Structural excitations that enable interbasin (IB) barrier crossings on a potential energy landscape are thought to play a facilitating role in the relaxation of liquids. Here, we show that the population of these excitations exhibits the same density scaling observed for α relaxation in propylene carbonate, even though they are heavily influenced by intramolecular modes. We also find that IB crossing modes exhibit a Grüneisen parameter (γG) that is approximately equivalent to the density scaling parameter γTS. These observations suggest that the well-documented relationship between γG and γTS may be a direct result of the pressure dependence of the frequency of unstable (relaxation) modes associated with IB motion.

For many decades, there has been evidence that discrete molecular rearrangements, so-called “hops,” contribute significantly to relaxation in viscous liquids.1,2 At reduced temperatures, it is thought that molecules become trapped in a cage formed by their neighbors so that relaxation requires collective rearrangements of neighborhood particles3 that are thought to be facilitated by structural excitations.4,5

The case for the importance of excitations in influencing liquid dynamics is supported through their ability to account for super-Arrhenius relaxation,6 the breakdown of the Stokes–Einstein relation,7 the appearance of dynamic heterogeneity,8,9 and the behavior of specific heat at the calorimetric glass transition.10 If excitations are a causal factor in liquid dynamics, one would also expect a connection to the ubiquitous scale-invariance observed in liquid dynamics,11–25 but this connection has not been explored.

As pointed out by Roland et al.,15 expectation of scale-invariance arises when the potential that controls system behavior is well-approximated by a pairwise additive inverse power law (IPL).16,19 Dyre and co-workers17,19,20,26 have expanded these ideas, demonstrating pressure and temperature invariance of dynamic, structural, and thermodynamic properties of model and real liquids when length and energy are properly scaled.20 These properties are invariant along “isomorphs” of constant ργ/T, where γ = n/3 and n is the exponent of the IPL pair potential. In systems where only the IPL potential is important, the isomorph scaling exponent γ is related to instantaneous fluctuations in energy and the first pressure virial.17–20 

When the dominating interactions in a liquid do not conform to IPL potentials, the entire gamut of isomorph behavior can be lost. This is observed in systems with multiple interaction length scales, dominant electrostatic forces, or those with distinct directionality such as hydrogen bonding.19 For example, molecular dynamics (MD) simulations of the SPC water model have shown that there is essentially no virial–energy correlation and that isomorphic state points do not exist.20,26 Similar results are found in experimental studies of heterogeneous hydrogen bonding (e.g., polyalcohol) systems.27,28

When only minor motional modes of a system (such as intramolecular or long-range intermolecular degrees of freedom) do not conform to IPL potentials, some aspects of isomorph behavior are lost, but many other aspects may be retained. These systems exhibit what is sometimes referred to as “pseudoisomorphs,”29 and properties associated with long timescales such as α relaxation may exhibit density scaling in that they are invariant at constant values of ργTS/T where γTS is an empirical density scaling exponent. Density scaling has been experimentally verified for a wide variety of systems, including van der Waals, and other nonassociating liquids, polymers, ionic liquids, and some associating and hydrogen bonding liquids.12–15,21–25 Several theoretical and experimental works19,30,31 have demonstrated a connection between γTS and the volume dependence of vibrational frequency for spatially extended collective modes of the liquid through the Grüneisen parameter.

In these pseudoisomorph systems, pressure–energy correlations are broken since the energy can partition to the non-IPL modes without strong coupling to changes in volume or pressure. The pressure of these fluids does not depend strongly on forces arising from angle vibrations and torsional rotations of at least some intramolecular modes.32 Olsen et al.33 found in model diatomic and oligomeric systems with harmonic intramolecular bonds that vibrational eigenmodes representing center-of-mass (COM) motion scale properly and could be used to find the scaling exponent but that the highest frequency eigenmodes did not scale. The number of non-scaling eigenmodes was equal to the number of intramolecular bonds, so all of the intramolecular modes for the diatomic model and at least 1/3 of the highest frequency intramolecular modes for the oligomeric system did not scale.

Thus, while low-frequency vibrational modes and slow dynamic processes follow density scaling, there appears to be some cutoff in timescale or length scale below which dynamic processes in the liquid no longer scale. Tölle et al. found that relaxation processes in ortho-terphenyl on timescales of 1 ns and longer follow density scaling.11 Hansen et al.23 and Ribeiro et al.34 demonstrated invariance down to 100 and 10 ps, respectively. Puosi et al.24 found density scaling at timescales of the “fast β” (βfast) relaxation, roughly 1 ps. In that work, Puosi et al. quantified the mean-squared displacement u2 of particles in course-grained polymer systems, finding excellent correspondence between γTS obtained for u2 and α relaxation, but only when a small term quadratic in (TργTS) was added to fit the latter.

We have recently shown that the Angstrom-length scale and picosecond timescale motion that has traditionally been classified as βfast relaxation and measured through u2 contains signatures of two separate processes.35,36 The faster of the two corresponds to elastic deformations of equilibrium local atomic structure, corresponding to so-called “inherent state” (IS) dynamics on a potential energy landscape (PEL).37 The slower of the two processes represents36,38 rearrangements of local atomic structure corresponding to interbasin (IB) barrier crossings on a PEL that lead to a new IS and the instantaneous population of IB crossing events is equivalent to the excitation population.39 We have further found evidence that intramolecular modes may figure prominently in excitations,38 leading us to question whether excitations and all the components of βfast should be expected to have the same γTS as the α process or whether involvement of intramolecular modes in these short-timescale, local motions breaks this scaling.

To investigate to what extent excitations obey density scaling, we have conducted molecular dynamics simulations of propylene carbonate (PC) over a range of pressures and temperatures. PC was previously shown to follow density scaling31,40 and provides an interesting case study for this investigation due to its strong electrostatic interactions (ɛ ∼ 65), molecular anisotropy, and the considerable flexibility of its ring, all of which affect its liquid state properties.41–43 

Our simulations utilize an ab initio, polarizable force field capable of predicting properties that are in excellent agreement with experimental measurements near STP. By searching its phase diagram, we identify two P–T curves in PC that exhibit largely invariant structural and dynamical properties. We then quantify excitation populations at these phase points, utilizing a previously benchmarked approach.35,36 We conclude that excitations fundamentally involve the atomic-scale intramolecular structure of PC and are not well characterized by center-of-mass motion alone. Furthermore, we demonstrate that excitation populations in this liquid obey density scaling as they are constant across the P–T curves exhibiting invariant structure and dynamics. We also show that the motions associated with excitation are the fastest dynamic processes that can be considered to follow density scaling.

Molecular dynamics (MD) simulations were conducted for propylene carbonate (PC) utilizing the OpenMM simulation package.44 We utilize the atomistic, polarizable symmetry adapted perturbation theory-force field (SAPT-FF), which is an ab initio force field developed on the basis of symmetry adapted perturbation theory (SAPT) calculations.45 We supplement these existing force field parameters with atomic charges and dihedral potentials that are specifically parameterized for PC utilizing additional density functional theory (DFT) calculations as described in the supplementary material. Polarization is incorporated via a Drude oscillator model utilizing a dual-Langevin thermostat as implemented in OpenMM, with a 1 ps−1 friction coefficient used for both thermostats.46 The particle–mesh Ewald method (PME)47 is used for electrostatics and van der Waals (VDWs) interactions are computed up to a 1.4 nm cutoff. Each simulation consisted of 400 propylene carbonate molecules initially constructed using Packmol.48 Equilibration consisted of 1 ns NPT simulations conducted with a Monte Carlo Barostat, followed by subsequent 5 ns NVT simulations at the equilibrated density. All simulations utilized a 1 fs time step. Because OpenMM utilizes a Monte Carlo Barostat, the virial is not directly accessible. We have thus utilized Gromacs to conduct an additional MD simulation in the NVT ensemble at 300 K to compute energy–virial correlation, shown in the supplementary material, Fig. S1.49 The simulation details for the Gromacs simulation are largely identical to the OpenMM simulations, except that in the former Drude oscillator positions are treated strictly adiabatically (minimized at every timestep). See the supplementary material for more details about the simulation settings.

Finding isomorph families for Lennard-Jones (LJ) fluids and other simple systems for which pairwise IPL potentials dominate is accomplished by analyzing the virial–energy correlation function.20 For IPL potentials, the virial and potential energy are perfectly correlated, so that the normalized correlation function is unity.26 This is not possible for systems with significant intramolecular degrees of freedom. For organic liquids with molecular constituents of increasing size and complexity, it is expected that isomorphic behavior will not be found based on analysis of virial–energy correlation.23,24,29 This is because intramolecular rather than intermolecular interactions dominate the potential energy of the system, with the former growing as 3N − 6 as the size “N” of the molecules gets larger. The intramolecular interactions do not follow inverse power laws; additionally, fluctuations in the intramolecular energy are not expected to correlate with virial fluctuations. The latter is inferred from the fact that the pressure of fluids does not depend on forces arising from angle vibrations and torsional rotations.32 For molecules of large size and complexity, near zero correlation of the virial and energy is expected due to the dominance of the intramolecular interactions. Experimentally, Hansen et al. have shown that density scaling breaks down for dynamic signatures indicative of intramolecular motion.23 

We have computed the virial–energy correlation from an MD simulation of PC at 300 K, 1 atm; the results are shown in the supplementary material, Fig. S1. We plot the data as normalized fluctuations of each quantity, in analogy to previous work.24,26,29 As expected from the dominance of the intramolecular energetics of PC and similar to results for other flexible molecules,24,29 we find essentially no correlation between the virial and energy fluctuations.

We note that if it were possible to explicitly separate the intramolecular and intermolecular energy contributions, then one could analyze virial correlations with just the latter. However, such a separation is generally not tractable within a molecular dynamics code. The reason being that electrostatic interactions, which contribute to intramolecular energy for 1–4 and longer atom separations, are computed with an Ewald/PME sum and not easily separated on a pairwise interaction basis. Furthermore, the electrostatic energy is formally many-body for explicitly polarizable force fields, as utilized in this work.

Because of these considerations, in this work we utilize a brute-force approach to search for isomorph curves. We searched for two isomorph curves in the P–T phase diagram, shown in Fig. 1. We initially ran a 1 ns NPT simulation of 400 PC molecules at temperature/pressure values of 300 K/1 atm; simulation settings are the same as noted above. We then performed a 20 ns NVT simulation. We calculated the diffusion coefficient (DT) and the inverse rotational correlation time (1/τrot) for this state point. DT was calculated using the Einstein relation given by

(1)

and calculated τrot by integrating the rotational correlation function as follows:

(2)

with tmax equal to 1 ns and zCO equal to the carbonyl bond vector. As isomorph properties are only apparent in reduced units, the calculated values of DT and τrot were converted to reduced unit values D̃T and τ̃rot.20 We performed a series of 1 ns NPT/20 ns NVT simulations at temperatures of 325, 350, 375, and 400 K and a range of pressures. D̃T and 1/τ̃rot were computed at each state point. A point was classified as an isomorph point if the calculated values agreed with D̃T and 1/τ̃rot at 300 K/1 atm (within statistical uncertainty).

FIG. 1.

Demonstration of ργTS/T=constant for the two isomorph curves, with γTS = 3.7 as determined from the work of Pawlus et al.40 plotted as dashed lines. The solid line denotes the liquid/glass boundary computed from state points taken from the work of Bonetti and Dubois.52 

FIG. 1.

Demonstration of ργTS/T=constant for the two isomorph curves, with γTS = 3.7 as determined from the work of Pawlus et al.40 plotted as dashed lines. The solid line denotes the liquid/glass boundary computed from state points taken from the work of Bonetti and Dubois.52 

Close modal

We followed a very similar procedure for the second isomorph. Here, we initially ran a simulation at a temperature/pressure value of 250 K/1 atm. As the dynamics are noticeably slower at this state point compared to 300 K/1 atm, we ran 50 ns NVT simulations after the initial 1 ns NPT simulation; both D̃T and 1/τ̃rot were then computed at this P–T point from the NVT simulation. We then followed a similar procedure as above, running simulations at varying pressures at the temperature points 275, 300, 325, and 350 K. The temperature/pressure points making up both isomorphs along with their calculated values of D̃T and 1/τ̃rot are listed in Table I.

TABLE I.

D̃T and 1/τ̃rot for isomorph 1. Uncertainties in both D̃T and 1/τ̃rot are ∼8% due to uncertainties in the predicted density at a given T/P state point.

Isomorph pointD̃T(103)1/τ̃rot(101)
300 K/1 atm 4.1 0.95 
325 K/1000 atm 4.1 0.98 
350 K/3200 atm 3.9 1.02 
400 K/4600 atm 3.9 1.03 
250 K/1 atm 1.2 0.31 
275 K/1300 atm 1.2 0.32 
300 K/2400 atm 1.1 0.32 
325 K/4000 atm 1.1 0.32 
350 K/5600 atm 1.1 0.33 
Isomorph pointD̃T(103)1/τ̃rot(101)
300 K/1 atm 4.1 0.95 
325 K/1000 atm 4.1 0.98 
350 K/3200 atm 3.9 1.02 
400 K/4600 atm 3.9 1.03 
250 K/1 atm 1.2 0.31 
275 K/1300 atm 1.2 0.32 
300 K/2400 atm 1.1 0.32 
325 K/4000 atm 1.1 0.32 
350 K/5600 atm 1.1 0.33 

Isomorph and density scaling behavior present in reduced units of time, length, and energy as described by Gnan et al.20 The reduced units of length and time used here are ρ−1/3 and m/kBT/ρ1/3, respectively. Here, we take the mass “m” as the molecular mass of PC, but note that most scaling arguments are derived for atomic fluids. We compute scaled diffusion coefficients as D̃=ρ1/3m/kBTD. We compute scaled time constants τ using the reduced time unit. Dynamic scattering functions are given in real length/time units when comparing to experiment, and reduced length/time units otherwise, and this convention will be explicitly stated in context of discussion.

We note that experimental characterization of density scaling has primarily utilized real rather than reduced units. Formally, the latter should always be used, but in practice use of real units is largely inconsequential when relaxation times are plotted over many orders of magnitude. For example, Pawlus et al.40 demonstrated density scaling for propylene carbonate by characterizing relaxation times over ten orders of magnitude, at different phase points. While relaxation times were plotted in real units, the conversion to reduced units differs by order unity for the investigated phase space and is thus not observable over plots that span many decades in time. We show a plot comparing reduced vs real units for DT vs τ in Fig. S2 for the PC pseudoisomorph points.

In computer simulations, excitation populations are generally quantified by searching for particles that undergo sufficiently large spatial displacements that persist in time, usually on the order of a picosecond.5,50 In such an analysis, one must average out fast vibrations that correspond to motion within the same inherent structure, and one must justify the spatial and temporal cutoff values applied to discriminate hops.5,50 Instead, we quantify excitation populations through analysis of the self-intermediate scattering function F(q, t)35,36 by modeling it as a double-Gaussian in q as follows:

(3)

This approach has the advantage of being applicable to both experimental and computational characterization, and it requires no a priori assumptions about length scales or timescales of the excitations since the length scale and timescale signatures of excitations and their associated hops come directly from fitting to the model. Here, Φ(t) signifies the fraction of particles (H atoms or molecular centers of mass) that have undergone an excursion of length scale σIB up to time t. Φ(t ≈ 1 ps) is the instantaneous population of particles at an IB barrier and thus the instantaneous excitation population. Moreover, at t ≈ 1 ps, σIS and σIB are the characteristic length scales for inherent state elastic deformation and for hopping, respectively. In Eq. (3), we have neglected a diffusive term, so, at later times, σIS and σIB begin to grow slowly as they acquire diffusive character. This form provides excellent fits for experimental and simulation data as is shown in Figs. 35. The time-dependence of fit parameters σIS, σIB, and Φ is shown.

We first briefly discuss the accuracy of the PC force field utilized for MD simulations as this dictates the accuracy of all the simulation results presented in the manuscript. Compared to experiment, errors in the predicted density and enthalpy of vaporization are 3% and 9%, respectively, at 300 K, 1 atm; such accuracy is comparable to that of previous property predictions of organic liquids with the SAPT-FF force field.51 Additionally, the computed diffusion coefficients and rotational correlation times are within 10% of the experimental values over the temperature range of 250–350 K at 1.0 atm (Fig. S8). Because there are no empirical parameters in the model, similar accuracy is expected throughout the studied temperature–pressure phase diagram.

Utilizing a brute-force screening approach (Sec. II B), we were able to locate two curves on the temperature–pressure phase diagram that display largely invariant structural and dynamical properties (computed in reduced units). The curves that we located are shown in Fig. 1, in which we have mapped the pressure to the equilibrium density. Also shown in Fig. 1, for perspective, is the glass transition curve.52 One pseudoisomorphic curve starts at 300 K, 1 atm and extends to higher temperature/pressure, and the other starts at 250 K, 1 atm; by intention, both curves were chosen to be far away from the glass transition.

Our isomorphic curves display identical density scaling to that previously determined experimentally for propylene carbonate. Pawlus et al.40 found that relaxation times and dc conductivity in propylene carbonate solutions measured across the phase diagram collapsed to a common curve when plotted against V−3.7/T.40 This scaling exponent of γTS = 3.7 agrees with our computed isomorph curves as shown in Fig. 1. All points on a given isomorph curve were fit to the form ργTS/T=constant. A best fit of our data gave a value of γTS = 3.75 ± 0.09, which is consistent with the findings of Pawlus.

In Table I, we give calculated diffusion coefficients (D̃T) and inverse rotational relaxation times (1/τ̃rot), both in reduced units, for the two pseudoisomorphic curves. These computed dynamical properties are invariant along each pseudoisomorph to within statistical uncertainty. In addition to invariance to these dynamic measures, we expect structural properties of the fluid to be invariant along pseudoisomorphs. To analyze the liquid structure, we compute the radial distribution functions (RDFs) between propylene carbonate molecules in the liquid as a function of center-of-mass positions. The RDFs are shown in Fig. 2(a) (top panel) and in Fig. S3 of the supplementary material as computed for all phase points on each pseudoisomorph curve. The RDFs for all points on each pseudoisomorph collapse to a single curve when the length scale is plotted in reduced units. In the inset, we highlight the similarity across phase points with a high resolution scale of the first RDF peak. For comparison, we show RDFs computed from 1 atm isobar phase points spanning the temperature range 300–350 K in Fig. 2(b). It is clear from the insets in Figs. 2(a) and 2(b) that there is structural deviation (albeit small) between the isobar points, in contrast with the identical structure among pseudoisomorph phase points.

FIG. 2.

Center-of-mass structural and dynamic metrics at pseudoisomorphic phase points compared to isobar phase points. Top: RDFs computed between propylene carbonate center-of-mass positions and plotted as a function of reduced distance for (a) all points along the 300 K pseudoisomorph and (b) for 1 atm isobar points at several temperatures. Bottom: Comparison of F(q,t) at selected wavevectors, computed based on center-of-mass positions for (c) 300 K, 1 atm pseudoisomorph phase points and (d) for 1 atm isobar points at several temperatures.

FIG. 2.

Center-of-mass structural and dynamic metrics at pseudoisomorphic phase points compared to isobar phase points. Top: RDFs computed between propylene carbonate center-of-mass positions and plotted as a function of reduced distance for (a) all points along the 300 K pseudoisomorph and (b) for 1 atm isobar points at several temperatures. Bottom: Comparison of F(q,t) at selected wavevectors, computed based on center-of-mass positions for (c) 300 K, 1 atm pseudoisomorph phase points and (d) for 1 atm isobar points at several temperatures.

Close modal

In the bottom panel of Fig. 2, we show the self part of the intermediate scattering function F(q, t), defined as

(4)

where q is the momentum transfer wavevector and ⟨⋯⟩ denotes an ensemble average. In Fig. 2(c), we show Fs(q, t) computed for molecular center-of-mass (COM) positions, and in Fig. S3 we show Fs(q, t) calculated for hydrogen atom (H-atom) positions [i.e., the sum in Eq. (4) runs over COM or H-atom indices, respectively]. Here and for the remainder of the manuscript, all reference to the scattering function concerns the self part (and not the distinct contribution), and the quantity in Eq. (4) will be referred to simply as Fq,t (“s” subscript dropped). The q values shown describe dynamics on length scales ranging from 70% to 2.5 times the average intermolecular spacing (qmax,COM = 5.76). At each of these length scales, COM dynamics among all the phase points investigated are identical to within uncertainty. These combined dynamic and structural analyses indicate that our identified phase point curves (Fig. 1) indeed behave as pseudoisomorphs. For comparison, in Fig. 2(d), we show Fs(q, t) computed from 1 atm isobar phase points spanning the temperature range 300–350 K (computed for COM positions) As expected, there is significant temperature dependence of dynamic relaxation for the isobar points; in contrast, the pseudoisomorph phase points exhibit essentially identical COM dynamics as indicated by Fs(q, t) in Fig. 2(c).

Based on our simulation analysis, we find that propylene carbonate exhibits density scaling such that structural and dynamic properties are functions of ρ3.7/T, in agreement with previous experimental findings.40 

Interbasin barrier crossing events in atomic model systems are local cooperative particle reorganizations3,53 and represent fundamental relaxation events. We have recently shown that motion associated with IB crossing and IS basin exploration can be quantified from the scattering function.36 Specifically, we showed that experimental S(q, ω) [the time Fourier transform of F(q, t)] from quasi-elastic neutron scattering (QENS) on propylene carbonate is optimally fitted with three Lorentzians, having q-dependent amplitudes consistent with a diffusive process and two localized processes. The latter two processes had time-dependent length scale characteristics corresponding to localized processes, consistent with inherent state (IS) vibrational motion (σv) and interbasin (IB) hopping motion (σh) on a PEL. In order to characterize IB and IS motions, in this study we focus on F(q, t) calculated from Eq. (4). For comparison with QENS data from PC that were previously acquired by one of us,36 we sum over H atoms because incoherent scattering from hydrogen atoms dominates the QENS signal in simple organics.

In Fig. 3, we compare the scattering function computed from our simulations to that previously measured by QENS.36 There is generally good agreement between the simulated and experimental scattering functions, attesting to the accuracy of the employed force field. Since we do not include coherent scattering in Eq. (4), we expect deviations of a few percent at low q values, which are not visible in Fig. 3. However, there are prominent discrepancies between simulated and experimental Fq,t at q > qmax = 1.1 Å−2 and at short times and high q values.

FIG. 3.

(Top) Scattering function Fs(q, t) computed from MD simulation at 300 K, 1 atm based on H-atom positions. (Bottom) Scattering function Fs(q, t) measured by incoherent QENS from PC.36 

FIG. 3.

(Top) Scattering function Fs(q, t) computed from MD simulation at 300 K, 1 atm based on H-atom positions. (Bottom) Scattering function Fs(q, t) measured by incoherent QENS from PC.36 

Close modal

These discrepancies are not surprising, considering the approximations of the intramolecular interaction potential (harmonic bonds, angles, etc.) used in the simulations (supplementary material). For example, the largest plotted wavevector q2 = 5 Å−2 corresponds to less than 3 Å real space distance, which is smaller than the molecular diameter of PC. Thus, Fq,t at this wavevector is expected to be quite sensitive to the torsional potentials and even methyl group rotation, the latter of which is not expected to be accurately modeled by the force field. In contrast, the scattering at q < qmax (distances longer than intermolecular separation) is dictated largely by intermolecular forces and the intermolecular potential utilized in the MD simulations is expected to be quite accurate, as confirmed by the good agreement with experiment for Fq,t in this regime.

We obtain parameters for excitations, IB and IS motion by fitting F(q, t) with Eq. (3). A straight line in a ln(F(q, t)) vs q2 plot signifies a single relaxation process that is Gaussian in q. The bilinear form of lnF(q,t) vs q2 illustrates the two-step relaxation process. From a data-fitting perspective, Φ(t) serves as a time-dependent non-Gaussian parameter. From a particle dynamics perspective, Φ(t) represents the fraction of particles that have “hopped” (undergone a non-reversing IB crossing) up to time t. Bearing in mind that hops constitute significantly larger excursions than IS cage distortions, it is clear that until a molecule executes a hop, the q-dependence of its scattering signature will be characteristic of only small length scale σIS. Once a hop occurs, that signature will change to larger length scale for all subsequent times.

Figure 4 shows F(q, t) calculated from hydrogen atoms, fits to Eq. (3), and fit parameters. As shown in the top panel, the model fits quite well, fully parameterizing F(q, t) for this q and t range. The middle and bottom panels in Fig. 4 show excellent agreement in the time-dependence of model fits for each of the 300 K, 1 atm pseudoisomorphs. In the middle panel, Φ(t) increases monotonically as expected,36 with (1 − Φ) exhibiting an initial rapid drop during the period of quasi-ballistic motion when atoms are exploring the cage formed by their neighbors. As we have explained elsewhere,39 just after this initial drop in (1 − Φ), Φ(t ≈ 1 ps) represents the instantaneous fraction of H atoms involved in large-length scale motion characteristic of structural excitations. After this initial period of cage exploration, (1 − Φ) decays almost exponentially as excitations visit new regions of the sample and facilitate hops of an increasingly large fraction of particles. In the bottom panel, the characteristic length scales σIS and σIB initially show signatures of localized motion during the period of cage exploration, reaching quasi-asymptotic values of ≈0.1 and ≈0.2 respectively, consistent with IS vibration and IB crossing.53 The values of these fit parameters continue to grow, taking on a diffusive character at longer times in this model because, for simplicity’s sake, we do not explicitly include diffusion through a separate parameter.

FIG. 4.

F(q, t) and fits. (Top) F(q, t) at 300 K, 1 atm based on H-atom positions, plotted against q2 at evenly spaced times between 0.5 and 6 ps. Black points are calculated from simulation and dashed red lines are fits to Eq. (3). (Middle and Bottom) Time-dependence of model fits shown for each pseudoisomorph of 300 K, 1 atm.

FIG. 4.

F(q, t) and fits. (Top) F(q, t) at 300 K, 1 atm based on H-atom positions, plotted against q2 at evenly spaced times between 0.5 and 6 ps. Black points are calculated from simulation and dashed red lines are fits to Eq. (3). (Middle and Bottom) Time-dependence of model fits shown for each pseudoisomorph of 300 K, 1 atm.

Close modal

The high degree of overlap among the fit parameters at each of the pseudoisomorphic phase points signifies excellent correspondence between F(q, t) for H atoms at each phase point and demonstrates that excitation populations as well as dynamic timescales and length scales associated with IS vibrations and IB crossing obey density scaling. We have previously shown that the IS and IB modes are related in a simple way to α relaxation and translational diffusion,35,36,38 so the fact that they exhibit scaling is consistent with Hansen’s conclusion (based on longer timescale data) that motion coupled to α relaxation should exhibit density scaling.23 

While the dynamics encoded by the diffusion coefficient, rotational relaxation time (Table I), and F(q, t) for COM integrate over intramolecular motion, the atomic scattering function formally encompasses dynamical processes over all temporal and spatial scales relevant to IB and IS motion. The fast timescales and short length scales of IS and IB motion make them candidates for being associated with intramolecular modes. Comparing the top panels of Figs. 4 and 5 provides evidence that IS and IB modes in PC involve significant intramolecular motion. The top panel of Fig. 5 shows FCOM(q, t) calculated from molecular centers of mass for the 300 K, 1 atm phase point. Comparing with FH(q, t) from H atoms at the same phase point in Fig. 4, it is quite clear that the COM motion is of lower amplitude overall since, for all q and t, FH(q, t) ≤ FCOM(q, t). A similar relationship was found between COM and H-atom F(q, t) for glycerol.38 Molecular rotation will cause some differential decay of FH(q, t) compared to its COM counterpart, but a detailed comparison of the two scattering functions suggests another significant contribution. The two characteristic length scales for motion are similar for H-atom and COM fits, suggesting that the molecule moves in a quasi-rigid fashion, but ΦH ≥ ΦCOM at all times. This pronounced muting of the non-Gaussian nature of the COM F(q, t) with respect to the H-atom F(q, t) directly demonstrates that IS and IB motion executed by PC are largely intramolecular in nature (although they could in principle involve parts of more than one molecule). Hopping of individual H atoms—parts of the molecule—precedes similar length scale motion of the entire molecule reflected in the COM.

FIG. 5.

Top: F(q, t) computed from MD simulation at 300 K, 1 atm based on molecular centers of mass. Black points are calculated from simulation and dashed red lines are fits to Eq. (3). Middle and bottom: Time-dependence of model fits shown for each pseudoisomorph of 300 K, 1 atm.

FIG. 5.

Top: F(q, t) computed from MD simulation at 300 K, 1 atm based on molecular centers of mass. Black points are calculated from simulation and dashed red lines are fits to Eq. (3). Middle and bottom: Time-dependence of model fits shown for each pseudoisomorph of 300 K, 1 atm.

Close modal

We turn now to the question of which intramolecular modes are likely to contribute to IB and IS motions and why, if they contribute significantly, do they not spoil the density scaling demonstrated in Fig. 4. The origin of the effective IPL appears to be collective in nature since the strong energy–virial correlations are observed only when averaging over the ensemble in rigid particle models.19 Thus, we expect that if any intramolecular mode does contribute to density scaling, the contributing modes would most likely be low-frequency ones that involve significant portions of the molecule.

In the top left panel of Fig. 6, we overlay Raman spectra of PC from two sources. The red solid trace shows the lowest frequency intramolecular vibrational modes, obtained through spontaneous Raman scattering. These Raman modes at 191, 318, and 451 cm−1 are all molecularly extended ring deformation modes. The black trace is the Fourier transform of the βfast component from time-domain optical polarizability measured by Optical Kerr effect (OKE) spectroscopy.54 The βfast OKE signal contains a collective and librational part, which we demonstrated to have the same temperature-dependent frequencies and relative amplitudes as that of IB barrier crossing and IS motion measured by neutron scattering.54 The blue dashed line represents the spectrum of IB motion and the red dashed line represents the IS spectrum. The black line is the sum of the IS and IB spectra. In the PEL picture, both the IS and IB modes are coupled to relaxation in the sense that the IS modes assist in basin exploration but the molecular reorganizations that make up relaxation events proceed directly along IB mode coordinates. The intramolecular vibrational modes with frequencies below 500 cm−1 overlap spectrally with and thus may couple to the IS and IB modes.

FIG. 6.

Top left: Polarizability and Raman response in PC. Black line: βfast polarizability spectra from OKE measurements.54 Green dashed line: OKE spectrum due to IS motion (known as librations in OKE literature). Blue dashed line: OKE spectrum due to IB crossing (known as collective motion in OKE literature). Red solid line: Low-frequency component of a PC spontaneous Raman spectrum. Right top and bottom: Power spectra of velocity autocorrelation functions for all atoms (right top) and centers of mass (right bottom) at 300 K and the pressures indicated in the middle panel. Inset to right bottom panel: Grüneisen parameters calculated for each mode found in the power spectra. Left bottom: Model fit to the 300 K 1 atm centers-of-mass power spectrum.

FIG. 6.

Top left: Polarizability and Raman response in PC. Black line: βfast polarizability spectra from OKE measurements.54 Green dashed line: OKE spectrum due to IS motion (known as librations in OKE literature). Blue dashed line: OKE spectrum due to IB crossing (known as collective motion in OKE literature). Red solid line: Low-frequency component of a PC spontaneous Raman spectrum. Right top and bottom: Power spectra of velocity autocorrelation functions for all atoms (right top) and centers of mass (right bottom) at 300 K and the pressures indicated in the middle panel. Inset to right bottom panel: Grüneisen parameters calculated for each mode found in the power spectra. Left bottom: Model fit to the 300 K 1 atm centers-of-mass power spectrum.

Close modal

In the right top and bottom panels of Fig. 6, we overlay power spectra of velocity autocorrelation functions for ring atoms and molecular centers of mass, respectively, at the three simulated pressures indicated. The right top panel shows that the pressure dependence of the intermolecular mode characteristic frequencies drops with increasing frequency, indicating that higher frequency peaks contribute less to density scaling. Consistent with this, we found that eliminating the much higher frequency bond-stretching modes by constraining all bond lengths led to no change in density scaling (supplementary material, Fig. S4). Also consistent with these findings, Raman modes found below 800 cm−1 in organic molecules with more than one backbone atom typically involve molecular deformations and typically have larger Grüneisen parameters than do higher frequency modes involving bond stretches and H-X bending.55–57 

The broad βfast features below ≈150 cm−1 in the right bottom panel are due to rotational and translational motion, but they arise only from translational motion in the right bottom panel. In this spectral region, the IS and IB responses are not well-separated, but their characteristic frequencies appear to have relatively strong pressure dependencies. We quantify the characteristic frequencies of the IB and IS motion by fitting the COM power spectra to functions commonly used to characterize OKE data54 as described in the supplementary material. We find acceptable fits using functions corresponding to IB and IS motion and an additional Gaussian process at slightly higher frequency than the IS mode. In the left bottom panel of Fig. 6, we show the model fit to the 300 K, 1 atm COM power spectrum. Analogous fits for 1900 and 4600 atm are shown in supplementary material Fig. S5 and are of similar quality.

In the inset to the right bottom panel of Fig. 6, we show Grüneisen parameters for each of the modes found in the power spectra, calculated as

(5)

where i represents the ith vibrational mode. Among the modes characterized, we see a continuous progression of increasing γG with decreasing frequency. We emphasize that, although the IB and IS modes are distinct from intramolecular eigenmodes, the differences in FH(q, t) and FCOM(q, t) suggest they involve intramolecular motion and thus likely couple to these modes.

Finally, we note that γIBγTS. This suggests that it may be generally possible to find γTS in a simulated liquid by simply finding γIB. It also suggests that the relationship between γTS and the Grüneisen parameter may be more direct than previously thought. Associations between γTS and γG have been made theoretically through the latter’s connection to configurational entropy30 or through the kinetic energy derivative of the pressure in the harmonic limit,19 but the striking result in the inset to Fig. 6 suggests that the Grüneisen parameter may influence γTS directly through γIB.

The Grüneisen parameter formally characterizes the volume dependence of normal mode frequencies in a crystal, but, of course, there is no relaxation in an ideal crystal. Similarly, Debye–Waller (DW) factors strictly describe only harmonic vibrations but are used to describe liquid relaxation through the mean square displacement u2. An extensive literature, going back three decades35,58–61 has explored the relationship between α relaxation and u2 measured at ≈1 ps, the time of the βfast process. In these approaches, it is assumed either that the amplitude of harmonic motion describes a localization volume for particles trapped in a cage of their neighbors or is proportional to the probability of a local relaxation event. Recently, we have made a more direct connection between motions that constitute u2 in amorphous systems and relaxation processes by demonstrating that u2 contains signatures of both IS basin exploration and elemental relaxation events—the IB barrier crossing. Thus, when properly parsed, dynamic signatures of the elemental relaxation process are obtained directly from the βfast relaxation of u2 or F(q, t).35,36

In analogy with the idea that u2 is directly related to relaxation through its reporting on IB barrier crossings, we propose that a causal connection exists between γG and γTS through γIB as follows: The normal modes of the liquid that couple to the IB process are by definition found at the Frenkel frequency,62νF = 1/τ. At this frequency, the liquid supports longitudinal modes, but transverse modes are overdamped since the liquid can reorganize on the timescale of these local shear deformations, and this motion directly couples to relaxation. In an amorphous system, longitudinal and transverse modes will also be coupled since there is no well-defined extended lattice. Since the motion along these mode coordinates is coupled to elemental relaxation (IB crossing) events, the resonant frequency of the mode can be thought of as a relaxation attempt frequency. Thus, for a fixed barrier height, small changes in the frequency of these modes would lead to directly proportional changes in the rate of elemental relaxation events. We have previously demonstrated that for T > TB (TB being where the α and Johari–Goldstein β relaxations bifurcate), the timescales for IB crossing events and α relaxation have the same temperature dependence.36 Thus, for this temperature range, the Grüneisen parameter of the Frenkel modes (γIB), which describes the pressure dependence of their frequency, also describes the pressure dependence of the α relaxation time (γTS) through the pressure dependence of the elemental relaxation time. As we see in Fig. 6, γIBγG.

Above, we showed that high-frequency modes associated strictly with bond stretching have no impact on density scaling. Beyond arguing that the pressure dependence of lower frequency intramolecular modes and their spectral overlap with IB modes suggest that the former may contribute to IB barrier crossings and to density scaling, we can only speculate as to how important they may be. As we have shown above, hops associated with IB transitions appear to occur at the atomic rather than molecular level. Thus, intramolecular modes that conserve molecular center of mass may well contribute to hops of constituent atoms. However, to quantify this would require a series of simulations in which we systematically constrain these various intramolecular modes. While standard MD simulations are done using Cartesian equations of motion, an alternative approach for exploring the role of intramolecular modes on liquid dynamics would be to perform molecular dynamics in general/internal coordinates (or hybrid) to enable explicit interrogation of specific internal modes.63–65 This direction will be explored in the future work.

While evidence of density scaling in βfast at ∼1 ps timescales has been demonstrated previously, the relationship involved a quadratic correction term.24 Given that βfast relaxation has contributions from distinct modes of motion, some of them potentially intramolecular, and given that intramolecular vibrational modes can destroy the virial–energy correlation that is often the formal basis for density scaling arguments, it was unclear whether all aspects of βfast motion would couple or whether the necessity of a quadratic correction term signaled a partial decoupling. In this work, we have shown that the IB barrier crossing aspect of the βfast, a fundamental structural excitation mode, seems to obey the same density scaling, whereas the higher frequency IS mode is associated with a nearly twofold smaller scaling factor. Because structural excitations correspond to fundamental hopping (intrabasin barrier crossing) processes on the liquid potential energy landscape (PEL), our results suggest a novel microscopic perspective for density scaling. Interpretation of density scaling based on excitations may prove a promising general perspective applicable to molecular liquids for which virial–energy correlation or single inverse power law arguments may not hold due to the intramolecular modes or due to multiple effective potentials.

For propylene carbonate, we have demonstrated that the characteristic length scales σIS and σIB have the same values in F(q, t) computed for COM and to H-atom motion but that Φ(t) calculated for COM always trails that of H atoms. We interpret this as demonstrating that excitations must thus be interpreted as essentially atomic rather than molecular displacement events. This important conclusion has implications for generalizing excitation theories of atomic and model fluids5,50 to real molecular liquids. The fact that the Grüneisen parameter for collective IB modes is approximately equal to γTS suggests that finding these modes in molecular liquids will be the key to connecting excitation approaches in these liquids to those of atomic model systems.

The supplementary material includes figures and brief discussion concerning the following: the lack of virial–energy correlation for PC, a brief comparison of the use of real and reduced units for finding pseudoisomorphs in PC, a description of the radial distribution functions and intermediate scattering functions for pseudoisomorph families, the pressure dependence of bond-stretching vibrations and the fitting of velocity autocorrelation power spectra to find the pressure dependence of low-frequency intermolecular collective motion and intramolecular molecule deformation vibrations, and, finally, force field construction and benchmarks.

J.P.S. and J.G.M. acknowledge support from the Air Force Office of Scientific Research under Award No. FA9550-22-1-0025.

The authors have no conflicts to disclose.

John P. Stoppleman: Data curation (equal); Formal analysis (equal); Writing – original draft (equal). Jesse G. McDaniel: Funding acquisition (equal); Project administration (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal). Marcus T. Cicerone: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Project administration (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
H.
Miyagawa
,
Y.
Hiwatari
,
B.
Bernu
, and
J. P.
Hansen
, “
Molecular dynamics study of binary soft-sphere mixtures: Jump motions of atoms in the glassy state
,”
J. Chem. Phys.
88
,
3879
3886
(
1988
).
2.
C.
Bennemann
,
C.
Donati
,
J.
Baschnagel
, and
S. C.
Glotzer
, “
Growing range of correlated motion in a polymer melt on cooling towards the glass transition
,”
Nature
399
,
246
249
(
1999
).
3.
T. F.
Middleton
and
D. J.
Wales
, “
Energy landscapes of some model glass formers
,”
Phys. Rev. B
64
,
024205
(
2001
).
4.
D.
Chandler
and
J. P.
Garrahan
, “
Dynamics on the way to forming glass: Bubbles in space-time
,”
Annu. Rev. Phys. Chem.
61
,
191
217
(
2010
).
5.
A. S.
Keys
,
L. O.
Hedges
,
J. P.
Garrahan
,
S. C.
Glotzer
, and
D.
Chandler
, “
Excitations are localized and relaxation is hierarchical in glass-forming liquids
,”
Phys. Rev. X
1
,
021013
(
2011
).
6.
Y. S.
Elmatad
,
D.
Chandler
, and
J. P.
Garrahan
, “
Corresponding states of structural glass formers
,”
J. Phys. Chem. B
113
,
5563
5567
(
2009
).
7.
Y.
Jung
,
J. P.
Garrahan
, and
D.
Chandler
, “
Excitation lines and the breakdown of Stokes-Einstein relations in supercooled liquids
,”
Phys. Rev. E
69
,
061205
(
2004
).
8.
M.
Merolle
,
J. P.
Garrahan
, and
D.
Chandler
, “
Space–time thermodynamics of the glass transition
,”
Proc. Natl. Acad. Sci. U. S. A.
102
,
10837
(
2005
).
9.
L. O.
Hedges
,
L.
Maibaum
,
D.
Chandler
, and
J. P.
Garrahan
, “
Decoupling of exchange and persistence times in atomistic models of glass formers
,”
J. Chem. Phys.
127
,
211101
(
2007
).
10.
A. S.
Keys
,
J. P.
Garrahan
, and
D.
Chandler
, “
Calorimetric glass transition explained by hierarchical dynamic facilitation
,”
Proc. Natl. Acad. Sci. U. S. A.
110
,
4482
4487
(
2013
).
11.
A.
Tölle
, “
Neutron scattering studies of the model glass former ortho-terphenyl
,”
Rep. Prog. Phys.
64
,
1473
(
2001
); arXiv:cond-mat/0102352.
12.
R.
Casalini
and
C. M.
Roland
, “
Thermodynamical scaling of the glass transition dynamics
,”
Phys. Rev. E
69
,
062501
(
2004
).
13.
C.
Dreyfus
,
A.
Le Grand
,
J.
Gapinski
,
W.
Steffen
, and
A.
Patkowski
, “
Scaling the α-relaxation time of supercooled fragile organic liquids
,”
Eur. Phys. J. B
42
,
309
319
(
2004
).
14.
C. M.
Roland
,
S.
Hensel-Bielowka
,
M.
Paluch
, and
R.
Casalini
, “
Supercooled dynamics of glass-forming liquids and polymers under hydrostatic pressure
,”
Rep. Prog. Phys.
68
,
1405
1478
(
2005
).
15.
C. M.
Roland
,
S.
Bair
, and
R.
Casalini
, “
Thermodynamic scaling of the viscosity of van der Waals, H-bonded, and ionic liquids
,”
J. Chem. Phys.
125
,
124508
(
2006
).
16.
R.
Casalini
,
S.
Capaccioli
, and
C. M.
Roland
, “
What can we learn by squeezing a liquid?
,”
J. Phys. Chem. B
110
,
11491
11495
(
2006
).
17.
N. P.
Bailey
,
U. R.
Pedersen
,
N.
Gnan
,
T. B.
Schrøder
, and
J. C.
Dyre
, “
Pressure-energy correlations in liquids. I. Results from computer simulations
,”
J. Chem. Phys.
129
,
184507
(
2008
).
18.
N. P.
Bailey
,
U. R.
Pedersen
,
N.
Gnan
,
T. B.
Schrøder
, and
J. C.
Dyre
, “
Pressure-energy correlations in liquids. II. Analysis and consequences
,”
J. Chem. Phys.
129
,
184508
(
2008
).
19.
T. B.
Schrøder
,
N. P.
Bailey
,
U. R.
Pedersen
,
N.
Gnan
, and
J. C.
Dyre
, “
Pressure-energy correlations in liquids. III. Statistical mechanics and thermodynamics of liquids with hidden scale invariance
,”
J. Chem. Phys.
131
,
234503
(
2009
).
20.
N.
Gnan
,
T. B.
Schrøder
,
U. R.
Pedersen
,
N. P.
Bailey
, and
J. C.
Dyre
, “
Pressure-energy correlations in liquids. IV. ‘Isomorphs’ in liquid phase diagrams
,”
J. Chem. Phys.
131
,
234504
(
2009
).
21.
M.
Paluch
,
S.
Haracz
,
A.
Grzybowski
,
M.
Mierzwa
,
J.
Pionteck
,
A.
Rivera-Calzada
, and
C.
Leon
, “
A relationship between intermolecular potential, thermodynamics, and dynamic scaling for a supercooled ionic liquid
,”
J. Phys. Chem. Lett.
1
,
987
992
(
2010
).
22.
K.
Adrjanowicz
,
J.
Pionteck
, and
M.
Paluch
, “
Isochronal superposition and density scaling of the intermolecular dynamics in glass-forming liquids with varying hydrogen bonding propensity
,”
RSC Adv.
6
,
49370
49375
(
2016
).
23.
H. W.
Hansen
,
B.
Frick
,
S.
Capaccioli
,
A.
Sanz
, and
K.
Niss
, “
Isochronal superposition and density scaling of the α-relaxation from pico- to millisecond
,”
J. Chem. Phys.
149
,
214503
(
2018
).
24.
F.
Puosi
,
O.
Chulkin
,
S.
Bernini
,
S.
Capaccioli
, and
D.
Leporini
, “
Thermodynamic scaling of vibrational dynamics and relaxation
,”
J. Chem. Phys.
145
,
234904
(
2016
).
25.
I. H.
Bell
, “
Probing the link between residual entropy and viscosity of molecular fluids and model potentials
,”
Proc. Natl. Acad. Sci. U. S. A.
116
,
4070
(
2019
).
26.
U. R.
Pedersen
,
N. P.
Bailey
,
T. B.
Schrøder
, and
J. C.
Dyre
, “
Strong pressure-energy correlations in van der Waals liquids
,”
Phys. Rev. Lett.
100
,
015701
(
2008
).
27.
J. P.
Gabriel
,
M.
Tress
,
W.
Kossack
,
L.
Popp
, and
F.
Kremer
, “
Molecular heterogeneities in the thermal expansivity of polyalcohols
,”
J. Chem. Phys.
154
,
024503
(
2021
).
28.
C. M.
Roland
,
R.
Casalini
,
R.
Bergman
, and
J.
Mattsson
, “
Role of hydrogen bonds in the supercooled dynamics of glass-forming liquids at high pressures
,”
Phys. Rev. B
77
,
012201
(
2008
).
29.
A. A.
Veldhorst
,
J. C.
Dyre
, and
T. B.
Schrøder
, “
Scaling of the dynamics of flexible Lennard-Jones chains: Effects of harmonic bonds
,”
J. Chem. Phys.
143
,
194503
(
2015
).
30.
C. M.
Roland
and
R.
Casalini
, “
Entropy basis for the thermodynamic scaling of the dynamics of o-terphenyl
,”
J. Phys.: Condens. Matter
19
,
205118
(
2007
).
31.
S.
Bernini
,
F.
Puosi
, and
D.
Leporini
, “
Thermodynamic scaling of relaxation: Insights from anharmonic elasticity
,”
J. Phys.: Condens. Matter
29
,
135101
(
2017
).
32.
K. G.
Honnell
,
C. K.
Hall
, and
R.
Dickman
, “
On the pressure equation for chain molecules
,”
J. Chem. Phys.
87
,
664
674
(
1987
).
33.
A. E.
Olsen
,
J. C.
Dyre
, and
T. B.
Schrøder
, “
Communication: Pseudoisomorphs in liquids with intramolecular degrees of freedom
,”
J. Chem. Phys.
145
,
241103
(
2016
).
34.
M. C. C.
Ribeiro
,
T.
Scopigno
, and
G.
Ruocco
, “
Computer simulation study of thermodynamic scaling of dynamics of 2Ca(NO3)2 · 3KNO3
,”
J. Chem. Phys.
135
,
164510
(
2011
).
35.
M. T.
Cicerone
,
Q.
Zhong
, and
M.
Tyagi
, “
Picosecond dynamic heterogeneity, hopping, and Johari-Goldstein relaxation in glass-forming liquids
,”
Phys. Rev. Lett.
113
,
117801
(
2014
).
36.
M. T.
Cicerone
and
M.
Tyagi
, “
Metabasin transitions are Johari-Goldstein relaxation events
,”
J. Chem. Phys.
146
,
054502
(
2017
).
37.
F. H.
Stillinger
, “
A topographic view of supercooled liquids and glass formation
,”
Science
267
,
1935
(
1995
).
38.
M. T.
Cicerone
,
D.
Averett
, and
J. J.
de Pablo
, “
The role of hopping on transport above Tc in glycerol
,”
J. Non-Cryst. Solids
407
,
118
125
(
2015
).
39.
M. T.
Cicerone
and
K.
Badilla-Nunez
, “
Do crossovers in liquid transport mechanisms arise from percolation of discrete dynamic environments?
,” arXiv:2201.12593 [cond-mat.soft] (
2022
).
40.
S.
Pawlus
,
R.
Casalini
,
C. M.
Roland
,
M.
Paluch
,
S. J.
Rzoska
, and
J.
Ziolo
, “
Temperature and volume effects on the change of dynamics in propylene carbonate
,”
Phys. Rev. E
70
,
061501
(
2004
).
41.
J.-C.
Soetens
,
C.
Millot
,
B.
Maigret
, and
I.
Bakó
, “
Molecular dynamics simulation and X-ray diffraction studies of ethylene carbonate, propylene carbonate and dimethyl carbonate in liquid phase
,”
J. Mol. Liq.
92
,
201
216
(
2001
).
42.
X.
You
,
M. I.
Chaudhari
,
S. B.
Rempe
, and
L. R.
Pratt
, “
Dielectric relaxation of ethylene carbonate and propylene carbonate from molecular dynamics simulations
,”
J. Phys. Chem. B
120
,
1849
1853
(
2016
).
43.
V. A.
Koverga
,
I. V.
Voroshylova
,
Y.
Smortsova
,
F.-A.
Miannay
,
M. N. D. S.
Cordeiro
,
A.
Idrissi
, and
O. N.
Kalugin
, “
Local structure and hydrogen bonding in liquid γ-butyrolactone and propylene carbonate: A molecular dynamics simulation
,”
J. Mol. Liq.
287
,
110912
(
2019
).
44.
P.
Eastman
,
J.
Swails
,
J. D.
Chodera
,
R. T.
McGibbon
,
Y.
Zhao
,
K. A.
Beauchamp
,
L.-P.
Wang
,
A. C.
Simmonett
,
M. P.
Harrigan
,
C. D.
Stern
,
R. P.
Wiewiora
,
B. R.
Brooks
, and
V. S.
Pande
, “
OpenMM 7: Rapid development of high performance algorithms for molecular dynamics
,”
PLoS Comput. Biol.
13
,
e1005659
(
2017
).
45.
J. G.
McDaniel
and
J. R.
Schmidt
, “
Physically-motivated force fields from symmetry-adapted perturbation theory
,”
J. Phys. Chem. A
117
,
2053
2066
(
2013
).
46.
G.
Lamoureux
and
B.
Roux
, “
Modeling induced polarization with classical Drude oscillators: Theory and molecular dynamics simulation algorithm
,”
J. Chem. Phys.
119
,
3025
3039
(
2003
).
47.
U.
Essmann
,
L.
Perera
,
M. L.
Berkowitz
,
T.
Darden
,
H.
Lee
, and
L. G.
Pedersen
, “
A smooth particle mesh Ewald method
,”
J. Chem. Phys.
103
,
8577
8593
(
1995
).
48.
J. M.
Martínez
and
L.
Martínez
, “
Packing optimization for automated generation of complex system’s initial configurations for molecular dynamics and docking
,”
J. Comput. Chem.
24
,
819
825
(
2003
).
49.
B.
Hess
,
C.
Kutzner
,
D.
Van Der Spoel
, and
E.
Lindahl
, “
GROMACS 4: Algorithms for highly efficient, load-balanced, and scalable molecular simulation
,”
J. Chem. Theory Comput.
4
,
435
447
(
2008
).
50.
S. S.
Schoenholz
,
E. D.
Cubuk
,
D. M.
Sussman
,
E.
Kaxiras
, and
A. J.
Liu
, “
A structural approach to relaxation in glassy liquids
,”
Nat. Phys.
12
,
469
471
(
2016
).
51.
J. G.
McDaniel
and
J. R.
Schmidt
, “
First-principles many-body force fields from the gas phase to liquid: A ‘universal’ approach
,”
J. Phys. Chem. B
118
,
8042
8053
(
2014
).
52.
M.
Bonetti
and
A.
Dubois
, “
Isochronal superpositioning in the equilibrium regime of superpressed propylene carbonate to ∼1.8 GPa: A study by diffusivity measurement of the fluorescent probe Coumarin 1
,”
Eur. Phys. J. E
42
,
97
(
2019
).
53.
M.
Vogel
,
B.
Doliwa
,
A.
Heuer
, and
S. C.
Glotzer
, “
Particle rearrangements during transitions between local minima of the potential energy landscape of a binary Lennard-Jones liquid
,”
J. Chem. Phys.
120
,
4404
4414
(
2004
).
54.
J. S.
Bender
,
M.
Zhi
, and
M. T.
Cicerone
, “
The polarizability response of a glass-forming liquid reveals intrabasin motion and interbasin transitions on a potential energy landscape
,”
Soft Matter
16
,
5588
5598
(
2020
).
55.
K.
Goossens
,
L.
Smeller
, and
K.
Heremans
, “
Pressure tuning spectroscopy of the low-frequency Raman spectrum of liquid amides
,”
J. Chem. Phys.
99
,
5736
5741
(
1993
).
56.
H.
Shimizu
,
Y.
Ikeda
, and
S.
Sasaki
, “
High-pressure Raman study of liquid and crystalline carbonyl sulfide
,”
Chem. Phys. Lett.
175
,
349
353
(
1990
).
57.
K.
Aoki
,
Y.
Kakudate
,
S.
Usuba
,
M.
Yoshida
,
K.
Tanaka
, and
S.
Fujiwara
, “
High-pressure Raman study of liquid and crystalline C2H2
,”
J. Chem. Phys.
88
,
4565
4568
(
1988
).
58.
B.
Frick
,
D.
Richter
,
W.
Petry
, and
U.
Buchenau
, “
Study of the glass transition order parameter in amorphous polybutadiene by incoherent neutron scattering
,”
Z. Phys. B: Condens. Matter
70
,
73
79
(
1988
).
59.
M. T.
Cicerone
and
C. L.
Soles
, “
Fast dynamics and stabilization of proteins: Binary glasses of trehalose and glycerol
,”
Biophys. J.
86
,
3836
3845
(
2004
).
60.
L.
Larini
,
A.
Ottochian
,
C.
De Michele
, and
D.
Leporini
, “
Universal scaling between structural relaxation and vibrational dynamics in glass-forming liquids and polymers
,”
Nat. Phys.
4
,
42
45
(
2008
).
61.
D. S.
Simmons
,
M. T.
Cicerone
,
Q.
Zhong
,
M.
Tyagi
, and
J. F.
Douglas
, “
Generalized localization model of relaxation in glass-forming liquids
,”
Soft Matter
8
,
11455
11461
(
2012
).
62.
J.
Frenkel
,
Kinetic Theory of Liquids
, International Series of Monographs on Physics, edited by
R.
Fowler
,
P.
Kapitza
, and
N.
Mott
(
Oxford University Press
,
Oxford
,
1946
).
63.
D. J.
Tobias
and
C. L.
Brooks
, “
Molecular dynamics with internal coordinate constraints
,”
J. Chem. Phys.
89
,
5115
5127
(
1988
).
64.
S.
He
and
H. A.
Scheraga
, “
Macromolecular conformational dynamics in torsional angle space
,”
J. Chem. Phys.
108
,
271
286
(
1998
).
65.
N.
Vaidehi
and
A.
Jain
, “
Internal coordinate molecular dynamics: A foundation for multiscale dynamics
,”
J. Phys. Chem. B
119
,
1233
1242
(
2015
).

Supplementary Material