cond-mat0003368/tk.tex
1: \documentstyle[pre,aps,preprint,epsfig]{revtex}
2: \renewcommand{\thesubsection}{\arabic{subsection}} 
3:  \renewcommand{\theenumi}{\roman{enumi}}
4:  \renewcommand{\labelenumi}{(\theenumi)}   
5: \tightenlines   
6: \begin{document}
7: 
8: 
9: \title{THE VISCOUS SLOWING DOWN OF SUPERCOOLED LIQUIDS
10:  AND THE GLASS TRANSITION: PHENOMENOLOGY, CONCEPTS, AND MODELS}
11: \author{G. TARJUS}\address{
12: Laboratoire  de Physique Th{\'e}orique des  Liquides, Universit{\'e} Pierre et
13: Marie Curie, 4 Place Jussieu, 75252 Paris cedex 05, France}
14: \author{D. KIVELSON}
15: \address{Department of Chemistry and Biochemistry,
16: University of California, Los Angeles, CA 90095, USA}
17: \maketitle
18: \begin{abstract}
19: The viscous slowing down of  supercooled  liquids that leads to  glass
20: formation   can  be considered  as a   classical,  and  is assuredly a
21: thoroughly studied, example of a "jamming process". In this review, we
22: stress the distinctive features characterizing the phenomenon. We also
23: discuss  the  main  theoretical approaches, with   an  emphasis on the
24: concepts    (free   volume,  dynamic    freezing    and  mode-coupling
25: approximations, configurational    entropy and     energy   landscape,
26: frustration) that could  be  useful in  other areas of   physics where
27: jamming  processes are encountered.
28: \end{abstract}
29: \vspace{0.5cm}
30: 
31: To appear in:
32: {\it Jamming and Rheology: Constrained Dynamics on Microscopic and Macroscopic 
33: Scales}\\
34: {\it S. Edwards, A. liu, and S. Nagel Eds.}
35: 
36: 
37: \section{INTRODUCTION}
38: When cooling a liquid,  usually under isobaric P=1atm  conditions, one
39: can often  bypass   crystallization, thereby obtaining   a supercooled
40: liquid  that    is metastable   relative  to  the  crystal.   When the
41: temperature is further  lowered, the viscosity of  the liquid, as well
42: as the relaxation times associated  with the primary ($\alpha$) relaxation
43: of   all kinds  of structural,   dielectric,  macro- and  micro-scopic
44: observables, increase rapidly, until a temperature is reached at which
45: the liquid can no longer flow and equilibrate in the time scale of the
46: experiment.   The  system effectively   appears  as a  rigid amorphous
47: material and is then called a glass. Glass formation thus results from
48: the    strong  viscous  slowing down   of   a  liquid  with decreasing
49: temperature\footnote{Although not as widely used,  there are other ways
50: of generating  glassy  structures,  such  as  vapor  deposition,  in situ
51: polymerization  or chemical reactions.\cite{R3}} ,   a slowing down that
52: can  be considered as  a classical example of   a ``jamming process''. A
53: characteristic  of this process    that is unanimously recognized,   a
54: unanimity rare in  this otherwise quite  open and controversial field,
55: is that it  is a dynamic effect.  The  so-called ``glass transition'' is
56: not a {\it bona fide} thermodynamic phase transition, but represents a
57: crossover  below  which a liquid   falls  out of   equilibrium on  the
58: experimental time scale.  The transition temperature, $T_g$ depends on
59: this time scale,  set either by  the observation  time (corresponding,
60: for  instance, to a  relaxation  time of  $10^2$ or  $10^3$  sec or  a
61: viscosity of $10^{13}$ Poise) and/or by the cooling rate (in a typical
62: differential  scanning calorimetry  measurement, $10$  K  per minute).
63: The dependence on cooling rate is,  however, weak, a difference of few
64: K in $T_g$  for an order-of-magnitude change of  the rate; this  is so
65: because on further  lowering   of the temperature  the   viscosity and
66: $\alpha$-relaxation times rapidly become enormous and out  of reach of any
67: experimental technique. 
68: 
69: In  the following, we shall focus  on the jamming process occurring in
70: the       supercooled     liquid     state.\cite{R1,R2}     Both   the
71: crystal\footnote{Note  that  for       several    liquids,  such    as
72: m-fluoroaniline  and  dibutylphtalate  at  atmospheric   pressure  and
73: atactic polymers, crystallization has  never been observed.}, which is
74: the stable phase  below the melting point  $T_m$ but can be ignored in
75: discussing the  glass transition, and  the glassy state itself will be
76: excluded from  the  present discussion.  By  appropriately eliminating
77: the  crystal  (experimentally  as  well as theoretically),  metastable
78: supercooled  liquids can  be  treated by   equilibrium thermodynamics,
79: statistical  mechanics, and conventional linear-response   formalisms.
80: Glasses  on   the other  hand,    although mechanically  stable,   are
81: out-of-equilibrium states;  especially   near    $T_g$, they   display
82: nonlinear responses and relaxations  known as aging or annealing,  and
83: their properties depend on  their history of preparation. \cite{R4,R5}
84: These phenomena will not be considered in this review.
85: 
86: \section{SALIENT   PHENOMENOLOGY }
87: The distinctive feature   of  glass-forming liquids is  the  dramatic,
88: continuous  increase of   viscosity  and  $\alpha$-relaxation  times  with
89: decreasing  temperature.  This sort of  jamming is observed in a large
90: variety   of  substances:  covalently   bonded  systems  like $SiO_2$,
91: hydrogen-bonded    liquids,  ionic   mixtures,   polymers,   colloidal
92: suspensions, molecular van der Waals liquids,  etc.  The emphasis will
93: be placed on those liquids (the vast majority) that  do not form 2- or
94: 3-dimensional networks   of strong bonds because   they  show the most
95: striking behavior when  passing from the high-temperature liquid phase
96: to the deeply supercooled and very viscous regime.
97: 
98: \subsection{Strong, super-Arrhenius  T-dependence of
99: viscosity  and   $\alpha$-relaxation times    }  The viscosity   $\eta$  and
100: $\alpha$-relaxation  times can change  by $15$ orders   of magnitude for a
101: mere decrease of  temperature  by  a factor two\footnote{The    (shear)
102: viscosity   $\eta$   can be   related   to  a   time characteristic  of
103: $\alpha$-relaxation, the average  shear stress relaxation time  $\tau_s$, by
104: $\eta=G_\infty  \tau_s$,   where  $G_\infty$ is the infinite-frequency shear
105: modulus; $G_\infty$ is    typically   of   the order   of
106: $10^{10}-10^{11}$ erg.cm$^{-3}$, so that  a viscosity of $10^{13}$
107: Poise roughly
108: corresponds to a time  of  $10^2$  or  $10^3$ sec.}. Such  a  dramatic
109: variation is conveniently represented on a logarithmic plot of $\eta$ or
110: $\tau_\alpha$ versus $1/T,$  i.e., an Arrhenius plot:  see  Fig.   1a.  A
111: system like $GeO_2$, an   example  of  a network-forming   system,  is
112: characterized  by   an  almost linear  variation,  which  indicates an
113: Arrhenius temperature dependence.  For all other liquids on the figure
114: there     is   a  marked  upward     curvature,   which   represents a
115: faster-than-Arrhenius, or  super-Arrhenius, temperature dependence and
116: is often described by an empirical Vogel-Fulcher-Tammann (VFT) formula
117: (also called Williams-Landel-Ferry formula  in the context of polymers
118: studies\cite{R6}),
119: \begin{equation}
120: \label{eq:1}
121: \tau_\alpha =\tau_0\exp\left(D\frac{T}{T-T_0}\right),
122: \end{equation}
123: where   $\tau_0$, $D$ and  $T_0<T_g$  are adjustable  parameters. On the
124: basis of  such Arrhenius plots, with the  temperature scaled by $T_g$,
125: Angell   proposed  the  now   standard classification of glass-forming
126: liquids  into   {\it  strong}    (Arrhenius-like) and  {\it   fragile}
127: (super-Arrhenius)  systems;\cite{R7} in Eq.~(\ref{eq:1}), the smaller
128: the value of $ D$, the more fragile the liquid.  There are, of course,
129: alternative fitting  formulas that have been used,  some  which do not
130: imply a singularity at a nonzero temperature as does the expression in
131: Eq.~(\ref{eq:1}).\cite{R8}
132: 
133: A  different   way of representing  the    phenomenon is to   plot the
134: effective activation free energy for $\alpha$-relaxation, $E(T)$, obtained
135: from
136: 
137: \begin{equation}
138: \label{eq:2}
139: \tau_\alpha =\tau_{\alpha,\infty}\exp\left(\frac{E(T)}{k_BT}\right),
140: \end{equation}
141: where $k_B$ is  the Boltzmann constant and  $\tau_{\alpha,\infty}$ is a high-$T$
142: relaxation time, or from a similar equation for the viscosity. This is
143: illustrated in Fig. 1b, where the temperature has been scaled for each
144: liquid  to a  temperature  T* above  which  the dependence is  roughly
145: Arrhenius-like.  Although  the  determination of $T^*$ is   subject to
146: some  uncertainty,\cite{R8,R9} the  procedure emphasizes the crossover
147: from Arrhenius-like to super-Arrhenius behavior that is typical of and
148: quite distinct  in most supercooled  liquids.  The appreciable size of
149: the effective activation free  energies $E(T)$, namely, $40 k_BT_g$ at
150: the glass    transition,    is  indicative   of thermally    activated
151: dynamics.  Such a large   effective activation free energy  for weakly
152: bonded  fragile   molecular  liquids  such as  orthoterphenyl    is an
153: intriguing feature of the phenomenology.  Another peculiar property of
154: $E(T)$ for fragile systems is  that it increases significantly between
155: $T^*$ and $T_g$ (a factor $3$, i.e., a factor of  $5$ or $6$ in units
156: of  the  thermal energy $k_BT$,  for weakly   bonded fragile liquids).
157: Such a variation is not  commonly encountered.   For instance, in  the
158: field of critical phenomena, the slowing  down of dynamics that occurs
159: when approaching  the critical point  is  usually characterized  by  a
160: power law  growth  of  the  relaxation time;  in   terms of  effective
161: activation  free energy, this  corresponds to a logarithmic growth and
162: it  is  slower  than  the  variation described   by  the VFT  formula,
163: Eq.~(\ref{eq:1}).  Unusually strong  slowing  down, with exponentially
164: growing times similar to Eq.~(\ref{eq:1}), is found in some disordered
165: systems  like the  random   field Ising   model and  it  is known   as
166: ``activated dynamic    scaling''.\cite{R10}
167: 
168: \subsection{Nonexponential         relaxations}
169: In an ``ordinary'' liquid  above the melting point, relaxation functions
170: are usually  well described,  after some transient  time, by  a simple
171: exponential decay.  Deviations are  observed,   but they are   neither
172: systematic   nor  very   marked.   The    situation changes at   lower
173: temperatures, and the $\alpha$-relaxation is no longer characterized by an
174: exponential   decay.   A   better representation  is    provided  by a
175: ``stretched exponential'' (or Kohlrausch-Williams-Watts function),
176: 
177: \begin{equation}
178: \label{eq:3}
179: f_\alpha(t)\propto \exp\left[-\left(\frac{t}{\tau_\alpha}\right)^\beta \right],
180: \end{equation}
181: where $\beta$  is  the stretching parameter;   the smaller $\beta$  the more
182: ``stretched''  the relaxation.  Although  not unambiguously established,
183: the degree  of departure   from exponential behavior,   or stretching,
184: appears  to  increase   (i.e., $\beta$  decreases)   with  decreasing
185: temperature.  
186: 
187: Alternatively, in frequency space, the spectrum  of the imaginary part
188: of the susceptibility, which is characterized by a  peak at a frequency
189: $\omega_\alpha \propto 1/ \tau_\alpha $, tends to be broader (when  plotted as a function
190: of $log(\omega)$  ) than the  simple Lorentzian or  Debye spectrum that is
191: just the  Fourier transform of  the  time-derivative of an exponential
192: relaxation   function  (see  Fig.  2).   Fitting  formulas  related to
193: Eq.~(\ref{eq:3}),  formulas  like   the  Cole-Davidson function    for
194: frequency-dependent susceptibilities, $(1-i(\omega/\omega_\alpha ))^{-\beta'}$, are used
195: to fit  the spectroscopic data,  but similar trends  are observed: the
196: $\alpha$ peak, as  observed  for instance   in the imaginary  part of  the
197: dielectric susceptibility as a  function of $log(\omega)$, broadens as the
198: temperature  is  lowered  towards $T_g$,\cite{R11}   which   indicates
199: increasing   departure from  Debye/exponential  behavior.   Except for
200: network-forming  systems,  the stretching of  the   $\alpha$ relaxation is
201: significant ($\beta$ is  typically  between $0.3$  and $0.6$ for  fragile
202: liquids at $T_g$).  However, and  this point may  not have been  given
203: enough attention, the stretching, or broadening in frequency space, is
204: relatively small  when compared to the  extremely rapid variation with
205: temperature of  the  $\alpha$-relaxation  time  itself.   This  is  to  be
206: contrasted for instance  with the activated critical slowing discussed
207: above. There, the power law growth  of the activation free energy when
208: the  temperature is decreased toward the  critical  point comes with a
209: more striking stretching of  the relaxation function  that occurs on a
210: logarithmic scale:  in this case,  in place of a stretched exponential
211: behavior as in Eq.~(\ref{eq:3}), $\ln(f(t))$  goes  as some power of  
212: $(1/\ln(t))$.\cite{R12}
213: \subsection{No  marked  changes  in  structural  quantities}
214: 
215: It is tempting to associate the huge increase in $\alpha$-relaxation times
216: and  viscosity with  the growth   of  a {\it structural}   correlation
217: length.  However,   no such growth   has been   detected so   far   in
218: supercooled liquids.   Quite    to  the contrary,   the   variation of
219: structure in  liquids  and glasses, as measured  in  neutron and X-ray
220: diffraction experiments,  appear rather  bland\cite{R13,R14} (see Fig.
221: 3).  The ordinary,  high-temperature liquid has only short-range order
222: whose signature in the static structure  factor $S(Q)$ is a broad peak
223: (or a split peak for some molecular systems as illustrated in Fig.  3)
224: at a wave vector  $Q$ that roughly corresponds in  real space to  some
225: typical   mean   distance  between  neighboring   molecules.  As   the
226: temperature is lowered and   the supercooled regime is entered,  there
227: are small,  continuous variations of  the structure factor that mostly
228: reflect the change in density (typically, a $5\%$ change between $T_m$
229: and $T_g$) and, possibly,  some adjustments in the local  arrangements
230: of the molecules.  There  is  no sign,   however, of  a  significantly
231: growing correlation length, nor of the appearance of a super-molecular
232:  length. 
233: 
234: In  network-forming  and
235: $H$-bonded systems, an  additional ``pre-peak'' is sometimes detectable at
236: wave  vectors somewhat  lower than  that of the  main peak,  but it is
237: attributed  to specific effects  induced  by the strongly  directional
238: intermolecular  bonds and not to  a length  scale that would correlate
239: with  the   viscous   slowing  down.\cite{R15}
240: 
241: In
242: contrast to this lack of structural  signature for the existence of an
243: increasing  super-molecular   correlation     length   with decreasing
244: temperature, there  is significant evidence,  as discussed below, that
245: corresponding ``dynamical''  correlation length do  exist.
246: 
247: \subsection{Rapid entropy decrease and Kauzmann paradox }
248: 
249: 
250: The absence of marked changes  in the structure, {\it  at least at the
251: level  of two-particle density correlations}, or  of a strong increase
252: in  any directly measured static  susceptibility is a puzzling feature
253: of  the jamming  process  associated  with  glass formation. The  only
254: static quantity that  shows behavior  that might  be relevant is   the
255: entropy.  Below the  melting point, $T_m$,  the heat capacity $C_p(T)$
256: of a supercooled  liquid is  larger   than that of the   corresponding
257: crystal. (At $T_g$, the $C_p$ of  the liquid drops  to a value that is
258: characteristic of the glass and is close to the  $C_p$ of the crystal,
259: but this is  a  consequence of the   system no longer being   properly
260: equilibrated.) As a result of this ``excess'' heat capacity, the entropy
261: difference  between  the liquid  and     the crystal decreases    with
262: temperature, typically by a factor of $3$ between  $T_m$ and $T_g$ for
263: fragile liquids. The effect is illustrated in Fig.  4 and leads to the
264: famous Kauzmann   paradox:\cite{R16}  if  the  entropy   difference is
265: extrapolated  to  temperatures below  $T_g$,  its  extrapolated  value
266: vanishes  at   some nonzero temperature  $T_K$,   which results in the
267: unpleasant  feature that the  entropy of the  liquid  becomes equal to
268: that  of the  crystal (even  more  unpleasant: if the extrapolation is
269: carried to still lower temperatures, the entropy of the liquid becomes
270: negative, which violates the third law of thermodynamics). The paradox
271: is that this {\it extrapolated }entropy crisis is avoided for a purely
272: dynamic  reason, the intervention  of the glass transition: what would
273: occur if  one were  able to keep  the supercooled  liquid equilibrated
274: down to temperatures  below $T_g$?  There  are certainly  many ways to
275: answer the  question. The paradox  could be resolved  by the existence
276: between $T_g$ and $T_K$ of an  intrinsic limit of metastability of the
277: liquid\cite{R16} or of  a  second-order phase transition\footnote{Note
278: that a  low-$T$  first-order transition  does not resolve  the paradox
279: because it can be  supercooled.}  (a speculation that gains additional
280: credibility with the  observation  that the VFT  temperature  $T_0$ at
281: which the  extrapolated  viscosity and  $\alpha$-relaxation  times diverge
282: (see     Eq.~(\ref{eq:1}))      is     often    found    close      to
283: $T_K$\cite{R17,R18}).  Even  more  simply,  one might   find  that the
284: extrapolation breaks down above  $T_K$ and that the entropy-difference
285: curve levels off and  goes smoothly to zero  at zero $K$, in  much the
286: same  way as it  does in the Debye  theory of crystals.   These are of
287: course all speculations, but it remains that the rapid decrease of the
288: entropy of the   supercooled liquid relative   to that of  the crystal
289: represents an intriguing    aspect  of the phenomenology   of  fragile
290: glass-formers.
291: 
292: \subsection{Two-step relaxation and secondary processes}
293: As we stressed before, the salient features related to glass formation
294: concern the long-time (low-frequency) primary or  $\alpha$ relaxation.  As
295: the $\alpha$-relaxation time increases with decreasing temperature, so too
296: does the window between this  time and typical microscopic, picosecond
297: or  sub-picosecond times.  When   the  relaxation function  is plotted
298: against the logarithm of the time, one then observes what is sometimes
299: called a ``two-step  relaxation''.  An illustration is given  in Fig.  5
300: by the dynamic  structure  factor  of the fragile   ionic glass-former
301: $Ca_{0.4}K_{0.6}       (NO_3)_{1.4}$    obtained     by        neutron
302: techniques.\cite{R19} At  high temperature, the relaxation function is
303: essentially a  one-step process.  However,  as the liquid becomes more
304: viscous, the relaxation proceeds in two steps  separated by a plateau.
305: Although the terminology is  far from being universally  accepted, the
306: approach to the plateau from the short-time side  is often referred to
307: as $\beta$ or fast-$\beta$ relaxation.  If one is  to fit the long-time part
308: by a stretched exponential  (Eq.~(\ref{eq:3})), there is a large range
309: of ``mesoscopic'' times that is not adequately described and that widens
310: as the temperature is lowered.  Power  law functions of time are often
311: used to reproduce the relaxation function in this mesoscopic range.
312: 
313: This two-step relaxation feature is common to all fragile liquids.  In
314: addition,  there  may  also  appear   additional secondary  processes,
315: detected  first by  Johari  and Goldstein in  dielectric spectroscopy.
316: \cite{R20}  Such  secondary processes,  whose   presence and  strength
317: strongly   vary from  one   liquid  to  another,  have  characteristic
318: frequencies that  are  intermediate  between  those of the   $\alpha$  and
319: fast-$\beta$   relaxations.   They  are  denoted Johari-Goldstein-$\beta$,
320: slow-$\beta$,  or simply  $\beta$    processes. \cite{R2,R21}  To   make the
321: description more complete,  one  should  also mention  the   so-called
322: ``boson   peak'' that may     be  present on  the  high-frequency   side
323: ($\sim  10^2-10^3Ghz$) of light  and   neutron scattering (or   absorption)
324: spectra.\cite{R2,R22} Here we  do not discuss either the slow-$\beta$
325:   processes or the boson
326: peak.   
327: 
328: \section{ A SELECTION OF QUESTIONS }
329: After this brief review of the salient aspects of the phenomenology of
330: supercooled liquids as they  get glassy, we discuss  in more detail  a
331: number  of  questions,  whose    answers  give  justification or   put
332: constraints on the  theoretical picture one  can build to  explain the
333: viscous slowing down. 
334: \subsection{How  universal   is the  behavior   of  glass-forming
335: liquids ?}
336: 
337: Universality  is a  key concept  in  physics and it  has  proven to be
338: central  in the   field of  critical phenomena.   By  the standards of
339: critical phenomena  studies, the  observed  behavior  of glass-forming
340: liquids is not universal, the main reason being that no singularity is
341: detected  experimentally  (or approached   asymptotically  close),  as
342: stressed above.  However, if one is willing to take  a broader view of
343: the  notion of universality,  one  can find considerable generality or
344: ``universality''  in   the   properties, those  mostly   associated with
345: long-time and  low-frequency phenomena, that characterize the approach
346: to   the   glass  transition.    For  instance,  the   super-Arrhenius
347: $T$-dependence of   the  viscosity and  $\alpha$-relaxation times  and the
348: nonexponential character of the  relaxation function are  observed for
349: virtually all glass-formers, be they polymeric, $H$-bonded, ionic, van
350: der Waals,    etc.,  with  the exception  of    a minority   of strong
351: network-forming systems; and, for a given liquid, these properties are
352: found  by  a   large  variety  of experimental   techniques,   such as
353: dielectric  relaxation,  light and neutron  scattering, NMR, viscosity
354: measurements,  specific   heat    spectroscopy,  volume  and  enthalpy
355: relaxation, optical probe methods.
356: 
357: The presence of an underlying "universality" is  supported by the fact
358: that experimental data  covering a wide range   of temperatures and  a
359: great diversity of substances can be collapsed onto master curves with
360: only a  small number  of species-dependent adjustable   parameters.  A
361: good   example   is  provided    by    the  scaling   plot   of    the
362: frequency-dependent dielectric susceptibility  proposed  by  Nagel and
363: coworkers.\cite{R11} As  shown in Fig.  6, the  data taken over a $13$
364: decade range of  frequencies for many different  liquids can be placed
365: with good accuracy  onto a single curve after  scaling with only three
366: parameters associated     with  the $\alpha$-peak     position,  width and
367: intensity.  The  master-curve  for the  temperature  dependence of the
368: effective activation free energy for viscosity and $\alpha$ relaxation put
369: forward  by Kivelson {\it et al.}  \cite{R8} is another example. It is
370: nonetheless fair  to  say that  the fits  resulting from these various
371: scaling procedures are far from perfect, which  leaves room for debate
372: and conflicting interpretations. One can also ask the question whether
373: the universality holds only  up to the implied high-frequency  cut-off
374: of   the  susceptibility  scaling   curve  of   Nagel   {\it  et  al.}
375: (i.e., whether one should  be  focusing on slow behavior) or
376: whether it extends higher in light of the  fact that similarities have
377: also    been  observed     in    the  high-frequency susceptibilities.
378: \cite{R23}
379: \subsection{  Is   the $\alpha$-relaxation homogeneous or  heterogeneous?}
380: The  observation    stressed above     that the   $\alpha$-relaxation   is
381: nonexponential in the supercooled liquid  range and its representation
382: by a stretched   exponential  as in Eq.~(\ref{eq:2})  can  be formally
383: interpreted  in terms  of  a superposition  of  exponentially decaying
384: functions with a distribution of relaxation times;  but, this {\it per
385: se} does  not guarantee that  the dynamics be  ``heterogeneous'', in the
386: sense that relaxation of the molecules differs from one environment to
387: another  with  the   environment life time   being    longer than  the
388: relaxation time.   An alternative explanation  can be offered within a
389: ``homogeneous'' picture in which  relaxation of the molecules everywhere
390: in the liquid is intrinsically nonexponential.
391: 
392: In  recent  years,   there  has     been   mounting evidence      that
393: heterogeneities,    sufficiently long-lived to   be   relevant to  the
394: $\alpha$-relaxation   and  to be   at  least  partly  responsible  for its
395: nonexponential feature, do exist in supercooled liquids.\cite{R24} The
396: heterogeneous character of the slow  dynamics has been demonstrated in
397: several experiments:  multi-dimensional NMR, \cite{R25} photobleaching
398: probe rotation measurements,  \cite{R26}  nonresonant  dielectric hole
399: burning.  \cite{R27} These  techniques   involve  the selection  of  a
400: sub-ensemble  of molecules  in the  sample  that is characterized by a
401: fairly  narrow distribution  of   relaxation times (and  in general  a
402: relaxation  slower  than average)  and the further  monitoring  of the
403: gradual return to the  equilibrium situation.  Additional  evidence of
404: the spatially   heterogeneous   nature  of the   dynamics   in fragile
405: supercooled  liquids is provided by   the  so-called breakdown of  the
406: Stokes-Einstein relation  between the translational diffusion constant
407: and the viscosity, and the concomitant ``decoupling'' between rotational
408: and   translational    time  scales:\cite{R28,R29} see  Fig.7.
409: 
410: The  size of the  heterogeneities  is not  directly observable in  the
411: above mentioned experiments, but  various estimates, obtained, e.  g.,
412: from optical studies of the rotational relaxation of probes of varying
413: size,  \cite{R30} NMR   measurements, \cite{R31} the   study of excess
414: light scattering,  \cite{R32}  and  the  influence of  a  well-defined
415: $3$-dimensional confinement \cite{R33} lead   to a typical length   of
416: several nanometers in different fragile liquids near $T_g$. One should
417: recall that these signatures  are all dynamical  and that no signature
418: at such a length scale has been detected so far in small-angle neutron
419: and X-ray diffraction data.    If the heterogeneous character  of  the
420: $\alpha$-relaxation  appear  reasonably  well  established,  at  least for
421: deeply  supercooled  fragile  liquids,  several points  concerning the
422: lifetime, the size, and the  nature of the heterogeneities need  still
423: be settled.
424: 
425: \subsection{Is  density or temperature the dominant control variable?}
426:   The phenomenon of viscous
427: slowing down and glass  formation as it  is  studied most of the  time
428: (and  described in the  preceding sections) takes place under isobaric
429: $P=1atm$ conditions. As a consequence, when  the temperature is lowered,
430: there is also an increase of the density of the liquid.  This increase
431: is     small   (a typical  variation   of     $5\%$  between   $T_m$
432: and $T_g$),   but it could still  have  a
433: major influence   on the dynamics.    Actually,  there are theoretical
434: models of  jamming, such as  those  based on free  volume concepts and
435: hard   sphere  systems, that  attribute   the spectacular  increase of
436: viscosity and $\alpha$-relaxation
437: times of    fragile glass-formers (almost)    entirely to  the density
438: changes. It  is  thus important to  evaluate the  role of density  and
439: temperature in driving  the jamming process  that  leads to the  glass
440: transition {\it  at 1 atm}. 
441: 
442: Basic  models and theories are  usually formulated  in terms of either
443: density  or temperature as     control variable, but experiments   are
444: carried out with  pressure   and   temperature as external     control
445: variables.   The  data must be   converted,  when enough  experimental
446: results are available, in order to analyze the influence of density at
447: constant temperature and that of  temperature at constant density, for
448: a  range of density  and  temperature  that  is characteristic of  the
449: phenomenon  {\it at  1 atm}.   Extant  analyses\cite{R34} are far from
450: exhaustive.  However, as  illustrated  in Fig.  8, the  characteristic
451: super-Arrhenius   T-dependence     of  the   viscosity,   $\eta$,    and
452: $\alpha$-relaxation  times,  $\tau_\alpha$, appears   predominantly due  to  the
453: variation of temperature and not to  that of density.  This conclusion
454: is confirmed  by a comparative  study of the  contributions induced by
455: density variations  (at    constant temperature)  and   by  temperature
456: variations (at constant  density) to the rate of  change  of $\eta$ and
457: $\tau_\alpha$   at constant (low) pressure  in  the viscous liquid regime of
458: several molecular and polymeric glass-formers. \cite{R34}
459: 
460: How general   is  the   above  result? Temperature
461: appears  to  be the    dominant   variable controlling  the    viscous
462: super-Arrhenius slowing down  of supercooled liquids at  low pressure,
463: but  this may not be  the case at  much  higher pressure (although not
464: much data  are presently available  to confirm this  point), and it is
465: most  likely     not true  for    describing the  concentration-driven
466: congestion  of dynamics in colloidal suspensions.  In the absence of a
467: ``super-universal'' picture of   the jamming  associated  with
468: glass formation, we shall  restrict  ourselves, as we have  implicitly
469: done above, to the consideration of supercooled  liquids at 1 atm.
470: 
471: \subsection{What  are  the relevant characteristic temperatures?}
472: 
473: There  is   no  unbiased way  of    presenting  the  phenomenology  of
474: glass-forming liquids. Choices must be made  about the emphasis put on
475: the  different aspects, about  the  best graphic representations,  and
476: about the way in which one analyzes  experimental data.  To make sense
477: out of the wealth of  observations and measurements,  it is natural to
478: look for characteristic temperatures about which to organize and scale
479: the   data.  However, since no singularity   is directly detected, the
480: selection  of  one   or several relevant    temperatures  is far  from
481: straightforward.   The temperatures   that  can be  easily  determined
482: experimentally are the boiling point, $T_b$, the melting point, $T_m$,
483: and the  glass transition  temperature(s), $T_g$.  Unfortunately,  the
484: former  two are  generally   considered as irrelevant  to the  jamming
485: phenomenon, and    the   latter has  an   operational  rather   than a
486: fundamental  nature (see the introduction). 
487: 
488: Several other candidates have  been suggested, that  can be split into
489: two groups.  First, there are ``extrapolation temperatures'' {\it below}
490: $T_g$, i.e.,  temperatures dynamically  inaccessible to supercooled
491: liquids, at which  extrapolated behavior diverges or becomes singular.
492: This  is the case  of the VFT temperature  $T_0$ (see Eq.   1) and the
493: Kauzmann  temperature $T_K$  (see section II-4  and  Fig.  4).  In the
494: second group are ``crossover temperatures'' {\it above} $T_g$ at which a
495: new  phenomenon seems     to  appear (decoupling    of  rotations  and
496: translations, emergence    of  a secondary    $\beta$  process,  etc.), a
497: crossover of behavior or a change of 
498: $\alpha$-relaxation  mechanism seems  to take place   (passage from
499: Arrhenius  to super-Arrhenius  $T$-dependence, arrest  of the relaxation
500: mechanisms described by the  mode-coupling theory, putative  emergence
501: of activated barrier crossing  processes,  etc.).  A variety  of  such
502: temperatures for the fragile  glass-former $OTP$ are shown  in Fig.   9:
503: the   location  of   the   different characteristic    temperatures is
504: illustrated  on an Arrhenius plot  of the viscosity. It is interesting
505: to note that most putative crossover behaviors occur  in the region of
506: strong curvature where $\eta \sim  1-10^2$ Poise or $\tau_\alpha \sim 10^{-10}-10^{-8}$
507: sec,  while,   on  the  other   hand,  the   temperatures obtained  by
508: extrapolation of data to low T lie fairly close to each other, some 40
509: K below $T_g$.
510: 
511:  \subsection{What  can be learned from computer simulations?}
512: 
513: Computer  simulation studies, \cite{R39}  in particular those based on
514: Molecular Dynamics  algorithms,   have proven extremely   valuable  in
515: investigating the structure and  dynamics of simple, ordinary liquids.
516: Their  contribution  to the  understanding of  the glass transition is
517: unfortunately limited, the  main reason being  the restricted range of
518: lengths  and times  that    are   accessible to Molecular     Dynamics
519: simulations:  typical   simulations on    atomistic   models  consider
520: $10^3-10^4$  atoms and can follow  relaxations for less than $10^{-8}$
521: sec  (when expressing the elementary  time step in terms of parameters
522: characteristic of simple liquids).   As a result, the viscous,  deeply
523: supercooled  regime   of  real  glass-forming  liquids,   where strong
524: super-Arrhenius   behavior,   heterogeneous   dynamics,    and   other
525: significant features associated  with the jamming process develop,  is
526: out of reach, as  is the laboratory glass transition  that occurs on a
527: time  scale of $10^2$   or $10^3$ sec.  Simple   liquid models do form
528: ``glasses'' on  the observation, i.e.,  simulation, time with  many of
529: the attributes of the laboratory   glass transition: abrupt change  in
530: the thermodynamic coefficients, dependence on  the cooling rate, aging
531: effects,   etc\footnote{ However,    these   models   are   effectively
532: high-temperature  structures,  since     they correspond   to   liquid
533: configurations that are kinetically arrested at a temperature at which
534: the primary relaxation time is of the order  only of nanoseconds. Even
535: with systems   specially designed to  avoid  crystallization,  such as
536: binary Lennard-Jones  mixtures, the cooling  rates  to prepare glasses
537: ($10^8$ to $10^9$   K/sec) are orders  of  magnitude higher than  those
538: commonly used in experiments.}.  However, these supercooled simulation
539: liquids  are  not   truly    fragile   (the  $T$-dependence  of    the
540: $\alpha$-relaxation   time shows  only    small departure  from  Arrhenius
541: behavior) and   are not deeply supercooled so    that one{\'{}}s ability to
542: extract   insights into  the  deeply   supercooled fragile liquids  is
543: questionable.
544: 
545: Computer   simulations  can  be  useful   in  studying the  moderately
546: supercooled liquid region, where one can  observe the onset of viscous
547: slowing down (see for  instance Fig.  9). The  major interest  of such
548: studies is    that  static  and   dynamic   quantities that   are  not
549: experimentally accessible can  be   investigated, such  as  multi-body
550: (beyond  two-particle) correlations  involving  a variety of variables
551: and microscopic mechanisms  for  transport and  relaxation.  They also
552: allow for  testing  in  detail  theoretical  predictions  made in  the
553: relevant   window of times   and    lengths (e.   g.,   those of   the
554: mode-coupling theory) and   for analyzing properties   associated with
555: configurational or phase   space (see below).  
556: 
557: In addition to  the  much studied   one- or two-component  systems  of
558: spheres  with spherically  symmetric interaction potentials,\cite{R39}
559: the  models investigated in  computer simulations  can be divided into
560: two main  groups: on one hand, more   realistic microscopic models for
561: molecular   glass-formers  that attempt   to describe species-specific
562: effects;\cite{R40}   on  the  other   hand,  more  schematic  systems,
563: coarse-grained representations, \cite{R41} lower-dimensional systems
564: \cite{R42} or toy-models,\cite{R43}
565: that  bear less detailed resemblance with  real glass-formers, but can
566: be studied on much longer time scales and with bigger system sizes.
567: 
568: 
569:  
570: \section{THEORETICAL APPROACHES.}
571: There is  a  large number  of  theories,  models, or simply  empirical
572: formulae that  attempt to  reproduce  pieces  of the phenomenology  of
573: supercooled liquids.  There  are   fewer  approaches,  however,   that
574: address the question of why  and how the  viscous slowing down leading
575: to the glass transition, with its salient characteristics described in
576: the preceding sections, occurs  in liquids as they  are cooled. In the
577: following,  we shall briefly review  the  main theoretical approaches,
578: with an emphasis on the concepts and  methods that may prove useful in
579: other areas of physics  where  some sort  of  jamming process is  also
580: encountered\footnote{Models addressing  more specific questions,  such
581: as the  ``coupling model'' \cite{R44}  or the ``continuous  time random
582: walk'', approach\cite{R45}  are   discussed in the   reviews  cited in
583: \cite{R1}.}.   More  specifically, we  shall  discuss phenomenological
584: models  based  on free    volume   and configurational   entropy,  the
585: description of  a purely dynamic   arrest resulting from mode-coupling
586: approximations, ideas relying on the  consideration of the topographic
587: properties of   the  configurational  space   (energy and  free-energy
588: landscapes) or on the analogy with  generalized spin glass models, and
589: approaches centered on the concept of frustration.
590: 
591: \subsection{Free volume } 
592: Free-volume models rest on the assumption  that molecular transport in
593: viscous fluids occurs only when voids having a  volume large enough to
594: accommodate   a molecule form   by  the  redistribution of  some ``free
595: volume'', where  this latter is  loosely defined as some surplus volume
596: that is not  taken up by the  molecules.  In the standard presentation
597: by Cohen and Turnbull,\cite{R46} a molecule in a   dense fluid is mostly
598: confined to a cage  formed by its  nearest neighbors.  The  local free
599: volume,  $v_f$,  is roughly that part  of  a cage space
600: which  exceeds that taken  by a molecule.  It  is assumed that between
601: two events contributing to  molecular transport, a reshuffling of free
602: volume among  the cages occurs  at  no cost of  energy.  Assuming also
603: that the  local  free  volumes  are  statistically  uncorrelated,  one
604: derives a   probability distribution, $P(v_f)$, which is exponential,
605: \begin{equation}
606: \label{eq:4}
607: P(v_f)\propto \exp\left(-\gamma \frac{v_f}{\overline{v_f}}\right),
608: \end{equation}
609: where $\overline{v_f}$ is the average free volume per  molecule and $\gamma$ is
610: a  constant  of  order  $1$. Since    the limiting mechanism   for the
611: diffusion of a molecule is  the occurrence of a void,  i.e., a  local
612: free  volume  $v_f$  larger  than some   critical value, $v_0$, that is
613: approximately equal to  the molecular  volume, the diffusion  constant
614: $D$ is given by  the  probability of  finding  a free volume equal  to
615: $v_0$; this leads  to an expression for  $D$, and by extension for the
616: viscosity $\eta$,
617: 
618: \begin{equation}
619: \label{eq:5}
620: \eta\propto 1/D\propto \exp\left(\gamma \frac{v_0}{\overline{v_f}}\right),
621: \end{equation}
622: which is similar to the formula first proposed by Doolittle.\cite{R47}
623: 
624: In   the Cohen-Turnbull  formulation,  the  average   free volume  per
625: molecule is given by $\overline{v_f}=v-v_0$, where $v=1/ \rho$ is the average
626: total volume per molecule.  The  free-volume concept, in zeroth order,
627: relies on  a hard-sphere picture in  which thermal activation plays no
628: role.  For application to real liquids, temperature enters through the
629: fact  that molecules, or molecular segments  in the  case of polymers,
630: are  not truly ``hard'' and   that, consequently, the  constant-pressure
631: volume is temperature-dependent,
632: \begin{equation}
633: \label{eq:6}
634: \overline{v_f(T)}\propto  \alpha_P(T-T_0),
635: \end{equation}
636: where $\alpha_P$  is the coefficient of  isobaric expansivity and $T_0$ is
637: the temperature  at which all free volume  is consumed, i.e., $v=v_0$.
638: Inserting     the  above   equation   in    the   Doolittle   formula,
639: Eq.~(\ref{eq:4}),   gives the VFT   expression, Eq.~(\ref{eq:1}).   An
640: unanswered, but fundamental  question associated with Eq.~(\ref{eq:6})
641: is why  the free volume should  be consumed at a  nonzero temperature,
642: $T_0$?   An extended version   of  the free-volume approach has   been
643: developed by  Cohen   and Grest, in which   the  cages or  ``cells'' are
644: divided into two groups, liquid-like and solid-like, and concepts from
645: percolation  theory are included to  describe  the dependence upon the
646: fraction of liquid-like  cells.  \cite{R48}  (See  also the model  for
647: molecular diffusivity in  fluids of long rod  molecules by Edwards and
648: Vilgis.\cite{R49})
649: 
650: The main criticisms  of the free volume models are
651: (i) that the concept of free volume is ill-defined, which results in a
652: variety   of  interpretations and   difficulty    in finding a  proper
653: operational procedure even for simple model systems, and (ii) that the
654: pressure dependence of  the viscosity (and  $\alpha$-relaxation  times) is  not adequately
655: reproduced.    This   latter  feature  has  been   emphasized  in many
656: studies,  \cite{R17,R34,R50} and     it is a
657: consequence of the observation made above (see  II-3) that the viscous
658: slowing down of  glass-forming liquids at  1 atm and more generally at
659: low pressure is primarily controlled by temperature and not by density
660: or  volume.     Glass  formation  in  supercooled   liquids  does  not
661: predominantly results from  the drainage of   free volume, but  rather
662: from thermally activated processes.
663: 
664: \subsection{Mode-coupling approximations}
665: The   theory   of  glass-forming liquids that      has had the highest
666: visibility  for  more than    a      decade is  the  mode     coupling
667: theory. \cite{R51}   It predicts a  dynamic   arrest of the
668: liquid structural  relaxation  without any  significant  change in the
669: static properties. All  structural  quantities are  assumed  to behave
670: smoothly and jamming results from  a nonlinear feedback mechanism that
671: affects the relaxation  of the   density fluctuations. Formally,   the
672: theory involves an analysis of a set of nonlinear integro-differential
673: equations describing  the evolution of  pair  correlation functions of
674: wave-vector-  and time-dependent  fluctuations that  characterize  the
675: liquid.    These equations have     the form  of  generalized Langevin
676: equations, and  they    can be  derived   by  using  the  Zwanzig-Mori
677: projection-operator formalism. The equation for  the quantity of prime
678: interest in  the theory, the (normalized)  correlation function of the
679: density fluctuations,
680: 
681: \begin{equation}
682: \label{eq:7}
683: \phi_Q(t)= \frac{<\rho_Q(t)\rho_Q^*(0)>}{<|\rho_Q(0)|^2>},
684: \end{equation}
685: where $\rho_Q(t)=\sum_j\exp(i{\bf Q}{\bf r}_j)$
686:  and ${\bf r}_j$
687: denotes the position of the $j$th particle, can be written as
688: \begin{equation}
689: \label{eq:8}
690: \frac{d^2}{dt^2}\phi_Q(t)+\Omega^2_Q\phi_Q(t)+\int_0^tdt'm_Q(t-t') \frac{d}{dt'}\phi_Q(t')=0,
691: \end{equation}
692: where  $\Omega_Q$ is  a microscopic  frequency obtainable  from the static
693: structure   factor,  $S(Q)\propto<~|\rho_Q(0)|^2> $,    and  $m_Q(t)$   is the
694: time-dependent   memory function that   is  formally  related to   the
695: correlation function of a Q-dependent random force. The above equation
696: being exact, the crux    of  the mode-coupling approach   consists  in
697: formulating an approximate expression for $m_Q(t)$.  The mode-coupling
698: scheme has  been implemented for  liquids both  in  the frame  of  the
699: kinetic   theory  of fluids \cite{R51}   and  that of  the fluctuating
700: nonlinear hydrodynamics.   \cite{R52} It   essentially boils down   to
701: approximating the  memory function  $m_Q(t)$  as the   sum of a   bare
702: contribution  coming    from   the fast    relaxing variables   and  a
703: mode-coupling  contribution coming  from the slowly  decaying bilinear
704: density modes,
705: \begin{equation}
706: \label{eq:9}
707: m_Q(t)=\gamma\delta(t)+\sum_{{\bf Q'}}V_{\bf QQ'}\phi_{\bf Q'}(t)\phi_{|{\bf Q-Q'}|}(t),
708: \end{equation}
709: where the vertices $V_{\bf QQ'}$  can be expressed  in terms of the static
710: structure factor. The self-consistent solution of the resulting set of
711: nonlinear  equations  predicts a  slowing of   the relaxation  that is
712: attributed, within a purely homogeneous picture (see  II-2), to a cage
713: effect and to  the feedback mechanism  above mentioned.  This solution
714: exhibits a dynamic arrest at a critical point, $T_c$, which represents
715: a transition from an ergodic to a nonergodic state with no concomitant
716: singularity in  the thermodynamics and structure   of the system.  The
717: main achievements of   the  mode-coupling approach are the   predicted
718: anomalous increase in relaxation time and the appearance of a two-step
719: relaxation process with  decreasing temperature, as indeed observed in
720: real  fragile glass-formers  (compare  Fig.   10 to  Fig.   7)  and in
721: molecular dynamics simulations.  \cite{R39} Early  on, however, it was
722: realized,  both  from empirical fits   to  experimental data and  from
723: comparison  to simulation data    on model systems,  that  the dynamic
724: arrest at  $T_c$  did not describe the   observed glass  transition at
725: $T_g$ nor  the transition to an ``ideal  glass'' at  a temperature below
726: $T_g$.   This  is  illustrated in    Fig.  11.  Thus,  the  $T_c$  was
727: interpreted as a temperature above $T_g$.  The singularity at $T_c$ is
728: avoided because   of    the breakdown   of  the  simple  mode-coupling
729: approximation, Eq.  9, and the $T_c$ of what is called the ``idealized''
730: mode-coupling  theory is taken  as  a crossover below which additional
731: relaxation mechanisms, such  as  activated processes, presumably  take
732: over.  Unfortunately, beyond some empirical introduction,
733: \cite{R51,R52} activated processes are not theoretically described by
734: mode-coupling approaches, and so the theory of $\alpha$ relaxation has not
735: been extended  to  temperatures below $T_c$.   To  draw once   again a
736: parallel with critical phenomena (where  a singularity occurs at $T_c$
737: in the structure and the thermodynamics  of the system), mode-coupling
738: approximations, as formulated for  instance by Kawasaki, are known  to
739: describe  quite well the standard critical  slowing down,  but not the
740: activated  dynamic scaling such as that  observed in  the random field
741: Ising  model  (see section   I-1).  This  failure  is related  to  the
742: underlying nature of the approximation  that corresponds to a one-loop
743: self-consistent resummation scheme in a perturbative treatment
744: \cite{R52,R54} (see also below in III-4 the parallel with spin glass models).
745: 
746: Mode-coupling approaches can thus   describe at best the  dynamics  of
747: moderately supercooled liquids\footnote{It is  possible that  they are
748: also    applicable  to  the     fast-$\beta$    relaxations  even   below
749: $T_c$.\cite{R51}}  (see Fig.     11).   Because of the many    detailed
750: predictions it makes  in this  regime,  the mode-coupling theory   has
751: stimulated the  use  and the  development of  experimental techniques,
752: such as neutron and   depolarized   light scattering, and    molecular
753: dynamics  simulations that are  able to probe   the early stage of the
754: viscous slowing down;  but, the very  fact that the predicted  dynamic
755: singularity is not observed makes it difficult  to reach any clear-cut
756: conclusion about the quantitative adequacy of the theory, and this has
757: led to  much debate in   recent years. \cite{R55}
758: 
759: \subsection{ Configurational entropy  and (free) energy landscape }
760: The existence of a crossover temperature in the moderately supercooled
761: liquid  region   where $\alpha$-relaxation  times  are    of the order  of
762: $10^{-9}$ sec (hence in the same region as  the $T_c$ predicted by the
763: mode-coupling   theory)   was      advocated    30  years    ago    by
764: Goldstein.\cite{R56} Goldstein  argued that  below this crossover flow
765: is dominated by  potential energy barriers that  are  high compared to
766: thermal  energies and slow relaxation  occurs as a result of thermally
767: activated  processes   taking the system   from  one   minimum of  the
768: potential energy  hypersurface to another.    The idea that  molecular
769: transport in viscous liquids approaching the glass transition could be
770: best described by invoking motion of the representative state point of
771: the system on   the   potential energy   hypersurface had  also   been
772: suggested by  Gibbs. \cite{R57}  In his  view,  the   slowing down of
773: relaxations  with decreasing temperature  is  related to a decrease of
774: the number of available minima  and to  the increasing difficulty  for
775: the system to find such  minima.  The viscous  slowing down would thus
776: result from the decrease  of some ``configurational  entropy'' that is a
777: measure  of the  number   of accessible minima.   These  two concepts,
778: potential energy hypersurface,   also denoted ``energy  landscape'', and
779: ``configurational entropy'', have   gained a renewed interest in  recent
780: years, boosted  by  the analogy  with   the  situation encountered  in
781: several generalized spin   glass models (see  below).  
782: 
783: The Adam-Gibbs approach \cite{R58}
784: represents a phenomenological attempt to
785: relate the  $\alpha$-relaxation time of a glass-forming liquid
786: to  the ``configurational entropy''.   In  the picture,
787:  $\alpha$-relaxation takes  place by increasingly
788: cooperative rearrangements of groups  of  molecules.  Any such  group,
789: called a ``cooperatively   rearranging  region'', is  assumed to   relax
790: independently  of the    others.  It   is    a  kind   of   long-lived
791: heterogeneity.  Molecular    motion is  activated   and the  effective
792: activated free  energy is  equal  to  the  typical energy barrier  per
793: molecule, which is taken as   independent  of temperature, times   the
794: number   of molecules that   are  necessary  to  form a  cooperatively
795: rearranging   region whose   size    permits a  transition   from  one
796: configuration to  a  new one independently  of  the environment.  This
797: latter number goes as the  inverse of the configurational entropy  per
798: molecule,  $S_c(T)/N$, where  $N$ is the total number
799: of molecules in the sample.  Since $S_c$
800:  decreases with decreasing temperature, the reasoning leads
801: to an effective   activation  free energy that  grows  with decreasing
802: temperature, i.  e., to a super-Arrhenius behavior,
803: \begin{equation}
804: \label{eq:10}
805: \tau_\alpha=\tau_0\exp\left(\frac{C}{TS_c(T)}\right), 
806: \end{equation}
807: where $C$ is proportional to $N$ times  the typical energy barrier per
808: molecule.  If  the    configurational entropy vanishes   at a  nonzero
809: temperature,   an   assumption    somewhat  analogous   to     that in
810: Eq.~(\ref{eq:6}) for the  free volume model, but  one that is inherent
811: for instance in the  Gibbs-di Marzio approximate mean  field treatment
812: of a lattice  model of linear  polymeric  chains, \cite{R59} then  the
813: $\alpha$-relaxation times diverge  at  this same nonzero temperature.   In
814: particular if the configurational entropy is identified as the entropy
815: difference     between     the    supercooled     liquid    and    the
816: crystal\footnote{This  phenomenological choice  for   the   entropy of
817: configuration has been criticized  by Goldstein who showed for several
818: glass-formers that only half   of the entropy difference  between  the
819: liquid and the crystal comes  from strictly ``configurational'' sources;
820: the remainder comes  mostly from changes in  vibrational anharmonicity
821: or differences  in the number  of molecular  groups able to  engage in
822: local motions.  \cite{R60}},  the   Adam-Gibbs theory allows   one  to
823: correlate  the  extrapolated divergence   of the $\alpha$-relaxation times
824: with  the Kauzmann paradox (see I-4):   the Kauzmann temperature $T_K$
825: would then signal   a singularity  both in  the  dynamics and  in  the
826: thermodynamics of  a supercooled liquid\footnote{A recent careful, but
827: conjectural analysis of dielectric relaxation data suggests that these
828: data  are  consistent with the  existence  of  a critical  point, both
829: structural  and dynamical, at the approximate  $T_0$  specified by the
830: VTF expression in  Eq.~(\ref{eq:1}).\cite{R61}}.   Note  also that  by
831: using a hyperbolic temperature dependence to fit the experimental data
832: on the heat  capacity difference between the  liquid and  the crystal,
833: $\Delta C_P(T)=K/T$,  and using
834: this  formula to extrapolate the  configurational  entropy down to the
835: Kauzmann temperature, one converts Eq.~(\ref{eq:10}) to a VFT formula,
836: \begin{equation}
837: \label{eq:11}
838: \tau_\alpha=\tau_0\exp\left(\frac{CT_K}{K(T-T_K)}\right), 
839: \end{equation}
840: with  the  VFT  temperature $T_0$ equal    to the Kauzmann temperature
841: $T_K$.      When      comparing    to    experimental     data,    the
842: configurational-entropy based  expressions  provide a good description
843: at  least over  a  restricted temperature   range, but  the  resulting
844: estimates  for   the   critical   number  of molecules    composing  a
845: cooperatively   rearranging region is  often  found to be unphysically
846: small (only a few molecules at $T_g$).
847: \cite{R1}
848: 
849: Building  upon the early suggestion  made by Goldstein, \cite{R62} and
850: others proposed that the apparent  passage with decreasing temperature
851: from flow dynamics described by  a mode-coupling approach to activated
852: dynamics such as pictured  by  the configurational-entropy theory   of
853: Adam  and Gibbs  could be rationalized   by considering the physics of
854: exploration  of the  energy   landscape:  see  Fig.  12.  The   energy
855: landscape is the potential energy  in configurational space. It can be
856: envisaged as an incredibly complex, multi-dimensional ($3N$ dimensions
857: for a   system of  $N$  particles)  set  of hills,   valleys,  basins,
858: saddle-points, and passage-ways around  the hills.  At constant volume
859: and constant number of  particles,  this landscape is  independent  of
860: temperature.   However, the fraction of   space that is  statistically
861: accessible to the  representative state point  of the system decreases
862: with decreasing  temperature, and  the  system becomes constrained  to
863: deeper  and deeper wells. (Recall  that   below the melting point  the
864: deepest   energy minima  corresponding   to  the crystalline   part of
865: configurational  space must be excluded  when studying the supercooled
866: liquid.)  At low enough  temperature, when the representative point of
867: the supercooled   liquid is  mostly  found in   fairly deep and narrow
868: wells, it seems reasonable to  define a ``configurational entropy'' that
869: is proportional  to  the logarithm of  the number  of minima  that are
870: accessible   at  a   given   temperature.  The  liquid  configurations
871: corresponding to  these accessible minima   have been called ``inherent
872: structures''       and Stillinger  and      coworkers   have devised  a
873: gradient-descent mapping procedure to find the inherent structures and
874: study   their   properties   in    computer  simulations.   \cite{R63}
875: Interestingly, Stillinger    has  also  shown,  with    fairly general
876: arguments, that   if one  is    to use  the  above defined   notion of
877: configurational entropy,   an ``ideal glass   transition'' of  the  type
878: commonly  associated   with    the Kauzmann  paradox,    i.   e.,  one
879: characterized by the vanishing   of the configurational entropy at   a
880: nonzero temperature, cannot occur  for  systems of limited   molecular
881: weight  and short-range   interactions. \cite{R64} 
882: 
883: It may be  more  fruitful to investigate  in  place  of the  potential
884: energy landscape a free-energy landscape. Such a landscape can only be
885: defined if  one is  able to  construct a free-energy  functional  by a
886: suitable coarse-graining procedure, as can be done for instance in the
887: case  of mean-field  spin glass  models   (see below).  A  free-energy
888: landscape is temperature-dependent, and  it is important to note  that
889: the ``configurational  entropy'',  also  called  ``complexity'',\cite{R65}
890: that  one  can define  from the   logarithm of  the number  of
891: accessible free-energy  minima  differs  from   the   ``configurational
892: entropy'' computed from the potential energy landscape.  In particular,
893: the   behavior of the complexity is   not restricted by the Stillinger
894: arguments given above. 
895: 
896: The ``landscape paradigm''  is  very  appealing  in  rationalizing  many
897: observations    on  liquids and    glasses and,   more   generally, in
898: establishing a framework   to describe qualitatively  slow dynamics in
899: complex systems that span a wide range of scientific fields.\cite{R66}
900: It has been used to motivate, in addition to the Adam-Gibbs theory and
901: other   phenomenological approaches  like  the  soft-potential  model,
902: \cite{R67} simple   stochastic  models  of    transport based   on  master
903: equations.  \cite{R68} Nevertheless, it has not so far offered a
904: way for elucidating the  {\it physical} mechanism that is  responsible
905: for  the  distinctive    features of the   viscous   slowing   down of
906: supercooled liquids.  
907: 
908: \subsection{Analogy with generalized spin glass models}
909: If  one  takes   seriously  the  observation   that  the  extrapolated
910: temperature dependence   of  both the  viscosity  (and $\alpha$-relaxation
911: times) and  the ``configurational'' entropy (taken  as the difference of
912: entropy    between the liquid and   the   crystal) become divergent or
913: singular at essentially the same temperature $T_0\simeq T_K$(see Fig.  9),
914: one is naturally led to postulate the existence at this temperature of
915: an underlying thermodynamic  transition,  usually referred  to  as the
916: ``ideal  glass  transition''.   Looking    for  analogies  with    phase
917: transitions in  spin glasses is  then appealing.  However, the kind of
918: dynamic  activated scaling   that  would be  required to  describe the
919: slowing   down  of relaxations    when   approaching the   ideal glass
920: transition  (see I-1)  is not  found in  the  most studied Ising  spin
921: glasses.  \cite{R69} Kirkpatrick, Thirumalai,  and Wolynes argued that
922: generalized spin  glass models, such as Potts  glasses and random {\it
923: p}-spin  systems,  would   be  better candidates.   \cite{R70,R71} The
924: random  {\it p}-spin model, for instance,  is defined by the following
925: hamiltonian:
926: \begin{equation}
927: \label{eq:12}
928: H=\sum_{i_1<i_2<\ldots<i_p}J_{i_1i_2\ldots i_p}\sigma_{i_1}\sigma_{i_2}\ldots\sigma_{i_p},
929: \end{equation}
930: where the    $\sigma_{i}$'s   are  Ising  variables   and    the couplings
931: $J_{i_1i_2\ldots  i_p}$'s are quenched  independent  random variables that
932: can take positive and negative values according to a given probability
933: distribution.    The behavior of  these  systems,  {\it at least  when
934: solved in  the mean-field limit where   the interactions between spins
935: have  infinite range},  bears many  similarities  with the theoretical
936: description    of  glass-forming  liquids    outlined above.   Indeed,
937: mean-field Potts glasses (with a number of states strictly larger than
938: 4) and mean-field  p-spin models (with $p\geq  3$) have essentially  the
939: following characteristics:
940: \begin{enumerate}
941: \item
942:  At
943: high temperature, the system  is in a fully  disordered (paramagnetic)
944: state. At   a  temperature   $T_D$,   there  appears   an
945: exponentially large number of metastable ``glassy'' states whose overall
946: contribution to the  partition  function  is  equal  to that of    the
947: paramagnetic minimum. The free energy and all other static equilibrium
948: quantities are fully regular at   $T_D$, but  the dynamics have a
949: singularity of the exact same type as that  found in the mode-coupling
950: theory of     liquids.  At   $T_D$,  the system is trapped in
951: one of the metastable free-energy  minima, and ergocity is broken.
952: 
953: \item   Below   $T_D$ ,  a peculiar   situation  occurs.  The
954: partition  function  has  contributions from,   both, the paramagnetic
955: state  and  the exponentially  large  number of ``glassy'' (free-energy)
956: minima, the logarithm of which  defines the configurational entropy or
957: complexity.  This latter  decreases as the temperature is lowered.
958: \item
959:  At   a nonzero temperature  $T_s<T_D$,
960:  the configurational entropy  vanishes. The system undergoes a
961: {\it bona  fide }thermodynamic transition to a  spin-glass phase.  The
962: transition has been termed ``random first order'' 
963: \cite{R70,R71} because it is  second-order  in the usual thermodynamic sense
964: (with, e. g., no  latent heat), but  shows a discontinuous jump in the
965: order  parameter.  (Technically,   within the  replica  formalism,  it
966: corresponds     to   a one-step  replica   symmetry    breaking with a
967: discontinuous  jump  of  the Edwards-Anderson   order
968: parameter.\cite{R69,R70,R71})
969: \end{enumerate}
970: 
971: 
972: These   mean-field systems  are  the  simplest, analytically tractable
973: models found so  far  that  display a high-temperature   mode-coupling
974: dynamic  singularity,   a   nontrivial  free-energy   landscape, and a
975: low-temperature  ideal  (spin)  glass  transition    with an  ``entropy
976: crisis''\footnote{They  are also  aging phenomena, as  discussed in Ref
977: \cite{R5}.}.   Analyzing   them   sheds light   on   the  mode-coupling
978: approximation, whose  validity for fluid  systems is otherwise hard to
979: assess.    The  mode-coupling  approximation   becomes  exact  in  the
980: mean-field  limit,   because the  barriers   separating the metastable
981: minima diverge  (in the thermodynamic limit)  at and below  $T_D$ as a
982: result of the assumed infinite  range of the interactions. One expects
983: that in  a finite-range model,  provided the  same type of free-energy
984: landscape  is  still encountered, barriers are    large but finite and
985: ergodicity is  restored by thermally activated processes. Accordingly,
986: the  dynamic transition is  smeared out, and  the activated relaxation
987: mechanisms that take over must be  described in a nonperturbative way,
988: as suggested for instance by Kirkpatrick, Thirumalai, and Wolynes
989: \cite{R71}   in  their  dynamic   scaling approach  based  on entropic
990: droplets.  
991: 
992: An   advantage of  an   analogy  between   glass-forming liquids  and
993: generalized spin  glasses\footnote{See also the frustrated percolation
994: model\cite{R72}.} is that the powerful tools  that have been developed
995: in the theory of spin-glass models to characterize the order parameter
996: and the properties associated with the  existence of a large number of
997: metastable glassy states, among which the replica formalism,
998: \cite{R69} can be used {\it  mutatis mutandis }to{\it }study liquids and
999: glasses.   \cite{R73} However, to  make the analogy really successful,
1000: one must still find a short-range model (even more convincing would be
1001: a model  without  quenched disorder) that  actually displays activated
1002: dynamic scaling and a random  first-order transition and make progress
1003: in describing the slow relaxation.
1004: 
1005: 
1006: \subsection{Intrinsic frustration  without randomness }
1007: Spin glasses, and  related systems like orientational  glasses, vortex
1008: glasses  and  vulcanized  matter,  \cite{R74}   owe their  fascinating
1009: behavior   to  two  main ingredients:  {\it    randomness}, namely the
1010: presence    of an   externally  imposed quenched   disorder, and  {\it
1011: frustration}, which expresses   the impossibility   of  simultaneously
1012: minimizing  all the interaction terms  in  the energy  function of the
1013: system.  Liquids and glasses (sometimes called "structural glasses" to
1014: stress the difference with spin  glasses) have no quenched randomness,
1015: but frustration has been  suggested as  a  key feature to  explain the
1016: phenomena  associated with glass formation. \cite{R75,R76,R77,R78,R79}
1017: Frustration in  this context is attributed to  a competition between a
1018: short-range tendency  for the extension of  a  locally preferred order
1019: and global constraints that preclude  the periodic tiling of the whole
1020: space with the  local structure. 
1021: 
1022: The best  studied example  of  such an intrinsic frustration  concerns
1023: single-component  systems  of   spherical   particles interacting with
1024: simple  pair  potentials.   What is   usually  called  "geometric"  or
1025: "topological"  frustration can be  more easily understood by comparing
1026: the situations encountered in  $2$ and $3$ dimensions  \cite{R76} (see
1027: Fig.  13).  In  $2$  dimensions,  the  arrangement  of disks   that is
1028: locally preferred, in  the  sense that  it maximizes the  density  and
1029: minimizes the energy, is  a hexagon of  6 disks around a central  one,
1030: and  this hexagonal structure  can be  extended to the  whole space to
1031: form a triangular  lattice.  In $3$ dimensions, as  was shown long ago
1032: by Frank, \cite{R80} the locally  preferred cluster  of spheres is  an
1033: icosahedron; however, the  $5$-fold rotational symmetry characteristic
1034: of icosahedral order is incompatible  with translational symmetry, and
1035: formation of a periodic  icosahedral crystal is forbidden.   Geometric
1036: or topological frustration is thus  absent in the $2$-dimensional case
1037: but present in  the $3$-dimensional case.   A consequence of this, for
1038: instance,  is that      crystallization is  continuous,    or   weakly
1039: first-order, in  $2$   dimensions  (with some  subtleties   related to
1040: ordering  in   $2$  dimensions  \cite{R81}) whereas  it    is strongly
1041: first-order in $3$  dimensions and accompanied  by the breaking of the
1042: local  icosahedral  structure  to   make the    face-centered-cubic or
1043: hexagonal-close-packed order that  allows to tile space  periodically.
1044: (In contrast, aligned cubes  in 3 dimensions  have no frustration  and
1045: undergo a  continuous   freezing transition to  a  crystalline  state.
1046: \cite{R82}).  The   geometric  frustration   that  affects spheres  in
1047: $3$-dimensional Euclidean space can be relieved in curved space with a
1048: specially  tuned  curvature;   the  creation  of  topological  defects
1049: (disclination lines)  can then be viewed  as the result of forcing the
1050: ideal icosahedral ordering into "flat" space.   This picture of sphere
1051: packing disrupted by frustration  has been further developed in models
1052: for simple atomic systems and metallic glasses, \cite{R75,R76} and the
1053: slowing down of relaxations  has  been tentatively attributed  to  the
1054: topological constraints that   hinder  the kinetics of  the  entangled
1055: defect    lines; \cite{R76} however,  the  treatment  remains only
1056: qualitative and incomplete.
1057: 
1058: A significant  difficulty in   applying  the concept of  geometric  or
1059: topological frustration  to supercooled liquids   is that real fragile
1060: glass-formers  are in  general   either mixtures  or  single-component
1061: systems  of nonspherical  molecules with  a variety of  shapes, all of
1062: which obscures the  detailed   mechanisms  and constraints that    are
1063: responsible  for  the  frustration.  Attempts  have been   made to get
1064: around this problem by  proposing a more coarse-grained description of
1065: frustration\footnote{In addition, several ``toy models''possessing
1066: frustration, but no quenched disorder have been studied by computer
1067: simulation: see for instance Ref.\cite{R83}.}. \cite{R77,R78,R79}
1068: 
1069: In Stillinger's "tear and  repair" mechanism for relaxation and  shear
1070: flow   \cite{R78}    and   in     the   more      recently  introduced
1071: "frustration-limited     domain    theory",\cite{R79}  frustration  is
1072: described  as the source  of a  strain  free  energy that  opposes the
1073: spatial extension  of   the  locally  preferred structure   and  grows
1074: super-extensively with system size.  It  results in the breaking up of
1075: the  liquid  into   domains, whose size   and  growth  with decreasing
1076: temperature are limited by frustration, the weaker the frustration the
1077: larger the domains.  The super-Arrhenius temperature dependence of the
1078: viscosity and $\alpha$-relaxation times   and the heterogeneous nature  of
1079: the   dynamics  are  attributed     to  these  domains   (see     also
1080: Ref.\cite{R77}).  Progress has been  made along these lines, by making
1081: use of  a  scaling approach based on  the  concept of avoided critical
1082: behavior\footnote{This   approach differs from   both those based upon
1083: spin-glass  analogies  and  those  in  which   the slow kinetics   are
1084: attributed to frustration-induced entangled defect lines in that these
1085: others scale about a  low-temperature characteristic point  signifying
1086: ultimate  slowing  down, whereas  in  the  frustration-limited  domain
1087: theory    the scaling is    carried   out  about  a   high-temperature
1088: characteristic  point  signifying the initiation  of  anomalously slow
1089: dynamics.  For the  same reason, it also  differs from the  domain (or
1090: cluster) picture that  has been proposed on  the  basis of  an analogy
1091: between a supercooled liquid  approaching the glass transition  and a
1092: mean-field  model   with   purely  repulsive interactions    near  its
1093: spinodal.\cite{R85}}.    \cite{R79,R84}   However, the  putative order
1094: variable characterizing the locally  preferred structure of the liquid
1095: has  not yet been   properly identified, and,   as in the  case of the
1096: generalized spin-glass  models  discussed above,  one must  still give
1097: convincing  evidence that  the $3$-dimensional  statistical-mechanical
1098: frustrated models   that have been  suggested  as  minimal theoretical
1099: descriptions do show the expected activated dynamics.
1100: 
1101: \section{CONCLUSION}
1102: The viscous  slowing down of supercooled liquids  that leads  to glass
1103: formation  can be  considered as  a  classical  and thoroughly studied
1104: example of a "jamming process".  In this review,  we have stressed the
1105: distinctive  features     characterizing  the   phenomenon:    strong,
1106: super-Arrhenius temperature  dependence  of  the  viscosity  and   the
1107: $\alpha$-relaxation  times,  nonexponential and heterogeneous character of
1108: the $\alpha$  relaxation,  absence  of  marked changes   in the structural
1109: (static) quantities, rapid decrease of the  liquid entropy relative to
1110: that of the crystal, appearance of a sequence of steps (or regimes) in
1111: the relaxation   functions.    These  features are   common   to  most
1112: glass-forming liquids (with the exception of  systems forming $2$- and
1113: $3$-dimensional  networks of  strong  intermolecular  bonds).  We have
1114: also discussed the main theoretical approaches that have been proposed
1115: to describe the origin and the nature of the  viscous slowing down and
1116: of the glass transition. We have emphasized the concepts, such as free
1117: volume,    dynamic      freezing  and  mode-coupling   approximations,
1118: configurational entropy and  (free) energy landscape, and frustration,
1119: that could be useful in other areas of physics where jamming processes
1120: are encountered.
1121: \begin{references}
1122: 
1123: \bibitem{R1} Other aspects of supercooled liquids and glasses are
1124: presented in previous reviews:J. J{\"a}ckle, Rep. Prog. Phys. {\bf 49},
1125: 171 (1986),C. A. Angell, J. Non-Cryst. Solids {\bf 131-133}, 13
1126: (1991), M. D. Ediger, C. A. Angell, and S. R. Nagel,
1127: J. Phys. Chem. {\bf 100}, 13200 (1996).
1128: 
1129: \bibitem{R2} More specific and detailed presentations can be found
1130: for  instance in the  {\it Proceedings  of the International Meetings on
1131: ``Relaxations in Complex Systems''}: (i) K. L.  Ngai and G. B. Wright
1132: Eds., Office of Naval Research, Arlington (1985); (ii)  K. L. Ngai and
1133: G.  B.  Wright Eds.,{\bf  }J. Non-Cryst.  Solids {\bf 131-133} (1991);
1134: (iii)  K.   L.  Ngai,  E.  Riande, and  G.  B.   Wright Eds.,{\bf  }J.
1135: Non-Cryst.  Solids   {\bf 172-174}  (1994); (iv) K.   L. Ngai Ed.,{\bf
1136: }J. Non-Cryst. Solids {\bf 235-238} (1998).
1137: 
1138: \bibitem{R3} C. A. Angell, Science {\bf 267}, 1924 (1995). 
1139: \bibitem{R4} L. C. E. Struik, {\it Physical Aging in Amorphous Polymers
1140: and Other Materials}  (Elsevier, Amsterdam, 1978).  G.  W. Scherer,
1141: {\it Relaxation in Glasses and   Composites } (John Wiley, New  York,
1142: 1986).  I. M. Hodge, J. Non-Cryst. Solids {\bf 169}, 211 (1994).
1143: \bibitem{R5} J. P. Bouchaud, L. Cugliandolo, J. Kurchan, and
1144: M. Mezard,   in {\it Spin Glasses  and  Random Fields},  A.  P. Young
1145: Ed. (World  Scientific, Singapore,  1998), pg.  161. J.  Kurchan, this
1146: volume.
1147: \bibitem{R6} J. D. Ferry, {\it Viscoelastic Properties of Polymers
1148: } (John Wiley, New York, 3$^{rd}$ Edition, 1980).
1149: \bibitem{R7} C. A. Angell, in {\it Relaxation in Complex Systems
1150: }, K. L. Ngai and G. B. Wright Eds.{\bf }(Office of Naval Research,
1151: Arlington, 1985), pg. 3.
1152: \bibitem{R8} D. Kivelson, G. Tarjus, X-L Zhao, and S. A. Kivelson, 
1153: Phys. Rev. E {\bf 53}, 751 (1996).
1154: 
1155: \bibitem{R9} F. Stickel, E. W. Fisher, and R. Richert,
1156:  J. Chem.  Phys. {\bf 102}, 6251 (1995);  {\it ibid} {\bf  104}, 2043
1157:  (1996).
1158: 
1159: \bibitem{R10} D. S. Fisher, G. M. Grinstein, 
1160: and A. Khurana, Physics Today Dec. 1988, 56 (1988).
1161: 
1162: \bibitem{R11} P.K. Dixon, L. Wu, S.R. Nagel, B.D. Williams and
1163: J.P.  Carini, Phys.    Rev. Lett.  {\bf  65},  1108  (1990).   L.  Wu,
1164: P.  K. Dixon,   S.  R. Nagel,  B. D.   Williams,  and  J.  P.  Carini,
1165: J.  Non-Cryst. Solids  {\bf 131-133},   32  (1991).  S.  R. Nagel,  in
1166: {\it "Phase Transitions and Relaxation in  Systems with Competing Energy
1167: Scales"},  T. Riste   and   D. Sherrington Eds.  (Kluwer  Academic,
1168: Netherlands, 1993), pg. 259.
1169: 
1170: \bibitem{R12} J. Villain, J. Phys. (Paris) {\bf 46}, 1843 (1985);
1171:  D. S. Fisher, Phys. Rev. Lett. {\bf 56},  416 (1986).  A. T. Ogielski
1172:  and D. A. Huse, Phys. Rev. Lett. {\bf 56}, 1298 (1986). 
1173: 
1174: \bibitem{R13} A. T{\"u}lle, H. Schober, J. Wuttke,
1175:  and F. Fujara, Phys. Rev. E. {\bf 56}, 809 (1997).
1176: 
1177: \bibitem{R14} See also: B. Frick, D. Richter, and Cl. Richter,
1178:  Europhys. Lett.  {\bf 9},  557  (1989). E.  Kartini, M.   F. Collins,
1179:  B. Collier, F. Mezei, and E. C. Swensson, Phys. Rev. B {\bf 54}, 6292
1180:  (1996).  R. Leheny, N. Menon, S. R. Nagel, K. Volin, D. L. Price, and
1181:  P. Thiyagarjan, J. Chem. Phys. {\bf 105}, 7783 (1996).
1182: 
1183: \bibitem{R15} S. R. Elliot, Nature {\bf 354}, 445 (1991); J. Phys. Condens. Matter {\bf 4}, 7661 (1992). D. Morineau, C. Alba-Simionesco, M.-C. Bellissent-Funel, M.-F. Lauthie, Europhys. Lett. {\bf 43}, 195 (1998).
1184: 
1185: \bibitem{R16} W. Kauzmann, Chem. Rev. {\bf 43}, 219 (1948).
1186: 
1187: \bibitem{R17} C. A. Angell, J. Res. Natl. Inst. Stand. Technol. {\bf 102}, 171 (1997). 
1188: \bibitem{R18}{\bf }R. Richert and C. A. Angell, J. Chem. Phys. {\bf 108}, 9016 (1998).
1189: \bibitem{R19} W. Knaak, F. Mezei, and B. Farago, Europhys. Lett. {\bf 7}, 529 (1988).
1190: \bibitem{R20} G. P. Johari and M. Goldstein, J. Chem. Phys. {\bf 53},
1191: 2372 (1970).
1192: 
1193: \bibitem{R21} E. R{\"o}ssler, Phys. Rev. Lett. {\bf 69},
1194:  1620 (1992). E. R{\"o}ssler, U. Warschewske, P. Eiermann, A. P. Sokolov, D. Quitmann, J. Non-Cryst. Solids, {\bf 172-174}, 113 (1994). E. Bartsch, F. Fujara, B. Geil, M. Kiebel, W. Petry, W. Schnauss, H. Sillescu, and J. Wuttke, Physica A {\bf 201}, 223 (1993).
1195: 
1196: \bibitem{R22} E. R{\"o}ssler, V. N. Novikov, and A.P. Sokolov, Phase Transitions {\bf 63}, 201 (1997), and references therein.
1197: 
1198: \bibitem{R23} U. Schneider, P. Lukenheimer, R. Brand, and A. Loidl, J. Non-Cryst. Solids {\bf 235-237}, 173 (1998).
1199: 
1200: \bibitem{R24} See for instance the review by H. Sillescu, J. Non-Cryst. Solids, in press (1999).
1201: 
1202: \bibitem{R25} K. Schmidt-Rohr and H.W. Spiess, Phys. Rev. Lett.
1203:  {\bf 66}, 3020 (1991). A. Heuer, M. Wilhelm, H. Zimmermann, and
1204: H. W. Spiess,
1205:  Phys. Rev. Lett.{\bf 75}, 2851 (1995). R. B{\"u}hmer, G. Hinze,
1206: G. Diezemann, 
1207: B. Geil, and H. Sillescu, Europhys. Lett. {\bf 36}, 55 (1996).
1208:  R. B{\"u}hmer, G. Diezemann, G. Hinze, and H. Sillescu,
1209: J. Chem. Phys. {\bf 108}, 890 (1998); 
1210: 
1211: \bibitem{R26} M. T. Cicerone and M. D. Ediger, J. Chem. Phys. {\bf
1212: 103}, 5684 (1995); C.-Y. Wang and M. D. Ediger, J.  Phys. Chem. B , in
1213: press (1999).
1214: 
1215: \bibitem{R27} B. Schiener, R. B{\"u}hmer, A Loidl, and R. V. Chamberlin,
1216:  Science{\bf  274},  752   (1996);  B.  Schiener,   R.  V. Chamberlin,
1217:  G. Diezemann, and R. B{\"u}hmer, J. Chem. Phys.{\bf 107}, 7746 (1997).
1218: \bibitem{R28} F. Fujara, B. Geil, H. Sillescu and G. Fleischer, 
1219: Z.   Phys.  B-Condensed  Matter {\bf   88},    195 (1992).  I.  Chang,
1220: F.    Fujara,     G.   Heuberger,  T. Mangel,       and   H. Sillescu,
1221: J. Non-Cryst. Solids {\bf 172-174}, 248 (1994).
1222: 
1223: \bibitem{R29} M. T. Cicerone and M. D. Ediger, J. Chem. Phys. {\bf
1224: 104},   7210 (1996). F.  Blackburn,   C.-Y. Wang,  and  M.  D. Ediger,
1225: J. Phys. Chem. {\bf 100}, 18249 (1996).
1226: \bibitem{R30} M. T. Cicerone, F. R. Blackburn, 
1227: and M. D. Ediger, J. Chem. Phys. {\bf 102}, 471 (1995). M. T. Cicerone
1228: and M. D. Ediger, J. Non-Cryst. Solids {\bf 235-238}, (1998).
1229: \bibitem{R31} U. Tracht, M. Wilhelm,
1230:  A.  Heuer,    H.    Feng,  K.  Schmidt-Rohr,      and  H.W.   Spiess,
1231:  Phys. Rev. Lett., 2727 (1998).
1232: \bibitem{R32} C. T. Moynihan and J. Schroeder,
1233:  J. Non-Cryst. Solids {\bf 160}, 52 (1993).
1234: \bibitem{R33} G. Barut, P. Pissis, R. Pelster,
1235:  and G. Nimtz, Phys. Rev. Lett.{\bf 80}, 3543 (1998).
1236: \bibitem{R34} M. L. Ferrer, C. Lawrence, 
1237: B.  G. Demirjian,  D.   Kivelson, C.  Alba-Simionesco, and  G. Tarjus,
1238: J. Chem. Phys. {\bf 109}, 8010 (1998), and references therein.
1239: \bibitem{R35} W. T. Laughlin and  D. R.    Uhlmann,      J.  Phys.   Chem.  {\bf    76},      2317
1240:  (1972).     M.  Cukierman, J.  W.   Lane,  and  and  D.  R.  Uhlmann,
1241:  J. Chem.  Phys. {\bf 59}, 3639 (1973).   R. G. Greet and D. Turnbull,
1242:  J. Chem. Phys. {\bf 46}, 1243 (1967).
1243: \bibitem{R36} C. Hansen,
1244:  F. Stickel,   T.  Berger,   R.     Richert, and  E.   W.     Fischer,
1245:  J. Chem. Phys. {\bf 107}, 1086 (1997)
1246: \bibitem{R37} R. R{\"o}ssler and
1247:  A.   P. Sokolov, Chem.  Geol.  {\bf  128}, 143   (1996); E.  R{\"o}ssler,
1248:  K.-U.  Hess, and V.  N. Novikov, J.  Non-Cryst. Solids {\bf 223}, 207
1249:  (1998).
1250: \bibitem{R38} W. Petry, E. Bartsch, F. Fujara, M. Kiebel, H. Sillescu, and B. Farago, Z. Phys.B {\bf 83}, 175 (1991).
1251: 
1252: \bibitem{R39} See for instance the various reviews: J. L. Barrat and
1253: M. L. Klein, Annu. Rev. Phys. Chem. {\bf 42}, 23 (1991). J. P. Hansen,
1254: Physica A {\bf 201}, 138 (1993).  W. Kob, J. Phys. Condens. Matter, in
1255: press (1999), as  well as the special  issue: {\it "Glasses
1256: and the  glass  transition.   Challenges   in  Materials  Theory   and
1257: Simulation"},{\it } S.   Glotzer Ed.  (Computational  Materials Science
1258: {\bf 4}, 1995).
1259: 
1260: \bibitem{R40} P. Sindzingre and M. L. Klein, J. Chem. Phys. {\bf 96},
1261: 4681 (1992). P. H. Poole, F. Sciortino, U. Essmann, and H. E. Stanley,
1262: Nature   {\bf  360}, 324 (1992).     L.  J.  Lewis  and G.  Wahnstrom,
1263: Phys. Rev.  E {\bf 50},  3865  (1994). M. Wilson   and P.  A.  Madden,
1264: Phys. Rev.  Lett.  {\bf 72}, 3033   (1994). J.  Horbach, W.  Kob,  and
1265: K. Binder, Phil. Mag. B {\bf 77}, 297 (1998).
1266: 
1267: \bibitem{R41}{\bf }K. Binder, J. Baschnagel, W. Paul, H. P. Wittmann,
1268: and W. Wolfgardt, Computational Materials Science {\bf 4}, 309 (1995),
1269: and references therein.
1270: 
1271: \bibitem{R42}{\bf }T. Muranaka and Y. Hiwatari, Phys. Rev. E {\bf 51},
1272: R2735  (1995); D. N.  Perera and P. Harrowell, Phys.  Rev. E {\bf 52},
1273: 1694 (1995); R.  Yamamoto and A. Onuki,  J. Phys. Soc. Jpn. {\bf  66},
1274: 2545  (1997). A. I.   Mel’cuk, R. A.   Ramos, H. Gould, W. Klein,  and
1275: R. D. Mountain, Phys. Rev. Lett. {\bf 75}, 2522 (1995).
1276: 
1277: \bibitem{R43}{\bf }G. H. Fredrickson and H. C. Andersen,
1278: Phys.  Rev.  Lett.   {\bf   53}, 1244  (1984).   T.  A.  Weber and  F.
1279: H. Stillinger, Phys.   Rev. B   {\bf  36}, 7043  (1986).  F.   Ritort,
1280: Phys. Rev. Lett.  {\bf 75}, 1190 (1995). M. Foley and P. Harrowell, J.
1281: Chem.  Phys.    {\bf  98}, 5069 (1993).      See also J.  J{\"a}ckle,
1282: Prog. Theor. Phys. Suppl. {\bf 126}, 53 (1997).
1283: 
1284: \bibitem{R44} K. L. Ngai and 
1285: R.  W. Rendell,   in {\it   Supercooled Liquids:   Advances and  Novel
1286: Applications}, J. Fourkas, D. Kivelson,  U. Mohanty,  and K. A.  Nelson
1287: Eds. (ACS Symposium Series 676, Washington, 1997), pg. 45.
1288: \bibitem{R45} J. T. Bendler and M. F. Schesinger, 
1289: in {\it ``Relaxations  in Complex Systems''} K.  L.   Ngai and G.  B.
1290: Wright Eds. (Office of Naval Research, Arlington, 1985), pg. 261.
1291: \bibitem{R46} 
1292: M. H. Cohen and D. J. Turnbull, J. Chem. Phys.  {\bf 31}, 1164 (1959);
1293: D. Turnbull  and M. H. Cohen,  {\it ibid} {\bf   34}, 120 (1961), {\bf
1294: 52}, 3038 (1970).
1295: \bibitem{R47} A. K. Doolittle, J. Appl. Phys. {\bf 22}, 1471 (1951).
1296: \bibitem{R48} G. S. Grest and M. H. Cohen, 
1297: Adv. Chem. Phys. {\bf 48}, 455 (1981).
1298: \bibitem{R49} S. F. Edwards and Th. Vilgis, Phys. Scripta T {\bf 13},
1299: 7 (1986).
1300: 
1301: \bibitem{R50} M. Goldstein, J. Phys. Chem. {\bf 77}, 667
1302: (1973). J. P. Johari and E. Whalley, Faraday Symp. Chem. Soc. {\bf 6},
1303: 23  (1973). S. A.  Brawer,   {\it Relaxation  in Viscous  Liquids  and
1304: Glasses} (American Ceramic Society,  Colombus, 1985).  D. M.  Colucci,
1305: G.  B. McKenna, J. J.  Filliben, A. Lee,  D. B. Curliss, K. B. Bowman,
1306: and  J. D.  Russel, J. Polym.  Sci.  B:  Polym. Phys.  {\bf 351}, 1561
1307: (1997).
1308: 
1309: \bibitem{R51} W. G{\"o}tze, in {\it "Liquids, Freezing, and the
1310: Glass Transition"},  J. P.  Hansen,  D.  Levesque, and  J. Zinn-Justin
1311: Eds.  (North   Holland, Amsterdam,  1991),  pg.   287.  W.  G{\"o}tze  and
1312: L. Sj{\"o}gren, Rep. Prog. Phys. {\bf 55}, 241 (1992). W. G{\"o}tze, J. Phys.:
1313: Cond. Mat. {\bf 11}, A1 (1999).
1314: \bibitem{R52} S. P. Das and G. F. Mazenko, Phys. Rev. A {\bf 34},
1315:  2265 (1986). B. Kim and G.F. Mazenko, Adv. Chem. Phys. {\bf 78}, 129 (1980).
1316: \bibitem{R53} K. Kawasaki,
1317:  in {\it Phase  Transitions  and  Critical  Phenomena}, C.   Domb  and
1318:  M. S. Green Eds; (Academic Press, New York, 1976), vol. 5a.
1319: 
1320: \bibitem{R54} J. P. Bouchaud, L. Cugliandolo, J. Kurchan,
1321:  and M. Mezard,  Physica A {\bf   226}, 243 (1996).  C. Z.-W.  Liu and
1322:  I. Oppenheim, Physica A {\bf 235}, 369 (1997).
1323: \bibitem{R55} For a sample of viewpoints and contributions, see:
1324: G. Li, W. M. Du, X.K. Chen, H. Z. Cummins, and N. J. Tao, Phys. Rev. A
1325: {\bf 45}, 3867 (1992); G.  Li, W. M. Du, A.  Sakai, and H. Z. Cummins,
1326: Phys. Rev. A {\bf 46}, 3343 (1992); H. Z.  Cummins, W. M. Du,M. Fuchs,
1327: W. G{\"o}tze, S. Hildebrand, G. Li, and N. J.  Tao, Phys. Rev. E {\bf 47},
1328: 4223 (1993).  X.  C. Zeng, D. Kivelson, and  G.  Tarjus, Phys. Rev.  E
1329: {\bf 50},  1711  (1994); P.  K.  Dixon,  N.  Menon, and  S. R.  Nagel,
1330: Phys.  Rev. E {\bf 50}, 1717 (1994).  A.  P.  Sokolov, W. Steffen, and
1331: E.   R{\"o}ssler, Phys.  Rev.   E {\bf  52},  5105 (1995).   F.  Mezei and
1332: M. Russina, J. Phys.: Cond. Mat.  {\bf 11}, A341 (1999).  J. Gapinski,
1333: W. Steffen, A. Patkowski, A.  P. Sokolov, A. Kisliuk, U.  Buchenau, M.
1334: Russina, F. Mezei, and A. Schober, preprint (1998).
1335: 
1336: \bibitem{R56} M. Goldstein, J. Chem. Phys. {\bf 51}, 3728 (1969).
1337: \bibitem{R57} J. H. Gibbs, in {\it Modern Aspects of the Vitreous State}, J. D. Mackenzie Ed. (Butterworths, London, 1960), vol. 2, pg. 152.
1338: \bibitem{R58} G. Adam and J. H. Gibbs, J. Chem. Phys. {\bf 43}, 139 (1965).
1339: 
1340: \bibitem{R59} J. H. Gibbs and E. A. Di Marzio,
1341:  J. Chem. Phys. {\bf 28}, 373 (1958).
1342: \bibitem{R60} M. Goldstein, J. Chem. Phys. {\bf 64}, 4767 (1976).
1343: \bibitem{R61} N. Menon and S. R. Nagel, Phys. Rev. Lett. {\bf 74}, 1230 (1995).
1344: \bibitem{R62} C. A. Angell, J. Phys. Chem. Solids {\bf 49}, 863 (1988).
1345: \bibitem{R63} F.  H.  Stillinger and T.  A. Weber,  Science  {\bf  225}, 983 (1983);
1346: F.  H.  Stillinger, {\it ibid} {\bf 267},  1935 (1995); S.  Sastry, P.
1347: G.   Debenedetti,    and F.  H.    Stillinger,   Nature {\bf 393}, 554
1348: (1998). See also R. J.  Speedy and P. G.  Debenedetti, Mol. Phys. {\bf
1349: 88}, 1293 (1996). T. Keyes, J. Chem. Phys. {\bf 101}, 5081 (1992).
1350: 
1351: \bibitem{R64} F. H. Stillinger, J. Chem. Phys. {\bf 88}, 7818 (1988).
1352: \bibitem{R65} R. G. Palmer, Adv. Phys. {\bf 31}, 669 (1982).
1353: 
1354: \bibitem{R66} C. A. Angell, Nature {\bf 393}, 521 (1998).
1355: \bibitem{R67} U. Z{\"u}rcher and T. Keyes, in {\it Supercooled
1356: Liquids: Advances and Novel Applications}, J. Fourkas et al. Eds. (ACS
1357: Symposium Series 676, Washington, 1997), pg. 82.
1358: 
1359: \bibitem{R68} S. A. Brawer, J. Chem. Phys. {\bf 81}, 954 (1984);
1360:  H.  B{\"a}ssler, Phys.  Rev.  Lett. {\bf  58},  767  (1987); J.  C. Dyre,
1361:  Phys. Rev. Lett. {\bf 58}, 792 (1987); U.  Mohanty, I. Oppenheim, and
1362:  C.  H.  Taubes,  Science   {\bf  266}, 425   (1994);  G.   Diezemann,
1363:  J. Chem. Phys. {\bf 107} (1997).
1364: 
1365: \bibitem{R69} See for instance: K. Binder and A. P. Young,
1366: Rev. Mod. Phys.  {\bf 58}, 801 (1986);
1367: 
1368: {\it Spin Glass Theory and Beyond},  M. Mezard, G. Parisi,
1369: and M. A.  Virasoro   Eds. (World Scientific, Singapore,  1987);  
1370: {\it Spin Glasses  and    Random  Fields}, A. P.     Young
1371: Ed. (World Scientific, Singapore, 1998).
1372: 
1373: \bibitem{R70} T. R. Kirkpatrick and P. G. Wolynes, Phys. Rev. A{\bf
1374: 35}, 3072 (1987); Phys. Rev. B{\bf 36}, 8552 (1987); T. R. Kirkpatrick
1375: and D. Thirumalai, Phys. Rev. Lett. {\bf  58}, 2091 (1987); J. Phys. A
1376: {\bf 22}, L149 (1989).
1377: \bibitem{R71} T. R. Kirkpatrick, D. Thirumalai, and P. G. Wolynes,
1378: Phys. Rev. A {\bf 40}, 1045 (1989).
1379: \bibitem{R72} M. Nicodemi and A. Coniglio,
1380: J. Phys. A {\bf 30}, L187 (1997); A. Coniglio, A. de Candia, and
1381: M. Nicodemi, J. Phys.: Cond. Mat. {\bf 11}, A167 (1999).
1382: \bibitem{R73} G. Parisi,
1383: in  {\it  Supercooled Liquids:  Advances  and  Novel  Applications}, J.
1384: Fourkas, D. Kivelson, U. Mohanty, and K. A. Nelson Eds. (ACS Symposium
1385: Series 676,  Washington, 1997),  pg.  110.   R.   Monasson, Phys. Rev.
1386: Lett. {\bf   75}, 2847 (1995).   S.  Franz and G.  Parisi, J. Physique
1387: (Paris) I {\bf 5}, 1401 (1995). M.  Mezard and G.  Parisi, J. Phys. A:
1388: Math. Gen. {\bf 29}, 6515 (1996); cond-mat/9812180 (1998).
1389: \bibitem{R74} See   the articles  collected in   
1390: {\it Spin   Glasses  and Random
1391: Fields},  A.  P. Young  Ed.   (World Scientific,  Singapore, 1998).
1392: 
1393: \bibitem{R75}
1394: M. Kleman and J.  F.  Sadoc, J. Phys.   Lett.  (Paris) {\bf 40},  L569
1395: (1979).   J.  P.   Sethna, Phys.     Rev.    Lett.  {\bf 51},     2198
1396: (1983). S. Sachdev  and D. R. Nelson,  Phys. Rev. Lett. {\bf 53}, 1947
1397: (1984); Phys. Rev. B {\bf 32}, 1480 (1985).   S. Sachdev, Phys. Rev. B
1398: {\bf 33}, 6395 (1986).
1399: 
1400: \bibitem{R76} D. R. Nelson and F. Spaepen, Solid State Phys. {\bf 42},
1401:  1 (1989).
1402: \bibitem{R77} J. P. Sethna, J. D. Shore, and M. Huang,
1403: Phys.  Rev. B {\bf 44}, 4943 (1991).
1404: \bibitem{R78}   F.   H.   Stillinger,
1405: J.  Chem.  Phys. {\bf     89},  6461  (1988);   F.H.   Stillinger  and
1406: J.  A. Hodgdon, Phys. Rev.  E  {\bf 50}, 2064  (1994).
1407: 
1408: \bibitem{R79} D. Kivelson,
1409: S. A. Kivelson, X.-L. Zhao, Z. Nussinov, and G. Tarjus, Physica A {\bf
1410: 219}, 27 (1995); G. Tarjus and D. Kivelson, J.  Chem. Phys. {\bf 103},
1411: 3071 (1995);   D. Kivelson and  G. Tarjus,  Phil. Mag. B{\bf  77}, 245
1412: (1998).
1413: \bibitem{R80}  F.C.  Frank,   Proc.  Royal Soc.  London {\bf    215A},  43
1414: (1952).
1415: \bibitem{R81}  K. J.      Standburg,  Rev.  Mod.   Phys.    {\bf   60},161
1416: (1988).
1417: \bibitem{R82} E.  A. Jagla, Phys. Rev. E  {\bf 58}, 4701 (1998).
1418: \bibitem{R83} L. Gu
1419: and B. Chakraborty, Mat. Res. Symp. {\bf 455},  229 (1997). B. Kim and
1420: S.  J. Lee, Phys.  Rev. Lett.  {\bf 79}, 3709   (1997). S. J.  Lee and
1421: B. Kim,  cond-mat/9901077.
1422: \bibitem{R84}
1423:  L. Chayes,  V. J.  Emery, S.A. Kivelson,
1424: Z. Nussinov,  and     G.   Tarjus,  Physica    A    {\bf 225},     129
1425: (1996).
1426: \bibitem{R85}
1427:   W.  Klein,   H.  Gould,  R. A.    Ramos,  I.  Clejan,  and
1428: A. I. Mel{\'{}}cuk, Physica A {\bf 205}, 738 (1994).
1429: 
1430: \end{references}
1431: 
1432: \begin{figure}
1433: 
1434: \caption[99]{
1435: Super-Arrhenius   $T$-dependence  of     the     viscosity $\eta$    and
1436: $\alpha$-relaxation times $\tau_\alpha  $ in  glass-forming  liquids. { \bf  a)}
1437: Logarithm   (base 10)  of  $\eta$  and $\tau_\alpha $   versus reduced inverse
1438: temperature $T_g/T$ for several liquids. For $GeO_2$, a system forming
1439: a  network of strong intermolecular  bonds,  the  variation is  almost
1440: linear, whereas    the  other  liquids   (glycerol,   m-toluidine, and
1441: ortho-terphenyl) are characterized, below some temperature $T^*$, by a
1442: strong  departure from linear    dependence:  the behavior  is  nearly
1443: Arrhenius in the former case and super-Arrhenius  in the latter. (Data
1444: taken   from references      cited  in     Ref.\cite{R8}  and     from
1445: C. Alba-Simionesco, private  communication.)   { \bf b)  }   Effective
1446: activation free energy  $E(T)$, obtained from data  shown in {\bf a)},
1447: as a function of inverse temperature. Both $E(T)$  and $T$ are divided
1448: by the crossover temperature $T^*$ shown in { \bf a) }.}
1449: 
1450: \end{figure}
1451: 
1452: \begin{figure}
1453: \caption[99]{
1454: Imaginary  part  of  the dielectric  susceptibility $\chi''$ of liquid
1455: m-toluidine            versus           $ log_{10}(\omega)$ for several
1456: temperatures close to $T_g$ ($T_g=183.5 K$). The inset
1457: shows       that  the $\alpha$ peak  is  broader than a Debye spectrum
1458: that     would correspond  to    a  purely   exponential relaxation in
1459: time.     (Data   from     C.    Tschirwitz,    E. R{\"o}ssler,   and
1460: C.   Alba-Simionesco,  private  communication.)}
1461: \end{figure}
1462: 
1463: \begin{figure}
1464: \caption[99]{Static
1465: structure  factor   $S(Q)$  of  liquid  (deuterated) ortho-terphenyl  at
1466: several temperatures from just  below melting ($T_m=329 K$) to
1467: just above the glass transition ($T_g=243 K$). The inset shows
1468: the weak  variation with  temperature of  $S(Q)$ for three  values  of $Q$
1469: indicated by the symbols above  the $Q$-axis. (From Ref.\cite{R13}.)}
1470: \end{figure}
1471: \begin{figure}
1472: \caption[99]{
1473: Kauzmann's representation of the ``entropy paradox'': entropy difference
1474: $\Delta S$ between the liquid and the crystal (normalized by its value $\Delta
1475: S_m$ at the  melting  point) versus  reduced temperature  $T/T_m$. The
1476: break in the slopes of the full lines signals  the glass transition at
1477: $T_g$.    The dashed lines indicate    an extrapolation of the entropy
1478: difference curves below $T_g$.  Except for the strong, network-forming
1479: liquid $B_2O_3$,   the extrapolated entropy   difference vanishes at a
1480: nonzero temperature $T_K$.  (Data from Refs.\cite{R16},\cite{R18}, and
1481: from H.  Fujimori and C.  Alba-Simionesco, private communication.)}
1482: \end{figure}
1483: \begin{figure}
1484: \caption[99]{Time  dependence  of the (normalized)
1485: dynamic       structure     factor       $S(Q,t)/S(Q)$      of    liquid
1486: $Ca_{0.4}K_{0.6}(NO_3)_{1.4}$  (for
1487: $Q\simeq 1.9{\AA}^{-1} $) at various
1488: temperatures.  The  continuous   lines    are   obtained  by   Fourier
1489: transforming neutron  time-of-flight data  and  the symbols  represent
1490: neutron spin-echo results. When $T$ decreases, two  steps separated by a
1491: plateau    appear  in the  relaxation.     (From Ref.\cite{R19}.)}
1492: \end{figure}
1493: \begin{figure}
1494: \caption[99]{
1495: Scaling plot  for the     imaginary  part $\chi''$ of  the    dielectric
1496: susceptibility  of  several  glass-forming  liquids  (glycerol, salol,
1497: propylene     glycol,      dibutyl-phtalate,     $\alpha$-phenyl-o-cresol,
1498: ortho-terphenyl,  ortho-phenylphenol). Experimental   data similar  to
1499: those shown in Fig. 2, but for 13 decades  of frequency, are collapsed
1500: for    all temperatures  and    all  liquids  onto  the   master-curve
1501: $w^{-1}log_{10}(\chi''\omega_\alpha/  \Delta\chi\omega)$ vs $w^{-1}(1+w^{-1})log_{10}(\omega/
1502: \omega_\alpha)$; $\omega_\alpha$ is the $\alpha$-peak position, $w$ is a shape factor that
1503: characterizes  the deviation from Debye  behavior,  and $\Delta\chi$ is  the
1504: static susceptibility. (From Ref.\cite{R11}.) }
1505: \end{figure}
1506: \begin{figure}
1507: \caption[99]{``Decoupling'' between rotational and
1508: translational  time scales: logarithm    (base 10) of the   rotational
1509: ($D_r$)  and   translational  ($D_t$)  diffusion     coefficients  for
1510: ortho-terphenyl  as a  function of  temperature  and viscosity.  $D_r$
1511: (diamonds) follows the   viscosity at all temperatures whereas   $D_t$
1512: (squares, triangles, and dots) departs from viscosity (and from $D_r$)
1513: below $T\simeq 290 K$.  (From Ref.\cite{R28}).}
1514: \end{figure}
1515: \begin{figure}
1516: \caption[99]{
1517: Relative influence of temperature  ($T$)  and density  ( $\rho$) on  the
1518: viscous  slowing    down of  liquid   triphenyl  phosphite.   { \bf a) }
1519: $log_{10}(\eta)$  versus $\rho$  at   constant $T$ for several  isotherms.
1520: Also shown    are the isobaric   data  at $P=1atm$  (under atmospheric
1521: pressure,  the   glass  transition takes  place at    $T_g\simeq 200K$ and
1522: $\rho_g\simeq1.275g/cm^3$).  {\bf b)} $log_{10}(\eta)$  versus $T$ at constant
1523: $\rho$ for several isochores. Note that the  change of viscosity is much
1524: smaller with density (at constant $T$) than it is with temperature (at
1525: constant $\rho$) for the range of temperature and density characteristic
1526: of  the liquid and supercooled  liquid phases at atmospheric pressure.
1527: (From Ref.\cite{R34}.) }
1528: \end{figure}
1529: \begin{figure}
1530: \caption[99]{Illustration of the various
1531: choices  of   characteristic  temperature for  describing  the viscous
1532: slowing down  of  liquid ortho-terphenyl.  $log_{10}(\eta)$   is plotted
1533: versus inverse temperature.  ``Extrapolation'' temperatures:  $T_0(VFT$,
1534: see Eq.  1)$=200-202  K$ \cite{R18,R35}, $T_K$(Kauzmann, see I-4)=$204
1535: K$  \cite{R18}.   ``Crossover''   temperatures: $T^*$(see   Fig.   1 and
1536: III-5)=$350 K$\cite{R8}, $T_c$(MCT, see III-2)=$276-290 K$
1537: \cite{R38}, $T_A=455 K$ \cite{R36}, $T_B=290 K$ \cite{R37}, $T_X=289
1538: K$\cite{R37}.  Also shown    are the experimentally measured   boiling
1539: ($T_b=610  K$), melting ($T_m=311 K$),  and glass transition ($T_g=246
1540: K$) temperatures.  (Viscosity data from Ref.
1541: \cite{R38}.) }
1542: \end{figure}
1543: 
1544: \begin{figure}
1545: \caption[99]{Mode-coupling scenario of kinetic freezing and appearance
1546: of   a  $2$-step  relaxation:  time dependence   of   the (normalized)
1547: density-density   correlation function for  a  schematic mode-coupling
1548: model.  The curves  from $A$ to $G$  correspond to the approach toward
1549: the dynamic  singularity  from  the ergodic  state.  The other  curves
1550: correspond to the nonergodic state. (From Ref.\cite{R51}.)}
1551: \end{figure}
1552: \begin{figure}
1553: \caption[99]{Breakdown of
1554: the  ``idealized''     mode-coupling theory   illustrated   on  $log_{10}(\eta)$ 
1555:  versus $1/T$  for liquid ortho-terphenyl
1556: (see caption to Fig. 9). The  continuous line is the mode-coupling fit
1557: to the experimental data. The predictions break  down below a point at
1558: which  the viscosity  is  of the   order  of  10  Poise;   the dynamic
1559: singularity is not observed,  and  $T_c$ is interpreted as   a
1560: crossover    temperature.}
1561: \end{figure}
1562:  \begin{figure}
1563: \caption[99]{Illustration of  the  putative relation
1564: between the rapid increase of $\alpha$-relaxation time ({\bf a}), the decrease
1565: of  the entropy difference  between the liquid  and  the crystal ({\bf
1566: b}), and the characteristic energy level  on the (schematic) potential
1567: energy landscape   ({\bf c})  for  a  fragile glass-former  at various
1568: temperatures. The   ideal   glass level  corresponds   to the Kauzmann
1569: temperature  (see I-4) and     to an extrapolated  divergence   of the
1570: relaxation  time;   Goldstein's   crossover (see   III-3)   takes place
1571: somewhat  below point $2$.  (From Ref.\cite{R17}.)}
1572:  \end{figure}
1573:  \begin{figure}
1574: \caption[99]{Illustration  of    geometric
1575: frustration for spherical particles. {\bf a)} Packing in 2 dimensions:
1576: equilateral triangles  are preferred locally  and  combine  to form  a
1577: hexagonal local cluster that can tile space to generate a close-packed
1578: triangular  lattice. {\bf b)}  Packing in $3$ dimensions: tetrahedra are
1579: preferred  locally  and combine (with slight   distortions)  to form a
1580: regular icosahedral cluster; however,  the $5$-fold symmetry axes of the
1581: icosahedron preclude a simple icosahedral space-filling lattice. (From
1582: Ref. \cite{R76}.) }
1583:  \end{figure}
1584: \end{document}
1585: