Quick viewing(Text Mode)

Relaxational Dynamics in the Glassy, Supercooled Liquid, and Orientationally Disordered Crystal Phases of a Polymorphic Molecular Material

Relaxational Dynamics in the Glassy, Supercooled Liquid, and Orientationally Disordered Crystal Phases of a Polymorphic Molecular Material

PHYSICAL REVIEW B VOLUME 59, NUMBER 14 1 APRIL 1999-II

Relaxational dynamics in the glassy, supercooled liquid, and orientationally disordered crystal phases of a polymorphic molecular material

M. Jime´nez-Ruiz, M. A. Gonza´lez, and F. J. Bermejo Instituto de Estructura de la Materia, Consejo Superior de Investigaciones Cientı´ficas, Serrano 123, E-28006 Madrid, Spain

M. A. Miller and Norman O. Birge Department of Physics and Astronomy and Center for Fundamental Materials Research, Michigan State University, East Lansing, Michigan 48824

I. Cendoya and A. Alegrı´a Departamento de Fı´sica de Materiales, Facultad de Quı´mica, Apartado 1072, 20080 San Sebastia´n, Spain ͑Received 14 September 1998͒ The relaxational dynamics of the ambient pressure phases of ethyl alcohol are studied by means of mea- surements of frequency dependent dielectric susceptibility. A comparison of the ␣ relaxation in the super- cooled liquid and in the rotator phase crystal indicates that the molecular rotational degrees of freedom are the dominant contribution to structural relaxation at temperatures near the transition, the flow processes having lesser importance. Below the a secondary ␤ relaxation is resolved for the orientational and structural . Computer molecular-dynamics results suggest that localized molecular librations, strongly coupled to the low-frequency internal molecular motions, are responsible for this secondary relax- ation. ͓S0163-1829͑99͒00914-5͔

I. INTRODUCTION Most studies to date have been conducted on materials that can be studied within the supercooled liquid ͑SCL͒ state The dynamics of glass-forming materials in the highly for a reasonable long period of time. This excludes simple viscous regime characteristic of the supercooled liquid have binary alloys, which may be described in terms of soft been extensively studied during recent years from theoretical spheres. Instead, materials with a complexity ranging from and experimental points of view. Empirically, some regulari- ionics to high polymers have been scrutinized4 and the re- ties have been found in a vast number of glasses, which lead sults have sometimes been interpreted within the framework to their identification as fingerprints of glassy dynamics. of theories couched for simple hard-sphere fluids. This prob- Mechanical or dielectric spectroscopies are sensitive to lem is only recently being addressed with the introduction of mode coupling approximations which take explicit account motions with frequencies stretching from macroscopic to 5 mesoscopic scales. This makes them especially suited for the of the nonsphericity of the particles and orientation depen- exploration of the dynamics of glass-forming liquids within dent forces. While the rotational dynamics in low molecular their highly viscous regimes ͑i.e., shear about liquids is usually understood on hydrodynamic grounds 102 –1010 P). The strongest peak in the dielectric spectrum ͑Stokes-Einstein-Debye͒,6 such approximations usually known as the relaxation shows some features regarded as ␣ break down within the SCL regime. This happens as a con- ‘‘universal.’’ These include a non-Debye behavior of its line sequence of the sluggish motions which take place when the shape and a temperature dependence of the relaxation time shear viscosity exceeds some tens of poises, which makes ␶(T) strongly departing from Arrhenius as the temperature is any description in terms of uncoupled rotations and center- decreased towards the glass transition temperature Tg , and of-mass diffusion of scarce value. In actual fact, on strict showing a divergence at temperatures below Tg . Such a de- hydrodynamic grounds, the same dependence on viscosity pendence is usually parameterized by means of the empirical ͑and temperature͒ is expected for both mass diffusion and 1 Vogel-Tamman-Fulcher law. Secondary, ␤, or sub-Tg relax- rotational coefficients. Therefore a test of the hydrodynamic ations are detectable at higher frequencies, usually but not predictions could be carried out by studying a single system exclusively at temperatures well below Tg . They commonly where one of the two kinds of motions can be cleanly iso- show Arrhenius behavior and have been measured for many lated by experimental means. glassy systems, albeit reliable data are scarce since the sig- As described previously, ethyl alcohol can be prepared in nals are very weak and broad. Whether this relaxation arises two phases showing quantitatively close ‘‘glassy’’ behaviors from intramolecular motions which remain active below Tg , at the same temperatures and very close densities, one of a view defended by Wu2 and others, or it is an intrinsic these phases with topological disorder ͑liquid and glassy͒ property of glasses due to local rearrangements taking place and the other showing orientational disorder ͑rotator crystal within some minima of the potential energy,3 still needs to be and orientational glass phases͒.7 By orientationally disor- clarified. dered crystals, we refer to those crystalline solids where the

0163-1829/99/59͑14͒/9155͑12͒/$15.00PRB 59 9155 ©1999 The American Physical Society 9156 M. JIME´ NEZ-RUIZ et al. PRB 59 molecular centers of mass sit at lattice positions of a long- ethyl alcohol ͑ethanol͒ obtained from Quantum Chemical range periodic array, while molecular orientations remain Company of Tuscola, Illinois. To avoid contamination with disordered. Such disorder may be of dynamic nature as hap- water vapor, filling of the capacitor was carried out at room pens in the rotator phase crystal ͑RP͒ where the molecules temperature in a dry nitrogen atmosphere. For the low- execute whole body rotations,8,9 or static as in the orienta- temperature measurements (TϽ110 K) a sealed coaxial ca- 16 tional glass ͑OG͒ state which is attained by freezing the ro- pacitor identical in design to that used in our previous work tations into random orientations. Both phases thus provide a was used. After filling the capacitor, the sample cell was way to separate the contributions of the rotational and center- frozen, evacuated of remaining nitrogen gas, and sealed. The sample cell was then mounted inside a nested cryocan ar- of-mass motions to the dynamics. 16 The RP OG rotational freezing transition exhibits most rangement described elsewhere. Finally, the entire cryocan of the characteristics→ of a purely dynamical phenomenon. It assembly was immersed in a double-walled glass dewar. This design provided a temperature control within 0.05 K shows a jump in the specific heat across the transition10 con- over the range 4–300 K, and minimized the occurrence of comitant with a jump in the thermal expansivity,11 without temperature gradients across the sample. For the measure- any other change attributable to a structural transition. Also ments at TϾ150 K, in the liquid state, and high frequencies, and in parallel with the canonical liquid glass, the transi- → 12 1 MHz–1 GHz, a parallel plate sample capacitor was tion can be induced by application of moderate pressures. mounted as part of the inner conductor of the coaxial mea- To investigate in more detail the close similarity between surement line. The temperature control in this case was the relaxation processes of ethanol in the amorphous and achieved by employing a nitrogen-jet heating/cooling system orientational glass form, low-temperature specific-heat mea- with temperature stability better than 0.1 K. surements and neutron time-of-flight ͑TOF͒ spectroscopy The sample preparation regarding the normal glass phase have been carried out, and the results reveal that at low tem- and the rotator phase ͑RP͒ crystal is straightforward since it perature the vibrational density of states of these two phases only involves a quench below Tgϭ97 K at a rate greater is very similar.14 Such dynamic proximity has also been ex- than 6 K/min ͑glass͒ and a subsequent at 110 K tended to macroscopic scales by means of dielectric spectros- during half an hour ͑RP͒. The orientational glass ͑OG͒ is 15 OG copy measurements, and the results lead to the consider- formed from the RP by cooling down below Tg ϭ97 K. ation of this material as an ideal candidate for studies of the The measurements of the ␣ relaxation span the frequency glass transition. range from 1 mHz–10 MHz, of the liquid state from 1 Here, our purpose is to provide as detailed a comparison MHz–1 GHz, and the ␤ relaxation from 10 Hz–10 MHz. as possible of the relaxational processes taking place at both A number of computer simulations of the glass, super- sides of the glass transitions in glassy and orientationally cooled liquid, and crystalline modifications of the material disordered crystalline ethanol by means of dielectric spec- have been carried out.17 The results relevant for this work are troscopy. Our aim goes well beyond that of our previous those obtained from a simulation using the optimized poten- report15 since a far wider spectral band, comprising both ␣ tial model for liquid simulations ͑OPLS͒,18 which describes and ␤ relaxations is considered. Since preliminary data15 in- reasonably well the static and some dynamic properties of dicated a close analogy between the main relaxation appear- the normal-liquid phase.19 The system consisted of 216 mol- ing in both SCL and RP phases, here we pursue a study of ecules at densities of 0.941 g cmϪ3 and a temperature of 80 the relaxation detectable below freezing in both solids as K which corresponds to the glass phase, and 0.888 g cmϪ3 well as of the high-frequency wing of the spectra of both and 180 K, which, within the computer model corresponds to SCL and RP. The motivation behind this is the possibility the SCL. The initial configuration was obtained from a offered by such a chemically simple material to unravel quench of the liquid at 298 K at a cooling rate of dT/dt some of the details of the dynamics at microscopic scales ϭ0.1 K psϪ1. Additional details about the model and the which lead to the observation of macroscopic relaxations, by simulations ͑of the liquid, supercooled liquid, and glass͒ will means of the concurrent use of experimental and computer be given elsewhere.17 simulation approaches. The functions evaluated from the computed trajectory which are relevant for the present study are the standard orientational correlation functions for a vector fixed to some II. EXPERIMENTAL AND COMPUTATIONAL DETAILS molecular frame. Two choices were made. One of them lo- cates the vector parallel to the molecular dipole moment Measurements of the frequency and temperature- Ϫ30 dependent complex dielectric susceptibility were carried out (␮ϭ2.22 DϷ7.405ϫ10 C m), whereas the other posi- using three different schemes depending upon the frequency tions it along the O–H molecular bond. Also, the Fourier range under consideration. In the range 1 mHz–10 kHz, the transforms of the angular velocity autocorrelation of the OH in-phase and out-of-phase currents through the capacitor and ␮ vectors ͓P␮(␻) and POH(␻)] were evaluated. For a were measured using a Stanford SR850 dual phase lock-in solid or a highly viscous liquid these functions provide a amplifier together with a Stanford SR570 current-to-voltage measure of the rotational components of the weighted vibra- amplifier. From frequencies of 1 kHz–10 MHz, a Hewlett- tional density of states, and therefore can be related to mea- surements of the experimental Z(␻) frequency distribution Packard 4192A impedance analyzer was used in a standard 10,12 four-terminal pair configuration to measure the capacitance derived from neutron inelastic scattering. and conductance of the sample. From 1 MHz to 1 GHz a III. RESULTS Hewlett-Packard 4191A rf impedance analyzer was used, with a coaxial measurement line. In what follows the results concerning the two spectral The sample material consisted of 200-proof, dehydrated regions are described separately. PRB 59 RELAXATIONAL DYNAMICS IN THE GLASSY,... 9157

FIG. 1. The real ͑a͒ and imaginary ͑b͒ parts of the dielectric FIG. 2. The real ͑a͒ and imaginary ͑b͒ parts of the dielectric susceptibility vs frequency at temperatures of 96 K ͑glass͒,98K, susceptibility as a function of frequency at temperatures of 96 K and 100 K ͑supercooled liquid͒. The solid lines are fits to two ͑orientational glass͒ and from 98–108 K ͑rotator phase crystal͒. The Cole-Davidson functions with different relaxation times. solid lines are fits to two Cole-Davidson functions.

22 A. ␣ relaxation materials. In particular, ⑀Љ grows linearly with frequency below the peak, then decreases with a power law somewhat 15 In our previous report the main relaxational processes in less than 1 above the peak, and finally flattens out to a more both phases were compared in terms of the real and imagi- shallow power law starting at a frequency about two decades nary parts of the dielectric susceptibility, which were fitted above the peak. The presence of a universal curve implies 20 simultaneously to the empirical Cole-Davidson form. Al- certain relationships between the two power laws above the though this provides good fits for the main relaxation, the peak, as well as between the peak frequency and the fre- model function shows a clear deviation from the experimen- quency at which the slope changes. Rather than scaling our tal data at higher frequencies. In particular, data for the data onto the ‘‘Dixon-Nagel scaling plot,’’ we prefer to ex- imaginary part of the dielectric susceptibility lie systemati- amine the raw data on a simple log-log plot where the power cally above the Cole-Davidson curves for all frequencies laws are most clearly visible. In the search for a reliable way starting about two decades above the peak frequency. Fig- to parameterize the data, we found that the sum of two Cole- ures 1͑a͒ and 1͑b͒ show the real (⑀Ј) and imaginary (⑀Љ) Davidson functions provides excellent fits to the data over parts of the susceptibility of the supercooled liquid ͑SCL͒ for the entire frequency range, while ensuring that the simulta- temperatures of 96, 98, and 100 K. Figures 2͑a͒ and 2͑b͒ neous fits to ⑀Ј and ⑀Љ obey the Kramers-Kro¨nig relation- show the corresponding quantities for the rotator phase ͑RP͒ ship. The fitting function is given by crystal for temperatures from 96 to 108 K. The logarithmic plots of ⑀Љ show a broad tail starting about two decades 2 ͑⑀0Ϫ⑀ϱ͒i above the peak frequency and extending to the highest ex- ⑀ϭ⑀ϱϩ , ͑1͒ ͚ϭ CD ␣i plored frequencies. We have followed two different routes to i 1 ͑1Ϫi␻/␻pi ͒ interpret this tail. The first considers that the susceptibility curves shown in Figs. 1 and 2 are characterized by a single where (⑀0Ϫ⑀ϱ)i is the relaxation strength, ␻pi is the peak Ϫ1 relaxation whose shape is nearly universal, as first suggested frequency ͑related to the mean relaxation time, ␶iϰ␻pi ), ⑀ϱi 21 by Dixon et al. Those authors showed that dielectric data is the susceptibility at high frequencies and ␣i is the width from several different materials at many temperatures could parameter23 which usually varies between 0 and 1. The De- be made to coincide when plotted on a particular type of bye ͑Lorentzian͒ function is approached as ␣i 1. logarithmic scaling plot. The ‘‘universal’’ curve they found Using the Kramers-Kro¨nig relationship,→ the real and on the scaling plot corresponds to the features seen in raw imaginary parts of the dielectric susceptibility, ⑀Ј(␻) and data such as that shown in Figs. 1 and 2, or from other ⑀Љ(␻) are fitted simultaneously: 9158 M. JIME´ NEZ-RUIZ et al. PRB 59

2 cos͓␣iarctan␻␶i͔ ⑀Јϭ⑀ϱϩ ͑⑀0Ϫ⑀ϱ͒i , ͑2͒ ͚ϭ 2 2 ␣i/2 i 1 ͓1ϩ␻ ␶i ͔

2 ͑⑀0Ϫ⑀ϱ͒isin͓␣iarctan␻␶i͔ ⑀Љϭ . ͑3͒ ͚ϭ 2 2 ␣i/2 i 1 ͓1ϩ␻ ␶i ͔

The fit determines the seven parameters, ␻p , ⑀0Ϫ⑀ϱ , and ␣ of the two different relaxations, and ⑀ϱ . These parameters fully characterize the shape of the relaxation in the frequency range from mHz to MHz and are temperature dependent. The first Cole-Davidson function fits the main part of the peak, and departs only mildly from a Debye relaxation (␣ Ϸ0.8). The second Cole-Davidson function fits the very shallow power law at high frequency, hence it departs strongly from Debye relaxation (␣Ϸ0.3). To minimize the CD number of adjustable parameters, the relaxation times ␶i of the Cole-Davidson functions were constrained to be equal. A further examination of the exponents ␣1 and ␣2 shows that they obey the relationship implied by the Dixon-Nagel scal- ing form. Further details of this analysis can be found in Ref. 24. An alternative view follows the approach of Garg et al. and others,25–29 which analyzed the spectrum of monohydric alcohols in terms of three different dispersion processes. The rationale behind such a treatment is to enable a connection with the high-temperature liquid, where as detailed below, the whole relaxation spectrum can be understood on the basis FIG. 3. The real ͑a͒ and imaginary ͑b͒ parts vs frequency in the 27–29 of phenomenological treatments or fully atomistic liquid phase of ethanol. Solid lines are fits to the Debye function. simulations.17,19 To proceed, we assume that there exist ͑at least͒ two different relaxation processes, a main relaxation (␣) and a secondary contribution which appears at higher frequencies. Operationally, the only difference between this fitting scheme and the previous one is that now we let ␶2 be a free parameter independent of ␶1. The real and imaginary parts have thus been fitted to two Cole-Davidson forms given by Eqs. ͑2͒ and ͑3͒ labeling iϭ1 the main relaxation and i ϭ2 the secondary relaxation which could not be tracked at all temperatures within the normal-liquid phase. Figures 3͑a͒ and 3͑b͒ show the real (⑀Ј) and imaginary (⑀Љ) parts of the dielectric susceptibility of the normal-liquid phase of ethanol at several temperatures between 160 and 270 K. The solid lines represent the fits to two Cole- Davidson functions at 160 K and to one Cole-Davidson from 170–270 K ͑only one dispersion process is observed in the frequency range͒, plus a conductivity term in the imaginary part arising from the sample itself and sample cell which is subtracted ͓Eq. ͑3͔͒, ␴dc /␻. The results are summarized below: ͑i͒ Figures 4͑a͒ and 4͑b͒ show the temperature depen- dence of the Cole-Davidson width parameters of the main and secondary relaxations for both the supercooled liquid ͑SCL͒ and rotator phase ͑RP͒. The decrease of both ␣1 and ␣2 with decreasing temperature signals a deviation from the Debye behavior. The fact that at the same temperature ␣(SCL) Ͼ␣(RP) for both relaxations merits some remarks. The stronger deviation from exponentiallity in the spectra of the RP can be thought of as due to orientational correlations FIG. 4. The temperature dependence of the width parameters ␣1 between neighbor units ͑presumed more important in the RP and ␣2 of the main relaxation for the supercooled liquid glass and 8 → crystal than in the structural glass ͒. The inset of Fig. 4͑a͒ rotator phase orientational glass transitions. The inset shows ␣ → 1 shows the temperature dependence of ␣1 of the liquid phase. for the liquid phase. PRB 59 RELAXATIONAL DYNAMICS IN THE GLASSY,... 9159

FIG. 6. The mean relaxation times vs 1000/T for the ␣ ͑liquid phase, SCL, and RP crystal phases͒, secondary ͑RP͒ and ␤ ͑struc- tural glass and OG͒ relaxations. The solid lines are fits to the Vogel- Tamman-Fulcher form for the ␣ relaxation and to the Arrhenius form for the ␤ and secondary ones. The parameters obtained after the fitting are listed in Table I.

Figure 6 also shows the time scale of the ␤ relaxation of the corresponding low-temperature phases ͑structural glass and orientational glass͒, but further comments are deferred to the FIG. 5. The temperature dependence of the relaxation strengths next section. From Fig. 6, it is clearly seen that the relaxation times in ⑀0Ϫ⑀ϱ of the parts of the main relaxation for the supercooled liquid ͑SCL͒ and rotator phase ͑RP͒ crystal. the SCL and RP phases are about the same at Tϭ96 K, and as the temperature increases a shift to a longer relaxation The values are almost equal to 1, a behavior indistinguish- times in the RP relative to the SCL is observed. This behav- able from isotropic rotational diffusion ͑Debye͒. ior can be explained considering the degrees of freedom ͑ii͒ Figures 5͑a͒ and 5͑b͒ show the temperature depen- available through the glass transition in each phase. The dence of the the relaxation strength ⑀ Ϫ⑀ for the two re- OG RP transition involves the melting of rotational de- 0 ϱ → laxations. The decrease in the relaxation strength below 104 grees of freedom, while the glass SCL transition involves → K in both the RP and SCL data was noted in our previous rotational and translational degrees of freedom. As the tem- work.15 The decrease cannot be attributed to partial crystal- perature increases above 96 K, the additional translational lization of the sample because the SCL data were taken on degrees of freedom present in the SCL provide the molecules warming of the sample, and the RP data were confirmed to with a less restrictive local environment, leading to increas- be reproducible on cooling and warming. The data may in- ingly shorter relaxation times in the SCL relative to the RP dicate a slight tendency toward antiparallel alignment of for a given temperature. This interpretation is in agreement neighboring molecules as the temperature is lowered. with the observed values for the jump in specific heat at both 8 ͑iii͒ The dependence of the mean relaxation times ␶1 with transitions, the RP OG transition involves a ᭝C P of about Ϫ1 Ϫ1 → inverse temperature, obtained from fits to the Cole-Davidson 22 J mol K and the canonical glass transition SCL function of data for the RP and SCL phases and to the Debye glass is accompanied by a jump in ᭝C P of function in the liquid phase, is shown in Fig. 6. The solid line →31 J molϪ1 KϪ1. The extra 9 J molϪ1 KϪ1 can be assigned shows the fit to the Vogel-Tamman-Fulcher form: to the translational degrees of freedom that are not activated in the OG RP transition. These results suggest that the → ␶ϭ␶ eE0 /͑TϪT0͒, ͑4͒ rotational degrees of freedom are the dominant contributor to 0 structural relaxation processes in ethanol, whereas mass dif- fusion associated with translational degrees of freedom in the where ␶0 is the attempt relaxation time and T0 is the Vogel temperature. The fitting parameters for both SCL and RP phases are summarized in Table I. TABLE I. Values obtained by fitting the temperature depen- Also plotted in Fig. 6 are data for the dependence with dence of the mean relaxation time in the RP crystal and SCL phases temperature of the second relaxation time ␶ corresponding to the Vogel-Tamman-Fulcher equation, and in the glass and OG to 2 the Arrhenius law. The f value is defined as 1/(2␲␶ ). to the high-frequency wing of the ␣ peak if considered as an 0 0 additional relaxation process ͑see above͒. A significant set of Relaxation Phase ␶ ͑sec͒ f ͑Hz͒ E ͑K͒ T data was obtained only for the RP crystal and this is shown 0 0 0 0 in Fig. 6. The point that merits some attention concerns the ␣ SCL 2.3ϫ10Ϫ11 2.7ϫ1011 693 73.1 Ϫ8 8 disparate behavior with temperature of ␶2 and the main ␣ RP 3.2ϫ10 2.0ϫ10 525 73.4 Ϫ14 12 relaxation time ␶1. The former seems to follow a well de- ␤ Glass 6.5ϫ10 2.4ϫ10 941 0 fined Arrhenius behavior with an activation energy of 3220 OG 1.6ϫ10Ϫ14 1.0ϫ1013 1203 0 K, in contrast to the Vogel-Fulcher behavior of the latter. 9160 M. JIME´ NEZ-RUIZ et al. PRB 59

FIG. 7. Imaginary ⑀Љ part of the dielectric susceptibility of the FIG. 9. Comparison of the temperature dependence of the width structural glass phase of ethanol at several temperatures for the ␤ of the energy barrier distribution between the glass and OG phases relaxation. The continuous lines are fits using Eq. ͑5͒. of ethanol in the ␤ relaxation. The solid and dashed lines are the linear fits of the glass and OG, respectively. SCL contributes to a lesser extent. This result is also sup- ported from the values of T0 in Table I, almost equal for both The shape of the ␤ relaxation peaks can be described with phases, which provide an additional indication of the fact a log-normal distribution, a form which has been used for that at low temperatures the rotations dominate the dynamics some other glassy crystals,30 which reads of the SCL. ͑iv͒ As also shown in Fig. 6, the dynamics of the SCL and RP take place at scales which approach each other as mo- 2 Ϫ [log␻Ϫlog␻p] ⑀0 ⑀ϱ Ϫ tions get progressively frozen. Below Tg , motions giving ⑀͑␻͒ϭ e W2 , ͑5͒ rise to dielectric response should be originated by sequential ͱ␲W molecular reorientations. These should be able to survive after whole molecule large-angle rotations freeze out at tem- where the fitting parameters are the peak frequency of the peratures below T and will give rise to additional relax- g relaxation, ␻ ͑related to the mean relaxation time ␶ ational response as described in the next section. This agrees p ϭ Ϫ1 Ϫ with work of Ramos et al. where they conclude, by means of ␻p ); the relaxation strength ⑀0 ⑀ϱ , and the full width at low-temperature specific-heat measurements and neutron half maximum in decades W. Its functional form can be de- 14 rived if one postulates the presence of a Gaussian distribu- TOF spectroscopy, that the low-frequency dynamics of the 30,31 glass and the OG are remarkably close. tion of energy barrier heights, and is written as

B. ␤ relaxation 1 ⑀͑␻,T͒ϭ⑀ϱϩ͑⑀0Ϫ⑀ϱ͒ dE Figures 7 and 8 show the imaginary parts (⑀Љ) of the ͵ ͱ␲␴ susceptibilities for the structural and orientational glass ͑OG͒ phases at several temperatures well below the glass transi- 2 2 1 ϫeϪ͑EϪE0͒ /␴ . ͑6͒ tion. Data concerning the real part were too weak to be sig- E/KT nificant. 1Ϫi␻/␻0e

The width of the energy barrier distribution is given by ␴ϭWKT ln10 and the peak energy E0 from the slope of the Arrhenius plot. ͑i͒ Figure 6 shows the temperature dependence of log(␶), obtained after fitting the imaginary part to Eq. ͑5͒, for the structural glass and OG. Again, one can see the closeness of behavior below Tg for both phases, as was discussed in the previous section. The solid lines represent the corresponding E0 /KT fits with the Arrhenius law ␶ϭ␶0e , where E0 is the activation energy of the process. The parameters obtained from the Arrhenius fits for both phases are given in Table I. The graph also shows the ␣ relaxation and its fit to the VTF law. ͑ii͒ In Fig. 9 we plot the temperature dependence of the FIG. 8. Imaginary ⑀Љ part of the dielectric susceptibility of the distribution of energy barriers, ␴, obtained from the peak OG of ethanol for the ␤ relaxation. The lines correspond to the width W. We find that the values for both phases have a fitting to the Eq. ͑5͒. roughly linear dependence with temperature, ␴ϭ␴0Ϫ␤T, PRB 59 RELAXATIONAL DYNAMICS IN THE GLASSY,... 9161

ing process than the motions related to mass diffusion which, because of the high viscosities involved, are energetically more costly to execute. At higher frequencies, the first point to consider is whether the trailing shoulder of the ␣ peak can be sensibly ascribed to an additional relaxation. As stated above, two mutually exclusive rationalizations can be brought to the fore. Our data, although not conclusive, show some evidence signaling a dependence with temperature of the high- frequency wing significantly different from that of the main peak. If this were the case, then the relaxation at such fre- quencies could possibly be ascribed to some of the sources of relaxation discussed in the past by a number of authors ͑molecular clusters of varying sizes͒.19,25,26,29 The present findings provide however clear hints showing that such as- FIG. 10. Temperature dependence of the relaxation strength signments are inadequate. First, the fact that the high- (⑀ Ϫ⑀ ) in the ␤ relaxation for the glass and OG phases of etha- 0 ϱ frequency tail appears in the RP seems to invalidate the nol. view25 attributing this relaxation to reorientations of free molecules ͑‘‘monomers’’ at the end of a hydrogen-bonded where ␴0 is the zero-temperature width of the energy barrier 8 distribution, which we obtain by assuming a linear extrapo- ‘‘chain’’͒. In fact, as shown previously, within the RP crys- tal all molecules must be considered on the same footing, lation to Tϭ0K. which makes the distinction between ‘‘monomers’’ and as- ͑iii͒ The temperature dependence of the ␤ relaxation sociated units of scarce utility. Rather, a combination of strength (⑀ Ϫ⑀ ) is shown in Fig. 10. The absolute values 0 ϱ mechanisms involved in the ␣ and those for the ␤, which are for both phases are small if compared with the ␣ relaxation. described below, seems physically more sound. The ␤ relaxation strength in the OG is about 1.3 times larger At temperatures well below T molecular motions in the than in the glass. Such a difference in relaxation amplitudes g glass and OG phases involving whole body rotations are fro- between the structural and orientational glasses can be un- zen and therefore the dielectric response cannot arise from derstood on the basis of the existence in the latter of a crys- molecular large-angle reorientations. A way to ascertain the talline bcc lattice. To see this one only has to recall that the origin of such a relaxation would be to extrapolate the value orientational glass structure is basically a snapshot of that of of its relaxation time up to normal-liquid temperatures where the RP crystal, and that in the latter molecular rotations are the presence of several relaxation processes has been docu- significantly less hindered than in the SCL, as noted above. mented for monohydric alcohols.26 In doing so, a value of Upon freezing the large-angle, whole molecule rotations one 1.66 ps is found for room temperature. This value is not too would expect that segmental and librational motions will far from the lifetime of a hydrogen bond in water, ␶HB take place in the orientational glass phase with less hindrance 32 than in the SCL. This point constitutes a dynamical correlate ϭ0.54 ps, deduced from Rayleigh scattering experiments, and is very close to that deduced by Barthel et al.26 for liquid of what is found in calorimetric measurements of C (T),8 P ethanol at room temperature. In fact values within the inter- which show that for a range of temperatures below the two val within 0.22 psр␶ р1.81 ps have been reported19 from transitions (86рTр95 K) COGϾCglass , that is a larger 3 P P studies combining dielectric and infrared spectroscopies number of particles participate in the motions within the dis- where ␶ , the shortest relaxation time of the three distin- ordered crystal. 3 guishable relaxation steps was acribed to the relaxation of the hydroxyl group. On such grounds one can explain why IV. DISCUSSION the ␤ relaxation is observed at the same frequency for a At temperatures where the ␣ relaxation dominates the given temperature in the different phases, but the relaxation spectrum, the main contribution to the dielectric spectra of strength, ⑀0Ϫ⑀ϱ , is higher in the OG phase. both supercooled liquid and plastic-crystalline ethanol should be ascribed to molecular reorientations of some kind. A. Origin of the observed relaxations Such an assertion is based upon: ͑a͒ the Rietveld analysis of In what follows a tentative assignment of the relaxations the powder diffraction pattern of the rotator phase crystals observed within the normal liquid as well as below T to which indicate that single molecules reorient quasi- g underlying molecular motions is proposed. The relaxation isotropically, ͑b͒ the comparison of the relaxation times for dynamics within the supercooled liquid and rotator phase both SCL and RP referred to above, and ͑c͒ results of recent crystals reveals substantially more complicated features than neutron scattering measurements13 which show that the rota- those observed in these two limiting cases. tional melting involves fast motions on a scale of about a picosecond. The result seems to be of relevance for studies of the glass transition, where it is often assumed that the 1. High-temperature regime freezing the SCL phase is mostly dominated by the arrest of The presence in liquid monohydric alcohols at tempera- translational degrees of freedom ͑i.e., identification of the ␣ tures within the normal-liquid range of several dynamical relaxation with some flow process͒. Our results show that processes which contribute to the quasielastic neutron and reorientational motions play a more important role in freez- light-scattering spectrum ͑i.e., arising from stochastic mo- 9162 M. JIME´ NEZ-RUIZ et al. PRB 59 tions͒ seems now well established.27,28 In fact, several spec- frequency data for room-temperature ethanol reported by 26 tral components are needed to account for the observed line Barthel et al. The required values for p0,2 and ␶HB are taken 19,29 shape of the Sq.el(Q,␻) quasielastic neutron scattering spec- from molecular dynamics results, and rough estimates trum. The spectrum shows pronounced broadenings of the for the rotational coefficients for each ith state DR(␩i) were quasielastic ͑i.e., centered at zero frequency͒ peak with line- derived on hydrodynamic grounds, that is using the Stokes- widths stretching over the experimentally accessible fre- Einstein-Debye formula for the corresponding molecular quency window ͑0.25 GHz–2 THz͒. A comparison with volumes, temperature, and viscosity. Using the constraint measurements of the high-frequency dielectric response ͑i.e., ¯ ¯ DRϭ͚iϭ0,2f iDR(␩i), where DR was identified with the ex- measurable by infrared techniques͒ may be achieved after perimental value for the rotational coefficient measured by converting the neutron spectra into generalized susceptibili- NMR ͑Ref. 33͒ enables one to put the hydrodynamic esti- ties, that is ␹(Q,␻)ϭ␲Sq.el(Q,␻)͓1ϩn(␻)͔, the term mates into an absolute scale. There, f i stands for the fraction within square brackets being the Bose factor. In addition, of molecules in a given state and the only free parameter left computer molecular-dynamics19 calculations of the dielectric is ␶ f . Solution of Eq. ͑8͒ with such parameters leads to an function for the high-temperature liquid show that the ob- estimate of the three relaxation times as 160, 3.5, and 1.8 ps, served spectra may be understood in terms of three distinct 26 which compares with 163, 8.97, and 1.91 ps, as reported in relaxation steps which can be reproduced on semiquantita- Ref. 26. That is, it correctly predicts the relaxation time of tive grounds using a moderately sized simulation box about ͑ the main ␣ dielectric band in the high-temperature liquid, 23 Å on a side͒. showing that it roughly coincides with the time spent by a The presence of several relaxations was rationalized in the molecule in the structured part of the liquid, and accounts for past by recourse to some phenomenological models such as 29 the presence of additional relaxation processes. that due to Bertolini et al. which pictures the deviations Such a picture, although physically appealing, loses its from ideality ͑exponential relaxation͒ of both the rotational validity within the deeply supercooled liquid. In fact, at tem- and translational diffusion processes as driven by the hydro- peratures well below Ϸ140 K the predicted behavior of gen bond dynamics. In the case of dielectric relaxation, the ␶(T) for the ␣ relaxation lies substantially below experi- characteristic times are calculated by solution of an equation ment. The interesting point is that this temperature, which of motion for the correlation function for the total dipole (i) also marks the limit of stability of the supercooled liquid, moment ⌽␮ of a given molecule found in the ith environ- signals a strong departure from Arrhenius behavior of a num- ment. That is, molecules within the sample are assumed to be ber of momentum and charge transport properties.34 The found in a discrete number of states, which are here inter- model cannot follow the strong temperature dependence of preted as the number of bonds to a neighboring molecule. 17,19,29 ␶(T) but still predicts the presence of additional relaxations Data from computer molecular dynamics identify three which show an Arrhenius behavior of their frequencies, and possible states for these alcohols corresponding to particles show characteristic residence times ␶ within 100–350 ps. with 0 – 2 bonds to neighbors ͑in a time-average sense͒, HB 29 The failure of Eq. ͑8͒ to account for the dynamics of the SCL respectively. The equation of motion then reads is understood as a breakdown of the validity of the approxi- mations upon which such treatment was built. In a nutshell, d⌽͑i͒ 2 it simply illustrates the breakdown of the Brownian- ␮ ͑i͒ ͑j͒ ϭϪDR͑␩i͒⌽␮ ͑t͒ϩ ͑K␩͒ij⌽␮ ͑t͒, iϭ0,2, dynamics regime, a departure which is also witnessed by dt j͚ϭ0, several transport properties.34 ͑7͒ where DR(␩i) stands for the rotational diffusion coefficient 2. ␤ relaxation for molecules in the ith state, and (K␩)ij is a matrix of ki- The most striking results coming from the measurements netic coefficients connecting the states iϪ j which is given in below the two glass transitions concern the closeness of the terms of the probabilities for each ith state. The relevant activation energies E0 and widths of the barrier distributions parameters to specify such transition rates are the ␴ for both the structural and orientational glasses. The latter ‘‘hydrogen-bond’’ lifetime ␶HB , the time ␶ f which one bond quantities can be extrapolated to Tϭ0K,␴(T)ϭ␴0Ϫ␤T may last as broken and finally the probabilities p0,2 of mol- yielding values of 14.5 and 11.1 THz, respectively. Such ecules in each state. All such parameters are in principle figures and especially the ratios ␴ 0 /E0ϭ 0.74 and 0.45 are obtainable from molecular simulations and, as we will see comparable to others measured in the glassy crystal, such as below, only a rough estimate of them is required to interpret (KBr)1Ϫx(KCN)x with ‘‘defect’’ concentrations within its the general trends of the dielectric relaxation spectra. ‘‘glassy’’ range.35 Motivated by this analogy, we have fol- Solution of Eq. ͑7͒ is easily achieved in terms of the ei- 36 29 lowed the steps delineated by Grannan et al. with the aim genvalue equation of identifying the microscopic entities giving rise to fre- quency dependent dielectric signals. det͉͉͓DR͑␩i͒Ϫ␭ϩK␩͔I͉͉ϭ0 ͑8͒ Well below the glass transitions, rigid body molecular reorientations involving large amplitude angular excursions from where the relaxation times are determined from the which would give rise to dielectric response arising from calculated eigenvalues. motions of the molecular dipoles seem highly unlikely. In- An assessment of the capability of Eq. ͑8͒ to reproduce stead, the motions involving some molecular segment carry- the most salient features of the dielectric spectrum was car- ing a large local dipole moment may indeed take place as a ried out using present results supplemented with the high- consequence of a thermally activated mechanism. From mi- PRB 59 RELAXATIONAL DYNAMICS IN THE GLASSY,... 9163 croscopic modeling of the dynamics of the solids whether in orientationally ordered, orientationally disordered, or glassy phases10 the two most likely candidates to give rise to low- frequency motions not involving large rearrangements of neighboring particles are torsional oscillations about the C–C and C–O molecular bonds. The characteristic harmonic frequencies for such librations were estimated from ab initio Hartee-Fock calculations as 7.3 and 9.1 THz, respectively, and found to correspond to definite features in the experi- mental densities of states.10 Although those frequencies are far too high to contribute to relaxation within the 10Ϫ8 –10Ϫ3 s scale, the point which merits attention con- cerns the strong hybridization between these internal and ‘‘lattice’’ motions which will provide a coupling between FIG. 11. The low-frequency part of the P (␻) and P (␻) these and low-frequency vibrations such as long-wavelength ␮ OH distributions of librational frequencies for a vector directed along hydrodynamic phonons. Such hybridization occurs as a con- the molecular dipole ͑solid line͒ and along the O–H bond ͑dashes͒, sequence of the lack of time-scale separation between low- respectively. The glass ͑80 K͒ and supercooled liquid ͑180 K͒ show energy molecular and phonon modes ͑in fact, the experimen- 10 very close frequency distributions. The inset shows the distribution tal density of states up to 20 THz lacks any clear gap ͒. for the whole range of frequencies. Explicit evaluation of such effects in a glassy material37 shows how hybridization alters dramatically the density of A glance at the graphs shown in Fig. 11 unravels the vibrational states of a solid composed by rigid molecules. strong mixing of modes referred to above. In other words, Indeed, the resulting frequency distribution is remarkably although the strongest peak appearing in P (␻) and P (␻) stretched towards low and high frequencies. In other words, ␮ OH are located at fairly high frequencies ͑about 20 THz͒, a broad such a mode mixing results in a large increase of the vibra- band is seen at low frequencies as a result of the strongly tional density of states at low frequencies which arises from collective nature of most motions taking place in these the admixture of acoustic phonons with the molecular inter- phases. That is, most motions including those below 1 THz nal motions. The latter thus become strongly coupled to the which are dominated by long-wavelength phonons are strain fields and consequently, any excitation of such fields strongly coupled to movements which will generate a re- taking place at relatively low frequency is expected to in- sponse of the polarization field. As expected, both P (␻) volve a substantial molecular deformational ͑i.e., involving ␮ and P (␻) show well defined maxima within frequencies low-energy molecular internal coordinate͒ component. Such OH covered by the broad maxima of the vibrational frequency a mechanism is thus able to explain why the relaxations distributions.12 Such a situation is reminiscent of that of sampled at macroscopic scales can follow the motions of mixed KBr:KCN halide crystals36 where harmonic excita- these high-frequency modes. tions in addition to acoustic phonons are found. These show Between the two plausible candidates referred to above, a density of states g (␻) peaked at about 1 THz and cor- only motions along the C–O bond are expected to be active lib respond to librations of the CN group, whose presence serves in a dielectric relaxation experiment because of the relatively to explain the peak in C (T)/T3, the low-temperature spe- strong OH C local dipole. If this were the case then the P cific heat, and the ‘‘plateau’’ in the thermal conductivity. harmonic frequency→ given above should not be far from Such ‘‘additional harmonic excitations’’ are also well known ␻ ϭͱ(2E /I), where E stands for the barrier height harm 0 0 from studies on canonical glasses such as vitreous SiO ,38 and I is the effective moment of inertia of the reorienting 2 and indeed, show a peaked density of states at about 2 THz. entity. Using data from Table I for E and the value for the 0 The shape of both P (␻) and P (␻) displayed in Fig. moment of inertia of the free molecule, Iϭ2.84 ␮ OH 11 shows the same extrema albeit with rather different ϫ10Ϫ45 Kg m2 one gets frequencies of 3.06 and 3.18 ␻harm weights. This should not come as a surprise in view of the THz for the glass and OG crystal, respectively. The role of a strong contribution of the O–H fragment to the molecular quantity such as ␻harm is to provide a rational basis to un- dipole moment. The P␮(␻) distribution seems however derstand the origin of the large values for the f 0 ‘‘attempt more meaningful if the same arguments given in Ref. 36 are frequencies’’ given in Table I. Using the same arguments as brought to the fore. That is, if one identifies P(␻) as the those given in Ref. 35, one finds that the ‘‘attempt frequen- function giving rise to the maxima in C (T)/T3 at about 5–7 cies’’ f which are about 2.4 THz for the glass and 10 THz P 0 K,10 then it should show a low frequency part not too differ- for the glassy crystal, can be qualitatively interpreted in ent from that shown by the total density of states Z(␻).10,12 terms of the characteristic frequency of some microscopic On such a basis, the P (␻) distribution seems to be the process, such as the libration of some molecular dipole. On ␮ sought quantity. such a basis, we have evaluated two of the most relevant functions which will give rise to finite frequency peaks in the dielectric function. They are the librational frequency distri- 3. Dynamics of the supercooled liquid butions ͑calculated from the autocorrelation of angular ve- The fact that the dynamics of the bcc crystal lies so close locities͒ of a vector parallel to the total molecular dipole to that of the SCL serves to scrutinize the various explana- 3 moment P␮(␻), and of a vector parallel to the O–H bond tions offered in the literature concerning the origin of the POH(␻). The results are shown in Fig. 11. main relaxation. These go from those assuming individual 9164 M. JIME´ NEZ-RUIZ et al. PRB 59

C1(t) for both ␮ and O–H vectors is not too different, ex- ception made of the absolute values as well as of a stronger initial decay of the latter. The results shown in Fig. 12 thus illustrate the presence of different time scales in the orientational dynamics of this material. The strong decay at long times appearing in the SCL can be tentatively identified with the tail of the main ␣ relaxation ͑the simulations are too short to sample the re- quired time scales͒ whereas the region at intermediate times ͑1–100 ps͒, which is common to SCL and glass should be related to the higher frequency relaxation. Its slow decay will translate into a broad spectrum in the frequency domain which will be very weak if compared with that originated by the decay at long times. In fact, the Fourier transforms of the correlation functions show three separate regions corre- sponding to a low-frequency broad band which dominates the spectrum, a second region stretching from about 1–10 THz, and a well defined microscopic peak at a higher fre- quency.

V. CONCLUSIONS The dielectric measurements of the ␣ relaxation have

FIG. 12. The C1(t) reorientational correlation functions derived demonstrated that although the glass transition is regarded as from the computer molecular-dynamics calculations. Solid traces a structural transition, the glasslike transition in RP ͑with show the reorientation of a vector ␮ directed along the molecular translational order͒ exhibits all the characteristics observed dipole moment and dot-dashes depict the function for a vector lo- in the canonical transition SCL glass: a decrease of the cated along the O–H bond. The inset shows an enlargement of the width parameter as the glass transition→ is approached and a decay at short times. relaxation described by the Vogel-Tamman-Fulcher form. The relaxation times above 96 K display the effect of the molecular dipole reorientations to those involving molecular additional translational freedom available to the liquid result- clusters.39 The former possibility is ruled out by consider- ing in faster molecular reorientation in the SCL relative to ation of data from NMR and neutron quasielastic scattering the RP crystal for a given temperature. We can therefore measurements13,14,40 which show that within the rotator crys- conclude that orientational relaxations play an important role tal whole molecule quasi-isotropic8 reorientations within the in the glass transition which has been considered up to recent picosecond scale take place ͑about ten orders of magnitude times as a purely translational transition. faster than those sampled by dielectric measurements͒.On In addition, our results serve to illustrate how molecular the other hand the latter seems wholly incompatible with the motions of different kinds ͑i.e., pure reorientations and mass very existence of the crystalline bcc lattice. diffusion͒ exhibit disparate temperature dependences when To understand the origin of the relaxation under high vis- the macroscopic viscosity reaches the large values character- cosity conditions some help can be sought from the calcu- istic of the supercooled liquid. This can be viewed as a tran- lated computer molecular dynamics trajectories. Specifically, sition from a high-temperature regime where collective ef- we aim to derive some relevant conclusion from comparison fects are reduced to distances comparable to few particle between the reorientational correlation functions diameters, to another where such motions involve meso- or macroscopically large volumes. One then expects that the Brownian dynamics approximation ͑i.e., same temperature Cl͑t͒ϭ͗Pl͓cos␪͑t͔͒͘ ͑9͒ dependence for all the diffusions coefficients͒ will only hold within the normal-liquid phase. To put it somewhat differ- at both sides of the glass transition. Here, Pl stands for the ently, the large number of degrees of freedom which are lth Legendre polynomial and ␪(t) is the angle through which thermally activated in the latter phase are seen as a ‘‘thermal a molecule fixed vector ͑either along ␮ or the O–H bond͒ bath’’ by the reorienting dipoles. In contrast, within the SCL rotates in time t. Figure 12 provides such a comparison. and lower temperature phases collective effects are felt with There, the functions C1(t) which are relevant for dielectric increasing strength as the temperature is lowered leading to a relaxation are shown for the two vectors referred to above for complete absence of separation of the time scales character- Tϭ80 K ͑glass͒ and Tϭ180 K ͑still in the supercooled liq- istic of ‘‘fast’’ ͑i.e., the thermal bath͒ and ‘‘slow’’ motions. uid range for the computer glass͒. All the curves shown there The results also lend additional support to the idea put exhibit an initial fast decay with superimposed oscillations forward from recent studies,10–12 which attributes a purely which last about 0.5 ps followed by a flat or a slowly decay- dynamic character to the RP OG transition. This makes ing regime stretching to several tens of picoseconds. As a the orientationally disordered→ crystals ideal candidates for main distinctive feature of the SCL state a second strong the study of the glass transition, and also enables one to draw decay is seen after about 100 ps or so. Again, the shape of a parallelism with the well studied cases of orientational PRB 59 RELAXATIONAL DYNAMICS IN THE GLASSY,... 9165 glasses.41 In both cases, the presence of long order positional the structural glass because in the fully disordered phase the periodicity enables one to understand the freezing phenom- molecular orientations are strongly correlated with the posi- enon on the basis of some minimal model, as that of hard tions of the molecular centers of mass. This result agrees 42 needles on a fcc lattice, explored by Renner et al. It con- with the excess in specific heat C P observed in previous tains the basic ingredients of the ‘‘glass transition.’’ Indeed, work, in which the observed C P below the glass transition in it exhibits a kinetic glass transition when lϭL/a, the needle the OG is higher than in the glass.8 length to lattice constant ratio, exceeds about 2.7, and is Finally, one of the most interesting aspects in the com- simple enough to a allow numerical and analytical study. For parison of the dynamical behavior of the glassy and disor- a bcc lattice, such as that of the RP and OG crystal phases, a dered crystal states of this material would be to compare the ‘‘computer glass transition’’ is observed for a ratio of temperature and frequency dependence of their elastic con- 13 3.3–3.4. stants, which are expected to show marked anomalies with The origin of the ␤ relaxation is ascribed to thermally respect to those shown by the ordered crystal. activated local molecular librations. Our analysis shows that such motions are the result of a strong coupling between molecular internal and collective modes, and some points of ACKNOWLEDGMENTS contact are established with work performed on KBr:KCN ͑Ref. 36͒ mixed crystals where the flipping motion of the CN This work was performed in part under DGICYT ͑Spain͒ dipoles is postulated as the origin of this relaxational peak. Grant Nos. PB95-0075-C03-01 and PB94-0468. We thank R. The intensity of the ␤ relaxation is higher in the OG than in Ferna´ndez-Perea for helpful discussions.

1 H. Vogel, Z. Phys. 22, 645 ͑1961͒; G.S. Fulcher, J. Am. Chem. Phys. Rev. B 57, R13 977 ͑1998͒. Soc. 8, 339 ͑1925͒; G. Tamman and W. Hesse, Z. Anorg. Allg. 15 D.L. Leslie-Pelecky and N.O. Birge, Phys. Rev. B 50,13250 Chem. 156, 245 ͑1926͒. ͑1994͒; Phys. Rev. Lett. 72, 1232 ͑1994͒. 2 L. Wu, Phys. Rev. B 43, 9906 ͑1991͒. 16 M.A. Gonza´lez, E. Enciso, F.J. Bermejo, and M. Be´e, J. Chem. 3 A. Kudlik, C. Tschirwitz, S. Benkhof, T. Blochowicz, and E. Phys. ͑to be publsihed͒. Ro¨ssler, Europhys. Lett. 40, 649 ͑1997͒. 17 W.L. Jorgensen, J. Phys. Chem. 90, 1276 ͑1996͒. 4 Dynamics of Disordered Materials II, edited by A.J. Dianoux, W. 18 L. Saiz, J.A. Padro´, and E. Gua`rdia, J. Phys. Chem. B 101,78 Petry, and D. Richter ͑North Holland, Amsterdam, 1993͒. ͑1997͒; also, L. Saiz, Ph.D. dissertation, University of Barce- 5 R. Schilling and T. Scheidsteger, Phys. Rev. E 56, 2932 ͑1997͒; lona, 1998. S. Ka¨mmaner, W. Kob, and R. Schilling ͑unpublished͒. 19 D.W. Davidson and R.H. Cole, J. Chem. Phys. 18, 1417 ͑1951͒. 6 M. Evans, G.J. Evans, W.T. Coffey, and P. Grigolini, Molecular 20 P.K. Dixon, L. Wu, S.R. Nagel, B.D. Williams, and J.P. Carini, Dynamics and Theory of Broadband Spectroscopy ͑Wiley, New Phys. Rev. Lett. 65, 1108 ͑1990͒. York, 1982͒, Chap. 2. 21 P. Lunkenheimer, A. Pimenov, M. Dressel, Y.G. Goncharov, R. 7 O. Haida, H. Suga, and S. Seki, J. Chem. Thermodyn. 9, 1133 Bo¨hmer, and A. Loidl, Phys. Rev. Lett. 77, 318 ͑1996͒. ͑1977͒; M. Hugo, K. Schlu¨ter, and A. Wu¨rflinger, Z. Phys. 22 R.H. Cole, in Molecular Liquids—Dynamics and Interactions, ed- Chem. ͑Munich͒ 175, 235 ͑1992͒. ited by A.J. Barnes, W.J. Orville-Thomas, J. Yarwood, Vol. 135 8 F.J. Bermejo, A. Criado, R. Fayos, R. Ferna´ndez-Perea, H.E. of NATO Advanced Study Institute, Series C: Mathematical and Fischer, E. Suard, A. Guelylah, and J. Zu´˜niga, Phys. Rev. B 56, Physical Sciences ͑Reidel, Dordrecht, 1983͒,p.91. 11 536 ͑1997͒; F.J. Bermejo, H.E. Fischer, M.A. Ramos, A. de 23 M.A. Miller, Ph.D. thesis, Michigan State University, 1998. Andre´s, J. Dawidowski and R. Fayos, in Complex Behaviour of 24 S.K. Garg and C.P. Smyth, J. Phys. Chem. 69, 1294 ͑1965͒. Glassy Systems, edited by M. Rubı´ and C. Pe´rez-Vicente, 25 J. Barthel, K. Bachhuber, R. Buchner, and H. Hetzenauer, Chem. Springer Lecture Notes in Physics, Vol. 492 ͑Springer, Berlin, Phys. Lett. 165, 369 ͑1990͒. For recent results see J.T. Kindt and 1997͒,p.44. C.A. Schmuttenmaer, J. Phys. Chem. 100,10373͑1996͒. 9R. Fayos, F.J. Bermejo, J. Dawidowski, H.E. Fischer, and M.A. 26 F.J. Bermejo, F. Batalla´n, W.S. Howells, C.J. Carlile, E. Enciso, Gonza´lez, Phys. Rev. Lett. 77, 3823 ͑1996͒. M. Garc-Herna´ndez, M. Alvarez, and J. Alonso, J. Phys.: Con- 10 C. Talo´n, M.A. Ramos, S. Vieira, G.J. Cuello, F.J. Bermejo, A. dens. Matter 2, 5005 ͑1990͒. Criado, M.L. Senent, S.M. Bennington, H.E. Fischer, and H. 27 M.P. Janelli, S. Magazu´, P. Migliardo, F. Alliotta, and E. Tetta- Schober, Phys. Rev. B 58, 745 ͑1998͒. manti, J. Phys.: Condens. Matter 8, 8157 ͑1996͒. 11 H.E. Fischer, F.J. Bermejo, G.J. Cuello, M.T. Ferna´ndez-Dı´az, J. 28 D. Bertolini, M. Cassettari, M. Ferrario, P. Grigolini, and G. Sal- Dawidowski, M.A. Gonza´lez, H. Schober, and M. Jime´nez- vetti, Adv. Chem. Phys. 62, 277 ͑1985͒. Ruiz, Phys. Rev. Lett. 82, 1193 ͑1999͒. 29 L. Wu, R.M. Ernst, Y.H. Jeong, S.R. Nagel, and S. Susman, Phys. 12 F.J. Bermejo, G.J. Cuello, J. Dawidowski, H.E. Fischer, H. Rev. B 37,10444͑1988͒. Schober, M.A. Gonza´lez, and S.M. Bennington, Physica B 241- 30 N.O. Birge, Y.H. Jeong, S.R. Nagel, S. Bhattacharya, and S. Sus- 243, 883 ͑1998͒. man, Phys. Rev. B 30, 2306 ͑1984͒. 13 M.A. Ramos, S. Vieira, F.J. Bermejo, J. Dawidowski, H.E. 31 M. Jime´nez-Ruiz, G.J. Cuello, F.J. Bermejo, F.R. Trouw, A. Fischer, H. Schober, M.A. Gonza´lez, C.K. Loong, and D.L. Criado, and H.L. Lo¨wen ͑unpublished͒. Price, Phys. Rev. Lett. 78,82͑1997͒. 32 O. Conde and J. Teixeira, J. Phys. ͑Paris͒ 44, 525 ͑1982͒. 14 M. Miller, M.A. Jime´nez-Ruiz, F.J. Bermejo, and N.O. Birge, 33 R. Ludwig and M.D. Zeidler, Mol. Phys. 82, 313 ͑1994͒. 9166 M. JIME´ NEZ-RUIZ et al. PRB 59

34 F. Stickel, E.W. Fischer, and R. Richert, J. Chem. Phys. 104, and W.A. Phillips, Phys. Rev. B 34, 5665 ͑1986͒. 2043 ͑1996͒. 39 C.A. Angell, in Hydrogen-Bonded Liquids ͑Kluwer Academic, 35 R.M. Ernst, L. Wu, S.R. Nagel, and S. Susman, Phys. Rev. B 38, Dordrecht, 1991͒,p.59. 6246 ͑1988͒. 40 T. Eguchi, G. Soda, and H. Chihara, Mol. Phys. 40, 681 36 E.R. Grannan, M. Randeria, and J.P. Sethna, Phys. Rev. B 41, ͑1980͒. 7784 ͑1990͒; 41, 7799 ͑1990͒. 41 A. Loidl, in Disorder Effects on Relaxation Processes, edited by 37 A. Criado, F.J. Bermejo, A. de Andre´s, and J.L. Martı´nez, Mol. R. Richert and A. Blumen ͑Springer, New York, 1994͒. Phys. 82, 787 ͑1994͒. 42 C. Renner, H. Lo¨wen, and J.L. Barrat, Phys. Rev. E 52, 5091 38 U. Buchenau, M. Prager, N. Nucker, A.J. Dianoux, N. Ahmad, ͑1995͒.