_a) \phi_{\ell 0} (k 0, t)$
925: for $a = A$ or $B$. 
926: As in the previous work \cite{Goetze00c}, only the diagonal correlators
927: shall be taken in the sum, i.e., 
928: the approximation will be used: 
929: $F_k^{a C} (t) \approx j_{0} (k z_{a}) F_k^{C C} (t)$. 
930: 
931: \section{RESULTS}
932: \label{sec:4}
933: 
934: A few concepts shall be mentioned
935: which were introduced \cite{Goetze91b} to describe the
936: MCT-liquid-glass-transition dynamics. In the space of control
937: parameters, a smooth function $\sigma$ is defined near the
938: transition points, called the separation parameter. Glass states
939: are characterized by $\sigma > 0$, liquid states by $\sigma < 0$,
940: and $\sigma = 0$ defines the transition hypersurface.
941: Suppose, only one control parameter, say the density $\rho$, is
942: varied near the transition point. Then one can write for small
943: distance parameters $\epsilon = (\rho - \rho_c) /\rho_c : \sigma =
944: C \epsilon$, $C > 0$. In addition to $C$, the transition point is
945: characterized by a time scale $t_0$ and by a number $\lambda$, 
946: $0 < \lambda < 1$.
947: The scale $t_0$ specifies properties of the transient dynamics,
948: and $\lambda$ is called the exponent parameter. 
949: The latter determines a certain
950: number $B > 0$, the critical exponent $a, \, 0 < a \leq 1/2$, and
951: the von-Schweidler exponent $b, \, 0 < b \leq 1$. There are two
952: critical time scales governing the 
953: bifurcation dynamics close to the transition:
954: \begin{equation}
955: t_{\sigma} = t_{0} \, / \, | \sigma |^{\delta}, \quad
956: t'_{\sigma} = t_{0} \, B^{-1/b} \, / \, | \sigma |^{\gamma}.
957: \label{eq:t-sigma}
958: \end{equation}
959: The anomalous exponents of the scales read: $\delta = 1/2 a, \,
960: \gamma = 1/2a + 1/2 b$. The hard-sphere-system (HSS) shall be
961: used as solvent. There is only one control parameter for the
962: equilibrium structure, which shall be chosen as the packing
963: fraction $\varphi$ of the particles with diameter $d, \, \varphi =
964: \pi \rho d^3/6$. The distance parameter shall be given by the
965: logarithm $x$ of $\mid \epsilon \mid$:
966: \begin{equation}
967: \epsilon = (\varphi - \varphi_{c}) \, / \, \varphi_{c} = \pm 10^{-x}.
968: \label{eq:epsilon-def}
969: \end{equation}
970: The structure factor $S_q$ is calculated within the Percus-Yevick
971: theory \cite{Hansen86}. The wave numbers are discretized to $M =
972: 100$ values with spacing $hd = 0.4$. For this solvent model,
973: results for the density correlators and their
974: spectra can be found in Ref.~\cite{Goetze00}. The glassy dynamics
975: is analyzed in Ref.~\cite{Franosch97}, from which one infers: $\varphi_c =
976: 0.516, \, C = 1.54, \, \lambda = 0.735,\, a = 0.312, \, b= 0.583$
977: and $B = 0.836$. 
978: Furthermore,  $t_0 = 0.0236 (d/v)$~\cite{Franosch98}.
979: 
980: Dumbbells of two fused hard spheres of diameters $d_A = d_B = d$
981: shall be used as solute. The elongation parameter 
982: $\zeta = L / d$ quantifies the bond length. The 
983: solute-solvent-direct-correlation functions are also
984: calculated within the Percus-Yevick theory. Within the
985: tensor-density description, the direct correlation functions
986: $c_\ell (q)$ have been determined in Ref.~\cite{Franosch97b}.
987: These results are substituted in the formulas of Appendix A, to
988: evaluate the equilibrium structure in the site representation. In all
989: summations over contributions due to various angular-momentum
990: indices $\ell$, a cutoff $\ell_{co} = 8$ is chosen. It was checked
991: for representative cases, that increasing the cutoff to $\ell_{co}
992: = 16$ does not significantly change the results to be discussed.
993: The discretization of the various wave-vector integrals is done
994: as specified above for the solvent. 
995: The results in Secs.~\ref{subsec:result-form-factors} and 
996: \ref{subsec:result-correlators} deal with a symmetric dumbbell
997: with $m_A = m_B = m$, and in Sec.~\ref{subsec:result-inertia}, 
998: the molecule
999: with $m_A = 10 m$, $m_B = m$ is considered. 
1000: 
1001: Throughout the rest of this paper, the particle diameter is chosen
1002: as unit of length, $d = 1$, and the unit of time is chosen so that
1003: the thermal velocity of the solvent is $v = 1$.
1004: 
1005: \subsection{Structural arrest}
1006: \label{subsec:result-form-factors}
1007: 
1008: There are two control parameters for the system, 
1009: namely the packing fraction $\varphi$ of the solvent and the
1010: elongation $\zeta$ of the solute molecule. Figure~\ref{fig:phase} 
1011: exhibits the phase diagram. 
1012: Phase I deals with states where $\varphi$ is
1013: below the critical value $\varphi_c$, i.e., the solvent is a liquid.
1014: In this case, the long-time limits of the mode-coupling
1015: kernels in Eqs.~(\ref{eq:MCT-u}) vanish. 
1016: All solute correlators relax to zero for long times, 
1017: and the molecule diffuses through the solvent.
1018: For $\varphi \geq \varphi_c$, the solvent is a glass. 
1019: Structural fluctuations behave nonergodically. In particular, a
1020: tagged solvent particle does not diffuse; rather it is localized.
1021: Since the atoms of the molecule with $d_{A} = d_{B} = d$ 
1022: experience the same interaction
1023: with the solvent as the solvent particles among each other, one
1024: expects the molecule to be localized as well. Indeed, Eq.~(\ref{eq:DW-NZ}) 
1025: yields for $\varphi \geq \varphi_c : f_q^N > 0$. If $\varphi$
1026: increases from below $\varphi_c$ to above $\varphi_c$, the
1027: long-time limit $\phi_q^N (t \to \infty)$ increases
1028: discontinuously at $\varphi_c$ from zero to $f_q^{Nc} > 0$. 
1029: Also, the quadrupole correlator exhibits
1030: nonergodic dynamics: $C_2 (t \to \infty) = f_2 > 0$. The cages
1031: surrounding the molecule cause such strong steric hindrance, that
1032: quadrupole fluctuations of the orientational vector $\vec e$
1033: cannot relax to zero. In this sense, the states $\varphi \geq
1034: \varphi_c$ are ideal glasses. 
1035: 
1036: There are two alternatives for the glass. Phase II deals
1037: with states for sufficiently small $\zeta$. There is
1038: such small steric hindrance for a flip of the molecule's axis
1039: between the two energetically equivalent positions 
1040: $\vec e$ and $- \vec e$ that Eq.~(\ref{eq:DW-NZ}) yields 
1041: $f_q^Z = 0$. The dynamics of the ``charge'' fluctuations is ergodic. In
1042: particular, the dipole correlator relaxes to zero: $C_1 (t \to
1043: \infty) = 0$. Phase II is an amorphous counterpart of a
1044: plastic crystal.
1045: For sufficiently large $\zeta$, steric
1046: hindrance for dipole reorientations is so effective, that also the
1047: ``charge'' fluctuations behave nonergodically. In this case, 
1048: Eq.~(\ref{eq:DW-NZ}) yields a positive long-time limit 
1049: $0 < f_q^Z = \phi_q^Z (t \to \infty)$. 
1050: In particular, dipole-disturbances do not relax to
1051: zero: $C_1 (t \to \infty) = f_1 > 0$. This phase III is a glass
1052: with all structural disturbances exhibiting nonergodic motion. 
1053: The two phases II and III are separated by transitions at 
1054: $\zeta = \zeta_c (\varphi)$, $\varphi \geq \varphi_c$. 
1055: With decreasing density,
1056: the steric hindrance for reorientations decreases. Thus, $\zeta_c
1057: (\varphi)$ increases with decreasing $\varphi$, as shown by the
1058: full line in Fig.~\ref{fig:phase}. The transition curve terminates with
1059: horizontal slope at the largest critical elongation  $\zeta_c =
1060: \zeta (\varphi_c) = 0.380$. 
1061: Function $\zeta_{c}(\varphi)$ was
1062: calculated before within the MCT based on the tensor-density
1063: description \cite{Franosch98b}, and the
1064: transition curve of this theory is added as dashed line in 
1065: Fig.~\ref{fig:phase}.
1066: The results of the two theories are in qualitative agreement.
1067: It would be interesting if molecular-dynamics studies would
1068: decide, which of the two theories leads closer to reality.
1069: The asymptotic laws for the transition from phase II to phase III
1070: have earlier been described as type-$A$ transition as
1071: can be inferred from Ref.~\cite{Franosch94} and the papers quoted
1072: there. At this transition, $C_1 (t \to \infty)$
1073: increases continuously with increasing $\zeta$.
1074: 
1075: The heavy full lines in Fig.~\ref{fig:fq-strong} exhibit critical nonergodicity
1076: parameters $f_q^{x \, c}$ for $\zeta = 0.80$, calculated from 
1077: Eq.~(\ref{eq:DW-NZ})
1078: for the liquid-glass transition point $\varphi = \varphi_c$.
1079: These quantities are Lamb-M\"ossbauer factors of
1080: the molecule. The function $f_{q}^{Nc}$ 
1081: can be measured, in principle, as cross section for incoherent
1082: neutron scattering from the solute, provided both centers $A$ and
1083: $B$ are identical atoms without spin. As expected for a localized
1084: probability-distribution Fourier transform, the
1085: $f_q^{x \, c}$--versus--$q$ curves decrease with increasing $q$. Most
1086: remarkable are the kinks exhibited by $f_q^{Nc}$ for wave numbers $q$
1087: near 5, 12.5 and 20, and by $f_q^{Zc}$ for $q$ near 10 and 17.5.
1088: The light full lines exhibit $f_{q}^{x \, c}$ calculated with 
1089: Eq.~(\ref{eq:Fab-site-tensor-2}) from
1090: the critical nonergodicity parameters $f^c (q \ell 0)$~\cite{Franosch97c}. 
1091: The results of both approximation theories are in semi-quantitative agreement,
1092: in particular concerning the position and size of the kinks. The
1093: $f^c (q \ell 0)$--versus--$q$ curves are bell shaped, close to
1094: Gaussians \cite{Franosch97c}. They enter Eq.~(\ref{eq:Fab-site-tensor-2}) 
1095: with prefactors $j_\ell (q \zeta /2)^2 = O (q^{2 \ell})$, so that the maximum of
1096: the contribution from angular-momentum-index $\ell$ occurs at some 
1097: $q^{\max}_{\ell}$ which increases with $\ell$. The separate contributions
1098: for different $\ell$ are shown as dotted lines in 
1099: Fig.~\ref{fig:fq-strong}. Thus,
1100: the kinks are due to interference effects of the  $f^c (q \ell 0)$
1101: with the intramolecular form factors $j_\ell (q \zeta / 2)$. 
1102: Let us add, that also the Lamb-M\"ossbauer factors of the atoms,
1103: $f_{q}^{a \, c}$, are well described by Gaussians for $q<10$;
1104: in particular these functions do not exhibit kinks.
1105: Figure~\ref{fig:fq-strong} demonstrates for a case of strong steric hindrance for
1106: reorientational motion that angular-momentum variables for $\ell$
1107: up to 6 are relevant to describe the arrested structure, and
1108: that the description of the molecule by site-density
1109: fluctuations properly accounts for the contributions with 
1110: $\ell \geq 2$.
1111: 
1112: Figure~\ref{fig:fq-weak} exhibits $f_{q}^{x \, c}$
1113: representative for weak
1114: steric hindrance for the reorientational dynamics. Naturally, the
1115: contributions due to the arrest of fluctuations of tensor densities
1116: with large $\ell$ are suppressed. The contributions for $\ell = 0$
1117: and $\ell = 2$ are sufficient to explain $f_q^{Nc}$, in particular
1118: its kink for $q$ near 12.5. Similarly, the contributions for $\ell
1119: = 1$ and $\ell = 3$ are necessary and sufficient to explain
1120: $f_q^{Zc}$ with its kink for $q$ near 17.5. 
1121: The dynamics is strongly influenced by precursor phenomena of the
1122: transition from phase II to phase III. This is demonstrated, for
1123: example, by the strong decrease of $f_1^c = f_{q \to 0}^{Zc}$ for
1124: the result shown in the lower panel of Fig.~\ref{fig:fq-weak} in comparison to the
1125: one shown in the lower panel of Fig.~\ref{fig:fq-strong}. The two
1126: approximation theories under discussion yield different numbers
1127: for the value $\zeta_c$ for the transition point.
1128: It is meaningless to compare different
1129: approximations for results near a singularity $\zeta_c$, referring
1130: to the same value $\zeta$. 
1131: It is more meaningful to compare results for
1132: the same relative distance from the critical point, 
1133: $(\zeta - \zeta_c) /\zeta_c$, as is done in Fig.~\ref{fig:fq-weak}. 
1134: Let us mention that $f_{q}^{Zc}$ shown by the heavy and light full lines would be
1135: somewhat closer, if one had compared elongations yielding the same
1136: value for $f_1^c$.
1137: 
1138: Figure~\ref{fig:fq-vs-zeta} exhibits critical Lamb-M\"ossbauer factors 
1139: $f_q^{x \, c}$ as function of the
1140: elongation. The lower panel demonstrates the transition from phase
1141: II for $\zeta < \zeta_c$ to phase III for $\zeta > \zeta_c$. 
1142: For strong steric
1143: hindrance, say $\zeta \geq 0.8$, $f_q^{Nc}$ is rather close to
1144: $f_q^{Zc}$ provided $q$ is not too small, say $q > 3$. 
1145: For $\zeta$ approaching $\zeta_c$, the $f_q^{Zc}$ fall below
1146: $f_q^{Nc}$. Most remarkable are the wiggles or even minima of the curves.
1147: These are the analogues of the kinks, discussed above in connection with 
1148: Figs.~\ref{fig:fq-strong} and \ref{fig:fq-weak}. 
1149: Again, these anomalies can be explained as interference
1150: effects between the geometric structure factors 
1151: $j_\ell (q \zeta /2)^2$ and the nonergodicity parameters $f^c (q \ell 0)$ 
1152: according to Eq.~(\ref{eq:Fab-site-tensor-2}). 
1153: Let us consider $f_q^{Nc}$ for
1154: an intermediate wave vector as shown for curves $b$ or $c$ in the
1155: upper panel of Fig.~\ref{fig:fq-vs-zeta}. For small $\zeta$, say $\zeta \leq 0.4$,
1156: the $\ell = 0$ contribution dominates the sum in 
1157: Eq.~(\ref{eq:Fab-site-tensor-2}), as
1158: can be inferred from Fig.~\ref{fig:fq-weak}. Function $f^c (q 0 0)$ reflects the
1159: isotropic part of the arrested fluctuations, and hence it is
1160: practically $\zeta$-independent 
1161: (as shown in Fig.~5 of Ref.~\cite{Franosch97c}). 
1162: Since $j_0 (q \zeta /2)^2 = 1 - \frac{1}{12}
1163: (q \zeta)^2 + O ((q \zeta)^4)$ decreases with increasing $\zeta$,
1164: the $f_q^{Nc}$--versus--$\zeta$ curve decreases too; and the
1165: decrease is stronger for larger $q$. The $f^c (q 2 0)$
1166: increase from 0 for $\zeta = 0$ to values near 0.5 for 
1167: $\zeta = 1$ 
1168: (as shown in Fig.~5 of Ref.~\cite{Franosch97c}). 
1169: Also, the geometric
1170: structure factor increases strongly with $\zeta: j_2 (q \zeta
1171: /2)^2 = ((q \zeta)^2 /60)^2 + O ((q \zeta)^6)$. The combined
1172: effect of both increases causes the increase of the
1173: $f_q^{Nc}$--versus--$\zeta$ curve for larger $\zeta$. The
1174: resulting minimum occurs for smaller $\zeta$ if $q$ is larger, 
1175: and this explains the difference between 
1176: the two curves $b$ and $c$.
1177: The theory produces the minima since it 
1178: accounts for the arrest of tensor-density fluctuations for $\ell
1179: \geq 2$.
1180: 
1181: \subsection{Correlation functions and spectra near the glass transition}
1182: \label{subsec:result-correlators}
1183: 
1184: Figure~\ref{fig:NN-ZZ-t} demonstrates the evolution of the dynamics for the
1185: correlators $\phi_q^N (t)$ and $\phi_q^Z (t)$
1186: for intermediate wave numbers $q$ near the transition from phase I
1187: to phase III. The oscillatory transient dynamics occurs within the
1188: short-time window $t < 1$. The control-parameter sensitive
1189: glassy dynamics occurs for longer times for packing
1190: fractions $\varphi$ near $\varphi_c$. At the
1191: transition point $\varphi = \varphi_c$, the correlators decrease
1192: in a stretched manner towards the plateau values $f_{q}^{x \, c}$ 
1193: as shown by the dotted lines. Increasing $\varphi$
1194: above $\varphi_c$, the long-time limits increase, as shown for
1195: $\phi_q^Z (t)$ for $q = 7.4$ and $\zeta = 0.80$. Decreasing
1196: $\varphi$ below $\varphi_c$, the correlators cross the plateau at
1197: some time $\tau_\beta$, and then decay towards zero. The
1198: decay from the plateau $f_q^{x \, c}$ to zero is the $\alpha$-process
1199: for $\phi_q^x (t)$. It is characterized by a time scale
1200: $\tau_\alpha$, which can be defined, e.g., by 
1201: $\phi_q^x (\tau_\alpha) = f_q^{x \, c} / 2$. 
1202: Upon decreasing $\varphi_c - \varphi$ towards zero, 
1203: the time scales $\tau_\beta$ and
1204: $\tau_\alpha$ increase towards infinity proportional to $t_\sigma$
1205: and $t_\sigma^\prime$, respectively, cited in 
1206: Eq.~(\ref{eq:t-sigma}). The figure
1207: exemplifies the standard MCT-bifurcation scenario. For small
1208: $\mid \varphi - \varphi_c \mid$, the results can be described
1209: in terms of scaling laws. This was explained in
1210: Refs.~\cite{Franosch97,Fuchs98} for the HSS, 
1211: and the discussion shall not be
1212: repeated here. 
1213: 
1214: One can deduce from Fig.~\ref{fig:fq-strong} 
1215: that for $\zeta = 0.80$ and $q \geq 5$
1216: the plateaus for both types of density fluctuations are very close
1217: to each other: $f_q^{Nc} \approx f_q^{Zc}$. The upper two panels
1218: of Fig.~\ref{fig:NN-ZZ-t} demonstrate that also the dynamics is nearly the same,
1219: $\phi_q^{N} (t) \approx \phi_q^{Z} (t)$. This means that
1220: for $q \zeta > 4$ and for strong steric hindrance, 
1221: the cross correlations $F_q^{AB} (t)$ are
1222: very small. The reason is that the
1223: intramolecular correlation factors $j_\ell (q \zeta /2)$ are
1224: small, and thus interference effects between the density
1225: fluctuations of the two interaction sites are suppressed.
1226: Coherence effects can be expected only for smaller wave vectors.
1227: For this case, the functions can be understood in terms of
1228: their small-$q$ asymptotes, 
1229: Eq.~(\ref{eq:Fab-small-q}). 
1230: 
1231: The lower two panels in Fig.~\ref{fig:NN-ZZ-t} deal with weak steric hindrance.
1232: In this case, the ``charge''-density
1233: fluctuations behave quite differently from the number-density
1234: fluctuations. The most important origin of this difference is the
1235: reduction of the mode-coupling vertices $V_{q,kp}^Z$ relative to
1236: $V_{q,kp}^N$ in Eq.~(\ref{eq:MCT-NZ-b}). For small elongations of the molecule,
1237: the effective solute-solvent potentials for reorientations are
1238: small. Therefore, the $f_q^{Zc}$ decrease strongly relative to
1239: $f_q^{Nc}$ for $\zeta$ decreasing towards $\zeta_c$, as is shown
1240: in Fig.~\ref{fig:fq-vs-zeta}. 
1241: Upon approaching $\zeta_{c}$, 
1242: the $\alpha$-peak strength of $\phi_q^Z (t)$ gets suppressed
1243: relative to the one of $\phi_q^N (t)$. Within phase II, 
1244: the ``charge''-density fluctuations
1245: relax to zero as in a normal liquid.
1246: This implies as precursor phenomenon, that
1247: the time scale $\tau_\alpha^Z$ of the
1248: ``charge''-density-fluctuation $\alpha$-process 
1249: shortens relative to the scale $\tau_\alpha^N$ for the
1250: number-density fluctuations. 
1251: Thus, the small-$\zeta$ behavior shown in the lower two panels
1252: of Fig.~\ref{fig:NN-ZZ-t} is due to disturbances of the standard MCT-transition
1253: scenario by the nearby type-$A$ transition.
1254: 
1255: The correlators $C_1 (t)$ and $C_2 (t)$ are
1256: shown in Fig.~\ref{fig:C1-C2} for the critical point $\varphi = \varphi_c$ and
1257: for two liquid states near the transition from phase I to phase
1258: III. For $\zeta = 0.80$, the anisotropic
1259: distribution of the solvent particles around the molecule leads to
1260: a stronger coupling to the dipole reorientations than to the
1261: reorientations for the quadrupole, and therefore the plateau for
1262: the former is higher than for the latter, $f_1^c > f_2^c$. A
1263: leading order expansion of the solutions of the equations of
1264: motion (\ref{eq:GLE-C1}) and (\ref{eq:GLE-C2}) 
1265: in terms of the small parameter $C_\ell (t) - f_\ell^c$ 
1266: leads to the factorization in the critical amplitude
1267: $h_\ell$ and a function $G (t)$ called the $\beta$-correlator,
1268: $C_{\ell}(t) - f_{\ell}^{c} = h_{\ell} G(t)$.
1269: The latter is the same for all correlation functions. It obeys the
1270: first scaling law of MCT,
1271: $G(t) = \sqrt{| \sigma |} \, g_{\pm} (t / t_{\sigma})$ for
1272: $\sigma \gtrless 0$.
1273: The master functions $g_{\pm} (\hat t)$ are determined 
1274: by $\lambda$. 
1275: They also describe the dynamics of the
1276: solvent in the window where 
1277: $\mid \phi_q (t) - f_q^c \mid \ll 1$~\cite{Franosch97}. In particular
1278: there holds $g_- (\hat t_-) = 0, \, \hat t_- = 0.704$, so that
1279: both correlators $C_\ell (t)$ cross their plateau at the same time
1280: $\tau_\beta = \hat t_- t_\sigma$. The non-linear mode-coupling
1281: effects request, that the correlators approach zero roughly at the
1282: same time. Thus one understands the general differences between the
1283: $\alpha$-processes, which were mentioned in the introduction:
1284: the $\alpha$-process for dipole relaxation is stronger, slower,
1285: and less stretched than the one for quadrupole relaxation. This
1286: finding is in qualitative agreement with the ones of
1287: the theory based on tensor-density representation of the
1288: structure \cite{Goetze00c}. There are, however, quantitative
1289: differences between the two approximation schemes. The plateaus
1290: $f_1^c = 0.905$ and $f_2^c = 0.674$ are smaller
1291: than the corresponding values 0.943 and 0.835 found in 
1292: Ref.~\cite{Goetze00c} and the amplitudes $h_1 = 0.19$ and $h_2 = 0.40$
1293: are bigger than the corresponding values 0.13 and 0.35 calculated
1294: previously \cite{Goetze00c}. The times $\tau_\alpha^\ell$
1295: characterizing the $\alpha$-process shall be defined by $C_\ell
1296: (\tau_\alpha^\ell) = f_\ell^c / 2$. They are marked by open
1297: squares in Fig.~\ref{fig:C1-C2}. 
1298: The values $\tau_{\alpha}^{1}=5.21 \times 10^{5}$, 
1299: $\tau_{\alpha}^{2}=1.64 \times 10^{5}$ for $x=3$ and $\zeta=0.80$
1300: are smaller than the ones
1301: reported in Ref.~\cite{Goetze00c}. 
1302: The present theory implies a somewhat weaker coupling of the
1303: reorientational degrees of freedom of the molecule to the
1304: dynamics of the solvent, than found earlier \cite{Goetze00c}. 
1305: This holds also for the small elongation $\zeta = 0.43$. The
1306: approach towards the transition from phase III to phase II, leads
1307: to a suppression of $f_1^c$, as discussed for the $f_q^{Zc}$ in
1308: Fig.~\ref{fig:fq-vs-zeta}. 
1309: The dipole relaxation speeds up for $\zeta \to \zeta_c$,
1310: as discussed for the lower panels of Fig.~\ref{fig:NN-ZZ-t}.
1311: This is reflected by an enhancement of $h_1 = 1.60$ relative to the
1312: amplitudes cited for $\zeta = 0.80$ but also relative to 
1313: $h_2 = 0.49$. 
1314: 
1315: One can perform 
1316: $\lim_{\sigma \to 0-} \phi_q (\tilde{t} t_\sigma^\prime) = 
1317: \tilde{\phi}_q (\tilde{t})$
1318: for the solutions of Eq.~(\ref{eq:GLE-v}), 
1319: where $\tilde{\phi}_q (\tilde{t})$ can be evaluated from 
1320: the mode-coupling functional at the critical point. It obeys the
1321: initial condition 
1322: $\tilde \phi_q (\tilde t) = f_q^c - h_q \tilde{t}^b + O (\tilde t^{2b})$. 
1323: Function $\tilde \phi_q (\tilde t)$ can
1324: be considered as shape function of the $\alpha$-process, and the
1325: result implies the second scaling law of MCT, also referred to as
1326: superposition principle: $\phi_q (t) = \tilde \phi_q (t /
1327: t_\sigma^\prime)$ for $\sigma \to 0-$ \cite{Goetze91b}.
1328: Corresponding laws hold for all functions, as is demonstrated in
1329: detail for the HSS in Refs.~\cite{Franosch97,Fuchs98}. 
1330: In particular, one gets for the reorientational correlators
1331: for $\sigma \to 0-$:
1332: \begin{mathletters}
1333: \label{eq:alpha}
1334: \begin{equation}
1335: C_{\ell}(t) = \tilde{C}_{\ell}(\tilde{t}), \quad
1336: \tilde{t} = t / t'_{\sigma}, \quad
1337: t_{\sigma} \ll t,
1338: \label{eq:alpha-a}
1339: \end{equation}
1340: and this corresponds to the $\alpha$-scaling law for the susceptibility
1341: spectra
1342: \begin{equation}
1343: \chi''_{\ell}(\omega) = \tilde{\chi}''_{\ell}(\tilde{\omega}), \quad
1344: \tilde{\omega} = \omega t'_{\sigma}, \quad
1345: \omega \ll 1/t_{\sigma}. 
1346: \label{eq:alpha-b}
1347: \end{equation}
1348: The initial decay of the master function $\tilde{C}_{\ell}(\tilde{t})$ 
1349: is described by von
1350: Schweidler's law
1351: \end{mathletters}
1352: \begin{mathletters}
1353: \label{eq:von}
1354: \begin{equation}
1355: \tilde{C}_{\ell}(\tilde{t}) = f_{\ell}^{c} - h_{\ell} \, \tilde{t}^{b}, \quad
1356: \tilde{t} \to 0,
1357: \label{eq:von-a} 
1358: \end{equation}
1359: which is equivalent to a power-law tail of the master spectrum
1360: $\tilde{\chi}''_{\ell}(\tilde{\omega})$:
1361: \begin{equation}
1362: \tilde{\chi}''_{\ell}(\tilde{\omega}) = 
1363: h_{\ell} \, \sin (\pi b / 2) \, \Gamma(1+b) \, / \, \tilde{\omega}^{b}, \quad
1364: \tilde{\omega} \to \infty. 
1365: \label{eq:von-b}
1366: \end{equation}
1367: The upper panel of Fig.~\ref{fig:alpha-spectra} 
1368: exhibits the $\alpha$-process master spectra for the
1369: reorientational processes for $\zeta = 0.80$ and 
1370: for the dimensionless longitudinal 
1371: elastic modulus $m_{q=0}(t)$ of the solvent.
1372: The latter can be measured by Brillouin-scattering spectroscopy.
1373: It probes a tensor-density fluctuation for 
1374: $\ell = 0$. 
1375: \end{mathletters}
1376: The von-Schweidler-law tails describe the spectra
1377: for frequencies exceeding the position $\tilde \omega_{\max}$
1378: of the susceptibility maximum by a factor of about 100, as shown
1379: by the dashed lines. Since $f_1^c > f_2^c$ and both plateau values are rather
1380: large, one understands from the theory for the leading corrections
1381: to Eq.~(\ref{eq:von-b}) \cite{Franosch97} that for decreasing $\tilde \omega$
1382: the von-Schweidler asymptote underestimates the spectrum, and this
1383: by larger values for $\ell = 1$ than for $\ell = 2$. 
1384: For smaller plateau values, the von-Schweidler asymptote may
1385: overestimate the spectrum, as is exemplified for the modulus.
1386: The lower panel of Fig.~\ref{fig:alpha-spectra}
1387: demonstrates that the $\alpha$-processes speed up if steric
1388: hindrance is decreased. As precursor of the transition to phase II,
1389: the spectrum for the dipole response is located at much higher frequencies
1390: than that for the quadrupole response. 
1391: Traditionally, dielectric-loss spectra have been fitted by the
1392: ones of the Kohlrausch-law 
1393: $\phi_K (\tilde{t}) = A \exp[ - (\tilde{t} B)^\beta]$. 
1394: Such fits also describe a major part of the
1395: spectra in Fig.~\ref{fig:alpha-spectra}, as shown by the dotted lines. The parameters
1396: $A$ and $B$ are adjusted to match the susceptibility maximum. The
1397: stretching exponent $\beta$ is chosen so that the spectrum is
1398: fitted at half maximum $\tilde\chi_{\ell}^{\prime \prime} (\tilde
1399: \omega_{\max}) / 2$. If one denotes the width in $\log_{10}
1400: \omega$ at half maximum by $W$, stretching means that this
1401: parameter is larger than the value $W_D = 1.14$ characterizing a
1402: Debye process, $\phi_D (\tilde t) = \exp( - \tilde{t})$. 
1403: The upper panel of Fig.~\ref{fig:alpha-spectra}
1404: quantifies the general results of the theory for strong steric
1405: hindrance: $\tilde \chi_1^{\prime \prime} (\tilde \omega_{\max}^1)
1406: > \tilde \chi_2^{\prime \prime} (\tilde \omega_{\max}^2)$,
1407: $\tilde \omega_{\max}^1 < \tilde \omega_{\max}^2$ and $\beta_1 >
1408: \beta_2$. It quantifies also the forth property cited in the
1409: introduction $\tilde \omega_{\max}^2 < \tilde \omega_{\max}^0$. 
1410: A further general property is $\beta_2 > \beta_0$.
1411: 
1412: The discussion of power-law spectra is done more
1413: conveniently in a double logarithmic diagram
1414: as shown in Fig.~\ref{fig:Nagel-plot} for normalized
1415: dipole-fluctuation-$\alpha$-process spectra 
1416: $\tilde{C}_{1}^{\prime \prime} (\tilde{\omega}) \tilde{\omega}_{\max} / f_{1}^{c} =
1417: \tilde{\chi}_1^{\prime \prime} (\tilde{\omega}) \tilde{\omega}_{\max} /
1418: f_{1}^{c} \tilde{\omega}$ as function of 
1419: $\tilde{\omega} / \tilde{\omega}_{\max}$.
1420: One notices, that there is a
1421: white-noise spectrum for $\tilde{\omega}$ below 
1422: $\tilde{\omega}_{\max}$. The high-frequency
1423: wing of the Kohlrausch-law fit decreases proportional to
1424: $\tilde{\omega}^{- \beta}$ and underestimates the
1425: spectrum $\tilde\chi_1^{\prime \prime}(\tilde{\omega})$ considerably. Because of
1426: the von-Schweidler asymptote, which is shown as dashed straight
1427: line, the spectrum exhibits an enhanced high-frequency wing.
1428: Dixon {\it et al.} \cite{Dixon90} made the remarkable observation that 
1429: their dielectric spectra could be
1430: collapsed on one master curve, if the vertical axis is rescaled by
1431: $w^{-1}$ and the horizontal axis by $w^{-1} (1 + w^{-1})$, where 
1432: $w = W / W_D$.
1433: In Fig.~\ref{fig:Nagel-plot} this scaling is used and the data for
1434: glycerol from Ref.~\cite{Dixon90} are included. The spectra 
1435: for molecules with $\zeta = 0.80$ and $\zeta=0.60$,
1436: which are relevant for the description of van der Waals 
1437: systems~\cite{Goetze00c}, follow the
1438: mentioned scaling surprisingly well.
1439: This finding appears non-trivial, since the scaling is not reproduced
1440: by the MCT results of the basic quantities $\phi_{q}(t)$~\cite{Fuchs92b}.
1441: The rescaled spectrum for $\zeta=0.43$ deviates from the scaling law
1442: for $w^{-1}(1+w^{-1}) \log_{10} (\tilde{\omega}/\tilde{\omega}_{max}) \geq 5$.
1443: 
1444: It might appear problematic that the dipole correlator was
1445: calculated within a different approximation scheme than the
1446: quadrupole correlator. But, there is no difficulty to also
1447: evaluate $C_1 (t)$ within the scheme explained in 
1448: Sec.~\ref{subsec:MCT-C2} for
1449: the evaluation of $C_2 (t)$. Figure~\ref{fig:compare-C1} 
1450: presents a comparison of 
1451: $C_1 (t)$ obtained along the two specified routes. 
1452: The two results for the small elongation $\zeta =
1453: 0.43$ are close to each other. The discrepancies are mainly due to the
1454: 7\% difference between the two plateau values $f_1^c$.
1455: With increasing $\zeta$, the discrepancies
1456: decrease. For the large elongation $\zeta = 0.80$, the results are
1457: practically undistinguishable.
1458: 
1459: \subsection{Structural relaxation versus transient dynamics}
1460: \label{subsec:result-inertia}
1461: 
1462: Let us introduce Fourier-Laplace transforms of functions of time,
1463: say $f (t)$, to functions of frequency, say $f (\omega)$, with the
1464: convention $f (\omega) = i \int_0^\infty dt \exp (izt) f (t)$, 
1465: $z = \omega + i0$. 
1466: The equations of motion (\ref{eq:GLE-u}) with the
1467: initial conditions from Eq.~(\ref{eq:Fab-short-time}) are transformed to
1468: \begin{equation}
1469: [\omega \, {\bf 1} + {\bf \Omega}_{q}^{2} \, {\bf m}_{q}(\omega)] \,
1470: [\omega \, {\bf F}_{q}(\omega) + {\bf w}_{q}] -
1471: {\bf \Omega}_{q}^{2} \, {\bf F}_{q}(\omega) = {\bf 0}.
1472: \label{eq:str-1}
1473: \end{equation}
1474: Within the glass, the long-time limits of ${\bf F}_q (t)$ and
1475: ${\bf m}_{q}(t)$ do not vanish, i.e., the transformed quantities exhibit
1476: zero-frequency poles. One gets, for example, $\omega {\bf F}_q
1477: (\omega) \to - {\bf F}_q^\infty$ for $\omega \to 0$, where the
1478: strength $- {\bf F}_q^\infty$ 
1479: of the poles follow from Eq.~(\ref{eq:DW-u}). Continuity of the
1480: solutions of the MCT-equation of motion implies that for vanishing
1481: frequencies and for vanishing distances from the transition points,
1482: ${\bf m}_q (\omega)$ becomes arbitrarily large. Hence, combinations like 
1483: $\omega + i \xi_q$ with constants $\xi_{q}$
1484: can be neglected compared to ${\bf m}_q (\omega)$.
1485: Therefore, in the region of glassy dynamics, Eq.~(\ref{eq:str-1}) can be
1486: modified to
1487: \begin{equation}
1488: {\bf F}_{q}(\omega) - {\bf m}_{q}(\omega) \, {\bf w}_{q} =
1489: i \,\mbox{\boldmath $\xi$}_{q} +
1490: \omega \, {\bf m}_{q}(\omega) \, {\bf F}_{q}(\omega).
1491: \label{eq:str-2}
1492: \end{equation}
1493: Let us assume, that this equation has a solution, to be denoted by
1494: ${\bf F}_q^{*} (\omega)$, which is defined for all frequencies,
1495: so that it can be back
1496: transformed to a function ${\bf F}_q^{*} (t)$, defined for all $t > 0$. 
1497: Choosing $\mbox{\boldmath $\xi$}_{q}$
1498: properly, Eq.~(\ref{eq:str-2}) can
1499: be written as 
1500: \begin{mathletters}
1501: \label{eq:str-3}
1502: \begin{equation}
1503: {\bf F}_{q}^{*}(t) - {\bf m}_{q}^{*}(t) \, {\bf w}_{q} =
1504: - (d/dt) \int_{0}^{t} dt' \, {\bf m}_{q}^{*}(t-t') \, {\bf F}_{q}^{*}(t').
1505: \label{eq:str-3-a}
1506: \end{equation}
1507: Similar reasoning leads from Eq.~(\ref{eq:GLE-v}) to
1508: \begin{equation}
1509: \phi_{q}^{*}(t) - m_{q}^{*}(t) = - (d/dt) \int_{0}^{t} dt' \,
1510: m_{q}^{*}(t-t') \, \phi_{q}^{*}(t').
1511: \label{eq:str-3-b}
1512: \end{equation}
1513: These formulas have to be supplemented with the MCT expressions
1514: for the kernels
1515: \begin{equation}
1516: {\bf m}_{q}^{*}(t) = \mbox{\boldmath ${\cal F}$}_{q}
1517: [ {\bf F}^{*}(t), \phi^{*}(t) ], \quad
1518: m_{q}^{*}(t) = {\cal F}_{q}[\phi^{*}(t)].
1519: \label{eq:str-3-c}
1520: \end{equation}
1521: Equations (\ref{eq:str-3}) for the glassy dynamics are scale
1522: invariant. 
1523: \end{mathletters}
1524: With one set of solutions $\phi_q^{*} (t)$ and ${\bf
1525: F}_q^{*} (t)$, also the set 
1526: $\phi_q^{* \, u} (t) = \phi_q^{*} (ut)$ and 
1527: ${\bf F}_q^{* \, u} (t) = {\bf F}_q^{*} (ut)$
1528: provides a solution for arbitrary $u > 0$. 
1529: To fix the solution uniquely, one can 
1530: introduce positive numbers $y_q$ and positive definite matrices
1531: ${\bf y}_q$ to specify the initial condition as
1532: power-law asymptotes~\cite{Fuchs99b}
1533: \begin{equation}
1534: {\bf F}_{q}^{*}(t) \, (t/t_{0})^{1/3} \to {\bf y}_{q}, \,\,
1535: \phi_{q}^{*}(t) \, (t/t_{0})^{1/3} \to y_{q}, \,\,
1536: (t/t_{0}) \to 0.
1537: \label{eq:str-4}
1538: \end{equation}
1539: 
1540: The theory of the asymptotic solution of the MCT equations for
1541: simple systems had been built on the analogue of Eq.~(\ref{eq:str-2}) with
1542: $\xi_q$ neglected \cite{Franosch97}. The present theory extends
1543: the previous one by the introduction of matrices. It seems obvious
1544: that the previous results for asymptotic expansions hold in a
1545: properly extended version. Let us only note 
1546: the formula for the long-time decay of the correlators at
1547: the critical point $\varphi = \varphi_c$
1548: \cite{Franosch97,Fuchs98}:
1549: \begin{equation}
1550: {\bf F}_{q}(t) = {\bf F}_{q}^{\infty \, c} +
1551: {\bf H}_{q} (t_{0}/t)^{a}
1552: [{\bf 1} + {\bf K}_{q} (t_{0}/t)^{a} + {\bf O}( (t_{0}/t)^{2a} )].
1553: \label{eq:str-5}
1554: \end{equation}
1555: The exponent $a$, mentioned above, can be calculated from
1556: the mode-coupling functional at the critical point. The same is
1557: true for the plateau values ${\bf F}_q^{\infty \, c}$, and the 
1558: amplitudes ${\bf H}_q$ and ${\bf K}_q$. 
1559: The dependence of the solution from the transient
1560: dynamics is given by the single number $t_0$. 
1561: Let us anticipate that
1562: Eq.~(\ref{eq:str-5}) and similar results can be extended to a complete
1563: solution. One concludes that the glassy dynamics is determined, up
1564: to a scale $t_0$, by the mode-coupling functionals in 
1565: Eq.~(\ref{eq:str-3-c}).
1566: 
1567: Equations (\ref{eq:BXX-1}) and (\ref{eq:BXX-2}) show, that the mode-coupling
1568: functionals ${\cal F}_q$ and ${\cal F}_{q}^{ab}$ 
1569: are specified by the
1570: density $\rho$, the static structure factor $S_q$, the direct
1571: correlation function $c_q$ of the solvent, and the
1572: solute-solvent direct correlation functions $c_q^a$,
1573: i.e., by equilibrium quantities. They are the same for
1574: systems with a Newtonian dynamics, as considered in this paper,
1575: and for a model with a Brownian dynamics, as is to be used for the
1576: description of colloidal suspensions. In particular, the 
1577: mode-coupling functionals are independent of the particle masses $m,
1578: m_A$ and $m_B$. Thus, the glassy dynamics of the
1579: molecule in the simple liquid does not depend on the 
1580: inertia parameters specifying the microscopic equations of motion.
1581: The same conclusions on the glassy dynamics, which were cited in
1582: the introduction for the basic version of MCT, hold for the model
1583: studied in this paper. Let us add, that neither the temperature
1584: $T$, nor the interparticle-interaction potentials $V$ of the
1585: solvent, nor the solute-solvent-interaction potentials $V^a$
1586: enter explicitly the mode-coupling functionals. These quantities
1587: only enter implicitly via $S_{q}$, $c_{q}$, 
1588: $c_{q}^{A}$ and $c_{q}^{B}$.
1589: 
1590: The independence of the glassy dynamics on the
1591: inertia parameters is demonstrated in Fig.~\ref{fig:C1-inertia} 
1592: for four states of
1593: the liquid. It is shown that the reorientational dynamics of the
1594: dipole does not change for $t > 1$ even if the mass ratio of the atoms
1595: $m_A /m_B$ is altered by a factor 10. The transient
1596: dynamics, which deals with overdamped librations, 
1597: exhibits an isotope effect. There is
1598: no fitting parameter involved in the shown diagram. The 
1599: scale $t_0$ neither depends on the density of the
1600: solvent nor on the elongation parameter, 
1601: and Eqs.~(\ref{eq:str-3}) describe the complete 
1602: control-parameter dependence of the glassy dynamics.
1603: 
1604: \section{SUMMARY}
1605: \label{sec:5}
1606: 
1607: The MCT for simple systems with a dilute solute
1608: of atoms has been generalized to the one with a dilute solute of
1609: diatomic molecules. The derived equations of motion generalize
1610: the ones for atoms in the sense that scalar functions
1611: are replaced by 2-by-2 matrix functions. 
1612: These generalizations result
1613: from the description of the position of the molecule in
1614: terms of interaction-site-density fluctuations. The numerical
1615: effort required for a solution of the equations is not seriously
1616: larger than the one needed to solve the corresponding equations
1617: for atomic solutes. This holds in particular for symmetric
1618: molecules where the matrix equations can be diagonalized
1619: by a linear transformation to number-density and
1620: ``charge''-density fluctuations. 
1621: The theory implies
1622: that the dynamics outside the transient regime is determined, up
1623: to an overall time scale $t_0$, by the equilibrium structure. In
1624: particular, it is independent of the inertia
1625: parameters of the system.
1626: 
1627: It was shown that the present theory reproduces all qualitative
1628: results obtained within the preceding much more involved theory based
1629: on a description of the solute by tensor-density fluctuations.
1630: Both theories yield a similar phase diagram, Fig.~\ref{fig:phase}. 
1631: The characteristic differences of the reorientational correlators
1632: exhibited for strong steric hindrance of rotations as opposed to
1633: weak steric hindrance are obtained here, Fig.~\ref{fig:C1-C2}, 
1634: as previously.
1635: There are systematic quantitative differences between the two
1636: approximation approaches in the sense, that the implications of
1637: the cage effect are weaker in the present theory than in the
1638: preceding work \cite{Franosch97c,Goetze00c}. The earlier findings
1639: on the differences of the spectra referring to
1640: responses with angular-momentum index $\ell = 0$, 1 and 2 have
1641: been corroborated by separating the $\alpha$-peaks from
1642: the complete spectra with the aid of the $\alpha$-scaling law,
1643: Fig.~\ref{fig:alpha-spectra}. 
1644: 
1645: The critical non-ergodicity parameters $f_q^{x \, c}$ have been
1646: calculated, which determine the form factors of
1647: quasi-elastic scattering from the liquid near the glass transition.
1648: Contrary to what one finds for atomic solutes, these are not
1649: bell-shaped functions of wave numbers $q$, rather they exhibit
1650: kinks, Fig.~\ref{fig:fq-strong}. The form factors for some 
1651: wave numbers $q$ 
1652: vary non-monotonically with changes of the elongation of the molecule, 
1653: Fig.~\ref{fig:fq-vs-zeta}.
1654: Analyzing these findings in terms of form factors defined for
1655: fixed angular-momentum index $\ell$, one can explain the
1656: results as due to intra-molecular interference effects 
1657: demonstrating that the theory accounts for reorientational
1658: correlations with angular-momentum index $\ell \geq 2$.
1659: 
1660: It was shown that the $\alpha$-relaxation spectra for dipole
1661: reorientations with strong steric hindrance obey the scaling law
1662: proposed by Dixon {\it et al.} \cite{Dixon90} within the window and
1663: within the accuracy level considered by these authors. There is no
1664: fitting parameter involved in the construction of 
1665: Fig.~\ref{fig:Nagel-plot}, which
1666: demonstrates this finding for the two elongation parameters
1667: $\zeta=0.60$ and 0.80. Thus, it is not
1668: justified to use the cited empirical scaling as an argument against the
1669: applicability of MCT for a discussion of dielectric-loss spectra.
1670: 
1671: \acknowledgments
1672: 
1673: We thank M.~Fuchs, M.~Sperl and Th.~Voigtmann for many
1674: helpful discussions and suggestions. 
1675: We are grateful to A.~Latz and R.~Schilling for constructive
1676: critique of our manuscript.
1677: S.-H.~C. acknowledges financial support from
1678: JSPS Postdoctoral Fellowships for Research Abroad. 
1679: This work was supported by 
1680: Verbundprojekt BMBF 03-G05TUM.
1681: 
1682: \appendix
1683: 
1684: \section{TENSOR-DENSITY REPRESENTATIONS}
1685: \label{appen:A}
1686: 
1687: Following the conventions of Refs.~\cite{Franosch97c} and
1688: \cite{Goetze00c}, normalized tensor-density fluctuations shall be used by
1689: decomposing the molecule's position variable in plane waves 
1690: $\exp( i {\vec q} \cdot {\vec r}_{C})$ for the center of mass $\vec r_C$ 
1691: and in spherical harmonics $Y_{\ell}^{m}({\vec e} \,)$ for the 
1692: orientation vector $\vec e$:
1693: \begin{equation}
1694: \rho_{\ell}^{m}({\vec q}) = i^{\ell} \, \sqrt{4 \pi} \, 
1695: \exp ( i {\vec q} \cdot {\vec r}_{\rm C}) \,
1696: Y_{\ell}^{m}({\vec e}).
1697: \label{eq:rho-tensor-def}
1698: \end{equation}
1699: The molecule-solvent interactions are specified by a set of
1700: direct correlation functions
1701: \begin{equation}
1702: c_{\ell}(q) = 
1703: \langle \rho_{{\vec q}_{0}}^{*} \, \rho_{\ell}^{0}({\vec q}_{0}) \rangle \, / \, 
1704: (\rho S_{q}); \quad
1705: {\vec q}_{0} = (0,0,q).
1706: \label{eq:cuv-tensor-def}
1707: \end{equation}
1708: The structure dynamics is described by the matrix of correlators,
1709: defined by
1710: \begin{equation}
1711: \phi_{\ell k}(qm,t) = \langle \rho_{\ell}^{m}({\vec q}_{0},t)^{*} 
1712: \rho_{k}^{m}({\vec q}_{0}) \rangle.
1713: \label{eq:tensor-correlator}
1714: \end{equation}
1715: 
1716: Since the position vectors of the interaction sites can be written as 
1717: $\vec r_a = \vec r_C + z_a \vec e$, 
1718: the Rayleigh expansion of the exponential in Eq.~(\ref{eq:rho-def}) 
1719: yields the formula
1720: \begin{equation}
1721: \rho^{a}_{{\vec q}_{0}} = \sum_{\ell} \sqrt{2 \ell + 1} \, 
1722: j_{\ell}(q z_{a}) \, \rho_{\ell}^{0}({\vec q}_{0}).
1723: \label{eq:rho-site-tensor}
1724: \end{equation}
1725: Substitution of this expression into Eq.~(\ref{eq:huv-def}) leads with 
1726: Eq.~(\ref{eq:cuv-tensor-def})
1727: to an expression connecting the direct correlation functions
1728: $c_q^a$ with $c_{\ell}(q)$:
1729: \begin{equation}
1730: \sum_{b} w^{ab}_{q} \, c^{b}_{q} = 
1731: \sum_{\ell} \sqrt{2 \ell + 1} \, j_{\ell}(q z_{a}) \, c_{\ell}(q).
1732: \label{eq:cuv-site-tensor}
1733: \end{equation}
1734: Substitution of Eq.~(\ref{eq:rho-site-tensor}) into Eq.~(\ref{eq:Fab-def}) 
1735: leads with Eq.~(\ref{eq:tensor-correlator}) to the
1736: expression of the site-density correlators in terms of the
1737: tensor-density correlators:
1738: \begin{equation}
1739: F^{ab}_{q}(t) = \sum_{\ell k} \sqrt{(2 \ell + 1)(2 k + 1)} \, 
1740: j_{\ell}(qz_{a}) \, j_{k}(qz_{b}) \,
1741: \phi_{\ell k}(q0,t).
1742: \label{eq:Fab-site-tensor}
1743: \end{equation}
1744: For the long-time limits of the correlators $F_q^{ab}(t)$ one gets a
1745: combination of the non-ergodicity parameters $f_{\ell k} (q m) =
1746: \phi_{\ell k} (q m, t \to \infty)$. If one uses the diagonal
1747: approximation $f_{\ell k} (q, m) = \delta_{\ell k} f(q \ell m)$,
1748: one can write
1749: \begin{equation}
1750: F^{ab}_{q}(t \to \infty) = \sum_{\ell} ( 2 \ell + 1) \, 
1751: j_{\ell}(qz_{a}) \, j_{\ell}(qz_{b}) \,
1752: f(q \ell 0).
1753: \label{eq:Fab-site-tensor-2}
1754: \end{equation}
1755: 
1756: Substituting the Rayleigh expansion of Eq.~(\ref{eq:rho-def}) into 
1757: Eq.~(\ref{eq:m2-MCT-a}),
1758: one can express the pair modes in terms of those formed with
1759: tensor-density fluctuations:
1760: \begin{equation}
1761: A^{a}({\vec k}, {\vec p} \, ) = \sqrt{4 \pi} \sum_{\ell m}
1762: j_{\ell}(k z_{a}) \, Y_{\ell}^{m}({\vec k}) \,
1763: [\, \rho_{\ell}^{m}({\vec k}) \, \rho_{\vec p} \, / \sqrt{N S_{p}} \, ].
1764: \end{equation}
1765: Therefore, the overlaps of the forces with the pair modes can be
1766: expressed as sums of the corresponding quantities calculated in
1767: Ref. \cite{Franosch97c}. One finds for the mode-coupling
1768: coefficients in Eq.~(\ref{eq:m2-final}):
1769: \begin{mathletters}
1770: \label{eq:m2-exp}
1771: \begin{equation}
1772: V_{2}^{ab}(k) = [ k^{2} \rho S_{k} / 2 \pi^{2}] 
1773: \sum_{c d} ({\bf w}_{k}^{-1})^{ac} \, D_{c}(k) \, D_{d}(k) \,
1774: ({\bf w}_{k}^{-1})^{db}.
1775: \label{eq:m2-exp-a}
1776: \end{equation}
1777: Here one gets for $a = A, B$ or $C$
1778: \bea
1779: D_{a}(k) &=& 
1780: {\textstyle \frac{1}{12}} \sum_{\ell J}
1781: (-1)^{\frac{1}{2} (\ell + J)} \, (2 \ell + 1) \, \sqrt{2J + 1} \,
1782: j_{\ell}(k z_{a}) \, c_{J}(k) 
1783: \nonumber \\
1784: &\times&
1785: [ \, J ( J + 1 ) + 6 - \ell ( \ell + 1 ) \, ]
1786: \left(
1787: \begin{array}{ccc}
1788: 2 & \ell & J \\
1789: 0 & 0    & 0 
1790: \end{array}
1791: \right)^{2},
1792: \label{eq:m2-exp-b}
1793: \eea
1794: where the last factor denotes Wigner's 3-$j$ symbol. 
1795: \end{mathletters}
1796: 
1797: \section{Mode-Coupling Coefficients}
1798: \label{appen:B}
1799: 
1800: Mode-coupling equations based on a description of the molecules
1801: by site-density fluctuations have been derived in Ref.
1802: \cite{Chong98b} by extending the procedure used originally for
1803: atomic systems \cite{Goetze91b}. But the reported formulas
1804: \cite{Chong98b} do not seem appropriate, since they do not reduce
1805: to the ones for tagged particle motion if the limit of a vanishing
1806: elongation parameter $\zeta$ is considered. Therefore, an
1807: alternative derivation will be presented which starts from the
1808: theory developed by Mori and Fujisaka \cite{Mori73} for an
1809: approximate treatment of nonlinear fluctuations. The application
1810: of this theory will be explained by rederiving the equations for
1811: the solvent before the ones for the solute are worked out.
1812: 
1813: \subsection{The Mori-Fujisaka equations}
1814: \label{subsec:B1}
1815: 
1816: Let us consider a set of distinguished dynamical variables
1817: $A_\alpha, \alpha = 1,2, \cdots$, defined as functions on the
1818: system's phase space. The time evolution is generated by the
1819: Liouvillian, $A_\alpha (t) = \exp (i {\cal L} t) A_\alpha$. Using
1820: the canonical averaging to define scalar products in the space of
1821: variables, $(A, B) = \langle A^* B \rangle$, the Liouvillian is
1822: Hermitian. It is the goal to derive equations of motion for the
1823: matrix of correlators $C_{\alpha\beta} (t) = (A_\alpha (t),
1824: A_\beta)$. The initial condition is given by the positive definite
1825: matrix $g_{\alpha\beta} = (A_\alpha, A_\beta)$. The theory starts
1826: with a generalized Fokker-Planck equation for the distribution
1827: function $g_a = \Pi_\alpha \, \delta (A_\alpha - a_\alpha), \, a =
1828: (a_1, a_2, \cdots)$. It is assumed that the time scales for the
1829: fluctuations of the $A_\alpha$ and their products is larger than
1830: the ones for the Langevin fluctuating forces. The spectra of the
1831: latter can then be approximated by a constant matrix
1832: $\Gamma_{\alpha\beta}$. It is assumed furthermore, that the
1833: $\Gamma_{\alpha\beta}$ are independent of the distinguished
1834: variables, and that the equilibrium distribution of the latter is
1835: Gaussian:
1836: \begin{equation}
1837: \langle g_{a} \rangle = C \,
1838: \exp \biggl( - \frac{1}{2} \sum_{\alpha \beta}
1839: a_{\alpha} \, g_{\alpha \beta}^{-1} \, a_{\beta}^{*} \biggr).
1840: \label{eq:B1}
1841: \end{equation}
1842: In the cited Fokker-Planck equation there occurs the 
1843: streaming velocity $v_\alpha (a)$ given by
1844: \begin{equation}
1845: v_{\alpha}(a) \, \langle g_{a} \rangle = 
1846: \langle \dot{A}_{\alpha}^{*} \, g_{a} \rangle =
1847: k_{B}T \sum_{\beta} \frac{\partial}{\partial a_{\beta}}
1848: \langle \{ A_{\alpha}^{*}, A_{\beta} \} \, g_{a} \rangle.
1849: \label{eq:B2}
1850: \end{equation}
1851: 
1852: The Fokker-Planck equation is now reduced by projecting out the subspace
1853: of distinguished variables. There appears the frequency matrix
1854: specifying the linear contribution to the streaming term,
1855: \begin{equation}
1856: \Omega_{\alpha \beta} = \sum_{\gamma}
1857: (A_{\alpha}, {\cal L} A_{\gamma}) \, g_{\gamma \beta}^{-1}.
1858: \label{eq:B3}
1859: \end{equation}
1860: The nonlinear contributions enter as combination
1861: \begin{equation}
1862: f_{\alpha} = \int da \, v_{\alpha}(a) \, g_{a} -
1863: i \sum_{\beta} \Omega_{\alpha \beta} A_{\beta}.
1864: \label{eq:B4}
1865: \end{equation}
1866: They determine the relaxation kernel as
1867: \begin{equation}
1868: M_{\alpha \beta}(t) = \sum_{\gamma} (f_{\alpha}(t), f_{\gamma}) \,
1869: g_{\gamma \beta}^{-1}.
1870: \label{eq:B5}
1871: \end{equation}
1872: The time evolution is generated by the reduced Liouvillian ${\cal
1873: L}^\prime = {\cal Q L Q}, \, f_\alpha (t) = \exp (i {\cal
1874: L}^\prime t) f_\alpha$, where $\cal Q$ is the projector on the
1875: space perpendicular to the one spanned by the distinguished
1876: variables. The result, which is equivalent to 
1877: Eq.~(2.15) in Ref.~\cite{Mori73}, reads
1878: \bea
1879: \partial_{t} C_{\alpha \beta}(t) &=& 
1880: - \sum_{\gamma} [ 
1881: (i \Omega_{\alpha \gamma} + \Gamma_{\alpha \gamma}) \, C_{\gamma \beta}(t) 
1882: \nonumber \\
1883: & &
1884: \quad \quad
1885: + \,
1886: \int_{0}^{t} dt' \, M_{\alpha \gamma}(t-t') \, C_{\gamma \beta}(t')].
1887: \label{eq:B6}
1888: \eea
1889: 
1890: \subsection{MCT equations for simple systems}
1891: 
1892: To get a description of slowly varying structural fluctuations, the
1893: original reasoning of MCT shall be adopted, and as distinguished
1894: variables the density fluctuations $\rho_{\vec q}$ and the
1895: longitudinal current fluctuations 
1896: $j_{\vec q} = \sum_{\kappa} v_{\kappa,z} \exp(i {\vec q} \cdot {\vec r}_{\kappa})$
1897: will be chosen. The variable label of the
1898: preceding subsection consists of two bits, 
1899: $\alpha = (\lambda, \vec q \,)$, $\lambda = 1,2$, so that 
1900: $A_{1 \vec q} = \rho_{\vec q}$, 
1901: $A_{2 \vec q} = j_{\vec q}$. 
1902: Notations follow the ones of
1903: the first paragraph of Sec.~\ref{sec:2}. 
1904: One gets $g_{1 {\vec q} \, 1 {\vec q}} = NS_q$,
1905: $g_{2 {\vec q} \,  2 {\vec q}} = Nv^2$,
1906: $\Omega_{1 {\vec q} \, 2 {\vec q}} = q$,
1907: $\Omega_{2 {\vec q} \, 1 {\vec q}} = \Omega_q^2 / q$, and all the
1908: other elements of the matrices $g$ and $\Omega$ are zero. Because
1909: of translational invariance, one gets 
1910: $C_{\lambda {\vec q} \, \mu {\vec k}}(t) = 0$ unless $\vec q = \vec k$, and rotational
1911: invariance implies that $C_{\lambda {\vec q} \, \mu {\vec q}} (t)$
1912: depends on the modulus $q$ only. The same holds for the matrices
1913: $\Gamma$ and $M$, so that Eq.~(\ref{eq:B6}) reduces to a two-by-two matrix
1914: equation with $q$ appearing as a parameter. 
1915: Since ${\cal L} \rho_{\vec q}$ 
1916: is an element of the distinguished set of
1917: variables, the kernels $\Gamma_{\alpha\beta}$ and $M_{\alpha\beta}
1918: (t)$ vanish unless $\alpha = \beta = (2, {\vec q} \,)$. The latter
1919: shall be denoted by $\Gamma_q$ and $M_q (t)$, respectively.
1920: Equation (\ref{eq:B6}) can then be reduced to the equation of motion for the
1921: normalized density correlator 
1922: $\phi_q (t) = C_{1 {\vec q} \, 1 {\vec q}} (t) / NS_q$:
1923: \bea
1924: & &
1925: \partial_{t}^{2} \phi_{q}(t) + \Gamma_{q} \, \partial_{t} \phi_{q}(t) 
1926: \nonumber \\
1927: & &
1928: \qquad
1929: + \,
1930: \Omega_{q}^{2} \, \phi_{q}(t) +
1931: \int_{0}^{t} dt' \, M_{q}(t-t') \, \partial_{t'} \phi_{q}(t') = 0.
1932: \label{eq:B7}
1933: \eea
1934: The application of the Mori-Fujisaka formalism can be summarized
1935: as follows. The relaxation kernel $\Omega_q^2 m_q (t)$ of the
1936: exact Eq.~(\ref{eq:GLE-v}) is approximately split into a white noise
1937: contribution $2 \Gamma_q \delta (t)$ and a remainder $M_q (t)$,
1938: where well defined formulas for the two contributions are
1939: available. In this paper, the kernel $\Gamma_q$ is neglected and 
1940: Eq.~(\ref{eq:B5}) for the kernel,
1941: \begin{equation}
1942: M_{q}(t) = (f_{\vec q}(t), f_{\vec q}) \, / \, N v^{2}, \quad
1943: f_{\vec q} = f_{2 {\vec q}}, 
1944: \label{eq:B8}
1945: \end{equation}
1946: shall be approximated further.
1947: 
1948: If one writes $a_{1 \vec q} = \tilde\rho_{\vec q}$ and $a_{2 \vec
1949: q} = \tilde j_{\vec q}$, one can denote the equilibrium distribution
1950: $w (a) = \langle g_a \rangle$ for the solvent variables as
1951: \bea
1952: w(a) &=& C \, \exp
1953: \biggl\{
1954: - (1/2N) \sum_{\vec q} [
1955: (1 - \rho c_{q}) \, \tilde{\rho}_{\vec q} \, \tilde{\rho_{\vec q}}^{*} 
1956: \nonumber \\
1957: & &
1958: \qquad \qquad \qquad \qquad \qquad
1959: + \,
1960: (1/v^{2}) \, \tilde{j}_{\vec q} \, \tilde{j}_{\vec q}^{*}]
1961: \biggr\}.
1962: \label{eq:B9}
1963: \eea
1964: This is used to derive from Eqs.~(\ref{eq:B2}) and (\ref{eq:B4}):
1965: \begin{equation}
1966: f_{\vec q} = - i (\rho v^{2}/ N) \sum_{\vec k} ({\vec k} \cdot {\vec q}/q) \,
1967: c_{k} \, \rho_{\vec k} \, \rho_{{\vec q}-{\vec k}} + \delta f_{\vec q},
1968: \label{eq:B10}
1969: \end{equation}
1970: where $\delta f_{\vec q}$ denotes a term whose contribution to the memory
1971: kernel turns out to be irrelevant
1972: for the structural relaxation processes, and therefore shall be neglected.
1973: The remaining task is the evaluation of averages
1974: $(\rho_{\vec k}(t) \rho_{\vec p} (t), 
1975: \rho_{\vec k^\prime} \rho_{\vec p^\prime})$, 
1976: where the time evolution is generated by
1977: the reduced Liouvillian. Here the original MCT ansatz is used:
1978: $(\rho_{\vec k} (t) \rho_{\vec k^\prime}) 
1979: (\rho_{\vec p} (t) \rho_{\vec p^\prime}) + (k \leftrightarrow p)$. 
1980: As a result, one
1981: gets $M_q (t) = \Omega_q^2 m_q (t)$ with the well known expression
1982: for the mode-coupling functional in Eqs.~(\ref{eq:MCT-v}):
1983: \bea
1984: {\cal F}_{q}[\tilde{f}] &=& (\rho / 16 \pi^{3} q^{4})
1985: \int d{\vec k} \, S_{q} S_{k} S_{p} 
1986: \nonumber \\
1987: & &
1988: \qquad \qquad \qquad
1989: \times \,
1990: [{\vec q} \cdot {\vec k} \, c_{k} + {\vec q} \cdot {\vec p} \, c_{p}]^{2} \,
1991: \tilde{f_{k}} \, \tilde{f_{p}}, 
1992: \label{eq:BXX-1}
1993: \eea
1994: with $\vec p = \vec q - \vec k$.
1995: 
1996: \subsection{MCT equations for the solute molecule}
1997: 
1998: It is straightforward to generalize the preceding derivation to
1999: systems with a diatomic solute molecule provided the latter is
2000: considered as flexible. Therefore, let us use this modification of
2001: the problem. It will be assumed that the kinetic energy of the
2002: molecule is $\Sigma_a (m_a /2) \vec v_{a}^{\, 2}$ and that there is a
2003: binding potential $V^{f \ell} (\mid \vec r_A - \vec r_B \mid)$
2004: between the two interaction sites. The exact Eq.~(\ref{eq:GLE-u-a}) remains
2005: valid with the frequency matrix replaced by that of the flexible
2006: molecule $\Omega_q^{f \ell \, 2}$. It is defined via Eq.~(\ref{eq:GLE-u-b}) with
2007: the simple velocity correlator $J_{q}^{f \ell \, a b} = \delta^{a b} (k_B
2008: T /m_a)$.
2009: 
2010: The formulas of Sec.~\ref{subsec:B1} shall be applied with the extended set
2011: of densities $(\rho_{\vec q}, \rho_{\vec q}^A, \rho_{\vec q}^B)$
2012: and the corresponding longitudinal currents $(j_{\vec q}, j_{\vec
2013: q}^A, j_{\vec q}^B)$. Thus the index consists of three bits
2014: $\alpha = (\tau, \lambda, \vec q)$, where $\lambda = 1,2$
2015: discriminates between densities and currents, and $\tau = O, A, B$
2016: indicates solvent, atom $A$ and atom $B$, respectively. In
2017: the infinite dilution limit $N \to \infty$, the equations for the
2018: solute do not directly couple to those for the solvent. Therefore,
2019: Eq.~(\ref{eq:B7}) remains valid and one gets a modification of 
2020: Eq.~(\ref{eq:GLE-u-a}) for the solute:
2021: \bea
2022: & &
2023: \partial_{t}^{2} {\bf F}_{q}(t) + {\bf \Gamma}_{q} \, \partial_{t} {\bf F}_{q}(t) +
2024: {\bf \Omega}_{q}^{f \ell \, 2} \, {\bf F}_{q}(t) 
2025: \nonumber \\
2026: & &
2027: \qquad \qquad
2028: + \,
2029: \int_{0}^{t} dt' \, {\bf M}_{q}(t-t') \, \partial_{t'} {\bf F}_{q}(t') = {\bf 0}.
2030: \label{eq:B12}
2031: \eea
2032: The relaxation kernel reads
2033: \begin{equation}
2034: M_{q}^{ab}(t) = (f_{\vec q}^{a}(t), f_{\vec q}^{b}) \, (m_{b} / k_{B}T), \quad
2035: f_{\vec q}^{a} = f_{a 2 {\vec q}}.
2036: \label{eq:B13}
2037: \end{equation}
2038: 
2039: The determination of the Gaussian distribution of the extended
2040: distinguished variables requires the inversion of the
2041: three-by-three matrix $(A_{\tau 1 \vec q}, A_{\sigma 1 \vec
2042: q})$. Making use of the infinite dilution limit $N \to \infty$,
2043: one gets
2044: \bea
2045: \langle g_{a} \rangle &=& w^{O}(a) \,
2046: \exp \biggl\{
2047: (1/2) \sum_{\vec q} \sum_{ab} 
2048: [(2 \rho / N) \, \delta^{ab} \, 
2049: c_{q}^{a} \, \tilde{\rho}_{\vec q} \, \tilde{\rho}_{\vec q}^{a *} 
2050: \nonumber \\
2051: & &
2052: - \,
2053: ({\bf w}_{q}^{-1})^{ab} \, \tilde{\rho}_{\vec q}^{a} \, \tilde{\rho}_{\vec q}^{b *} -
2054: (m_{a}/k_{B}T) \, \delta^{ab} \, \tilde{j}_{\vec q}^{a} \, \tilde{j}_{\vec q}^{b *}]
2055: \biggr\},
2056: \label{eq:B14}
2057: \eea
2058: where $w^{O}(a)$ is the distribution for the solvent variables
2059: given by Eq.~(\ref{eq:B9}). This expression is used to work
2060: out the fluctuating force as explained in connection with 
2061: Eq.~(\ref{eq:B10}):
2062: \bea
2063: f_{\vec q}^{a} &=& - i (\rho/N) (k_{B}T/m_{a}) \sum_{\vec k}
2064: [({\vec q}-{\vec k}) \cdot {\vec q}/q] \, c_{k}^{a} \,
2065: \rho_{\vec k}^{a} \, \rho_{{\vec q}-{\vec k}} 
2066: \nonumber \\
2067: & &
2068: \qquad \qquad \qquad \qquad
2069: + \,
2070: \delta f_{\vec q}^{a}.
2071: \label{eq:B15}
2072: \eea
2073: As above, the contributions due to $\delta f_{\vec q}^a$ are
2074: neglected and the remaining pair correlations are factorized. This
2075: leads to
2076: \begin{equation}
2077: {\bf M}_{q}(t) = {\bf \Omega}_{q}^{f \ell \, 2} \, {\bf m}_{q}(t),
2078: \label{eq:B16}
2079: \end{equation}
2080: where the functional for the kernel in Eq.~(\ref{eq:MCT-u-a}) reads
2081: \bea
2082: {\cal F}^{ab}_{q}[\tilde{\textit{\textbf f}}, \tilde{f}] &=& q^{-2} 
2083: \sum_{c} w^{ac}_{q} \, (\rho / 8 \pi^{3})
2084: \int d{\vec k} \, ({\vec q} \cdot {\vec p} / q)^{2} 
2085: \nonumber \\
2086: & &
2087: \qquad \qquad \qquad \qquad
2088: \times \,
2089: S_{p} \, c_{p}^{c} \, c_{p}^{b} \, \tilde{f}_{k}^{cb} \, \tilde{f}_{p},
2090: \label{eq:BXX-2}
2091: \eea
2092: with $\vec p$ abbreviating $\vec q - \vec k$. Neglecting the
2093: friction term ${\bf \Gamma}_{q}$ in Eq.~(\ref{eq:B12}), 
2094: the MCT equations (\ref{eq:GLE-u}) and (\ref{eq:MCT-u}) 
2095: are derived for the flexible molecule.
2096: 
2097: Obviously, a theory for a rigid molecule can be obtained from one
2098: for a flexible molecule only within a quantum-mechanical
2099: approach. One has to consider the case where excitation energies
2100: for translational and rotational motion are small compared to the
2101: thermal energy, while the energies for
2102: vibrational excitations are large. Let
2103: us assume, that the formulas can be obtained by replacing all
2104: equilibrium averages in the preceding equations by the correct
2105: quantum mechanical ones, where the latter can be evaluated for the
2106: classical molecule model with five degrees of freedom. This
2107: amounts to replacing structure functions by the classical
2108: quantities, in particular the replacement of ${\bf \Omega}_q^{f
2109: \ell}$ by ${\bf \Omega}_q$. Thereby Eq.~(\ref{eq:B12}) produces 
2110: Eqs.~(\ref{eq:GLE-u}).
2111: 
2112: \begin{thebibliography}{10}
2113: 
2114: \bibitem{Megen95}
2115: W. van Megen, Transp. Theory Stat. Phys. {\bf 24},  1017  (1995).
2116: 
2117: \bibitem{Nauroth97}
2118: M. Nauroth and W. Kob, Phys. Rev. E {\bf 55},  657  (1997).
2119: 
2120: \bibitem{Gleim98}
2121: T. Gleim, W. Kob, and K. Binder, Phys. Rev. Lett. {\bf 81},  4404  (1998).
2122: 
2123: \bibitem{Gleim00}
2124: T. Gleim and W. Kob, Eur.\ Phys.\ J.\ B {\bf 13},  83  (2000).
2125: 
2126: \bibitem{Goetze92}
2127: W. G{\"o}tze and L. Sj{\"o}gren, Rep. Prog. Phys. {\bf 55},  241  (1992).
2128: 
2129: \bibitem{Goetze99}
2130: W. G{\"o}tze, J. Phys.: Condensed Matter {\bf 11},  A1  (1999).
2131: 
2132: \bibitem{Schilling97}
2133: R. Schilling and T. Scheidsteger, Phys. Rev. E {\bf 56},  2932  (1997).
2134: 
2135: \bibitem{Theis98}
2136: C. Theis and R. Schilling, J. Non-Cryst. Solids {\bf 235--237},  106  (1998).
2137: 
2138: \bibitem{Fabbian99b}
2139: L. Fabbian, A. Latz, R. Schilling, F. Sciortino, P. Tartaglia, and C. Theis,
2140: Phys. Rev. E {\bf 60},  5768  (1999).
2141: 
2142: \bibitem{Winkler00}
2143: A. Winkler, A. Latz, R. Schilling, and C. Theis, cond-mat/0007276.
2144: 
2145: \bibitem{Theis00}
2146: C. Theis, F. Sciortino, A. Latz, R. Schilling, and P. Tartaglia,
2147: Phys. Rev. E {\bf 62},  1856  (2000).
2148: 
2149: \bibitem{Letz00}
2150: M. Letz, R. Schilling, and A. Latz, Phys. Rev. E  , in print.
2151: 
2152: \bibitem{Franosch97c}
2153: T. Franosch, M. Fuchs, W. G{\"o}tze, M.~R. Mayr, and A.~P. Singh,
2154: Phys. Rev. E {\bf 56},  5659  (1997).
2155: 
2156: \bibitem{Goetze00c}
2157: W. G{\"o}tze, A.~P. Singh, and $\mbox{Th}$. Voigtmann, Phys. Rev. E {\bf 61},
2158:   6934  (2000).
2159: 
2160: \bibitem{Franosch98}
2161: T. Franosch, W. G{\"o}tze, M.~R. Mayr, and A.~P. Singh, J. Non-Cryst. Solids
2162:   {\bf 235--237},  71  (1998).
2163: 
2164: \bibitem{Fuchs99b}
2165: M. Fuchs and $\mbox{Th}$. Voigtmann, Philos.\ Mag.\ B {\bf 79},  1799  (1999).
2166: 
2167: \bibitem{Kawasaki97}
2168: K. Kawasaki, Physica A {\bf 243},  25  (1997).
2169: 
2170: \bibitem{Chong98b}
2171: S.-H. Chong and F. Hirata, Phys. Rev. E {\bf 58},  6188  (1998).
2172: 
2173: \bibitem{Chandler72}
2174: D. Chandler and H.~C. Andersen, J. Chem. Phys. {\bf 57},  1930  (1972).
2175: 
2176: \bibitem{Hansen86}
2177: J.-P. Hansen and I.~R. McDonald, {\em Theory of Simple Liquids}, 2nd ed.
2178:   (Academic Press, London, 1986).
2179: 
2180: \bibitem{Chong98}
2181: S.-H. Chong and F. Hirata, Phys. Rev. E {\bf 57},  1691  (1998).
2182: 
2183: \bibitem{Franosch97}
2184: T. Franosch, M. Fuchs, W. G{\"o}tze, M.~R. Mayr, and A.~P. Singh,
2185: Phys. Rev. E {\bf 55},  7153  (1997).
2186: 
2187: \bibitem{Fuchs98}
2188: M. Fuchs, W. G{\"o}tze, and M.~R. Mayr, Phys. Rev. E {\bf 58},  3384  (1998).
2189: 
2190: \bibitem{Hoye77}
2191: J.~S. H{\o}ye and G. Stell, J. Chem. Phys. {\bf 66},  795  (1977).
2192: 
2193: \bibitem{Sullivan81}
2194: D.~E. Sullivan and C.~G. Gray, Mol. Phys. {\bf 42},  443  (1981).
2195: 
2196: \bibitem{Goetze91b}
2197: W. G{\"o}tze,  in {\em Liquids, Freezing and Glass Transition}, edited by J.-P.
2198:   Hansen, D. Levesque, and J. Zinn-Justin (North-Holland, Amsterdam, 1991), p.\
2199:   287.
2200: 
2201: \bibitem{Goetze00}
2202: W. G{\"o}tze and M.~R. Mayr, Phys. Rev. E {\bf 61},  587  (2000).
2203: 
2204: \bibitem{Franosch97b}
2205: T. Franosch and A.~P. Singh, J. Chem. Phys. {\bf 107},  5524  (1997).
2206: 
2207: \bibitem{Franosch98b}
2208: T. Franosch and A.~P. Singh, J. Non-Cryst. Solids {\bf 235--237},  153  (1998).
2209: 
2210: \bibitem{Franosch94}
2211: T. Franosch and W. G{\"o}tze, J. Phys.: Condens. Matter {\bf 6},  4807  (1994).
2212: 
2213: \bibitem{Dixon90}
2214: P.~K. Dixon, L. Wu, S.~R. Nagel, B.~D. Williams, and J.~P. Carini,
2215: Phys. Rev. Lett. {\bf 65},  1108  (1990).
2216: 
2217: \bibitem{Fuchs92b}
2218: M. Fuchs, I. Hofacker, and A. Latz, Phys. Rev. A {\bf 45},  898  (1992).
2219: 
2220: \bibitem{Mori73}
2221: H. Mori and H. Fujisaka, Prog. Theor. Phys. {\bf 49},  764  (1973).
2222: 
2223: \end{thebibliography}
2224: 
2225: 
2226: \begin{figure}
2227: \centerline{\scalebox{0.55}{\includegraphics{figure1.ps}}}
2228: \caption{Phase diagram of a dilute solute of symmetric dumbbell
2229: molecules with elongation $\zeta$ consisting of two fused hard spheres
2230: which are immersed in a hard-sphere system (HSS) with packing
2231: fraction $\varphi$. The horizontal line marks the liquid-glass
2232: transition at the critical packing fraction
2233: $\varphi_c = 0.516$. The other full line is the curve $\zeta_c
2234: (\varphi)$ of critical elongations for a type-$A$ transition
2235: between phases II and III. In phase II dipole fluctuations of the
2236: solute relax to zero for long times, while they are frozen in
2237: phase III. The value $\zeta_c = \zeta_c (\varphi_c) = 0.380$ 
2238: is marked by an arrow. The dashed
2239: line is the corresponding transition curve calculated 
2240: in Ref.~\protect\cite{Franosch98b}; 
2241: it terminates at $\zeta_c^t = 0.297$.}
2242: \label{fig:phase}
2243: \end{figure}
2244: 
2245: 
2246: \newpage\noindent
2247: 
2248: 
2249: \begin{figure}
2250: \centerline{\scalebox{0.85}{\includegraphics{figure2.ps}}}
2251: \caption{
2252: Nonergodicity parameters $f_q^{x \,c}$ (heavy full lines)
2253: for the molecule's arrested number-density
2254: fluctuations $(x = N)$ and 
2255: ``charge''-density fluctuations $(x = Z)$ 
2256: for the critical packing fraction
2257: $\varphi = \varphi_c$.
2258: The elongation parameter $\zeta = 0.80$ is
2259: representative for strong steric hindrance for reorientational
2260: motion. The light full lines are evaluated with 
2261: Eq.~(\ref{eq:Fab-site-tensor-2}) with the
2262: nonergodicity parameters $f^c (q \ell 0)$ obtained in 
2263: Ref.~\protect\cite{Franosch97c} from a theory based on a
2264: tensor-density description. The dotted lines show the
2265: contributions to Eq.~(\ref{eq:Fab-site-tensor-2}) 
2266: from different angular-momentum index
2267: $\ell$. Here and in the following figures the diameter of the
2268: spheres is used as unit of length, $d = 1$. }
2269: \label{fig:fq-strong}
2270: \end{figure}
2271: 
2272: 
2273: \newpage\noindent
2274: 
2275: 
2276: \begin{figure}
2277: \centerline{\scalebox{0.85}{\includegraphics{figure3.ps}}}
2278: \caption{Results as in Fig.~\protect\ref{fig:fq-strong} 
2279: but for small elongations which are
2280: representative for weak steric hindrance for reorientational
2281: motion. The relative distance from the transition point between
2282: phases II and III is $(\zeta - \zeta_c) / \zeta_c = 0.347$. The
2283: heavy full line shows the result of the present theory for 
2284: $\zeta = 0.512$. The light full line
2285: shows the result for $\zeta=0.400$ based on 
2286: Ref.~\protect\cite{Franosch97c}.}
2287: \label{fig:fq-weak}
2288: \end{figure}
2289: 
2290: 
2291: \newpage\noindent
2292: 
2293: 
2294: \begin{figure}
2295: \centerline{\scalebox{0.85}{\includegraphics{figure4.ps}}}
2296: \caption{Critical nonergodicity parameters $f_q^{x \, c}$ for the solute
2297: as function of the elongation parameter $\zeta$ for the 
2298: wave numbers $q = 3.4 (a), \, 7.0 (b), \, 10.6
2299: (c), \, 14.2 (d)$, and 17.4 $(e)$. The arrow 
2300: marks the transition point from phase II to phase III at $\zeta_c
2301: = 0.380$. }
2302: \label{fig:fq-vs-zeta}
2303: \end{figure}
2304: 
2305: 
2306: \newpage\noindent
2307: 
2308: 
2309: \begin{figure}
2310: \centerline{\scalebox{0.65}{\includegraphics{figure5.ps}}}
2311: \caption{Correlators $\phi_q^N (t)$ (solid lines)
2312: and $\phi_q^Z (t)$ (dashed lines)
2313: for two intermediate wave numbers $q$ as function of the
2314: logarithm of time $t$. The decay curves at the
2315: critical packing fraction $\varphi_c$ for number-density and
2316: ``charge''-density correlators are shown as dotted lines and marked
2317: by $N$ and $Z$, respectively. Only a few solutions of glass states
2318: are shown for $\phi_q^Z (t)$ in order to avoid overcrowding of the
2319: figure. The distance parameter is 
2320: $\epsilon = (\varphi - \varphi_c) / \varphi_c = \pm 10^{-x}$.
2321: The full circles and squares mark the characteristic
2322: times $t_{\sigma}$ and $t'_{\sigma}$, respectively, according to
2323: Eq.~(\protect\ref{eq:t-sigma}) 
2324: for $x = 1$, 2, 3 and 4.
2325: The unit of time is chosen here and in the
2326: following figures such that the thermal velocity of the solvent
2327: reads $v = 1$.} 
2328: \label{fig:NN-ZZ-t}
2329: \end{figure}
2330: 
2331: 
2332: \newpage\noindent
2333: 
2334: 
2335: \begin{figure}
2336: \centerline{\scalebox{0.80}{\includegraphics{figure6.ps}}}
2337: \caption{Reorientational correlators $C_\ell(t)$ for $\ell = 1$
2338: (dashed lines) and $\ell = 2$ (full lines) 
2339: for two elongations $\zeta$. The correlators for the
2340: critical packing fraction $\varphi = \varphi_c$ are shown as
2341: dotted lines marked with $c_\ell$. The distance parameters
2342: $\epsilon = (\varphi - \varphi_c) / \varphi_c$ are -0.01 (faster
2343: decay) and -0.001 (slower decay). 
2344: The full circles and squares mark the
2345: corresponding time scales $t_\sigma$ and $t_\sigma^\prime$,
2346: respectively, from Eq.~(\protect\ref{eq:t-sigma}). 
2347: The open circles and squares on the curves mark the characteristic
2348: time scales $\tau_{\beta}^{\ell}$ and $\tau_{\alpha}^{\ell}$,
2349: respectively, 
2350: defined by $C_{\ell}(\tau_{\beta}^{\ell}) = f_{\ell}^{c}$ and
2351: $C_\ell (\tau_\alpha^\ell) = f_\ell^c / 2$. }
2352: \label{fig:C1-C2}
2353: \end{figure}
2354: 
2355: 
2356: \newpage\noindent
2357: 
2358: 
2359: \begin{figure}
2360: \centerline{\scalebox{0.80}{\includegraphics{figure7.ps}}}
2361: \caption{Susceptibility master spectra 
2362: $\tilde{\chi}^{\prime \prime} (\tilde{\omega})$ of
2363: the $\alpha$-process as function of the logarithm of the rescaled
2364: frequency $\tilde \omega = \omega t_\sigma^\prime$ (see text). 
2365: Upper panel: 
2366: the curves $\ell = 1$ and 2 refer to the response for the dipole
2367: and quadrupole, respectively, for elongation $\zeta = 0.80$. The
2368: curve $\ell = 0$ refers to the susceptibility master spectrum of 
2369: the dimensionless longitudinal elastic modulus $m_{q=0} (t)$ of the HSS.
2370: The dashed lines exhibit the von Schweidler tails, Eq.~(\ref{eq:von-b}). 
2371: The dotted lines are fits by Kohlrausch spectra 
2372: $\tilde \chi_K^{\prime \prime}(\tilde{\omega})$ 
2373: with stretching exponents $\beta = 0.97$, 0.88 and 0.63
2374: chosen for $\ell = 1, 2$ and 0, respectively, so that 
2375: the maximum and the full width
2376: at the half maximum $W$ in decades of 
2377: $\tilde \chi_K^{\prime \prime}(\tilde{\omega})$ agree
2378: with those of $\tilde \chi^{\prime \prime}(\tilde{\omega})$. 
2379: The position of the susceptibility maximum is $\tilde \omega_{\max} = 0.337$ (0.927,
2380: 2.69) and the width is $W = 1.17$ (1.28, 1.76) for $\ell = 1$ (2, 0).
2381: Lower panel: corresponding results for $\zeta = 0.43$.
2382: The stretching exponent $\beta$, the maximum position
2383: $\tilde \omega_{\max}$ and the width $W$ for $\ell=1$ ($\ell=2$) are
2384: $\beta = 0.79$ (0.71), 
2385: $\tilde \omega_{\max} = 17.5$ (2.11)
2386: and
2387: $W = 1.42$ (1.57), respectively.}
2388: \label{fig:alpha-spectra}
2389: \end{figure}
2390: 
2391: 
2392: \newpage\noindent
2393: 
2394: 
2395: \begin{figure}
2396: \centerline{\scalebox{0.70}{\includegraphics{figure8.ps}}}
2397: \caption{Double logarithmic presentation of the normalized
2398: fluctuation spectra of
2399: the dipole-reorientation $\alpha$-processes, 
2400: $\tilde{\chi}^{\prime \prime}_{1} (\tilde{\omega})
2401: \tilde{\omega}_{\max} / f_{1}^{c} \tilde{\omega}$, as function of 
2402: $\tilde{\omega} / \tilde{\omega}_{\max}$.
2403: Here $\tilde{\omega}_{\max}$ denotes the position of the susceptibility
2404: maximum. Following Dixon {\it et al.} \protect\cite{Dixon90} the vertical axis
2405: is rescaled by $w^{-1}$ and the horizontal one by $w^{-1} (1 +
2406: w^{-1})$, where $w = W / W_D$ is the ratio of the logarithmic full width
2407: at half maximum $W$ of the susceptibility peak to the same quantity $W_D$ of
2408: a Debye-peak. The open circles reproduce some of the dielectric-loss
2409: results for glycerol \protect\cite{Dixon90}. 
2410: The three full lines from the top
2411: are the results for $\zeta = 0.80$, 0.60 and 0.43,
2412: successively,
2413: although the upper two curves cannot be distinguished within the 
2414: resolution of the figure.
2415: The dotted and the dashed lines exhibit 
2416: the Kohlrausch fit with the stretching exponent
2417: $\beta = 0.97$ and the von-Schweidler-law tail, respectively,
2418: for the $\zeta = 0.80$--spectrum.}
2419: \label{fig:Nagel-plot}
2420: \end{figure}
2421: 
2422: 
2423: \newpage\noindent
2424: 
2425: 
2426: \begin{figure}
2427: \centerline{\scalebox{0.80}{\includegraphics{figure9.ps}}}
2428: \caption{Dipole correlators $C_1$(t) for the distance parameter
2429: $\epsilon = - 10^{-4}$ for three elongations $\zeta$. The full circle
2430: and square indicate the times $t_\sigma$ and $t_\sigma^\prime$,
2431: respectively, from Eq.~(\ref{eq:t-sigma}).
2432: The dashed lines are calculated from 
2433: Eqs.~(\ref{eq:GLE-C1}) and (\ref{eq:MCT-MSD-C1}). 
2434: The plateaus $f_1^c = 0.905$ (0.769, 0.376) for
2435: $\zeta = 0.80$ (0.60, 0.43) are shown by dashed horizontal lines. The
2436: full lines exhibit $C_1(t)$ for the same states, but evaluated from
2437: equations derived in analogy to 
2438: Eqs.~(\ref{eq:GLE-C2})--(\ref{eq:m2-final}),
2439: and their plateaus 
2440: $f_1^c = 0.907$ (0.782, 0.402) are indicated by full horizontal lines. }
2441: \label{fig:compare-C1}
2442: \end{figure}
2443: 
2444: 
2445: \newpage\noindent
2446: 
2447: 
2448: \begin{figure}
2449: \centerline{\scalebox{0.55}{\includegraphics{figure10.ps}}}
2450: \caption{Dipole correlators $C_1 (t)$ for a dumbbell of two fused
2451: hard spheres of diameters $d$ and distance $\zeta d$ between the
2452: centers moving in a liquid of hard spheres with diameter $d$
2453: for a distance parameter
2454: $(\varphi-\varphi_{c})/\varphi_{c} = - 10^{-x}$.
2455: The dashed lines reproduce the results from Fig.~\ref{fig:C1-C2} and 
2456: refer to a symmetric molecule with masses of the two atoms being 
2457: equal to the mass $m$ of the solvent particles $m_A = m_B = m$. 
2458: The full lines
2459: exhibit the results for an asymmetric dumbbell with 
2460: $m_A = 10 m$, $m_B = m$. 
2461: In the main frame the results for different states
2462: are successively shifted horizontally by 2 decades in order
2463: to avoid overcrowding.
2464: The inset shows the transient dynamics 
2465: on a linear time axis. }
2466: \label{fig:C1-inertia}
2467: \end{figure}
2468: 
2469: \end{document}
2470: