Marceau
Hénot
*a,
Pierre-Michel
Déjardin
b and
François
Ladieu
a
aSPEC, CEA, CNRS, Université Paris-Saclay, CEA Saclay Bat 772, 91191 Gif-sur-Yvette Cedex, France. E-mail: marceau.henot@cea.fr
bLaboratoire de Modélisation Pluridisciplinaire et Simulations, Université de Perpignan Via Domitia, 52 avenue Paul Alduy, F-66860 Perpignan, France
First published on 11th October 2023
The orientational dynamics of supercooled glycerol is probed using molecular dynamics simulations for temperatures ranging from 323 K to 253 K, through correlation functions of first and second ranks of Legendre polynomials, pertaining respectively to dielectric spectroscopy (DS) and depolarized dynamic light scattering (DDLS). The self, cross, and total correlation functions are compared with relevant experimental data. The computations reveal the low sensitivity of DDLS to cross-correlations, in agreement with what is found in experimental work, and strengthen the idea of directly comparing DS and DDLS data to evaluate the effect of cross-correlations in polar liquids. The analysis of the net static cross-correlations and their spatial decomposition shows that, although cross-correlations extend over nanometric distances, their net magnitude originates, in the case of glycerol, from the first shell of neighbouring molecules. Accessing the angular dependence of the static correlation allows us to get a microscopic understanding of why the rank-1 correlation function is more sensitive to cross-correlation than its rank-2 counterpart.
The complex dielectric permittivity obtained from DS measurements can be linked to the time-dependent equilibrium field free total dipole moment correlation function C1(t) through a Fourier–Laplace transform13–15
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
Recently, results from DS were compared to other techniques less sensitive to cross-correlations. The fluorescence response of a local probe diluted in a mono-alcohol was shown to be insensitive to the Debye relaxation of the liquid, allowing it to disentangle it from the other relaxation processes.18 Another technique that has been proven useful in that regard is depolarized dynamic light scattering (DDLS) which probes molecular orientations through the anisotropy of the polarizability. The relevant correlation function is given by eqn (2) with = 2. It follows that the technique does not distinguish between parallel and antiparallel alignments. There is strong experimental evidence that DDLS is insensitive to cross-correlations. For example, Gabriel et al.19 showed that in mono-alcohols DDLS displays an α peak but no Debye peak. In addition, in a non-associating liquid, Pabst et al.20 showed that progressively diluting the system in a non-polar solvent leads the DS spectra to look more and more alike the DDLS spectra. All of this illustrates the importance of cross-correlation effects in DS which can significantly broaden the α peak. Moreover, while the shape of the α peak in DS spectra is system dependent, it was shown in DDLS to follow a generic line shape of slope − 1/2 on the high-frequency flank.21 There is still debate, however, on whether this generic response reflects the true structural relaxation better than the dielectric one.22
When dealing with physical processes taking place at the nanometric scale, molecular dynamics (MD) simulation is an attractive method that can give access to microscopic observables that are otherwise hard, or impossible, to obtain experimentally. This method is however limited to high temperatures or simplified systems, due to its computational cost. To study the generic behavior of liquid glass-formers, model systems can be thought of being made of polydisperse beads interacting through a Lennard-Jones potential. This helped give information on the spatio-temporal nature of relaxations.7,23 Another approach, more suitable for direct comparison with experiments, is to rely on a more precise modelization of specific molecules, taking into account their dipolar nature and electrostatic interactions. This gives access to their dielectric response.24–28 Recently, MD simulations on a model dipolar system showed that, while the orientational = 1 correlation function of weakly polar systems is dominated by the self response, strongly polar liquids are much affected by cross-correlations.29
The wide variety of organic liquids available has led to the choice of some systems, considered as models or representatives. Glycerol, by its apparent simplicity and its low tendency to crystallize, has long been the subject of extensive studies, by various techniques including dielectric spectroscopy,2,11,30–32 neutron spectroscopy,33 nuclear magnetic resonance (NMR),34 DDLS35,36 and MD simulations.37–44 Its dynamics is, however, not particularly simple. As a tri-alcool, it is subject to H-bonds but does not display a Debye peak that would result from linear supramolecular chains. Shear mechanical spectroscopy has shown the existence of a low-frequency mode that is believed to result from the hydrogen-bonded network formed between molecules.45
In this article, we report a MD study of the orientational dynamics of glycerol, on a large temperature range (from 253 to 323 K) reaching the moderately supercooled regime, simulated from a model already widely used in the literature37–40,44 over durations of up to 7 s. We first compute the self response of the dipolar moment for ranks = 1 and 2 from which we deduce the loss function
for frequencies down to 200 kHz. We then analyze the cross-correlation and we exploit the possibility offered by MD to decompose this part of the response as a function of the relative distance and orientation of the dipoles. We compute the total loss function for both ranks as well as the part resulting from cross-correlations alone which allowed us to verify that cross-correlations play a major role in the
= 1 response while being almost negligible for
= 2. We compare these data to experimental DS and DDLS spectra and obtain similar temperature dependence for the relaxation time and slope of the high-frequency flank of the α peak. We discuss how the differences in the spectra associated with different ranks can be related to the underlying molecular relaxation mechanisms. Moreover, we show that, for glycerol, the net cross-correlation originates only from the first shell of neighouring molecules. This is the case for both
= 1 and 2 although their different sensitivity to orientational correlations leads, at the end, to significant differences in the importance of contributions coming from cross-correlations.
Cself![]() ![]() ![]() | (5) |
The mean self relaxation time is obtained from and is shown as a function of 1/T in blue in Fig. 6a. The relaxation times are shorter for
= 2 (empty markers) than for
= 1 (solid markers) and they both display a super-activated behaviour.
The self loss function was obtained, following eqn (1), by applying the fluctuation-dissipation theorem:50 where TF is the Fourier transform, computed using the fftlog algorithm adapted to log spaced data.51 The fact that the correlation function was averaged on long times (≈100τα) leads to a fairly low amount of noise on the spectra, shown in Fig. 1b. They were all rescaled by superimposing their microscopic peaks at 1013 Hz. The frequency at which the maximum of the α peak is reached was found to correspond (within the uncertainty) to 1/(2πτself
). On the low-frequency side, the spectra follow a power law with slope 1, as expected. On the high-frequency flank of each spectrum, there is a power law regime on one to two decades in frequency with a slope −βself
, interrupted by the fast process.3 The corresponding values of βself
are shown in blue in Fig. 6b. For
= 1 (solid markers), slope increases with temperature (ranging from 0.36 at 253 K to 0.46 at 323 K) while for
= 2 (empty markers), it is temperature independent and systematically smaller (≈0.27). These low β values are associated with the non-exponential nature of the relaxation process.
![]() | (6) |
We decompose the cross-correlation function Ccross(t) into contributions per unit distance denoted c
(r,t) depending on the distance r = ‖
j −
i‖ between the reference molecule i and all molecules within [r,r + dr]. This quantity, computed for slices of 0.5 Å and averaged over i and t0, is plotted in Fig. 2a for the static case (t = 0) for
= 1 and 2. The dipole density n(r) is plotted in black in Fig. 2b alongside with the parabola in red that would be obtained for a homogeneous system of the same average density. The difference between these two curves shows a series of maxima that corresponds to the first, second and third neighbour peaks. For
= 1, it appears that the cross-correlation contribution c1(r) (in blue) is maximum for the first neighbours and reaches zero before the second layers of neighbours. Fig. 2c represents, as a function of r, the total contribution
integrated within a sphere of radius r. We see that the cross-correlation reaches a plateau at r ≈ 7 Å while the cross-correlation coming from PCBs starts to be perceptible for r > 20 Å (see ESI†). It is also interesting to study the mean level of static cross-correlation Γ(r) = c(r)/n(r), plotted in Fig. 2d that shows that the cross-correlation per dipole is positive, is a strictly decreasing function of the distance and is only of the order of 5–15% in average for the nearest neighbours. From these data, we can deduce the Kirkwood factor
at 273 K, with rlim = 7.5 Å. This is smaller than the 2.6 ± 0.2 value reported in the literature and deduced from static permittivity measurements36 but it is compatible with previous numerical results on glycerol for T > 250 K.43 It is also interesting to note that our value of μ2gK (which is the quantity accessible from the experiments, see eqn (1)) matches exactly the experimental value and displays the same temperature dependence (cf. Fig. S4 in ESI†). In other words, the simulation gives the expected value of ε(0) in the whole temperature range. However, as the dipolar moment of glycerol has been measured with a reasonable accuracy in a non-polar solvent,48 it is likely that the MD simulation underestimate the value of gK. For
= 2, the cross-correlation (in orange) appears to be also due to the first layer of neighbours but is much less intense than for
= 1 and its integrated value saturates for a slightly higher radius. With rlim = 9.5 Å, the quantity analogous to the Kirkwood factor for
= 2 would be only 1.11 ± 0.01.
The dipolar interaction is anisotropic in nature and it makes sense, rather than averaging the cross-correlation over all dipoles situated at a given distance, to distinguish the contribution as a function of the relative orientation. Spherical coordinates (r,θ,ϕ) can be defined with respect to the reference dipole i where θ is the angle between i and
=
j −
i. By symmetry, the contribution should not depend on the azimuthal angle ϕ. The cross-correlation contribution c
(r,θ) per unit distance and solid angle is shown in the static case in Fig. 3a (
= 1) and b (
= 2) as a function of r and θ. For
= 1, similarly to a recent observation in water,28 the spatial distribution of cross-correlation appears strikingly different than the θ averaged curves shown previously. The correlation is positive in an angular sector situated above and below the dipole of reference |cos
θ| < cos
θlim and mostly negative on its sides. It is null on lines of constant |cos
θ| = cos
θlim shown in grey with θlim = 52°. The cross-correlation contributions summed over these two angular sectors are shown in Fig. 3c (top). These positive and negative contributions extend way over what is visible when considering their sum and, while decreasing, are far from negligible at r = 16 Å. However, these contributions cancel each other for r > 7.5 Å. This angular dependence of the cross-correlation is what is expected when considering the energy interaction between two electrostatic dipoles as a function of their relative orientation53 and drawings of the most favorable dipoles orientations are shown in Fig. 3a. The net contribution of cross-correlation comes from the first shell of neighbouring molecules. Close molecules situated above and below are strongly positively correlated (even more above than below) with an alignment rate reaching 50%. This correlation is favored by the dipole–dipole interaction although the top-bottom asymmetry illustrates that at such a close distance, it is not the only interaction playing a role. On the sides, molecules belonging to the first shell but further than 5.2 Å (corresponding to the first neighbor peak) are on average anti-aligned (again as favored by the dipole–dipole interaction), although not enough to compensate for the positive correlation. Finally, side molecules closer than 5.2 Å are positively aligned. This means that there exists other effects (that may be related to constraints on molecule conformation or to the presence of H-bonds) able to compensate for the a priori unfavorable situation of having barycenters of the same sign charges facing each other on average.
![]() | ||
Fig. 3 Relative orientation and distance dependence of the static dipole cross-correlation functions at T = 273 K. (a) Contribution c per unit distance and solid angle to the cross-correlation (![]() ![]() ![]() ![]() ![]() ![]() ![]() |
For = 2 also, a complex spatial dependence of the cross-correlation is visible in Fig. 3b. As this quantity is not sensitive to the correlation sign, it shows a very different behavior than
= 1. Some oscillations, that correlate well with the density inhomogeneity, are visible. Those also appear for
= 1 but dominate here. After the first layer of neighbours, the contribution from the two angular sectors are in anti-phase and cancel each other on average. Similarly to the
= 1 case, the net contribution comes from the first neighbour shell but it is interesting to observe, that its different sensitivity to orientational correlations makes this quantity significantly less sensitive to cross-correlations in the case of glycerol.
In the same way as for the self part, a loss function χ′′(f) can be computed from the cross part and for the total correlation. The results are plotted for 273 K in Fig. 4c ( = 1) and d (
= 2). It is well visible that the amplitude of the cross-correlation is of the same order as the self part for
= 1 while it is much smaller for
= 2. On each spectrum, the high frequency flank of the α peak has a power law slope −β such that βself
< βcross
< 1. The slope of the total loss function βtot
results from both previous slopes as well as the strength of the cross-correlation and takes an intermediate value. The total spectra for all temperatures are shown in Fig. 5a and the values of the slopes β are shown in Fig. 6b as a function of T: β1tot is increasing with T while βtot2 appears constant close to 0.3. This is already visible on the spectra of Fig. 5a but it is even clearer in Fig. 5b where the spectra are shown as a function of a dimensionless frequency ωτself1. The collapse of the high-frequency side is much better for the total spectra with
= 2 (although the α peak slightly broadens upon cooling) than for
= 1 (self or total).
![]() | ||
Fig. 5 Orientational loss function for all temperatures (same color scale than in Fig. 1) studied as a function of f (a) and ωτself1 (b). The total loss function is shown for ![]() ![]() ![]() ![]() ![]() |
![]() | ||
Fig. 6 (a) Mean relaxation time obtained by integrating the self (blue squares), cross (green triangles) and total (red circles) dipole correlation function for ![]() ![]() ![]() ![]() ![]() ![]() |
Contrary to DS, which probes the reorientational dynamics of the molecules by following the permanent dipolar moment, DDLS does it by following the anisotropy of the polarisability tensor. Glycerol molecules are believed to rotate as a rigid entity44 and we can reasonably assume that both techniques probe the same dynamics (although giving access to different ranks ). It is also worth noting that results from DDLS experiments can also be affected by a scattering mechanism called dipole-induced-dipole related to the fluctuation of the internal field. Cummins et al.56 have argued that this effect may be neglected in the experimental spectra of supercooled liquids. Here we follow these authors. The spectra
, from the simulation, should therefore be comparable with the Fourier transform of the DDLS correlation function reported by Gabriel et al.36 The authors concentrated mainly on frequencies under a few MHz and thus on temperatures smaller than 260 K (although they also performed a measurement at 323 K). Here also, the MD spectra show a good qualitative agreement with experiments. The DDLS mean relaxation time is shorter than in DS by an amount that appears similar in the experiments and in MD. The inequality βtot2 < βtot1 is also verified for both. In the DDLS experiments (see grey empty circles in Fig. 6d), given the uncertainty, the slope βtot2 does not appear to depend much on the temperature and it is also the case in the simulation (empty red circles). However, here again, the slopes are underestimated in the simulation.
The differences between the simulation and the real system must be weighed against the relative simplicity of the modeling. Indeed, force field parameters, that control intra and inter-molecular interactions were not adjusted specifically on glycerol but were designed to be applicable to the widest range possible of organic compounds. The only parameter that was adjusted specifically to glycerol was μ. Moreover, the partial charges in the molecule are considered as fixed point charges at the center of each atom, the electronic polarizability is not taken into account and H-bonds are mimicked only by electrostatic interactions of these fixed charges which can limit their strength.57 Nonetheless, it is already impressive that a classical model can reproduce well some aspects of the real system. With these limitations in mind, MD simulations can be used, as demonstrated recently by Becher et al.,44 to obtain information on microscopic observables or on the relative effect of external parameters such as the temperature.
For = 2, and in contrast to the
= 1 case, the effect of cross-correlation is very weak and the total loss function can reasonably be assimilated to the self loss function alone (see Fig. 4d). This is in complete agreement with the observations of Gabriel et al.36 and previous work by the same team19,20 on the ability for DDLS to give access to the self orientational dynamics. This is also in agreement with the recent MD observations on a model system of Koperwas et al.29 who compared the self and total correlation functions for
= 2 at long times (t ≥ τα).
The ratio of the self relaxation times τself1/τself2, shown in Fig. 6c, ranges from 1.4 to 2.0 and is increasing with temperature. This quantity is affected by the details of the molecular relaxation mechanism with two limiting cases leading to values of 3 for isotropic rotational diffusion and 1 for discrete jumps of random amplitude.5 However, a given value of this ratio cannot be associated with a typical orientational jump amplitude. For example, in a mean field model, tuning the interaction parameter allows to change continuously this ratio.58 In the present case, it is more likely that its value reflects a dynamics governed by rare relaxation events, that are less averaged for = 2 than for
= 1.59 In this case, its decrease upon cooling could be associated with rarer and rarer events. The slope β < 1 can also be linked with a dynamic governed by relatively rare events, spread on long time scales, that does not average enough to produce an exponential relaxation (β = 1).59 In this context, as C2 cancels out for smaller angles than C1, it is less averaged over relaxation events leading to β2 < β1, consistently with our observations.
In the DDLS works mentioned above,19,20,36 the authors suggested to use these measurements as a proxy to approach the DS self response, which is not accessible experimentally. This means admitting that the response is reasonably independent of the rank but this also requires a way to normalize the DDLS data so that they can be quantitatively compared to DS spectra. With a well-chosen, temperature independent, scaling factor, it is possible a get a perfect overlap of the DS and DDLS spectra on the excess wing flank at low temperature (T < 200 K). From this, it was shown that the DS signal could be well fitted by the sum of the DDLS signal and a cross-correlation term (described by a stretched exponential of fixed stretching parameter), with the relative weight of these two terms perfectly matching the Kirkwood correlation factor at all temperatures.36 This demonstrates the usefulness and relevance of the direct comparison between DDLS and DS spectra.
In our numerical work, we were not able to reach the range of low temperatures in which the excess wing appears. This prevented us from using the method described above to determine from scratch a scaling factor between the = 1 and
= 2 spectra. We instead had to rely on the experimental determination by Gabriel et al.36 Indeed, while the DS and DDLS spectra collapse on the excess wing for T < 200 K, they cross at a finite frequency fcross for higher temperatures. We determined a (T independent) scaling factor by making sure that the
= 1 and
= 2 total spectra cross at the same fcross/fα than the experiments for T = 323 K and T = 263 K (see vertical dashed lines in Fig. 5c). The result is shown in Fig. 5c and it appears that, for all temperatures, the α peak of the total
= 2 and the self
= 1 spectra match fairly well. The difference of slopes β is small enough so that the discrepancy remains low (<30%) under the frequency at which the fast process starts to be perceptible. This allows us to understand the success of the experimental approach consisting of a direct comparison between DS and DDLS data.19,20,36
Moreover, we took advantage of the possibility given by MD simulations to access molecular observables to discuss the microscopic origin of the cross-correlation observed in DS. We found out that the net cross-correlation originates from the first shell of neighbouring molecules that tend to positively align independently of their orientation. Investigating in more detail the molecular origin of the cross-correlations, and their link with molecular conformation and H-bonds will be the subject of future work.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp04578a |
This journal is © the Owner Societies 2023 |