Many ionic liquids show behavior similar to that of glassy systems, e.g., large and long-lasted deviations from Gaussian dynamics and clustering of “mobile” and “immobile” groups of ions. Herein a time-dependent four-point density correlation function—typically used to characterize glassy systems—is implemented for the ionic liquids, choline acetate, and 1-butyl-3-methylimidazolium acetate. Dynamic correlation beyond the first ionic solvation shell on the time scale of nanoseconds is found in the ionic liquids, revealing the cooperative nature of ion motions. The traditional solvent, acetonitrile, on the other hand, shows a much shorter length-scale that decays after a few picoseconds.

It is no surprise that the use of ionic liquids (ILs)—salts that exist in a molten state below 100 °C—has been tested and shows promise in a broad range of applications.1–5 Unusual and often desirable properties, such as chemical and thermal stability, high effective polarity, nonvolatility, wide electrochemical window, etc., make them ideal for a variety of specialized tasks. Furthermore, by judiciously engineering functional groups or exchanging ions entirely, the possibility arises for generating customizable ILs.

Parallel to a material’s applicability comes interest in the underlying molecular behavior. Many experimental6–23 and computational23–40 studies have reported the dynamic behavior of different ILs, revealing an array of interesting features, indicating their dynamics are heterogeneous. The examples include spatially correlated dynamics, sub-diffusive translational motions, non-diffusive reorientational dynamics, non-exponential solvent relaxation, etc. These dynamic behaviors, except for spatially correlated dynamics, have been analyzed using theoretical approaches based on two-point time correlation functions, such as van Hove functions.

In this paper, we extend previous computational efforts to gain further insight into heterogeneity of IL dynamics by using a time-dependent four-point density correlation function.41 This approach enables, for example, the investigation of the cooperative IL dynamics and the related dynamic length scale for spatial correlation, not accessible via two-point time correlation functions. While four-point time correlation functions have been used to characterize the dynamic behavior of structurally correlated species in glass-like systems,41–50 their application to ILs has been extremely limited.51 

Here, we study the time evolution of correlated motions in choline acetate ([Cho][Ac]) and 1-butyl-3-methylimidazolium acetate ([BMI][Ac]) (Fig. 1) as a protype of hydroxylic and non-hydroxylic ILs, respectively. For comparison, acetonitrile is considered as a standard nonglassy solvent.

FIG. 1.

Notation used throughout this work. The parenthetical letter in CAC(C/B) indicates if the acetate ion is paired with choline (CAC(C)) or BMI (CAC(B)).

FIG. 1.

Notation used throughout this work. The parenthetical letter in CAC(C/B) indicates if the acetate ion is paired with choline (CAC(C)) or BMI (CAC(B)).

Close modal

Details of the simulation systems are compiled in Table I. Three trajectories of each IL system listed there and one trajectory of acetonitrile were simulated with the GROMACS simulation package.52–55 As in Ref. 38, we used CHARMM force field parameters for ILs, generated via the ParamChem v.0.9.7 interface58 that automatically assigns atom types, force field parameters, and charges on molecules based on the 2b8 release of the CHARMM general force field.59–61 (See supplementary material for the complete list of force field parameters for BMI employed in the present study.) We note that the density of [BMI][Ac] determined with these parameters shows a very good agreement with the available experimental data (Table I). For acetonitrile, the potential model developed in Ref. 62 was employed. Electrostatic interactions were computed using the particle mesh Ewald (PME) method with a 1.4 nm cutoff. Nosé-Hoover temperature coupling and Parrinello-Rahman pressure coupling were used to maintain constant temperature and pressure of the system, respectively. For [Cho][Ac], initial configurations for each simulation were generated by annealing the system in the isothermal-isobaric (NPT) ensemble from 700 K and then equilibrating at production temperatures for 6 ns for 400 K and 350 K simulations and 2 ns for 600 K in the canonical (NVT) ensemble. The initial structures of [BMI][Ac] were prepared by annealing the system from 800 K in the NPT ensemble, followed by equilibration at production temperatures for 10 ns. The acetonitrile simulation was preceded by 0.2 ns of equilibration in the NVT ensemble at production temperatures. All production runs were performed in the NVT ensemble.

TABLE I.

Simulation systems: N refers to the number of ion pairs for ILs, while it refers to the number of molecules for acetonitrile; n is the average number density; ρ and ρexp are simulated and experimental densities; D is the diffusion coefficient of acetate anion and acetonitrile calculated using the Einstein equation; and time indicates the length of each production trajectory.

TnρρexpDTime
Solvent(K)N (nm−3)(g/cm3) (g/cm3) (10−12 m2/s) (ns)
[Cho][Ac] 350 256 4.149 1.124 … 0.13 200 
[BMI][Ac] 350 512 3.099 1.020 1.0192a 0.27 150 
[Cho][Ac] 400 256 4.079 1.106 … 3.0 30 
[BMI][Ac] 400 512 3.026 0.993 … 7.1 20 
[Cho][Ac] 600 256 3.737 1.013 … 4.4 × 102 10 
CH3CN 300 512 11.387 0.776 0.7857b 3.5 × 103 0.5 
TnρρexpDTime
Solvent(K)N (nm−3)(g/cm3) (g/cm3) (10−12 m2/s) (ns)
[Cho][Ac] 350 256 4.149 1.124 … 0.13 200 
[BMI][Ac] 350 512 3.099 1.020 1.0192a 0.27 150 
[Cho][Ac] 400 256 4.079 1.106 … 3.0 30 
[BMI][Ac] 400 512 3.026 0.993 … 7.1 20 
[Cho][Ac] 600 256 3.737 1.013 … 4.4 × 102 10 
CH3CN 300 512 11.387 0.776 0.7857b 3.5 × 103 0.5 
a

353 K, Ref. 56.

b

298 K, Ref. 57.

We start by considering the static structure of the system in the form of the radial distribution function (RDF), gAB(r), i.e.,

gAB(r)=VNANBiAjBδrrij,
(1)

where i and j represent individual particles and A and B denote species. The MD results for RDFs are displayed at 400 K for [Cho][Ac] and [BMI][Ac] and at 300 K for acetonitrile in Fig. 2. Higher temperature RDFs resemble those displayed in Fig. 2 but may show minor variations in peak positions and dimensions. For reference, the position (r′) and the height (g(r′)) of the first peak in each RDF are displayed in Table II.

FIG. 2.

RDFs for (a) [Cho][Ac] at 400 K, (b) [BMI][Ac] at 400 K, and (c) acetonitrile at 300 K.

FIG. 2.

RDFs for (a) [Cho][Ac] at 400 K, (b) [BMI][Ac] at 400 K, and (c) acetonitrile at 300 K.

Close modal
TABLE II.

Primary peak positions (r′) and heights (g(r′)).

RDFT (K)rg(r′)
NCHO⋯NCHO 350 0.60 2.2 
CAC(C)⋯CAC(C) 350 0.64 1.6 
CAC(C)⋯NCHO 350 0.43 3.0 
CBMI⋯CBMI 350 0.75 1.2 
CAC(B)⋯CAC(B) 350 0.66 1.4 
CAC(B)⋯CBMI 350 0.34 3.1 
NCHO⋯NCHO 400 0.61 2.1 
CAC(C)⋯CAC(C) 400 0.65 1.6 
CAC(C)⋯NCHO 400 0.44 3.0 
CBMI⋯CBMI 400 0.74 1.2 
CAC(B)⋯CAC(B) 400 0.65 1.4 
CAC(B)⋯CBMI 400 0.34 2.9 
NCHO⋯NCHO 600 0.63 1.9 
CAC(C)⋯CAC(C) 600 0.65 1.6 
CAC(C)⋯NCHO 600 0.46 2.9 
NAN⋯NAN 300 0.41 1.3 
CAN⋯CAN 300 0.44 1.5 
NAN⋯CAN 300 0.35 2.5 
RDFT (K)rg(r′)
NCHO⋯NCHO 350 0.60 2.2 
CAC(C)⋯CAC(C) 350 0.64 1.6 
CAC(C)⋯NCHO 350 0.43 3.0 
CBMI⋯CBMI 350 0.75 1.2 
CAC(B)⋯CAC(B) 350 0.66 1.4 
CAC(B)⋯CBMI 350 0.34 3.1 
NCHO⋯NCHO 400 0.61 2.1 
CAC(C)⋯CAC(C) 400 0.65 1.6 
CAC(C)⋯NCHO 400 0.44 3.0 
CBMI⋯CBMI 400 0.74 1.2 
CAC(B)⋯CAC(B) 400 0.65 1.4 
CAC(B)⋯CBMI 400 0.34 2.9 
NCHO⋯NCHO 600 0.63 1.9 
CAC(C)⋯CAC(C) 600 0.65 1.6 
CAC(C)⋯NCHO 600 0.46 2.9 
NAN⋯NAN 300 0.41 1.3 
CAN⋯CAN 300 0.44 1.5 
NAN⋯CAN 300 0.35 2.5 

The ILs show a prominent oscillatory behavior beyond 1 nm. This is generally attributed to an alternating-charge shell structure, i.e., cations tend to be surrounded by anions which, in turn, tend to be surrounded by cations.63 This oscillatory behavior can be found in acetonitrile to a slightly reduced degree—the amplitude of oscillations is smaller and they extend to shorter distances.

To obtain insight into the anisotropic character of relative distributions of ions, we considered three-dimensional spatial distribution functions (SDFs). The results for cation distributions around the acetate anion are presented as isodensity surfaces in Fig. 3 as an example. The isodensity values are chosen to highlight the first solvation shells.

FIG. 3.

Isodensity plots of SDFs for (a) NBMI and (b) CBMI of [BMI], and (c) CCHO of [Cho] around acetate at 400 K.

FIG. 3.

Isodensity plots of SDFs for (a) NBMI and (b) CBMI of [BMI], and (c) CCHO of [Cho] around acetate at 400 K.

Close modal

The SDF results indicate that cations—more precisely, their polar groups—are distributed preferentially around the C–O bonds of acetate. Though not shown here, the remaining solvation shell in the direction of the Ac methyl group is filled mainly by nonpolar groups of ions, viz., the butyl chain of BMI cations and the methyl group of other Ac anions. It is worthwhile to note that the absence of long aliphatic chains in choline reduces its amphiphilic nature considerably, compared to, e.g., BMI. As a result, the nonpolar hemisphere of the first solvation shell of acetate in [Cho][Ac] is filled predominantly by the methyl group of other Ac ions.38 The double peaks in Fig. 2(b) and the corresponding double shells in Fig. 3(b) are due to the interactions of CBMI of BMI cations with Ac anions in different directions, as shown in Fig. 4.

FIG. 4.

Ionic interactions of BMI and Ac ions.

FIG. 4.

Ionic interactions of BMI and Ac ions.

Close modal

A time-dependent order parameter to characterize the system “glassiness” for species A can be defined as

QAp(t)=i,jAδri(0)rj(t).
(2)

As discussed in Ref. 44, local vibrational dynamics, if present, can relax Qp(t) rapidly. Since our main focus is on correlations associated with systemic structural deformations as opposed to weakly correlated localized vibrations, we remove the latter by employing a coarse-graining function,42–44 

w(r)=1ifr<a,0ifr>a,
(3)

in place of the delta function in Eq. (2). This gives the species an effective volume of V0=43πa3 and leads to

QA(t)=i,jAwri(0)rj(t).
(4)

For IL systems, we employed a = 0.4 σ. For acetonitrile, a values for its carbon and nitrogen atoms, CAN and NAN, were 0.5 σ and 0.6 σ, respectively. Numerical values of a used for our analysis here can be found in Table III. These a values were selected to optimize the susceptibility associated with Q(t) (see below).44 It is important to note that Q(t) does not distinguish between particles of like species. This means that there is no contribution to the decay of Q(t) when, for example, two particles of species A exchange positions. Because of this lack of distinction between particles of like species and the finite values of a, we find that Q()/N=N0, where N0V0n and n is the number density of the species; n values are listed in Table I, and N0 values are listed in Table III.

TABLE III.

Q(t)/N fitting parameters for Q(t)/N=i=14AiQexp[t/τiQ]+N0, where N0=Q()/N. a is the parameter in w(r) function.

AtomT (K)a (nm)N0τ1Q (ns)A1Qτ2Q (ns)A2Qτ3Q (ns)A3Qτ4Q (ns)A4Q
NCHO 350 0.1320 0.039 97 0.078 0.04 2.339 0.14 19.467 0.66 … … 
CAC(C) 350 0.1424 0.050 18 0.084 0.03 2.581 0.14 19.467 0.67 … … 
CBMI 350 0.1567 0.049 95 0.009 0.04 0.116 0.07 1.172 0.13 10.118 0.53 
CAC(B) 350 0.1424 0.037 48 0.007 0.06 0.111 0.08 1.166 0.14 10.631 0.51 
NCHO 400 0.1320 0.039 30 0.046 0.14 0.575 0.36 2.794 0.43 … … 
CAC(C) 400 0.1424 0.049 34 0.049 0.13 0.630 0.36 2.864 0.44 … … 
CBMI 400 0.1567 0.048 77 0.002 0.07 0.025 0.12 0.154 0.22 0.895 0.51 
CAC(B) 400 0.1424 0.036 60 0.001 0.10 0.020 0.13 0.127 0.20 0.792 0.49 
NCHO 600 0.1320 0.036 00 0.001 0.35 0.006 0.38 0.023 0.24 … … 
CAC(C) 600 0.1424 0.045 20 0.001 0.35 0.006 0.38 0.024 0.27 … … 
NAN 300 0.1956 0.356 9 1.2 × 10−3 0.72 … … … … … … 
CAN 300 0.1700 0.234 3 1.2 × 10−3 0.85 … … … … … … 
AtomT (K)a (nm)N0τ1Q (ns)A1Qτ2Q (ns)A2Qτ3Q (ns)A3Qτ4Q (ns)A4Q
NCHO 350 0.1320 0.039 97 0.078 0.04 2.339 0.14 19.467 0.66 … … 
CAC(C) 350 0.1424 0.050 18 0.084 0.03 2.581 0.14 19.467 0.67 … … 
CBMI 350 0.1567 0.049 95 0.009 0.04 0.116 0.07 1.172 0.13 10.118 0.53 
CAC(B) 350 0.1424 0.037 48 0.007 0.06 0.111 0.08 1.166 0.14 10.631 0.51 
NCHO 400 0.1320 0.039 30 0.046 0.14 0.575 0.36 2.794 0.43 … … 
CAC(C) 400 0.1424 0.049 34 0.049 0.13 0.630 0.36 2.864 0.44 … … 
CBMI 400 0.1567 0.048 77 0.002 0.07 0.025 0.12 0.154 0.22 0.895 0.51 
CAC(B) 400 0.1424 0.036 60 0.001 0.10 0.020 0.13 0.127 0.20 0.792 0.49 
NCHO 600 0.1320 0.036 00 0.001 0.35 0.006 0.38 0.023 0.24 … … 
CAC(C) 600 0.1424 0.045 20 0.001 0.35 0.006 0.38 0.024 0.27 … … 
NAN 300 0.1956 0.356 9 1.2 × 10−3 0.72 … … … … … … 
CAN 300 0.1700 0.234 3 1.2 × 10−3 0.85 … … … … … … 

The MD results for Q(t)/N are displayed in Fig. 5. The relaxation of Q(t)/N was found to be well fitted by the sum of multiple exponential decays, i.e.,

Q(t)N=iAiQexp[t/τiQ]+N0.
(5)

Values for a, N0, τiQ, and AiQ can be found in Table III. [Cho][Ac] and [BMI][Ac] were fitted with three and four exponential functions, respectively. By contrast, a single exponential function well describes Q(t)/N of acetonitrile, indicating that underlying dynamics are Gaussian. As is expected, Q(t)/N decays faster at higher temperatures; all values of τiQs are smaller and the contribution of the fastest decay component, viz., A1Q, to the overall relaxation increases. At 350 K and 400 K, Q(t)/N for [Cho][Ac]—the more viscous of the two ILs—decays more slowly than Q(t)/N for [BMI][Ac]; at all simulated temperatures, both ILs decay much more slowly than acetonitrile.

FIG. 5.

Plots of Q(t)/N for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. For each IL system, cation⋯cation curves are blue and anion⋯anion curves are red; for acetonitrile, NAN⋯NAN is yellow and CAN⋯CAN is orange. The dashed line is just a guide to the eye.

FIG. 5.

Plots of Q(t)/N for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. For each IL system, cation⋯cation curves are blue and anion⋯anion curves are red; for acetonitrile, NAN⋯NAN is yellow and CAN⋯CAN is orange. The dashed line is just a guide to the eye.

Close modal

To understand the time evolution of the dynamic correlation between species A and B, we analyzed the four-point overlap correlation function g4ol(r,t) between the two species

g4ol(r,t)=VNANBijAklBδrrk(0)+ri(0)×wri(0)rj(t)wrk(0)rl(t).
(6)

For simplicity, we consider a radially averaged correlation function, g4ol(r,t), where r=r. We note that g4ol(r,t) is a direct extension of the full van Hove function; it describes the time evolution of correlation of density fluctuations occurring simultaneously at two different positions, located a distance of r apart. As such, it gauges the four-particle conditional probability, viz., given a particle of species A at position 1 and particle of species B at position 2 at the initial time t = 0, it measures (up to a constant) the probability that the same or another particle of A and the same or another particle of B will be found at positions very close to the original locations 1 and 2, respectively, at a later time. Like the order parameter described above, g4ol(r,t) does not distinguish between particles of the same species. This means that at time t = 0, g4ol(r,0) becomes the RDF, i.e., g4ol(r,0)=g(r). We also note that g4ol(,t)=QA(t)/NAQB(t)/NB because particles are of finite size and become spatially uncorrelated at large r. In other words, g4ol(r,t) does not decay to zero due to the finite value of a.

Analogous to the fitting for Q(t)/N in Eq. (5), we were able to fit g4ol(r,t) for ILs accurately with three exponential functions as

g4ol(r,t)=i=13Aiol(r)exp[t/τiol(r)]+b(r).
(7)

We found that g4ol(r,t) for acetonitrile is well described by a single exponential function. As mentioned above, b(r) in Eq. (7) becomes QA()/NAQB()/NB(=NA,0NB,0) at large r. Values for τiol, Aiol, and b for r = r′, i.e., positions of primary RDF peaks, are presented in Table IV (see Table II for values of r′ and g(r′)). The corresponding relaxation behaviors of g4ol(r,t) for the systems considered here are compared in Fig. 6. At a given temperature, [BMI][Ac] shows a faster decay of g4ol(r,t) than [Cho][Ac] just like the Q(t)/N case. The g4ol(r,t) relaxation in acetonitrile at 300 K is faster than that in [BMI][Ac] and [Cho][Ac] at 350 K by ∼4 orders of magnitude.

TABLE IV.

g4ol(r,t) fitting at r = r′.

Pair typesT (K)b(r)τ1ol(r) (ns)A1ol(r)τ2ol(r) (ns)A2ol(r)τ3ol(r) (ns)A3ol(r)
NCHO⋯NCHO 350 0.090 0.13 0.17 2.44 0.49 14.00 1.34 
CAC(C)⋯CAC(C) 350 0.063 0.10 0.11 2.44 0.42 14.59 1.03 
NCHO⋯CAC(C) 350 0.227 0.11 0.19 2.21 0.63 13.70 1.92 
CBMI⋯CBMI 350 0.092 0.10 0.26 1.12 0.26 7.95 0.58 
CAC(B)⋯CAC(B) 350 0.094 0.08 0.37 1.07 0.32 8.84 0.66 
CBMI⋯CAC(B) 350 0.380 0.09 0.69 0.98 0.57 8.20 1.58 
NCHO⋯NCHO 400 0.013 0.06 0.56 0.44 0.91 1.73 0.61 
CAC(C)⋯CAC(C) 400 0.010 0.06 0.43 0.54 0.78 1.97 0.39 
NCHO⋯CAC(C) 400 0.021 0.07 0.81 0.52 1.27 1.90 0.84 
CBMI⋯CBMI 400 0.020 0.01 0.40 0.11 0.37 0.57 0.42 
CAC(B)⋯CAC(B) 400 0.063 0.01 0.46 0.06 0.35 0.40 0.57 
CBMI⋯CAC(B) 400 0.181 0.01 0.81 0.07 0.72 0.45 1.33 
NCHO⋯NCHO 600 0.0033 0.95 × 10−3 0.99 4.30 × 10−3 0.71 15.93 × 10−3 0.19 
CAC(C)⋯CAC(C) 600 0.0041 0.08 × 10−3 0.64 4.15 × 10−3 0.65 16.18 × 10−3 0.18 
NCHO⋯CAC(C) 600 0.0088 0.80 × 10−3 1.35 4.34 × 10−3 1.20 16.24 × 10−3 0.36 
NAN⋯NAN 300 0.15 1.05 × 10−3 1.13 … … … … 
CAN⋯CAN 300 0.07 0.90 × 10−3 1.34 … … … … 
NAN⋯CAN 300 0.22 0.99 × 10−3 2.22 … … … … 
Pair typesT (K)b(r)τ1ol(r) (ns)A1ol(r)τ2ol(r) (ns)A2ol(r)τ3ol(r) (ns)A3ol(r)
NCHO⋯NCHO 350 0.090 0.13 0.17 2.44 0.49 14.00 1.34 
CAC(C)⋯CAC(C) 350 0.063 0.10 0.11 2.44 0.42 14.59 1.03 
NCHO⋯CAC(C) 350 0.227 0.11 0.19 2.21 0.63 13.70 1.92 
CBMI⋯CBMI 350 0.092 0.10 0.26 1.12 0.26 7.95 0.58 
CAC(B)⋯CAC(B) 350 0.094 0.08 0.37 1.07 0.32 8.84 0.66 
CBMI⋯CAC(B) 350 0.380 0.09 0.69 0.98 0.57 8.20 1.58 
NCHO⋯NCHO 400 0.013 0.06 0.56 0.44 0.91 1.73 0.61 
CAC(C)⋯CAC(C) 400 0.010 0.06 0.43 0.54 0.78 1.97 0.39 
NCHO⋯CAC(C) 400 0.021 0.07 0.81 0.52 1.27 1.90 0.84 
CBMI⋯CBMI 400 0.020 0.01 0.40 0.11 0.37 0.57 0.42 
CAC(B)⋯CAC(B) 400 0.063 0.01 0.46 0.06 0.35 0.40 0.57 
CBMI⋯CAC(B) 400 0.181 0.01 0.81 0.07 0.72 0.45 1.33 
NCHO⋯NCHO 600 0.0033 0.95 × 10−3 0.99 4.30 × 10−3 0.71 15.93 × 10−3 0.19 
CAC(C)⋯CAC(C) 600 0.0041 0.08 × 10−3 0.64 4.15 × 10−3 0.65 16.18 × 10−3 0.18 
NCHO⋯CAC(C) 600 0.0088 0.80 × 10−3 1.35 4.34 × 10−3 1.20 16.24 × 10−3 0.36 
NAN⋯NAN 300 0.15 1.05 × 10−3 1.13 … … … … 
CAN⋯CAN 300 0.07 0.90 × 10−3 1.34 … … … … 
NAN⋯CAN 300 0.22 0.99 × 10−3 2.22 … … … … 
FIG. 6.

Plots of g4ol(r,t) for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. For each IL system, cation⋯cation curves are blue, anion⋯anion curves are red, and cation⋯anion curves are purple; for acetonitrile, NAN⋯NAN is orange, CAN⋯CAN is yellow, and NAN⋯CAN is green. The dotted line is just a guide to the eye.

FIG. 6.

Plots of g4ol(r,t) for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. For each IL system, cation⋯cation curves are blue, anion⋯anion curves are red, and cation⋯anion curves are purple; for acetonitrile, NAN⋯NAN is orange, CAN⋯CAN is yellow, and NAN⋯CAN is green. The dotted line is just a guide to the eye.

Close modal

For further insight, we have examined the r-dependence of τiol(r) and their normalized contributions Aiol(r)/i=13Aiol(r). The results for [Cho][Ac] at 400 K are shown in Fig. 7. The different τiol(r) values correspond to different relaxation time-domains of g4ol(r,t). The fastest decay time, τ1ol(r), likely due to initial motions of ions within their cages, exhibits essentially no dependence on r [Fig. 7(a)]. We think that such motions are fairly localized and thus rapid density fluctuations arising from them are not affected by the motions of other ions. The slower decay times, especially τ3ol(r), on the other hand, show considerable r-dependence for r ≲ 1.3 nm. This means that slow density fluctuations of ions are likely influenced by and therefore correlated with the density fluctuations of other ions located within a distance of ∼1.3 nm. The results in Fig. 7(b) show that all three decay components make a significant contribution to the overall relaxation of g4ol(r,t). There are several noteworthy features. First, the normalized amplitudes of the two slow components, A2,3ol(r)/i=13Aiol(r), are strongly correlated while they show near perfect anticorrleation with the normalized amplitude of the fastest component. Second, the r-behavior of the amplitudes of the slow components is very similar to the corresponding RDF (cf. Fig. 2). We note that [BMI][Ac] shows behaviors similar to [Cho][Ac] and thus is not shown here.

FIG. 7.

Plots of τiol for CAC(C)⋯NCHO at 400 K are shown in (a), and plots of cofactors, Aiol(r), normalized to the sum of cofactors (Atotalol(r)=i=13Aiol(r)) are shown in (b), where i = 1 is shown in red, i = 2 is shown in blue, and i = 3 is shown in green.

FIG. 7.

Plots of τiol for CAC(C)⋯NCHO at 400 K are shown in (a), and plots of cofactors, Aiol(r), normalized to the sum of cofactors (Atotalol(r)=i=13Aiol(r)) are shown in (b), where i = 1 is shown in red, i = 2 is shown in blue, and i = 3 is shown in green.

Close modal

The susceptibility χ4(t), which describes the fluctuations of the order parameter of the system, is defined as

χ4(t)=βVNQ(t)2Q(t)2,
(8)

with β = 1/kBT. χ4(t) measures cooperative motions of ions at two different locations, averaged over the system. Figure 8 displays χ4(t) for CAC(C)⋯CAC(C) at 400 K determined with different values of a. As mentioned above where a was first introduced, the a value with the greatest susceptibility is used for calculations throughout this work.44 

FIG. 8.

Plots of χ4(t) for CAC(C)⋯CAC(C) in [Cho][Ac] at 400 K with different values of a.

FIG. 8.

Plots of χ4(t) for CAC(C)⋯CAC(C) in [Cho][Ac] at 400 K with different values of a.

Close modal

Figure 9 shows the MD results for the susceptibility for each system at the selected a value (Table III). As is expected from the temporal behaviors of both the order parameters and the four-point density time correlation functions, we find χ4(t) peaks at longer times for lower temperatures. Furthermore, χ4(t) for more viscous [Cho][Ac] is slower than that for [BMI][Cl]. We ascribe the molecular origin of this difference to the formation of hydrogen bond in the former.38 Interestingly, the susceptibility calculated between like species is greater than the susceptibility calculated between unlike species.

FIG. 9.

Plots of χ4(t) for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. For each IL system, cation⋯cation curves are blue, anion⋯anion curves are red, and cation⋯anion curves are purple; for acetonitrile, NAN⋯NAN is orange, CAN⋯CAN is yellow, and NAN⋯CAN is green.

FIG. 9.

Plots of χ4(t) for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. For each IL system, cation⋯cation curves are blue, anion⋯anion curves are red, and cation⋯anion curves are purple; for acetonitrile, NAN⋯NAN is orange, CAN⋯CAN is yellow, and NAN⋯CAN is green.

Close modal

In many prior studies,25,29,30,38,50,51,64,65 the deviation of the IL transport behavior from Gaussian dynamics is measured by

α2(t)=35Δr4(t)Δr2(t)21,
(9)

where Δr(t) is the displacement of ions at time t from their initial positions. Comparison of the results in Fig. 10 with those in Fig. 9 reveals that temporal behaviors of α2(t) and χ4(t), including their peak positions, are very similar for all temperatures and systems we considered. This indicates that the non-Gaussian behavior of single particle dynamics is directly related to time-dependent correlated motions of two particles, viz., cooperative dynamics.

FIG. 10.

Plots of the non-Gaussian factor, α2(t), for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile; cations are shown in blue, anions are shown in red, NAN is in orange, and CAN is in yellow.

FIG. 10.

Plots of the non-Gaussian factor, α2(t), for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile; cations are shown in blue, anions are shown in red, NAN is in orange, and CAN is in yellow.

Close modal

Finally, we proceed to the dynamic length scale associated with cooperative motions in ILs. We employ the approach widely used in the analysis of glass systems,44–48,50 viz., a small wave vector expansion of the four-point structure factor

S4ol(q,t)=1NAi,jAexpiqri(0)w(ri(0)rj(t))×k,lAexpiqrk(0)w(rk(0)rl(t)).
(10)

The inclusion up to the q2 terms in the expansion à la Ornstein-Zernike (OZ) theory yields44,45

S4ol(q,t)=S4ol(0,t)1+(qξ4(t))α
(11)

with α = 2. The corresponding r-space behavior is given by

g4ol(r,t)S4ol(0,t)rξ4(t)2exp[r/ξ4(t)].
(12)

While a more general expression with α ≠ 2 is often used in the analysis of glass systems,46–48,50 we employ here the original OZ form,44,45 i.e., α = 2 in Eq. (11), mainly due to the limited number of data points available for the numerical analysis. This limitation arises from the small system size we used to speed up the simulations because long trajectories are needed for proper sampling of highly viscous and thus slowly relaxing ILs. We computed the radially averaged structure factor, S4ol(q,t) (q=|q|), by including in the calculations only q vectors that satisfy the periodic boundary conditions of the simulation system and fitted the resulting S4ol(q,t) with Eq. (11) (α = 2) at the lowest five q values using the Newton method with S4ol(0,t) and ξ4(t) treated as fitting parameters.

The results for ξ4(t) for like ions (cation⋯cation and anion⋯anion) are shown in Fig. 11. As is expected, increasing temperature leads to shorter correlation lengths, i.e., lower peak heights of ξ4(t). This means that the spatial range of cooperative dynamics becomes shorter as the temperature rises. Another important feature is that the spatial range of cooperative dynamics varies with time. This can be understood as follows: A dynamic event occurring at, say, position 1 cannot be influenced by another dynamic event at a different location, position 2, if the two events occur at the same time because the effect of a dynamic event cannot propagate instantaneously. However, a dynamic event occurring at a later time at position 1 can be influenced by a dynamic event at an earlier time at position 2. We expect that with time, the propagation and strength of the effect of a dynamic event grows and diminishes, respectively. As a result, the corresponding spatial range of the influence of a dynamic event will exhibit a non-monotonic behavior in t, reaching a maximum at some point, followed by a decrease. χ4(t) measures this time-dependence via Eq. (12). Comparison with Fig. 9 shows that the peak positions of ξ4(t) are close to those of χ4(t).

FIG. 11.

Like-species plots of ξ(t) vs. time for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. Cation⋯cation curves are in blue and anion⋯anion curves are in red; NAN⋯NAN is in orange and CAN⋯CAN is in yellow.

FIG. 11.

Like-species plots of ξ(t) vs. time for (a) [Cho][Ac], (b) [BMI][Ac], and (c) acetonitrile. Cation⋯cation curves are in blue and anion⋯anion curves are in red; NAN⋯NAN is in orange and CAN⋯CAN is in yellow.

Close modal

The correlation length scale can be compared with the sizes of ions shown in RDFs, Fig. 2. For [Cho][Ac], the maximum ξ4(t) at 600 K is smaller than the ionic radii. As the temperature decreases, the correlation length approaches and goes slightly above the ionic radii for both [Cho][Ac] and [BMI][Ac]. The length scale of acetonitrile is much less than its molecular size, indicating no correlated motions. It also shows that the correlation length depends on the dynamics across the system rather than the size of the species.

It is worthwhile to note that the temperatures of our simulations are well above the reported glass transition temperatures of both systems. We expect that if simulations were performed at lower temperatures, the length scale would further increase, which implies that both systems would show a stronger glassy behavior at room temperature.

Spatially correlated dynamics in ILs are brought about, in part, by distinct spatial clustering of “mobile” and “immobile” ions. Studies of glass systems with similar behavior implement a four-point time-correlation function to simultaneously gauge the length-scale of correlated regions and the time associated with the correlation.

In this paper, we studied the time-dependent four-point density correlation function (g4ol(r,t)), its related susceptibility (χ4(t)), and the dynamic length scale of correlation (ξ4(t)) for the ILs [Cho][Ac] and [BMI][Ac]. We found that g4ol(r,t) exhibits r-dependent relaxation, revealing the influence of spatial correlation on collective motions of ions. ξ4(t) determined by fitting the four-point structure factor S4ol(q,t) in the low q region shows that the spatial range of collective motions initially increases with time but decreases after reaching a maximum. Compared with the traditional solvent, acetonitrile, correlation extends to greater distances in the studied ILs and is retained for significantly longer times.

Our results indicate that four-point time correlation functions can shed new light on IL dynamics, not accessible via two-point correlation functions, and thus provide a useful theoretical tool to study IL dynamics, especially their cooperative aspect. Therefore it would be worthwhile in the future to study structure-dynamics relationships in ILs, such as the influence of structural heterogeneity on collective dynamics, using the four-point time correlation functions and their variants.

See supplementary material for the force field parameters of BMI employed in the present study.

This work was supported in part by the National Science Foundation through NSF Grant No. CHE-1223988. H.J.K. thanks Professor Biman Bagchi and Professor Shinji Saito for suggesting four-point time correlation function analysis of ionic liquid dynamics.

1.
N. V.
Plechkova
and
K. R.
Seddon
, “
Applications of ionic liquids in the chemical industry
,”
Chem. Soc. Rev.
37
,
123
150
(
2008
).
2.
M.
Watanabe
,
M. L.
Thomas
,
S.
Zhang
,
K.
Ueno
,
T.
Yasuda
, and
K.
Dokko
, “
Application of ionic liquids to energy storage and conversion materials and devices
,”
Chem. Rev.
117
,
7190
7239
(
2017
).
3.
K. S.
Egorova
,
E. G.
Gordeev
, and
V. P.
Ananikov
, “
Biological activity of ionic liquids and their application in pharmaceutics and medicine
,”
Chem. Rev.
117
,
7132
7189
(
2017
).
4.
S.
Amiril
,
E.
Rahim
, and
S.
Syahrullail
, “
A review on ionic liquids as sustainable lubricants in manufacturing and engineering: Recent research, performance, and applications
,”
J. Cleaner Prod.
168
,
1571
(
2017
).
5.
V. B.
Agbor
,
N.
Cicek
,
R.
Sparling
,
A.
Berlin
, and
D. B.
Levin
, “
Biomass pretreatment: Fundamentals toward application
,”
Biotechnol. Adv.
29
,
675
685
(
2011
).
6.
A.
Triolo
,
O.
Russina
,
V.
Arrighi
,
F.
Juranyi
,
S.
Janssen
, and
C. M.
Gordon
, “
Quasielastic neutron scattering characterization of the relaxation processes in a room temperature ionic liquid
,”
J. Chem. Phys.
119
,
8549
8557
(
2003
).
7.
J. A.
Ingram
,
R. S.
Moog
,
N.
Ito
,
R.
Biswas
, and
M.
Maroncelli
, “
Solute rotation and solvation dynamics in a room-temperature ionic liquid
,”
J. Phys. Chem. B
107
,
5926
5932
(
2003
).
8.
D.
Chakrabarty
,
P.
Hazra
,
A.
Chakraborty
,
D.
Seth
, and
N.
Sarkar
, “
Dynamics of solvent relaxation in room temperature ionic liquids
,”
Chem. Phys. Lett.
381
,
697
704
(
2003
).
9.
P. K.
Chowdhury
,
M.
Halder
,
L.
Sanders
,
T.
Calhoun
,
J. L.
Anderson
,
D. W.
Armstrong
,
X.
Song
, and
J. W.
Petrich
, “
Dynamic solvation in room-temperature ionic liquids
,”
J. Phys. Chem. B
108
,
10245
10255
(
2004
).
10.
P. K.
Mandal
,
M.
Sarkar
, and
A.
Samanta
, “
Excitation-wavelength-dependent fluorescence behavior of some dipolar molecules in room-temperature ionic liquids
,”
J. Phys. Chem. A
108
,
9048
9053
(
2004
).
11.
H.
Shirota
,
A. M.
Funston
,
J. F.
Wishart
, and
E. W.
Castner
, Jr.
, “
Ultrafast dynamics of pyrrolidinium cation ionic liquids
,”
J. Chem. Phys.
122
,
184512
(
2005
).
12.
A.
Samanta
, “
Dynamic Stokes shift and excitation wavelength dependent fluorescence of dipolar molecules in room temperature ionic liquids
,”
J. Phys. Chem. B
110
,
13704
13716
(
2006
).
13.
S.
Arzhantsev
,
H.
Jin
,
N.
Ito
, and
M.
Maroncelli
, “
Observing the complete solvation response of DCS in imidazolium ionic liquids, from the femtosecond to nanosecond regimes
,”
Chem. Phys. Lett.
417
,
524
529
(
2006
).
14.
B.
Lang
,
G.
Angulo
, and
E.
Vauthey
, “
Ultrafast solvation dynamics of coumarin 153 in imidazolium-based ionic liquids
,”
J. Phys. Chem. A
110
,
7028
7034
(
2006
).
15.
A. M.
Funston
,
T. A.
Fadeeva
,
J. F.
Wishart
, and
E. W.
Castner
, “
Fluorescence probing of temperature-dependent dynamics and friction in ionic liquid local environments
,”
J. Phys. Chem. B
111
,
4963
4977
(
2007
).
16.
S.
Arzhantsev
,
H.
Jin
,
G. A.
Baker
, and
M.
Maroncelli
, “
Measurements of the complete solvation response in ionic liquids
,”
J. Phys. Chem. B
111
,
4978
4989
(
2007
).
17.
H.
Jin
,
X.
Li
, and
M.
Maroncelli
, “
Heterogeneous solute dynamics in room temperature ionic liquids
,”
J. Phys. Chem. B
111
,
13473
13478
(
2007
).
18.
H.
Jin
,
X.
Li
, and
M.
Maroncelli
, “
Heterogeneous solute dynamics in room-temperature ionic liquids
,”
J. Phys. Chem. B
114
,
11370
(
2010
).
19.
O.
Yamamuro
,
T.
Yamada
,
M.
Kofu
,
M.
Nakakoshi
, and
M.
Nagao
, “
Hierarchical structure and dynamics of an ionic liquid 1-octyl-3-methylimidazolium chloride
,”
J. Chem. Phys.
135
,
054508
(
2011
).
20.
M.
Kofu
,
M.
Nagao
,
T.
Ueki
,
Y.
Kitazawa
,
Y.
Nakamura
,
S.
Sawamura
,
M.
Watanabe
, and
O.
Yamamuro
, “
Heterogeneous slow dynamics of imidazolium-based ionic liquids studied by neutron spin echo
,”
J. Phys. Chem. B
117
,
2773
2781
(
2013
).
21.
J.
Guo
,
G. A.
Baker
,
P. C.
Hillesheim
,
S.
Dai
,
R. W.
Shaw
, and
S. M.
Mahurin
, “
Fluorescence correlation spectroscopy evidence for structural heterogeneity in ionic liquids
,”
Phys. Chem. Chem. Phys.
13
,
12395
12398
(
2011
).
22.
S.
Patra
and
A.
Samanta
, “
Microheterogeneity of some imidazolium ionic liquids as revealed by fluorescence correlation spectroscopy and lifetime studies
,”
J. Phys. Chem. B
116
,
12275
12283
(
2012
).
23.
E. C.
Wu
,
H. J.
Kim
, and
L. A.
Peteanu
, “
Spectroscopic and MD study of dynamic and structural heterogeneities in ionic liquids
,”
J. Phys. Chem. B
121
,
1100
1107
(
2017
).
24.
Y.
Shim
,
J.
Duan
,
M. Y.
Choi
, and
H. J.
Kim
, “
Solvation in molecular ionic liquids
,”
J. Chem. Phys.
119
,
6411
6414
(
2003
).
25.
M. G.
Del Pópolo
and
G. A.
Voth
, “
On the structure and dynamics of ionic liquids
,”
J. Phys. Chem. B
108
,
1744
1752
(
2004
).
26.
Y.
Shim
,
M. Y.
Choi
, and
H. J.
Kim
, “
A molecular dynamics computer simulation study of room-temperature ionic liquids. II. Equilibrium and nonequilibrium solvation dynamics
,”
J. Chem. Phys.
122
,
044511
(
2005
).
27.
Y.
Shim
,
D.
Jeong
,
M. Y.
Choi
, and
H. J.
Kim
, “
Rotational dynamics of a diatomic solute in the room-temperature ionic liquid 1-ethyl-3-methylimidazolium hexafluorophosphate
,”
J. Chem. Phys.
125
,
061102
(
2006
).
28.
M. N.
Kobrak
, “
Characterization of the solvation dynamics of an ionic liquid via molecular dynamics simulation
,”
J. Chem. Phys.
125
,
064502
(
2006
).
29.
Z.
Hu
and
C. J.
Margulis
, “
Heterogeneity in a room-temperature ionic liquid: Persistent local environments and the red-edge effect
,”
Proc. Natl. Acad. Sci. U. S. A.
103
,
831
836
(
2006
).
30.
C.
Cadena
,
Q.
Zhao
,
R. Q.
Snurr
, and
E. J.
Maginn
, “
Molecular modeling and experimental studies of the thermodynamic and transport properties of pyridinium-based ionic liquids
,”
J. Phys. Chem. B
110
,
2821
2832
(
2006
).
31.
Z.
Hu
and
C. J.
Margulis
, “
Room-temperature ionic liquids: Slow dynamics, viscosity, and the red edge effect
,”
Acc. Chem. Res.
40
,
1097
1105
(
2007
).
32.
Y.
Shim
and
H. J.
Kim
, “
Dielectric relaxation, ion conductivity, solvent rotation, and solvation dynamics in a room-temperature ionic liquid
,”
J. Phys. Chem. B
112
,
11028
11038
(
2008
).
33.
D.
Jeong
,
M. Y.
Choi
,
H. J.
Kim
, and
Y.
Jung
, “
Fragility, Stokes-Einstein violation, and correlated local excitations in a coarse-grained model of an ionic liquid
,”
Phys. Chem. Chem. Phys.
12
,
2001
2010
(
2010
).
34.
Z. L.
Terranova
and
S. A.
Corcelli
, “
On the mechanism of solvation dynamics in imidazolium-based ionic liquids
,”
J. Phys. Chem. B
117
,
15659
15666
(
2013
).
35.
J. C.
Araque
,
S. K.
Yadav
,
M.
Shadeck
,
M.
Maroncelli
, and
C. J.
Margulis
, “
How is diffusion of neutral and charged tracers related to the structure and dynamics of a room-temperature ionic liquid? Large deviations from Stokes–Einstein behavior explained
,”
J. Phys. Chem. B
119
,
7015
7029
(
2015
).
36.
V.
Lesch
,
A.
Heuer
,
C.
Holm
, and
J.
Smiatek
, “
Solvent effects of 1-ethyl-3-methylimidazolium acetate: Solvation and dynamic behavior of polar and apolar solutes
,”
Phys. Chem. Chem. Phys.
17
,
8480
8490
(
2015
).
37.
D.
Kim
,
S.-W.
Park
,
Y.
Shim
,
H. J.
Kim
, and
Y.
Jung
, “
Excitation-energy dependence of solvation dynamics in room-temperature ionic liquids
,”
J. Chem. Phys.
145
,
044502
(
2016
).
38.
J. A. L.
Willcox
,
H.
Kim
, and
H. J.
Kim
, “
A molecular dynamics study of the ionic liquid, choline acetate
,”
Phys. Chem. Chem. Phys.
18
,
14850
14858
(
2016
).
39.
S. D.
Verma
,
S. A.
Corcelli
, and
M. A.
Berg
, “
Rate and amplitude heterogeneity in the solvation response of an ionic liquid
,”
J. Phys. Chem. Lett.
7
,
504
508
(
2016
).
40.
R. P.
Daly
,
J. C.
Araque
, and
C. J.
Margulis
, “
Communication: Stiff and soft nano-environments and the ‘octopus effect’ are the crux of ionic liquid structural and dynamical heterogeneity
,”
J. Chem. Phys.
147
,
061102
(
2017
).
41.
C.
Dasgupta
,
A. V.
Indrani
,
S.
Ramaswamy
, and
M. K.
Phani
, “
Is there a growing correlation length near the glass transition?
,”
Europhys. Lett.
15
,
307
(
1991
).
42.
S. C.
Glotzer
,
V. N.
Novikov
, and
T. B.
Schrøder
, “
Time-dependent, four-point density correlation function description of dynamical heterogeneity and decoupling in supercooled liquids
,”
J. Chem. Phys.
112
,
509
512
(
2000
).
43.
N.
Lačević
,
F. W.
Starr
,
T. B.
Schrøder
,
V. N.
Novikov
, and
S. C.
Glotzer
, “
Growing correlation length on cooling below the onset of caging in a simulated glass-forming liquid
,”
Phys. Rev. E
66
,
030101
(
2002
).
44.
N.
Lačević
,
F. W.
Starr
,
T. B.
Schrøder
, and
S. C.
Glotzer
, “
Spatially heterogeneous dynamics investigated via a time-dependent four-point density correlation function
,”
J. Chem. Phys.
119
,
7372
7387
(
2003
).
45.
L.
Berthier
, “
Time and length scales in supercooled liquids
,”
Phys. Rev. E
69
,
020201
(
2004
).
46.
C.
Toninelli
,
M.
Wyart
,
L.
Berthier
,
G.
Biroli
, and
J.-P.
Bouchaud
, “
Dynamical susceptibility of glass formers: Contrasting the predictions of theoretical scenarios
,”
Phys. Rev. E
71
,
041505
(
2005
).
47.
L.
Berthier
,
G.
Biroli
,
J.-P.
Bouchaud
,
W.
Kob
,
K.
Miyazaki
, and
D. R.
Reichman
, “
Spontaneous and induced dynamic fluctuations in glass formers. I. General results and dependence on ensemble and dynamics
,”
J. Chem. Phys.
126
,
184503
(
2007
).
48.
L.
Berthier
,
G.
Biroli
,
J.-P.
Bouchaud
,
W.
Kob
,
K.
Miyazaki
, and
D. R.
Reichman
, “
Spontaneous and induced dynamic correlations in glass formers. II. Model calculations and comparison to numerical simulations
,”
J. Chem. Phys.
126
,
184504
(
2007
).
49.
S.
Karmakar
,
C.
Dasgupta
, and
S.
Sastry
, “
Growing length and time scales in glass-forming liquids
,”
Proc. Natl. Acad. Sci. U. S. A.
106
,
3675
3679
(
2009
).
50.
K.
Kim
and
S.
Saito
, “
Multiple length and time scales of dynamic heterogeneities in model glass-forming liquids: A systematic analysis of multi-point and multi-time correlations
,”
J. Chem. Phys.
138
,
12A506
(
2013
).
51.
T.
Pal
and
R.
Biswas
, “
Slow solvation in ionic liquids: Connections to non-Gaussian moves and multi-point correlations
,”
J. Chem. Phys.
141
,
104501
(
2014
).
52.
E.
Lindahl
,
B.
Hess
, and
D.
van der Spoel
, “
GROMACS 3.0: A package for molecular simulation and trajectory analysis
,”
J. Mol. Model.
7
,
306
317
(
2001
).
53.
H.
Berendsen
,
D.
van der Spoel
, and
R.
van Drunen
, “
GROMACS: A message-passing parallel molecular dynamics implementation
,”
Comput. Phys. Commun.
91
,
43
56
(
1995
).
54.
D.
Van Der Spoel
,
E.
Lindahl
,
B.
Hess
,
G.
Groenhof
,
A. E.
Mark
, and
H. J. C.
Berendsen
, “
GROMACS: Fast, flexible, and free
,”
J. Comput. Chem.
26
,
1701
1718
(
2005
).
55.
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
).
56.
K. G.
Bogolitsyn
,
T. E.
Skrebets
, and
T. A.
Makhova
, “
Physicochemical properties of 1-butyl-3-methylimidazolium acetate
,”
Russ. J. Gen. Chem.
79
,
125
128
(
2009
).
57.
CRC Handbook of Chemistry and Physics
, 95th ed., edited by
W. M.
Haynes
(
CRC Press
,
2014
).
58.
59.
K.
Vanommeslaeghe
,
E.
Hatcher
,
C.
Acharya
,
S.
Kundu
,
S.
Zhong
,
J.
Shim
,
E.
Darian
,
O.
Guvench
,
P.
Lopes
,
I.
Vorobyov
, and
A. D.
MacKerell
, Jr.
, “
CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force field
,”
J. Comput. Chem.
31
,
671
690
(
2010
).
60.
K.
Vanommeslaeghe
and
A. D.
MacKerell
, Jr.
, “
Automation of the CHARMM General Force Field (CGenFF) I: Bond perception and atom typing
,”
J. Chem. Inf. Model.
52
,
3144
3154
(
2012
).
61.
K.
Vanommeslaeghe
,
E. P.
Raman
, and
A. D.
MacKerell
, Jr.
, “
Automation of the CHARMM General Force field (CGenFF) II: Assignment of bonded parameters and partial atomic charges
,”
J. Chem. Inf. Model.
52
,
3155
3168
(
2012
).
62.
A. M.
Nikitin
and
A. P.
Lyubartsev
, “
New six-site acetonitrile model for simulations of liquid acetonitrile and its aqueous mixtures
,”
J. Comput. Chem.
28
,
2020
2026
(
2007
).
63.
Y.
Shim
,
M. Y.
Choi
, and
H. J.
Kim
, “
A molecular dynamics computer simulation study of room-temperature ionic liquids. I. Equilibrium solvation structure and free energetics
,”
J. Chem. Phys.
122
,
044510
(
2005
).
64.
A.
Rahman
,
K. S.
Singwi
, and
A.
Sjölander
, “
Theory of slow neutron scattering by liquids. I
,”
Phys. Rev.
126
,
986
996
(
1962
).
65.
J.
Colmenero
,
F.
Alvarez
, and
A.
Arbe
, “
Self-motion and the α relaxation in a simulated glass-forming polymer: Crossover from Gaussian to non-Gaussian dynamic behavior
,”
Phys. Rev. E
65
,
041804
(
2002
).

Supplementary Material