Molecular Dynamic Simulations and Vibrational ... - ACS Publications

Aug 5, 2013 - A complete list of force field parameters, autocorrelation functions for simulations at ...... Painter, P. C.; Zhao, H.; Park, Y. Vibrat...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCB

Molecular Dynamic Simulations and Vibrational Analysis of an Ionic Liquid Analogue Sasha L. Perkins, Paul Painter,* and Coray M. Colina* Department of Materials Science and Engineering, Pennsylvania State University, University Park, Pennsylvania 16802, United States S Supporting Information *

ABSTRACT: Deep eutectic solvents, considered ionic liquid (IL) analogues, show promise for many material science and engineering applications over typical ILs because they are readily available and relatively inexpensive. Atomistic molecular dynamics simulations have been performed over a range of temperatures on one eutectic mixture, 1:2 choline chloride/urea, using different force field modifications. Good agreement was achieved between simulated density, volume expansion coefficient, heat capacity, and diffusion coefficients and experimental values in order to validate the best performing force field. Atom−atom and center-of-mass radial distribution functions are discussed in order to understand the atomistic interactions involved in this eutectic mixture. Experimental infrared (IR) spectra are also reported for choline chloride−urea mixtures, and band assignments are discussed. The distribution of hydrogen-bond interactions from molecular simulations is correlated to curve-resolved bands from the IR spectra. This work suggests that there is a strong interaction between the NH2 of urea and the chlorine anion where the system wants to maximize the number of hydrogen bonds to the anion. Additionally, the disappearance of free carbonyl groups upon increasing concentrations of urea suggests that at low urea concentrations, urea will preferentially interact with the anion through the NH2 groups. As this concentration increases, the complex remains but with additional interactions that remove the free carbonyl band from the spectra. The results from both molecular simulations and experimental IR spectroscopy support the idea that key interactions between the moieties in the eutectic mixture interrupt the main interactions within the parent substances and are responsible for the decrease in freezing point.



INTRODUCTION The use of ionic liquids (ILs) has spread over the past two decades to include a wide range of applications as a result of their negligible vapor pressures, low melting points, and unusual solvent characteristics.1,2 Unfortunately, large-scale applications of these unusual solvents have been limited by their cost. Ionic liquid analogues, or deep eutectic solvents (DES), have emerged as more inexpensive alternatives to typical ILs while still showing similar properties. DES are a eutectic mixture most often consisting of a quaternary ammonium salt and a hydrogen-bond donor.2,3 In contrast to typical ILs that are made of discrete bulky ions, DES systems are interesting in that the components of the mixture have high melting points in their pure state and become liquid at room temperature after they are mixed. For example, choline chloride has a melting point of 302 °C and urea has a melting point of 133 °C. When urea and choline chloride are mixed at different molar ratios, the melting point of the mixture is depressed according to the amount of hydrogen-bond donor (HBD) present with a minimum at about 12 °C with 67 mol % of urea.4 Several review articles2,3,5,6 have been published recently highlighting the wide variety of combinations of HBD and quaternary ammonium salts that can be used to make DES with exceptional properties that can be used in many applications such as electrochemistry,7−9 catalysis,10−12 polymer synthesis,5 gas solubility,13 and purifications14 just to name a few. However, compared to typical ILs, there is a lack of experimental and © XXXX American Chemical Society

theoretical data on fundamental thermodynamic and transport properties for this family of solvents. Moreover, binary and ternary mixtures that form these eutectic mixtures are potentially numerous and are not straightforward to predict.15 Choline chloride is widely used to form DES as it is produced commercially in megaton quantities.2 Several choline chloridebased eutectic mixtures16−20 have been explored experimentally. For the choline chloride/urea system, it has been shown through fast atom bombardment mass spectroscopy (FAB MS) and NMR4,7 that the interactions involve both one and two urea molecules with one chlorine anion. It has been proposed2,19 that the depression in freezing temperature of these systems is related to several factors: (a) increased interactions between the anion and the HBD, (b) a decrease in the lattice energies of the systems as a result of the presence of bulky ions, and (c) the increase in entropy that occurs upon mixing. These factors can be further explored using the atomistic detail provided by molecular simulations. Molecular simulations complement experimental findings by predicting thermodynamic and transport properties as well as by adding atomistic understanding on the interactions that provide DES with their properties. ILs have been studied extensively Received: May 9, 2013 Revised: August 2, 2013

A

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Table 1. Force Field Parameters Studied in This Work type

electronic model (net charge for ionic species)

1Amb 0.9Amb 0.8Amb Mod1 Mod2

HF/6-31G*//HF/6-31G* (±1e) HF/6-31G*//HF/6-31G* (±0.9e) HF/6-31G*//HF/6-31G* (±0.8e) HF/6-31G*//HF/6-31G* (±1e) HF/6-31G*//B3LYP/6-31G* (±1e)

bonded terms all GAFF default values all GAFF default values all GAFF default values Morrow and Maginn for choline intramolecular parameters;49 Ozpinar et al. for urea intramolecular parameters50 ro and θo from geometry optimization; Morrow and Maginn Ch+ other intramolecular parameters;49 Ozpinar et al. urea other intramolecular parameters50

through molecular simulations,21 while DES simulations are just starting to emerge. Electrostatic effects and hydrogen-bonding interactions have been a particular focus.22−26 However, until very recently, less than a handful of computational analyses had been conducted on choline chloride-based eutectic solvents.27,28 Sun et al.28 examined the structure of choline chloride/urea systems at different molar ratios at the critical point of the eutectic mixture, 285 K, to probe structural differences, hydrogen-bond lifetimes, and interaction energies as a function of composition. From the experimental aspect, Yue et al.29 and Deng et al.30 performed some general band assignments of the infrared spectrum as this is a technique that is very sensitive to hydrogen bonding. In this work, a combination of both computational and experimental techniques is used to explore the interactions in these systems in order to gain a more thorough understanding of their behavior. Force field modifications are validated by comparing simulated density, thermodynamic, and transport properties of these mixtures to experimental values, where available, as well as a thorough hydrogen-bond analysis involving the geometry of the hydrogen bond and the interactions of functional groups involved. The results of the computational study and the experimental analysis are then compared in order to establish what main interactions occur within this DES and how they can affect the structure and, hence, the freezing point of this DES.

structure and transport properties in the simulations of DES, a technique recently implemented in typical ILs.24,26,40−43 Reduced charges of ±0.9e and ±0.8e for choline chloride were arbitrarily chosen on the basis of reduced values observed in the literature.43,44 For a choline chloride and urea system, several simulation boxes were created at low density with different initial configurations using a modified version of the POLYMATIC packing code45 with at least 500 urea molecules and 250 choline chloride ion pairs. Molecular dynamics (MD) simulations were performed using AMBER 10 and 12.36,46−48 The parameters used in this study are shown in Table 1 and in Tables S1−S8 of the Supporting Information. All force fields used the GAFF Lennard-Jones parameters. For all properties studied, three simulation boxes were averaged at each temperature. Simulation boxes were minimized using steepest descent method for 10 cycles, and then conjugate gradient method was used for 10 000 cycles. After minimization, the system was subject to a compression and decompression scheme51 that uses isothermal-isobaric (NPT) and canonical (NVT) ensemble cycles at different temperatures and pressures for at least 3 ns. This methodology allows for proper and efficient equilibration because of the slow dynamics of ILs and DESs. This scheme is one possible method that our group has used to successfully predict equilibrium properties for amorphous polymers and also efficiently works for these high viscosity systems. Alternatively, some authors (Liu et al.40 is an example) use an annealing process of raising and lowering the temperature to reach equilibrium properties faster than equilibrating at ambient conditions because of the sluggish dynamics of ILs. For NPT equilibration cycles and production runs, an isotropic position scaling with pressure relaxation times of 1 ps was used. The Langevin dynamics thermostat was used with a collision frequency of 5 ps−1 in order to ensure accurate temperature control.49,52 The SHAKE algorithm was applied to perform bond length constraints on all bonds attached to hydrogen atoms in the system. Periodic boundary conditions were applied to all simulations at 0.98 bar and 298, 315, and 330 K. A nonbonded interaction cutoff of 15 Å was used. The particle-mesh Ewald (PME) method was used to calculate long-range electrostatic interactions. NPT production runs of at least 20 ns after equilibration were used for average thermodynamic properties and structural analysis. One nanosecond trajectories with 2 ps between frames were used for hydrogen-bond analysis. The microcanonical (NVE) ensemble was used to study transport properties for 50 and 30 ns at 298 and 330 K, respectively. For higher temperatures, shorter times could be simulated since the diffusive regime was approached faster compared to lower temperatures. Smaller time steps and stricter PME and SHAKE tolerances of 10−6 and 10−8, respectively, were used for the NVE runs to properly conserve energy in the system and to obtain diffusion coefficients. A step size of 2 fs was used for NPT



METHODS A. Computational Details. Gaussian 0331 and R.E.D. Tools (along with the R.E.D. Server)32−34 were used to optimize the geometries of the isolated ions and molecules. For the choline cation, two conformation charge derivations were performed except in Mod2 where only one single conformation (optimized with a very tight convergence) was used since specific geometries were applied to this modified force field as shown in Table 1. Images of the multiconformation configurations of choline used are shown in Figure S1 of the Supporting Information. A single conformation was used for the urea molecule similar to that performed in the R.E.D. tools library.34 A vibrational frequency analysis was conducted in order to determine if a local minimum was reached. The molecular electrostatic potential on the isolated ion or the molecule optimized geometry was obtained at the HF/ 6-31G* level of theory, and partial charges were obtained using the restrained electrostatic potential (RESP) charge derivation method.35−38 The generalized amber force field (GAFF) functional form and parameters,39 also recently used to simulate other typical ILs,40,41 were used for these simulations, and partial charges obtained through ab initio calculations were added in hopes to provide a reliable alternative to the more computationally expensive polarizable force fields. Unity and scaled partial charges (implicitly including charge transfer and polarizability) were compared in order to study their effect on B

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

simulations, while a 1 fs step size was used for the subsequent NVE simulations. B. Experimental Section. Choline chloride and urea mixtures were formed by mixing choline chloride (Sigma Aldrich) and urea (IBI Scientific) in a 1:2 molar ratio. The mixture was stirred on a hot plate at around 80 °C for a few minutes until a clear solution was formed following the procedure described by Abbott et al.4 Adding water to the solid mixture, instead of applying heat, was also found to be a useful way to form these eutectic mixtures. The water can subsequently be evaporated in a vacuum oven for several hours prior to analysis. The DES liquid was then placed on polished KBr windows in order to record infrared spectra using a Nicolet 6700 Fourier transform infrared (FTIR) spectrometer. The spectra were obtained at a resolution of 2.0 cm−1 using 200 interferometer scans. Spectra of the solid samples were obtained using diffuse reflectance Fourier transform (DRIFT) on samples ground with 200 mg of KBr. Small amounts (∼1−3 mg) of the solid material were used to obtain these spectra.

The biggest differences among simulated densities observed in Figure 1 are due to the decrease in net charges of the ionic species although some slight changes are observed within unity charge force field modifications. As can be seen, reducing the net charge on the ionic species reduces the density, a trend also reported in the literature43 for certain ionic liquids, since decreasing the ionic charges expands the system because of lower cohesive energies. Density, although not a determining factor for force field validation,21 is a very important solvent physical property and thus serves as a starting point. Volume expansivity was calculated using eq 1 and was compared to experimental values. The volume expansivity was calculated from the slope of a molar volume versus temperature curve

αP =



1 ⎛ ∂V ⎞ ⎜ ⎟ V ⎝ ∂T ⎠

(1)

where V is the molar volume and T is the temperature. Experimental volume expansivities53 for choline chloride eutectic mixtures are in the range of (5−6) × 10−4 K−1, and thus we found that the 0.8Amb reduced charge force field performs best as shown in Table 2. It is

RESULTS AND DISCUSSION A. Computational Studies. Five different force field modifications were used, including the effect of charge transfer and polarization characteristics of DES to properly model structure and thermodynamic properties. Force Fields Validation. To validate the models, densities, volume expansivities, molar heat capacities, and self-diffusion coefficients were calculated and compared with the limited experimental data that is available. In Figure 1, the experimental

Table 2. Volume Expansivity for the Different Force Field Modifications Studied temperature range 298−330 K force field

volume expansivity αP × 104 (K−1)

Cres (J/mol K)

Cideal (J/mol K)

1Amb 0.9Amb 0.8Amb Mod1 Mod2 exp

2.74 4.16 5.32 2.69 2.88 5.0−6.0

37 64 76 33 46

108 108 108 108 108

CP (J/mol K) 145 172 184 141 154 181.4−186.4

interesting to observe that the volume expansion coefficients are smaller than those reported for organic liquids.56 Heat capacity and other thermodynamic properties of DESs and ILs can provide insight into solvent properties, such as for heat transfer,57 and also can provide insight into the origin of the low melting points of ILs.58 Experimental heat capacity data for DESs are scarce; thus, the only available data set57 was used for validation. Following Cadena and co-workers,59,60 the molar heat capacities were estimated by separating heat capacities into residual and ideal contributions. For the residual contribution, heat capacities and enthalpies were assumed to have a linear temperature relationship between 298 and 330 K given by

Figure 1. Density for 1:2 choline chloride urea eutectic mixture at 0.98 bar. Experimental values: •, Leron and Li;18 ▼, Ciocirlan et al.;53 ▲, Abbott et al.;54 ■, D’Agostino et al.16 Simulated values: Δ, 1Amb; ▽, 0.9Amb; ◊, 0.8Amb; ○, Mod1; □, Mod2. For simulated data, open symbols are connected by lines for clarity.

⎛ ∂H res ⎞ ⎟ CPres(T , P) = ⎜ ⎝ ∂T ⎠ P

densities16,18,53,54 of the 1:2 choline chloride/urea eutectic mixture are compared to the results of the five models. The Ciocirlan et al. values of density were plotted on the basis of a best fit line shown in Figure 1 of ref 53. A standard deviation of less than 0.003 g/cm3 between simulation boxes was observed for all densities studied but is not shown in the figure for clarity. These simulated density values were predicted by the simulation protocol used and were not fixed to the experimental densities. The results show that all force field modifications used in this work produce density values within 1−6% of the experimental values. There are differences in reported densities between experimental sources. This might be related to water content of the samples or the methods used to measure density. Choline chloride, for example, is known to be a very hygroscopic substance and can readily absorb water upon exposure to the atmosphere.2,55

(2)

Residual heat capacities were obtained from liquid molecular simulations at 298, 315, and 330 K and 0.98 bar as H res = U NB + PV − NAkT

(3)

NB

where U is the intermolecular potential energy (including 1−4 intramolecular nonbonded interactions), P is the pressure, NA is Avogrado’s number, and k is Boltzmann’s constant. The enthalpies were plotted against temperature, and the slope is the average residual heat capacity from 298 to 330 K. As mentioned by Cadena et al., ideal heat capacities can theoretically be obtained from these simulations, but these are overestimated for all force field modifications studied here C

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Figure 2. Average MSD as a function of time for one simulation box using 0.8Amb at (a) 298 K and (b) 330 K. Red, urea; blue, chlorine; orange, choline.

showing that some refinement of the bonded parameters should be applied. This overestimation has also been noted for other IL simulations.60 The ideal heat capacities were obtained after performing a vibrational frequency analysis on the isolated choline chloride and urea moieties using B3LYP/6-31G* at the conditions of interest and summing according to their respective mole fractions.61 No scaling factor was applied to the ideal heat capacity, although other groups have appplied a scale factor for vibrational frequencies performed at lower levels of theory.40 The average simulated heat capacities at constant temperature and pressure are reported for temperatures between 298 and 330 K and are presented in Table 2. A heat capacity between 141 and 184 J/(mol K) was calculated, depending on the force field modification, where a mol is a mole of DES with molar mass of 86.58 g/mol. Force field modifications with reduced charges gave values in good agreement with reported experimental values of 181.4−186.4 J/(mol K) for temperatures between 303.15 and 333.15 K.57 The values reported by the same authors57 for a different system (choline chloride/glycerol) are ∼50 J/(mol K) higher than those reported by another study.62 The importance of measuring transport properties for these eutectic mixtures has also been stressed in the literature.2 The diffusion mechanism of DES has been thought to be similar to those of ionic liquids with discrete ions,16 and thus similar procedures for calculating self-diffusion coefficients have been carried out here. Self-diffusion coefficients were calculated using the Einstein relation D=

1 d lim ⟨|ri(t ) − ri(0)|2 ⟩ 6 t →∞ dt

β (t ) =

d log10⟨Δr(t )2 ⟩ d log10 t

(5)

where Δr(t)2 is the average MSD of the species and t is time. If β < 1, the system is in the subdiffusive regime. At β =1, the calculated diffusion coefficient is in the diffusive regime, and thus, self-diffusion coefficients were taken in this region. For simulations using 0.8Amb at 298 K, β parameters approached 1 after 30 ns in contrast to simulations at 330 K, where unity values were reached after about 10 ns because of reduced viscosities at higher temperature. As seen in Figure 2, urea has a higher self-diffusion coefficient than choline consistent with findings in the literature,16 which indicate that the larger, higher molecular weight cation moves more slowly than the smaller hydrogen bond donor. Compared to experimental values, simulations at 298 K underestimate self-diffusivities from 25 to 51%, which may mean that still even longer simulations in the diffusive regime are required to achieve more accurate transport values at this temperature. However, excellent agreement with experimental values is seen at 330 K as shown in Table 3.16 Anion Table 3. Self-Diffusion Coefficients at 298 and 330 K for Different Simulation Boxes Using 0.8Amba temperature (K) 298 K

(4)

330 K

where i is the species and r is the vector coordinate of the atoms. The average mean square displacements (MSD) for choline, chlorine, and urea moieties in the eutectic mixture are plotted over time for trajectories obtained in the NVE ensemble using ptraj from AmberTools.46 To ensure that the simulations are in the diffusive regime, time autocorrelation functions were obtained using a vector from the nitrogen atom in choline to the oxygen atom in choline. For urea, the vector representing the carbonyl was used. Since 0.8Amb has provided the best values for density, thermal expansion coefficients, and heat capacity, 0.8Amb was used to validate self-diffusion coefficients for this DES. To determine an appropriate diffusive regime, autocorrelation curves at 298 and 330 K for 0.8Amb were analyzed and are shown in Figures S2 and S3 of the Supporting Information. The beta-parameter (β) was also calculated, as used by Del Pópolo and Voth63 to determine the location of the diffusive regime in other ILs, as given by

a

Sim1 Sim2 Sim3 experimental16 Sim1 Sim2 Sim3 experimental16 (328 K)

D+ (1011 m2/s)

D− (1011 m2/s)

DHBD (1011 m2/s)

0.17 (0.05) 0.21 (0.06) 0.26 (0.25) 0.35 2.18 (0.42) 1.80 (0.67) 1.73 (0.12) 2.10

0.25 (0.06) 0.26 (0.09) 0.34 (0.16)

0.39 (0.06) 0.43 (0.15) 0.47 (0.14) 0.66 3.67 (0.33) 3.26 (0.46) 3.31 (0.39) 3.55

2.65 (0.36) 2.88 (0.55) 2.59 (0.27)

Standard deviation in parentheses.

self-diffusion coefficients were also obtained, which have not to our knowledge been previously reported. This provides more information about the complexed anion’s transport behavior because self-diffusion coefficients are not the same for the hydrogen bond donor and the anion. This, in turn, implies that although there are strong hydrogen bonding interactions between the anion and urea, as discussed above, their motions are not tied, and urea molecules have greater mobility relative to the chlorine anion. Structure Comparisons. Coordination numbers are listed in Table 4 for cations and urea molecules around the anion. The area under the curve used for the estimation of the coordination number, shown in Table 4, was taken up to the first minima of each D

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

It is interesting that when using these force fields modifications, the urea/urea NH2/CO hydrogen bond distances are shorter than anion-NH2 hydrogen bonds. Similar RDFs were observed using all force fields, except Mod2, but only curves for 0.8Amb are shown in Figure 3. Additionally, the urea− urea interaction in our force field modifications is greater than that shown very recently by Sun et al.28 primarily because of the force field parameters used for urea although their work also found two hydrogen bond distances for the NH2--Cl and NH2-O (urea) interactions. Work to validate the force field modifications has been carried out in this study in order to ensure that the system is accurately simulated since these systems have not been thoroughly studied in-silico. Hydrogen Bond Analysis. A hydrogen bond analysis was performed for a 1 ns trajectory at the end of the production run using cpptraj from AmberTools 12 with 0.8Amb.46 The angles (where ∠DHA is the angle D−H--A when D is hydrogen donor and A is hydrogen acceptor) and distances between the anion and hydrogen atoms in urea and the hydroxyl group in choline were investigated. From an angle and distance distribution analysis, shown in Figure S7 of the Supporting Information, the heavy-to-heavy atom cutoff distance of 3.5 Å and an angle cutoff for ∠DHA of 150° were used as a strict hydrogen bond criterion in order to calculate the fraction of hydrogen bonds of each type in this system. These criteria fit into the typical hydrogen bond criteria used for other studies, including IL simulations.25,64,65 The relative fraction of the hydrogen bonds of each type was determined for every frame of a 1 ns molecular dynamics trajectory with 2 ps between frames (a total of 501 frames) also using cpptraj from AmberTools 12.46 Two types of NH groups are distinguished on each urea molecule, Htrans and Hcis, as illustrated in Figure 4 and labeled H1, H3 and H2, H4, respectively. This figure also illustrates the pattern of hydrogen bonding found in the urea crystal.66 Note the bifurcated hydrogen bonds formed between the Htrans and urea carbonyl. Figure 5a and b shows the fraction of N−H--OC and N− H--Cl− hydrogen bonds, respectively, with standard deviations shown in the Supporting Information. The data were obtained from three different simulations all under the same conditions and with the same force field (0.8Amb) but with different starting configurations. Curves are plotted as a function of frame number and show that the values remain fairly constant as the simulation progresses. Each frame is a snapshot of the number of hydrogen bonds present at that instant. The values shown in Figure 5a and b are consistent with the RDF curves, which show that there are more Hcis around the urea carbonyl than Htrans. On the other

center of mass (COM) radial distribution function (RDF) curve shown in Figures S4 and S5 of the Supporting Information (minima values are also listed in Table 4). Mod1 has a lower coordination Table 4. Average Number of Nearest Neighbors Up to First Minima (Using COM RDF)a first min position (Å) for force field Cl−Ch+ 1Amb 0.9Amb 0.8Amb Mod1 Mod2 a

4.78 4.88 4.88 4.83 4.83

Ch+ coordination number

first min position (Å) for Cl−urea

urea coordination number

1.42 (0.03) 1.61 (0.03) 1.54 (0.01) 1.48 (0.03) 1.41 (0.04)

5.18 5.38 5.48 4.13 5.13

3.08 (0.04) 3.88 (0.04) 3.87 (0.04) 1.44 (0.02) 3.67 (0.03)

Standard deviation in parentheses.

number because of the shoulder present at 4.13 Å. If longer r values are chosen similar to the other models, comparable coordination values are obtained. FAB MS measurements4 have provided values for the number of anions involved in complexing interactions, but coordination numbers were not measured, and cation−anion interactions are not detected using this technique. The simulations reported here indicated that a larger number of urea molecules interact with the anion compared to the choline cation with three to four urea molecules being located around chlorine anions relative to one to two choline cations. As might be expected, this is a consequence of molar composition of the system and hydrogen-bonding interactions, which we will consider in the next section. Figure 3a shows the atom−atom radial distribution function (RDF) for specific urea hydrogen interactions with Cl− anions. Hydrogen atoms were labeled on the basis of the positions at the beginning of the simulation as Hcis for atoms originally in the cis configuration and as Htrans for the trans configuration. The majority of the hydrogen atoms maintained their original configuration throughout the simulation. These RDFs show that Htrans is more likely to hydrogen bond to anions than Hcis. These configurations will be illustrated and discussed in more detail later. Figure 3b shows the RDFs for the NH2 hydrogen atoms in urea relative to urea carbonyl oxygen atoms. As with NH/anion interactions, the first peak of both cis and trans NH hydrogen atoms is at identical distance from the urea oxygen atom. The tallest peaks for the RDFs shown in Figure 3b are located near 2.4 Å (Hcis) and 3.0 Å (Htrans) corresponding to non-hydrogenbonded intramolecular distances, while the first peak, near 1.9 Å, corresponds to intermolecular hydrogen bond distances.

Figure 3. Atom−atom RDFs for (a) anion and hydrogen atoms in urea and (b) oxygen in urea and hydrogen atoms in urea for 0.8Amb. Blue dots and red line are for Htrans; gray line and green dots are for Hcis. E

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Figure 6. Average fraction of H-bonds of the type OH--X for all frames studies. No error bars are shown for clarity. Red: XCl− ; blue: Xurea CO; green: Xcholine OH group.

Figure 7 and illustrates the positioning of urea molecules around a Cl− ion such that bifurcated hydrogen bonds are formed. Also note the positioning of a Hcis such that it forms a hydrogen bond with a urea CO group. Figure 6 shows the fraction of hydrogen bonds formed between the choline OH group and the urea carbonyl, the Cl− anion, and other choline OH groups. There are very few OH-OH hydrogen bonds as most of the hydroxyls prefer to hydrogen bond to the urea carbonyl and the Cl− anion. However, a small number of Hcis groups were calculated to form hydrogen bonds to the oxygen of the choline cation (about 1%, plot not shown). Plots with average values and standard deviations are shown in Figures S8−S9 of the Supporting Information. The number of hydrogen bonds that can form between the donor and acceptor groups present in the mixture depends in a complex fashion on both the energy of each type of hydrogen bond and the ability of the molecules and ions to pack in such a fashion so as to maximize the total number of hydrogen bonds. The infrared spectroscopic results described in the following section show that in the 1:2 molar mixture (choline chloride:urea) there are no detectable free (non-hydrogenbonded) NH, OH, and carbonyl groups although at lower urea concentrations free carbonyl groups are observed. We will now consider these results. B. Experimental Spectroscopic Studies. Infrared spectroscopy is very sensitive to hydrogen-bonding interactions, and for choline chloride/urea mixtures, the most useful regions of the spectrum are the NH/OH stretching modes observed between 3600 and 3000 cm−1 and the NH2 bending and CO stretching bands observed between 1700 and 1600 cm−1. These two regions of the spectrum are shown in Figure 8a and b, respectively, where the spectrum of the DES mixture (1:2 molar ratio of choline chloride to urea) is compared to the spectra of urea and choline chloride. It can be seen that there are significant changes in both regions of the spectrum upon formation of the DES, but to interpret these, it is necessary to first consider band assignments in the spectra of the parent materials, urea and choline chloride. A detailed vibrational analysis of urea (as well as various urea complexes) has been reported by Keuleers et al.,67,68 who also comprehensively reviewed much of the preceding literature. They assigned bands near 3450 and 3345 cm−1 to the asymmetric and symmetric NH stretching modes, respectively. Because there are two NH2 groups per molecule, each of these modes could be split as a result of interactions into vibrations belonging to the A1 and B2 symmetry species of the C2v point group.67 In this regard, Derreumaux et al.69 performed a symmetry analysis of urea and

Figure 4. Hydrogen bonds formed between carbonyl and NH groups in urea. Bifurcated (linear, l) hydrogen bonds are formed between NH groups that are trans (t) to carbonyls, while NH groups that are cis form transverse (t) hydrogen bonds. Htrans groups are labeled 1, 3, while Hcis groups are labeled 2, 4. Hydrogen atoms are in white, carbon atoms are in cyan, nitrogen atoms are in blue, and oxygen atoms are in red.

Figure 5. Average fraction of H-bonds of the type (a) NH--OC and (b) NH--Cl− for all frames studied. No error bars are shown for clarity. Red: XH1; gray: XH2; blue: XH3; green: XH4 (see Figure 4).

hand, there are more Htrans than Hcis atoms around the Cl− anion. As we show in the Experimental Spectroscopic Section, hydrogen bonds between NH--Cl− hydrogen bonds are stronger than NH--OC hydrogen bonds, and so the former are preferred. In addition, the trans NH groups can form a bifurcated double hydrogen bond as in the urea crystal (see Figure 4), and so trans NH--Cl− hydrogen bonds are favored over cis. A snapshot of the groups surrounding a Cl− anion is shown in F

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Figure 7. Two snapshots showing examples of clusters surrounding a chlorine anion. Hydrogen atoms are in white, carbon atoms are in cyan, nitrogen atoms are in blue, chlorine is in pink, and oxygen atoms are in red.

Figure 8. Infrared spectra for choline chloride, urea, and the DES (a) NH/OH region and (b) CO/NH2 region.

established the correlation between the C2v point group and the space group showing that there are still two symmetric and asymmetric stretching modes allowed in the crystal. Indeed, the spectra in Figure 8a clearly show two asymmetric stretching modes near 3458 cm−1 and 3440 cm−1, while the 3345 cm−1 symmetric stretching mode has a clearly discernible shoulder near 3332 cm−1. Derreumaux et al.69 calculated only a small frequency difference of ∼2 cm−1 between the asymmetric A1 and B2 and ∼4 cm−1 for the symmetric A1 and B2 modes compared to the larger splittings experimentally observed in this study. We suggest that these larger splittings have their origin in a difference in the types of hydrogen bonds formed by NH groups that are cis and trans to carbonyl groups in the same molecule. The crystal structure of urea shows that NH groups that are trans to the carbonyl groups form what was termed linear hydrogen bonds labeled l in the schematic illustration shown in Figure 4.66 Both trans NH groups hydrogen bond to the same carbonyl forming a bifurcated structure and a chain of urea molecules linked in this fashion. The two cis NH groups hydrogen bond transversely to carbonyl groups on different chains of urea molecules (t in Figure 4). As a result, the carbonyl oxygen on each urea molecule is hydrogen bonded to four NH groups on three different urea molecules. It has been shown that the linear and transverse H-bonds have different lengths66 and, hence,

different strengths. Accordingly, in crystalline urea, the observed large splitting of the asymmetric and symmetric NH stretching modes could be related to perturbations associated with different hydrogen bond strengths. There are also several bands below 3300 cm−1 in the spectra of urea in Figure 8a, but these can be assigned to overtone/combination modes of the CO stretch and NH2 bending modes probably in Fermi resonance with the NH stretching modes. Keuleers et al.67 assign the band at 3264 cm−1 to a coupled combination mode of NH2 deformation and CO stretching vibrations. Urea has three bands in the 1700−1600 cm−1 region of the spectrum observed near 1689, 1631, and 1606 cm−1. The frequencies of these modes are slightly different from those reported in other studies29,67,69 presumably because we used a different sample preparation method, diffuse reflection, as opposed to KBr pellets. These bands have been assigned to symmetric and asymmetric NH2 deformation vibrations and a carbonyl stretching vibration, respectively. However, as Keuleers et al.67 point out, the symmetric (1689 cm−1) NH2 deformation and CO stretching modes (1606 cm−1) are highly coupled or mixed. Both belong to the same symmetry species (A1), but the asymmetric NH2 deformation belongs to a different symmetry species (B2) and does not appear to mix with the CO stretching mode. G

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Choline chloride also has an interesting spectrum with the OH stretching mode appearing near 3256 cm−1 (Figure 8a). This is consistent with the spectroscopic observations of Harmon and co-workers in studies of organic salts70 and choline fluoride,71 where a band near 3223 cm−1 was observed. Typically, OH--OH hydrogen-bonded groups in alcohols appear near 3350 cm−1, and so the much lower frequency of the OH stretching mode in choline chloride indicates that it is engaged in much stronger hydrogen bonds. The crystal structure of choline chloride, although difficult to interpret because of disorder effects, shows a close proximity of OH groups to the Cl− anion indicating that these groups are strongly hydrogen bonded.55 Upon forming the choline chloride/urea 1:2 (molar) mixture, there are significant changes in the NH2/OH stretching and NH2/CO bending regions of the spectrum as shown in Figure 8a and b, respectively. In the NH2/OH stretching region, two wellresolved bands can be observed near 3326 cm−1 and 3192 cm−1 with two prominent shoulders near 3256 cm−1 and 3400 cm−1, respectively. The 3192 cm−1 band is probably an overtone of the 1616 cm−1 band that is in Fermi resonance with a NH stretching mode resulting in an intensity enhancement and a shift to lower wavenumbers analogous to the amide A and B modes observed in polypeptides and proteins.72 The 3256 cm−1 band corresponds to the OH stretching band observed in crystalline choline chloride and presumably has the same origin of OH--Cl− hydrogen-bonded groups. The simulations reported above indicate that roughly half of the OH groups are hydrogen bonded to Cl− anions, while the other half are hydrogen bonded to urea CO. The OH stretching mode of the latter would appear at higher wavenumbers. However, it would seem likely that the 3326 cm−1 and 3400 cm−1 bands are due to the symmetric and asymmetric stretching modes of the urea NH2 groups, respectively. These are at lower frequencies than the corresponding modes in crystalline urea indicating that they are involved in stronger hydrogen bonds. Also, they do not appear to be split into two components as in the spectra of crystalline urea and are very broad suggesting that there is an equivalently broad distribution of hydrogen bonds with different hydrogen bond strengths. We suggest that the OH--OC hydrogen-bonded O−H stretching mode is buried in this profile. This conclusion is supported by the changes that occur in the CO stretching, NH2 bending region of the spectrum shown in Figure 8b. As discussed above, urea has three bands in this region of the spectrum observed near 1689, 1631, and 1606 cm−1. Upon forming the 1:2 deep eutectic mixture with choline chloride, these three bands are merged into what at first appears to be two major peaks near 1668 cm−1 and 1616 cm−1 initially suggesting that the groups are in a hydrogen-bonding environment with patterns of NH to CO and NH to Cl hydrogen bonds that are fairly uniform with some sort of strange merging of the NH2 deformation and C O stretching modes. However, a more detailed examination of this region of the spectrum together with the application of curveresolving techniques indicates a far more complex situation. The spectra of choline chloride/urea mixtures of different compositions in the 1800−1500 cm−1 region of the spectrum are compared in Figure 9. It can be seen that at molar ratios of 1:0.5 and 1:1 choline chloride:urea, a band near 1720 cm−1 can be observed. This can be assigned to free carbonyl groups, those that are not hydrogen bonded to NH or OH groups. This band is no longer apparent in the spectra of the 1:2 and 1:3 molar ratio mixtures. Essentially, in mixtures with lower concentrations of urea, NH2 groups favor formation of hydrogen bonds to Cl− or possibly the oxygen of choline chloride OH groups over urea

Figure 9. The spectra of choline chloride/urea mixtures as a function of composition. The spectra have been offset for clarity. From top to bottom, mixtures with a molar ratio of choline chloride to urea of 1:0.5, 1:1, 1:2, and 1:3.

carbonyl groups, more than likely the former. However, the most important observation is that all of these spectra show that both bands are asymmetrically broadened with the 1668 cm−1 band predominantly on the high frequency side, while 1616 cm−1 is clearly asymmetric on the low frequency side. This indicates the presence of many overlapping bands. We have curve-resolved the CO stretch and NH2 bending region of the spectrum, and the results for the 1:1 and 1:2 mixtures are shown in Figure 10. Curve resolving remains somewhat of an art. A good fit can always be obtained by including enough bands, but in the end, the character of the bands (particularly their width at half-height) and their assignment have to make spectroscopic sense. Our methodology is based on the approach73 we have used to analyze band shapes in the infrared spectra of polystyrene. The procedure involves an initial examination of a broad region of the spectrum to identify where a baseline can be placed. This is crucial in that a badly placed baseline can lead to significant distortions in band shape as the tails of Lorentzian bands extend to far beyond the tails of Gaussian bands. The baseline shown in Figure 10 was fixed at minima outside the region of the spectrum where there were no apparent absorption bands (in fact, the baseline was set very slightly below these values to account for the tails of Lorentzian bands, see ref 73). A brief summary on how the experimental spectra bands were curve-resolved can be found in the Supporting Information. We suggest that the six hydrogen-bonded bands are three pairs that correspond in origin to the three modes observed in urea: the coupled symmetric NH2 deformation and the CO stretching modes and the asymmetric NH2 deformation mode (coupled to C−N stretch according to the calculations of Keuleers et al.67). Each of these now has two components in the mixtures, one with a stronger average hydrogen bond strength than the other. We suggest that the pair of bands resolved near 1688 cm−1 and 1670 cm−1 correspond to the 1689 cm−1 mode observed in urea. The 1688 cm−1 mode is less intense than the 1670 cm−1 mode. The simulations reported above indicate that approximately 35% of the urea NH groups (both cis and trans) are hydrogen bonded H

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Figure 10. The spectrum of the choline chloride/urea spectrum scale expanded in the 1740−1560 cm−1 region of the spectrum together with bands identified by curve-resolving. Left, the spectrum of the 1:1 molar ratio choline chloride to urea mixture; right, the 1:2 molar ratio spectrum.

to Cl− anions, while about 52% of the NH groups are hydrogen bonded to urea carbonyl groups. This is consistent with the spectra in that if this mode has a larger contribution from NH2 deformation vibrations than CO stretch, as the calculations of Keuleers et al.67 suggest, then the more strongly hydrogen bonded species should be observed at higher frequency. The opposite should be true of the (predominantly) carbonyl-stretching mode, observed at 1606 cm−1 in urea, which now has a less intense component near 1592 cm−1 and a more intense component near 1613 cm−1, the higher frequency mode now corresponding to a distribution that is less strongly hydrogen bonded. The largely asymmetric deformation mode does not show this level of intensity difference, but these two weaker bands centered between more intense modes will be subject to the most error in curve-resolving.



eutectic mixture. Bands are assigned for the two most significant regions of the spectra, the OH/NH2 stretching and the CO/ NH2 stretching and bending regions of the spectrum. This latter region is curve-resolved providing a more detailed band assignment than has ever before been published in the literature to our knowledge. Comparing the relative contributions of hydrogen-bonding interactions observed in the simulations, the higher frequency curve-resolved band for the NH2 bending mode around 1688 cm−1 is assigned to the crucial NH2--Cl− interaction implying a strong interaction at these higher frequencies. Atomistic detail from the simulations shows that the anion prefers to interact with the Htrans in order to maximize the number of hydrogen bonds present in the system. Additionally, from the molar ratios studied through spectroscopy, the disappearance of the free carbonyl band with increasing urea concentrations suggests increasing NH2-CO or OH--CO interactions. This suggests that at low urea concentrations, interactions between NH2 groups with the anion are preferred, and as urea concentrations increase approaching the eutectic mixture, this interaction remains a crucial component of this mixture with other interactions also being present. In the 1:2 molar ratio spectrum, no free NH, OH, and CO bands are observed possibly suggesting that the system packs in a manner to maximize its hydrogen bond network between the different moieties. Both molecular simulations and vibrational spectroscopy are used to highlight key aspects of this eutectic mixture shining light toward the important interactions that are present in this system.

CONCLUSION

Hydrogen-bonding interactions of choline chloride/urea eutectic mixtures have been studied by a complementary simulation/ experimental approach. Simulated and experimental densities are in good agreement with all modified force fields used in this work although there is a broad range of experimental densities available in the literature. Simulated values for volume expansion coefficient and heat capacity are shown to better agree with experimental data for 8Amb. Not much improvement is observed when using Mod1 and Mod2 as they both have similar density, volume expansion coefficient, and heat capacity values as 1Amb. Reasonable agreement with experimental values is seen for diffusion coefficients at 298 K, and excellent agreement is seen at 330 K for the best performing force field modification. For structural properties shown in the RDFs, peak heights decrease and peak locations slightly increase with a reduction in net charge of the ionic species. From COM RDFs, a higher number of urea nearest neighbors around the anion is calculated compared to choline cations surrounding anions. Through a detailed hydrogen bond analysis, a relative percent of hydrogen-bonding interactions is calculated for the main types of hydrogen bonding. The results from simulations are compared to the infrared spectra to probe the interactions responsible for forming this



ASSOCIATED CONTENT

S Supporting Information *

A complete list of force field parameters, autocorrelation functions for simulations at 298 and 330 K, and an average COM RDF with error bars. This material is available free of charge via the Internet at http://pubs.acs.org



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (P.P.); [email protected] (C.C.). I

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

Notes

Based Deep Eutectic Solvents Studied by 1H Pulsed Field Gradient NMR Spectroscopy. Phys. Chem. Chem. Phys. 2011, 13, 21383−21391. (17) Leron, R. B.; Soriano, A. N.; Li, M.-H. Densities and Refractive Indices of the Deep Eutectic Solvents (Choline Chloride+Ethylene Glycol or Glycerol) and Their Aqueous Mixtures at the Temperature Ranging from 298.15 to 333.15 K. J. Taiwan Inst. Chem. Eng. 2012, 43, 551−557. (18) Leron, R. B.; Li, M. High-Pressure Density Measurements for Choline Chloride: Urea Deep Eutectic Solvent and Its Aqueous Mixtures at T = (298.15 to 323.15) K and Up to 50 MPa. J. Chem. Thermodyn. 2012, 54, 293−301. (19) Abbott, A. P.; Boothby, D.; Capper, G.; Davies, D. L.; Rasheed, R. K. Deep Eutectic Solvents Formed Between Choline Chloride and Carboxylic Acids: Versatile Alternatives to Ionic Liquids. J. Am. Chem. Soc. 2004, 126, 9142−9147. (20) Abbott, A. P.; Harris, R. C.; Ryder, K. S. Application of Hole Theory to Define Ionic Liquids by Their Transport Properties. J. Phys. Chem. B 2007, 111, 4910−4913. (21) Maginn, E. J. Molecular Simulation of Ionic Liquids: Current Status and Future Opportunities. J. Phys.: Condens. Matter 2009, 21, 373101/1−373101/17. (22) Schröder, C. Comparing Reduced Partial Charge Models with Polarizable Simulations of Ionic Liquids. Phys. Chem. Chem. Phys. 2012, 14, 3089−3102. (23) Dong, K.; Song, Y.; Liu, X.; Cheng, W.; Yao, X.; Zhang, S. Understanding Structures and Hydrogen Bonds of Ionic Liquids at Electronic Level. J. Phys. Chem. B 2011, 116, 1007−1017. (24) Dommert, F.; Wendler, K.; Berger, R.; Delle Site, L.; Holm, C. Force Fields for Studying the Structure and Dynamics of Ionic Liquids: A Critical Review of Recent Developments. ChemPhysChem 2012, 13, 1625−1637. (25) Kohagen, M.; Brehm, M.; Lingscheid, Y.; Giernoth, R.; Sangoro, J.; Kremer, F.; Naumov, S.; Iacob, C.; Kärger, J.; Valiullin, R.; et al. How Hydrogen Bonds Influence the Mobility of Imidazolium-Based Ionic Liquids. A Combined Theoretical and Experimental Study of 1-n-Butyl3-methylimidazolium Bromide. J. Phys. Chem. B 2011, 115, 15280− 15288. (26) Rigby, J.; Izgorodina, E. I. Assessment of Atomic Partial Charge Schemes for Polarisation and Charge Transfer Effects in Ionic Liquids. Phys. Chem. Chem. Phys. 2013, 15, 1632−1646. (27) Rimsza, J. M.; Corrales, L. R. Adsorption Complexes of Copper and Copper Oxide in the Deep Eutectic Solvent 2:1 Urea−Choline Chloride. Comput. Theor. Chem. 2012, 987, 57−61. (28) Sun, H.; Li, Y.; Wu, X.; Li, G. Theoretical Study on the Structures and Properties of Mixtures of Urea and Choline Chloride. J. Mol. Model. 2013, 19, 2433−2441. (29) Yue, D.; Jia, Y.; Yao, Y.; Sun, J.; Jing, Y. Structure and Electrochemical Behavior of Ionic Liquid Analogue Based on Choline Chloride and Urea. Electrochim. Acta 2012, 65, 30−36. (30) Deng, L.; Jing, Y.; Sun, J.; Ma, J.; Yue, D.; Wang, H. Preparation and Properties of a Moisture-Stable Ionic Liquid ChCl−ZnCl2−MgCl2· 2CH3COOCH2CH3·2H2O. J. Mol. Liq. 2012, 170, 45−50. (31) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A.; Vreven, J. T.; Kudin, K. N.; Burant, J. C.; et al. Gaussian 03, revision E.01; Gaussian, Inc.: Wallingford, CT, 2004. (32) Dupradeau, F.-Y.; Pigache, A.; Zaffran, T.; Savineau, C.; Lelong, R.; Grivel, N.; Lelong, D.; Rosanski, W.; Cieplak, P. The R.E.D. Tools: Advances in RESP and ESP Charge Derivation and Force Field Library Building. Phys. Chem. Chem. Phys. 2010, 12, 7821−7839. (33) Vanquelef, E.; Simon, S.; Marquant, G.; Garcia, E.; Klimerak, G.; Delepine, J. C.; Cieplak, P.; Dupradeau, F.-Y. R.E.D. Server: A Web Service for Deriving RESP and ESP Charges and Building Force Field Libraries for New Molecules and Molecular Fragments. Nucleic Acids Res. 2011, 39, W511−W517. (34) Dupradeau, F.-Y.; Cézard, C.; Lelong, R.; Stanislawiak, E.; Pêcher, J.; Delepine, J. C.; Cieplak, P. R.E.DD.B.: A Database for RESP and ESP Atomic Charges, and Force Field Libraries. Nucleic Acids Res. 2008, 36, D360−D367.

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported in part through the National Science Foundation (DMR-0901180) and Penn State’s MatSE ButtonWaller awards. High performance computational resources were provided by instrumentation funded by the National Science Foundation through grant OCI-0821527 as well as by Penn State’s Research Computing and Cyberinfrastructure Group. We also would like to thank Lauren Abbott for providing some of the source codes for data analysis.



REFERENCES

(1) Torimoto, T.; Tsuda, T.; Okazaki, K.; Kuwabata, S. New Frontiers in Materials Science Opened by Ionic Liquids. Adv. Mater. 2010, 22, 1196−1221. (2) Zhang, Q.; De Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep Eutectic Solvents: Syntheses, Properties and Applications. Chem. Soc. Rev. 2012, 41, 7108−7146. (3) Ruß, C.; König, B. Low Melting Mixtures in Organic Synthesis − an Alternative to Ionic Liquids? Green Chem. 2012, 14, 2969−2982. (4) Abbott, A. P.; Capper, G.; Davies, D. L.; Rasheed, R. K.; Tambyrajah, V. Novel Solvent Properties of Choline Chloride/Urea Mixtures. Chem. Commun. 2003, 70−71. (5) Carriazo, D.; Serrano, M. C.; Gutiérrez, M. C.; Ferrer, M. L.; Del Monte, F. Deep-Eutectic Solvents Playing Multiple Roles in the Synthesis of Polymers and Related Materials. Chem. Soc. Rev. 2012, 41, 4996−5014. (6) Tang, S.; Baker, G. A.; Zhao, H. Ether- and Alcohol-Functionalized Task-Specific Ionic Liquids: Attractive Properties and Applications. Chem. Soc. Rev. 2012, 41, 4030−4066. (7) Endres, F.; Abbott, A. P.; Macfarlane, D. R. Electrodecomposition from Ionic Liquids; John Wiley & Sons: Weinheim, Germany, 2008. (8) Abbott, A. P.; Capper, G.; McKenzie, K. J.; Ryder, K. S. Voltammetric and Impedance Studies of the Electropolishing of Type 316 Stainless Steel in a Choline Chloride Based Ionic Liquid. Electrochim. Acta 2006, 51, 4420−4425. (9) Abbott, A. P.; Capper, G.; Davies, D. L.; Rasheed, R. K.; Shikotra, P. Selective Extraction of Metals from Mixed Oxide Matrixes Using Choline-Based Ionic Liquids. Inorg. Chem. 2005, 44, 6497−6499. (10) Pawar, P. M.; Jarag, K. J.; Shankarling, G. S. Environmentally Benign and Energy Efficient Methodology for Condensation: An Interesting Facet to the Classical Perkin Reaction. Green Chem. 2011, 13, 2130−2134. (11) Phadtare, S. B.; Shankarling, G. S. Halogenation Reactions in Biodegradable Solvent: Efficient Bromination of Substituted 1-Aminoanthra-9,10-quinone in Deep Eutectic Solvent (Choline Chloride: Urea). Green Chem. 2010, 12, 458−462. (12) Sonawane, Y. A.; Phadtare, S. B.; Borse, B. N.; Jagtap, A. R.; Shankarling, G. S. Synthesis of Diphenylamine-Based Novel Fluorescent Styryl Colorants by Knoevenagel Condensation Using a Conventional Method, Biocatalyst, and Deep Eutectic Solvent. Org. Lett. 2010, 12, 1456−1459. (13) Li, X.; Hou, M.; Han, B.; Wang, X.; Zou, L. Solubility of CO2 in a Choline Chloride + Urea Eutectic Mixture. J. Chem. Eng. Data 2008, 53, 548−550. (14) Abbott, A. P.; Cullis, P. M.; Gibson, M. J.; Harris, R. C.; Raven, E. Extraction of Glycerol from Biodiesel into a Eutectic Based Ionic Liquid. Green Chem. 2007, 9, 868−872. (15) Maugeri, Z.; Domínguez de María, P. Novel Choline-ChlorideBased Deep-Eutectic-Solvents with Renewable Hydrogen Bond Donors: Levulinic Acid and Sugar-Based Polyols. RSC Adv. 2012, 2, 421−425. (16) D’Agostino, C.; Harris, R. C.; Abbott, A. P.; Gladden, L. F.; Mantle, M. D. Molecular Motion and Ion Diffusion in Choline Chloride J

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry B

Article

(35) Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; et al. AMBER 11, University of California, San Francisco, 2010. (36) Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Crowley, M.; Walker, R. C.; Zhang, W.; et al. Amber 10; University of California: San Francisco, CA, 2008. (37) Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Automatic Atom Type and Bond Type Perception in Molecular Mechanical Calculations. J. Mol. Graphics Modell. 2006, 25, 247−260. (38) Bayly, C. I.; Cieplak, P.; Cornell, W. D.; Kollman, P. A. A WellBehaved Electrostatic Potential Based Method Using Charge Restraints for Deriving Atomic Charges: The RESP Model. J. Phys. Chem. 1993, 97, 10269−10280. (39) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25, 1157−1174. (40) Liu, H.; Maginn, E. J.; Visser, A. E.; Bridges, N. J.; Fox, E. B. Thermal and Transport Properties of Six Ionic Liquids: An Experimental and Molecular Dynamics Study. Ind. Eng. Chem. Res. 2012, 51, 7242−7254. (41) Zhang, Y.; Maginn, E. J. A Simple AIMD Approach to Derive Atomic Charges for Condensed Phase Simulation of Ionic Liquids. J. Phys. Chem. B 2012, 33, 10036−10048. (42) Lynden-Bell, R. M.; Youngs, T. G. A. Simulations of Imidazolium Ionic Liquids: When Does the Cation Charge Distribution Matter? J. Phys.: Condens. Matter 2009, 21, 424120/1−424120/7. (43) Youngs, T. G. A.; Hardacre, C. Application of Static Charge Transfer within an Ionic-Liquid Force Field and Its Effect on Structure and Dynamics. ChemPhysChem 2008, 9, 1548−1558. (44) Bhargava, B. L.; Balasubramanian, S. Refined Potential Model for Atomistic Simulations of Ionic Liquid [Bmim][PF6]. J. Chem. Phys. 2007, 127, 114510/1−114510/6. (45) Abbott, L. J.; Hart, K. E.; Colina, C. M. Polymatic: A Generalized Simulated Polymerization Algorithm for Amorphous Polymers. Theor. Chem. Acc. 2013, 132, 1334/1−1334/19. (46) Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; et al. Amber 12; University of California: San Francisco, CA, 2012. (47) Götz, A. W.; Williamson, M. J.; Xu, D.; Poole, D.; Le Grand, S.; Walker, R. C. Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 1. Generalized Born. J. Chem. Theory 2012, 8, 1542−1555. (48) Le Grand, S.; Götz, A. W.; Walker, R. C. SPFP: Speed without CompromiseA Mixed Precision Model for GPU Accelerated Molecular Dynamics Simulations. Comput. Phys. Commun. 2012, 184, 374−380. (49) Morrow, T. I.; Maginn, E. J. Density, Local Composition and Diffusivity of Aqueous Choline Chloride Solutions: A Molecular Dynamics Study. Fluid Phase Equilib. 2004, 217, 97−104. (50) Ozpinar, G. A.; Peukert, W.; Clark, T. An Improved Generalized AMBER Force Field (GAFF) for Urea. J. Mol. Model. 2010, 16, 1427− 1440. (51) Larsen, G. S.; Lin, P.; Hart, K. E.; Colina, C. M. Molecular Simulations of PIM-1-like Polymers of Intrinsic Microporosity. Macromolecules 2011, 44, 6944−6951. (52) Morrow, T. I.; Maginn, E. J. Molecular Dynamics Study of the Ionic Liquid 1-n-Butyl-3-methylimidazolium Hexafluorophosphate. J. Phys. Chem. B 2002, 106, 12807−12813. (53) Ciocirlan, O.; Iulian, O.; Croitoru, O. Effect of Temperature on the Physico-chemical Properties of Three Ionic Liquids Containing Choline Chloride. Rev. Chim. (Bucharest) 2010, 61, 721−723. (54) Abbott, A. P.; Capper, G.; Gray, S. Design of Improved Deep Eutectic Solvents Using Hole Theory. ChemPhysChem 2006, 7, 803− 806. (55) Senko, M. E.; Templeton, D. H. Unit Cells of Choline Halides and Structure of Choline Chloride. Acta Crystallogr. 1960, 13, 281−285. (56) Caleman, C.; Maaren, P. J.; Van Hong, M.; Hub, J. S.; Costa, L. T.; van der Spoel, D. Force Field Benchmark of Organic Liquids: Density, Enthalpy of Vaporization, Heat Capacities, Surface Tension, Isothermal

Compressibility, Volumetric Expansion Coefficient, and Dielectric Constant. J. Chem. Theory Comput. 2012, 8, 61−74. (57) Leron, R. B.; Li, M.-H. Molar Heat Capacities of Choline Chloride-Based Deep Eutectic Solvents and Their Binary Mixtures with Water. Thermochim. Acta 2012, 530, 52−57. (58) Gardas, R. L.; Coutinho, J. A. P. A Group Contribution Method for Heat Capacity Estimation of Ionic Liquids. Ind. Eng. Chem. Res. 2008, 47, 5751−5757. (59) Cadena, C.; Maginn, E. J. Molecular Simulation Study of Some Thermophysical and Transport Properties of Triazolium-Based Ionic Liquids. J. Phys. Chem. B 2006, 110, 18026−18039. (60) Cadena, C.; Zhao, Q.; Snurr, R. Q.; Maginn, E. J. Molecular Modeling and Experimental Studies of the Thermodynamic and Transport Properties of Pyridinium-Based Ionic Liquids. J. Phys. Chem. B 2006, 110, 2821−2832. (61) Lagache, M.; Ungerer, P.; Boutina, A.; Fuchs, A. H. Prediction of Thermodynamic Derivative Properties of Fluids by Monte Carlo Simulation. Phys. Chem. Chem. Phys. 2001, 3, 4333−4339. (62) Pollet, B. G.; Hihn, J.-Y.; Mason, T. J. Sono-Electrodeposition (20 and 850 kHz) of Copper in Aqueous and Deep Eutectic Solvents. Electrochim. Acta 2008, 53, 4248−4256. (63) Del Pópolo, M. G.; Voth, G. A. On the Structure and Dynamics of Ionic Liquids. J. Phys. Chem. B 2004, 108, 1744−1752. (64) Zhao, W.; Leroy, F.; Heggen, B.; Zahn, S.; Kirchner, B.; Balasubramanian, S.; Müller-Plathe, F. Are There Stable Ion-Pairs in Room-Temperature Ionic Liquids? Molecular Dynamics Simulations of 1-n-Butyl-3-methylimidazolium Hexafluorophosphate. J. Am. Chem. Soc. 2009, 131, 15825−15833. (65) Méndez-Morales, T.; Carrete, J.; Cabeza, O.; Gallego, L. J.; Varela, L. M. Molecular Dynamics Simulation of the Structure and Dynamics of Water-1-Alkyl-3-methylimidazolium Ionic Liquid Mixtures. J. Phys. Chem. B 2011, 115, 6995−7008. (66) Shukla, A.; Isaacs, E.; Hamann, D.; Platzman, P. Hydrogen Bonding in Urea. Phys. Rev. B 2001, 64, 052101/1−052101/4. (67) Keuleers, R.; Desseyn, H. O.; Rousseau, B.; Van Alsenoy, C. Vibrational Analysis of Urea. J. Phys. Chem. A 1999, 103, 4621−4630. (68) Keuleers, R.; Janssens, J.; Desseyn, H. Thermal Analysis and Vibrational Spectroscopy of Mn(II)−Urea−Halide Complexes: Comparative Study of the Metal−Ligand Bond Strength. Thermochim. Acta 2000, 354, 125−133. (69) Derreumaux, P.; Vergoten, G.; Lagant, P. A Vibrational Molecular Force Field of Model Compounds with Biological Interest. I. Harmonic Dynamics of Crystalline Urea at 123 K. J. Comput. Chem. 1990, 11, 560− 568. (70) Harmon, K. M.; Cross, J. E.; Toccalino, P. L. Hydrogen Bonding: Part 25. The Nature of the Hydrogen Bond in Hydroxytropenylium Chloride (Tropone Hydrochloride). J. Mol. Struct. 1988, 178, 141−145. (71) Harmon, K. M.; Madeira, S. L.; Jacks, M. J.; Avci, G. F.; Thiel, A. C. Hydrogen Bonding: Part 18. The Nature of the OHF Hydrogen Bond in Choline Fluoride. J. Mol. Struct. 1985, 128, 305−314. (72) Krimm, S.; Bandekar, J. Vibrational Spectroscopy and Conformation of Peptides, Polypeptides, and Proteins. Adv. Prot. Chem. 1986, 38, 181−364. (73) Painter, P. C.; Zhao, H.; Park, Y. Vibrational Relaxation and Dynamical Transitions in Atactic Polystyrene. Macromolecules 2009, 42, 435−444.

K

dx.doi.org/10.1021/jp404619x | J. Phys. Chem. B XXXX, XXX, XXX−XXX