Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2022 Nov 3.
Published in final edited form as: J Chem Phys. 2014 May 28;140(20):204505. doi: 10.1063/1.4878116

On the collective network of ionic liquid/water mixtures. IV. Kinetic and rotational depolarization

Christian Schröder 1,a), Marcello Sega 1, Michael Schmollngruber 1, Elias Gailberger 1, Daniel Braun 1, Othmar Steinhauser 1
PMCID: PMC7613785  EMSID: EMS155866  PMID: 24880299

Abstract

Dielectric spectroscopy is a measure of the collective Coulomb interaction in liquid systems. Adding ionic liquids to an aqueous solution results in a decrease of the static value of the generalized dielectric constant which cannot be attributed to kinetic depolarization models characterized by the static conductivity and rotational relaxation constant. However, a dipolar Poisson-Boltzmann model computing the water depolarization in the proximity of ions is not only successful for simple electrolytes but also in case of molecular ionic liquids. Moreover, our simple geometric hydration model is also capable to explain the dielectric depolarization. Both models compute the dielectric constant of water and obtain the overall dielectric constant by averaging the values of its components, water and the ionic liquid, weighted by their volume occupancies. In this sense, aqueous ionic liquid mixtures seem to behave like polar mixtures.

I. Introduction

Ionic liquids and water have not always been best friends.1 Even if they are completely miscible with water, there are often issues of hydrolysis.2, 3 However, mixtures of ionic liquids and water may show physico-chemical properties which are not accessible by the pure compounds, thus enabling novel applications for many fields.1 Quite generally, in many applications ionic liquids suffer from their high viscosities. For example, 1-butyl-3- ethylimidazolium tetrafluoroborate [C4 mim][BF4] has a viscosity of approximately 200 mPa s which is more than two orders of magnitude higher compared to the value of water. Actually, one of the fluidest ionic liquids is 1-ethyl-3-methylimidazolium dicyanamide with a viscosity of roughly 20 mPa s at room temperature which is still far less fluid than water. This situation changes when mixing an ionic liquid with water. The viscosity4 and density5, 6 drop significantly with increasing water content whereas the static electric conductivity σ(0) shows a parabolic behavior as a function of the mole fraction.5, 7 Furthermore, the physico-chemical properties of the ionic liquid/water mixtures change with mole fraction and can be roughly classified into three regions:810 At low water mole fractions, xW < 0.2, the water self-diffusion coefficient stays nearly the same with increasing water concentration. These water molecules are diluted in the ionic liquid network. Between a mole fraction of 0.2 < xW < 0.8, water molecules seem to aggregate in small clusters, for example dimers and trimers, within the ionic liquid network. However, an increasing water content has a stronger effect on the diffusion of anions compared to that of cations indicating stronger interactions. Moreno et al. reported that the interaction between water and the ionic liquid ions is non-selective and water mainly initiates a swelling of the [C4mim][BF4] network.11 This corresponds to our former finding that the decreased interaction between cations and anions is not due to individual water molecules penetrating the IL network. Rather, the electric field of the water molecules screens the Coulomb interaction.12

At high water mole fractions, xW > 0.8, the swelling of the ionic liquid structure, i.e. the stretching of the chain-like structures, results in a breakdown of the highly branched ionic network and the mixture decomposes into water rich and ionic liquid rich regions. A turnover of the physico-chemical properties at this mole fraction was observed by Lopes et al.9, 10 and Voth et al.13, 14, but Maroncelli et al.15 found no evidence for abrupt changes taking place at a particular composition. The formation of micelles is only observed for long alkyl chain cations, e.g. C8 mim+.

In a series of papers,12,16,17 we analyzed the collective network of [C4 mim][BF4]/H2O mixtures with xW ranging from 0.425 to 0.967: The first manuscript12 dealt with the orientational structure. We found clear evidence of a strong anion-water network showing hydrogen bonding but weak interactions between the cations and water. Furthermore, the water molecules mainly act on the ions via their dipolar field and the anions play the role of a mediator between water and the cations. In the second publication,16 we juxtaposed the experimental and computed dielectric spectra of the mixture at water mole fractions xW from 0.768 to 0.967. At that time we could not resolve the translational contribution in detail and were restricted to a numerical Laplace transform of the auto-correlation function of the current, 〈J(0) · J(t)〉.16 In the present work we make up the leeway and present cosine-exponential based fits of the auto-correlation function 〈J(0)·J(t)〉.18 Moreover, we give here a representation of the frequency-dependent contribution of the ro-translational coupling ϵDJ(ω) between the col-lective rotational dipole moment MD(t) and the current J(t) to the generalized dielectric constant Σ0(ω). In the third paper17 the Voronoi decomposition proved to be helpful for the interpretation of the ionic liquid/water networks: We divided the C4 mim+-cations into head (methyl group), imidazolium ring and tail (butyl chain) and split the radial and orientational correlation functions into contributions from the first, second and third solvation shell. As expected, the tails turned out to be more or less hydrophobic and to prefer interactions with other tails. We also found a strong anion-water network as reported in literature.1923 Nevertheless, both species compete for their favorable positions in direct vicinity of the imidazolium ring hydrogens.

With the present work we want to draw the attention on the dielectric decrement of aqueous IL mixtures as a function of increasing ion content and its interpretation in terms of electrolyte theories,2438 as well as models of mixtures of polar solvents.3942 Experimental dielectric studies concerning aqueous IL mixtures16,4346 show a decreasing static dielectric constant Σ0 with increasing ion content. The depolarization effect is often partially attributed to the motion of the ions,3034 but we will show that the kinetic depolarization cannot explain the concentration dependence of Σ0(cIL) in our IL mixtures at ion concentrations above 1M and rotational depolarization plays a major role.

II. Methods

In this work we used the trajectories from our former studies concerning the mixtures of [C4 mim][BF4] with TIP3P water, but extended the simulation to 120 ns,12,16,17 which is more than one order of magnitude longer than other computational studies.911,13,14 Although the viscosity of these mixtures is below 6 mPa s,16 we deem long simulation periods essential for good statistics. In addition, not only the coordinates of the atoms at each time step but also their velocities were stored in order to compute more reliable data on the auto-correlation function of the electric current 〈J(0) · J(t)〉 and on the cross correlation between the collective rotational dipole moment MD(t) and the electric current J(t). These additional data improved the statistics of our investigated properties and lead to slightly changed values of the static electric conductivity σ(0) (see Table I) compared to those reported in our previous study.16

Table I.

Composition of the [C4 mim][BF4]/water mixtures. nIL and nW denote the number of the respective molecules. Even at high water mole fractions xW, the ionic liquid ions occupy a significant volume V of the simulation box as indicated by ΦIL = VIL/V. The volume fraction of water ΦW = Φbulk + Φhyd is splitted into bulk (Φbulk) and hy-dration (Φhyd) water based on the ratio of amplitudes of <MDW(0)MDW(t)>.

cIL = 0.00 M 1.25 M 2.52 M 3.77 M 5.31 M
n IL 0 55 111 166 234a
n W 2440b 1592 1147 548 0
x W 1.000 0.967 0.912 0.768 0.000
Φhyd 0.000 0.338 0.372 0.214 0.000
Φbulk 1.000 0.407 0.148 0.045 0.000
ΦIL 0.000 0.255 0.480 0.741 1.000
Σ0 102 55.4 38.5 24.0 12.5
σ(0) / S m−1 0.0 11.9 11.3 5.0 0.44c
a

Extrapolated number since Ref. 47 used a different box size.

b

Extrapolated number since Ref. 48 used a different box size.

c

see Ref. 47.

A detailed description of the simulation setup is given in Ref. 12 and summarized here briefly. Three independent molecular dynamics simulations at constant volume, V =(41.8 Å)3, with ionic liquid concentration cIL =1.25, 2.52, 3.77 M (see Table I) were performed at 300 K with a time step of 2 fs using particle mesh Ewald summation with a real-space cut-off of 10 Å and a screening parameter κ = 0.41 Å−1. The same cut-off was also used for dispersion forces. These mixtures are accompanied by the results of neat [C4mim][BF4] (Ref. 47) and pure TIP3P water48 in Table I.

III. Theory

A. Dielectric spectrum and its components

This present manuscript extends the dielectric theory already used in Ref. 16. The generalized dielectric constant, Σ0(ω) is composed of a dielectric permittivity, ε(ω), and dielectric conductivity, ϑ0(ω) and their ro-translational coupling εDJ(ω):18

Σ0(ω)=ϵ(ω)1+ϑ0(ω)+2ϵDJ(ω) (1)

Usually, the latter contribution εDJ(ω) is equally distributed between the dielectric permittivity and dielectric conductivity. Since we are interested in the coupling between the ionic motion characterized by the electric current J(t) and the rotation of dipoles MD(t) for the interpretation of the kinetic depolarization, this contribution is considered separately here. For the sake of simplicity, the static values of the dielectric properties in this work are denoted without the corresponding arguments, i.e.

limω0Σ0(ω)=Σ0(0)=Σ0. (2)

In Ref. 16, analytical representations of the collective dipole correlation functions for the cations <MDC(0)MDC(t)> and for water MDW(0)MDW(t),

fDD(t)=kAkexp(t/τk), (3)

were used to compute the dielectric permittivity ε(ω). Since these relaxation processes are of collective nature, an assignment of the kth-contribution to a distinct physical process is complicated. However, often the process with the longest τk is used to correlate single particle and collective rotation via the Kivelson-Madden equation.49

In the simulations, the dielectric permittivity ε(ω) can be decomposed in pure cationic, εCC(ω), and pure water, εWW(ω) contributions as well as their coupling, εWC(ω), using the collective dipole moments MDC(0) and MDW(0) for the auto-correlation functions, MDC(0)MDC(t) and MDW(0)MDW(t), or the cross-correlation function MDC(0)MDW(t), respectively. The anion BF4 has tetrahedral symmetry and therefore no permanent dipole moment. The transient dipole of BF4 due the vibrations is negligibly small to contribute to ε(ω).

The dielectric conductivity ϑ0(ω) is the translational contribution of the ions to the dielectric spectrum Σ0(ω). It is defined by the ratio of the frequency-dependent electric conductivity σ(ω) and the angular frequency ω:

ϑ0(ω)=4πiσ(ω)σ(0)ω (4)

In Ref. 16, the electric conductivity σ(ω) was only computed by numerically Laplace transforming the current correlation function 〈J(0) · J(t)〉. Afterwards, the dielectric conductivity ϑ0(ω) was obtained by eq. (4). This results in noisy spectra. The corresponding fit functions for 〈J(0) · J(t)〉 were first introduced in Ref. 18 and 50:

fJJ(t)=kAkcos(ωkt+δk)exp(t/τk). (5)

Slower translational relaxation processes were detected by the mean-squared displacement of the collective translational dipole moment ΔMJ2(t) and then added to the fit of 〈J(0)· J(t)〉 since d2ΔMJ2(t)/dt2=2J(0)J(t) as shown in Ref. 51. In the same reference we showed that all time constants τk are also valid for relaxation processes of the ion cage, i.e. the first shell around a central ion. These shells contain a significant number of ions with the same charge as the central ion.51 Cationcation contacts are still very prominent in our aqueous [C4mim][BF4] mixtures.17

The ro-translational coupling εDJ(ω) can be computed by the Laplace transform of the cross-correlation function 〈MD(0) · J(t)〉 which can be represented analytically by

fDJ(t)=kAktγk1expt/τk (6)

leading to a sum of Cole-Davidson peaks in the dielectric spectrum.52 Generally, the ro-translational contribution εDJ(ω) to the dielectric spectrum of pure ionic liquids is quite small due to moderate dipole densities and a small electric current J(t) of the viscous systems.47, 53 However, in ionic liquid/water mixtures, the strong dipolar density of water is combined with more mobile charge carriers, thus enhancing J(t). Consequently, the ro-translational contribution should be stronger than in pure ionic liquids.

B. Electrolyte theories

In a series of papers, Hubbard et al.30, 31 and Wolynes et al.3234 attributed part of the dielectric decrement to the kinetic depolarization of the ions moving through the water media which should be reflected by the ro- translational coupling εDJ. The general idea behind these models is the following: A moving ion enforces the rotation of the water dipoles which disturbs the normal alignment of water dipoles and results in a dielectric decrement. This is also achieved, if the rotation of the water molecules slows down the ionic motion. In case of the Hubbard model

ΔϵHub=pτDσ(0)ϵ0Σ0Σ0+1 (7)

we assume perfect slip (p = 2/3) boundary conditions between the ions and water. τD is the rotational relaxation constant of the auto-correlation MDW(0)MDW(t) of the collective rotational water dipoles, σ(0) the static conductivity and ε0 = 8.85 · 10−12As/Vm the dielectric constant in vacuo. Wolynes et al.33 developed a molecular picture of the ion-water interaction, that leads to a kinetic decrement

ΔϵWol=23τDσ(0)ϵ0i,jRiqiμjrijrij31nIL. (8)

Here, the index i runs over all ions and the index j over all water molecules with their dipolar vector μj. The vector rij points from the respective ion i to the respective water molecule j. The ion radius Ri can be easily approximated assuming a spherical shape and using the corresponding Voronoi volumes of the species.17 Nevertheless, the depolarization in both models depends linearly on the static conductivity σ(0) and the rotational relaxation constant τD

In contrast to the kinetic depolarization of Hubbard and Wolynes, also pure static approaches exist. For example, the dressed ion theory,54 which is an extension to the classical Debye-Hückel theory, computes the depolarization due to the ions in Fourier space as

ΔΣ0(k)=1k2(Σ0κD2+Szz(k)kBTϵ0), (9)

using the inversed Debye screening length κD and the charge-ordering function Szz(k). For very dilute electrolyte solutions, the charge-ordering function becomes negligible and the dressed ion theory simplifies to the result of Debye-Hückel.37, 55 However, the solvent enters these two theories only by its static dielectric constant and the complete effect on the overall static dielectric constant is attributed to ion-ion interactions with a solvent continuum as background.

Levy et al. studied the interactions of the water dipoles in the proximity of ions using a static dipolar Poisson- Boltzmann approach. Thus, they were able to reproduce the dielectric decrement of simple electrolyte solutions up to ion concentrations of 5M.38 The static value εDPB depends on the water concentration cW and a molecular water dipole μW.

ϵDPB1=cWμW23kBTϵ0 (10)

In order to yield the experimental dielectric permittivity of pure water ϵWexp=78.2 they used the dipole moment μW as a fit parameter which resulted in a value of μW = 4.80 D instead of the μWexp=1.85 D.38 This scaling was explained by missing correlation effects in the liquid phase. However, in contrast to their approach, we will use the static dielectric constant εW = 102 of pure TIP3P for the analysis of our computational mixtures. This corresponds to Σ0 at cIL = 0 M in Table I. Assuming point-shaped ions, the distance dependence of the dielectric permittivity from a central ion can be obtained by

ϵ(r)ϵW3h2(lh/r)+1 (11)

with an auxillary function h(…) (see Ref.38 for further details) and lh characterizing the shell thickness of water layers

lh=λBμWe10. (12)

Here, μW =2.35 D corresponds to the molecular dipole moment of TIP3P and λB is the Bjerrum length

λB=e24πϵ0ϵWkBT. (13)

The dependence of the static dielectric constant 〈ε(cIL)〉 on the concentration of ions cIL can be computed by spherical averaging,

ϵ(cIL)=0r(cIL)4π(r)2ϵ(r)dr43πr3(cIL), (14)

using a typical distance between two ions r(cIL)=12(2cIL)1/3.

Furthermore, Levy et al.38 proposed additional water and salt corrections to the dielectric permittivity based on the dipolar Poisson-Boltzmann approach. The first correction takes into account the fluctuation of water dipoles beyond their mean-field level. The second salt correction may be used to explain a dielectric increment with increasing ion concentrations. This phenomenon was found experimentally by Buchner et al. for [C4mim][BF4] mixtures with dichloromethane.56

C. Volume occupancies of the different species

The composition of our ionic liquid/water mixtures can be characterized by various properties, e.g. the water mole fraction 0.768≤ xW ≤0.967, the ion concentration 3.77M≥ cIL ≥ 1.25M or the volume occupancies ΦIL and 0.259≤ ΦW ≤0.745. Comparing the water mole fraction xW and the volume occupancy ΦW reveals that even the high water mole fraction xW =0.967 does not correspond to a dilute ion solution since only 74.5% of the volume is occupied by water molecules. This is due to the discrepancies of the molecular volumes of 270 Å3 and 33 Å3 of cations and water, respectively.17 The volume fractions of water, ΦW, and of the ionic liquid, ΦIL, were computed by Voronoi decomposition.17, 57

In a second step, the water volume fraction ΦW is further decomposed into bulk, Φbulk, and hydration water, Φhyd, based on a three exponential fit (k = 1 … 3) of the collective rotational dipole moment of all water molecules MDW(0)MDW(t) according to Eq. (3). Respective fit pa-rameters are given in the supplementary material. Since the rotation of hydration water is hindered, compared to bulk water, due to the interaction with the ionic liquid ions, the corresponding time constant τ3 is longer than τ2 ≃ 7.5 ps (the latter being attributed to bulk water), and is more or less independent of the IL concentration. The partitioning between hydration and bulk water is based on the relative amplitudes Ak:

Φhyd=A3A1+A2+A3ΦW (15)
Φbulk=A1+A2A1+A2+A3ΦW (16)

Overall, the ion concentrations considered here do not correspond to a highly diluted electrolyte solution anymore as pointed out by Bowers et al.58 as well as Singh and Kumar.59 Both groups observed significant aggregation of ions at cIL >0.8M and consequent changes in their physico-chemical properties. All mixtures considered here are above this threshold and overlap of hydration shells around different ions is expected.21,5860 Based on the coordination numbers reported in Ref. 17 it is obvious that at least the cations share some water molecules, since the number of ions, nC = nIL, times the coordination number of water around the cations exceeds the total number of water molecules, nW, in each of our simulation boxes. Since cation-cation contacts are very prominent in all our mixtures, the concept of solvent- separated ion pairs, solvent-shared ion pairs and contact ion pairs is difficult to apply.61

D. Geometric hydration model

It is interesting to note that the volume fraction Φhyd can be also well approximated using a simplified geometrical model: Let us assume that the hydrated ions are placed randomly in a simulation box of volume V with a density n/V = cIL. The volume of these ions and their hydration shell is Vhyd. The probability P in a Poisson process that one of the ions does not overlap with any other is

P=exp(Vhyd/V) (17)

This is also a reasonable estimate of the volume fraction of the unbound (bulk) water,

PΦbulk=1ΦILΦhyd. (18)

However, the ions are not exactly placed randomly since they cannot occupy the non-hydrated volume VIL. Using the effective volume Veff = VVIL in Eq. (17) instead of the total volume V yields

P=exp(Vhyd/V1VIL/V)=exp(ΦIL1ΦILVhydVIL) (19)

using ΦIL = VIL/V. Altogether, the volume fraction of the hydration water Φhyd results in

Φhyd(ΦIL)=1ΦILexp(ΦIL1ΦILVhydVIL) (20)

Of course, the volume fraction of the hydration water goes consistently to zero, both in the absence of ions (ΦIL = 0) and in the absence of water (ΦIL = 1). The ratio of the volumes Vhyd/VIL can be estimated by an average computed from the space occupied of water in the first Voronoi shell of a cation or anion and the space assigned to the ions. The ratio 〈Vhyd/VIL〉 at cIL=1.25 M is roughly 2.12 which means that the average hydration volume is more than twice the volume of the central ion. Although simple, the geometrical model is able to reproduce reasonably well the volume fractions without any fit parameters, as shown in Fig. 2 by the respective lines. Here, Φbulk (blue line) corresponds to the exponential part in Eq. (20). Please note that the volume occupancy ΦIL (red line) also depends linearly on the concentration of the ions cIL. The amount of bulk water drops more or less exponentially and the volume occupancy of the hydration water has a parabolic behavior: At low ion concentrations, cIL, only few possibilities for the water molecules exist to be the next neighbor of an ion. Although the number of ions increases with cIL, the competition for the water molecules to gain a place in the first solvation shell of an ion gets correspondingly harder. We will use this model later on to describe the dielectric decrement of the mixture.

Fig. 2.

Fig. 2

Volume fractions of the ionic liquid (ΦIL), bulk (Φbulk) and hydration water (Φhyd) obtained by Eq. (15) and (16) (see Table I) as well as their representation by the lines based on the geometrical model (see Eq. (20)).

IV. Results And Discussion

A. Computation of the dielectric spectrum

The frequency dependent, generalized dielectric constant Σ0(ω) emerges from the collective rotation of the dipoles within the system and the collective translational behavior of the ions. In principle, the rotational part in aqueous electrolyte solutions is dominated for dilute solutions by the auto-correlation function of the dipole moment of water molecules yielding εWW(ω), since atomic ions possess no dipole moment and the translational-rotational coupling εDJ(ω) is usually small. Increasing the concentration of the ions leads to an additional peak in the dielectric spectrum at slightly higher frequencies due to the dielectric conductivity ϑ0(ω).16,18,50

In case of molecular ions, e.g. ionic liquids, this situation gets a little more complicated since (1) the ions may have dipole moments as well. In the present work, each C4mim+-cation has an average gas-phase dipole moment μC of 5.7 D which is more than double the value of TIP3P water model (μW =2.35 D).16 For simplicity the anion BF4 was chosen to have no net dipole moment on average. (2) In simple electrolytes the rotational and translational contribution to Σ0(ω) stem from different species since the ions have no dipole moment and the water molecules have no net charge. In case of ionic liquids, however, the ions contribute to both. Nevertheless, the ro-translational coupling, εDJ = -0.3, is small in neat ionic liquids.47 On the one hand the viscosity is high, hampering the mobility of the charge carrier. On the other hand, the alignment of dipole moments is not very strong and the overall dipole density is low. (3) However, this cross contribution can be enhanced in ionic liquid/water mixtures since the viscosity of the system is lowered exponentially with increasing water content and the dipole density of water is significantly higher than that of common ionic liquids.

Based on the computational dielectric theory presented in Ref. 18, the auto-correlation functions of the collective rotational dipole moment MD(t) and of the electric current J(t) yield the dielectric permittivity ε(ω) and dielectric conductivity ϑ0(ω), respectively. For the three ionic liquid/water mixtures at cIL =1.25, 2.52 and 3.77M the corresponding frequency dependent dielectric spectra are depicted in Fig. 3a-c). The corresponding fit parameters of the respective constituting correlation functions (see Eq. (3)-(6)) are listed in the supplementary material. The self-contributions from the cations (CC, red area) and water (WW, blue area) as well as the dielectric conductivity (orange) and the ro-translational coupling (green) are shown separately. The agreement with the experimental results (black triangles) is quite good.16 In contrast to our former work,16 the frequencydependent dielectric conductivity ϑ0(ω) (orange area in Fig. 3a-c) is not only computed from cosine exponential fits of 〈J(0) · J(t)〉 (see Eq. (5)) but also confirmed by the fit of the initial region of the mean-squared displacement ΔMJ2 which yields slower processes with higher accuracy.51 In particular, the major contribution to ϑ0 stems at all ion concentrations from the slowest process (k = 4) of 〈J(0) · J(t)〉 with a time constant τ4 of roughly 10 ps (see supplementary material). Since the amplitude of this contribution is rising with increasing ion concentration, ϑ0 also rises. The oscillatory behavior of 〈J(0) · J(t)〉, on the other hand, modelled by k = 1, 2, contributes only marginally to the static value ϑ0. Nevertheless, the characteristic oscillatory shape of 〈J(0) · J(t)〉 is mainly determined by the first (k = 1) cosine exponential of fJJ(t) and the slowest contribution (k = 4) is almost invisible at first inspection.51 Interestingly, the frequency ω1 ≃12.3 ps−1 of the k = 1 oscillation does not change with increasing ion content and corresponds to a wave number ν˜=65cm1 which is much lower than ν˜exp=94cm1 as observed for the pure liquid in experiments.62 This finding is interpreted as an uniform weakening of the cation-anion interactions by water molecules and points towards a stable an- ion/water network persisting at all concentrations cIL investigated here. The strength of these networks is known in literature.12,1923,59

Fig. 3.

Fig. 3

Decomposition of the frequency dependent dielectric spectra of [C4mim][BF4]-water mixtures at various IL concentrations. Pure water and cationic contributions to the dielectric permittivity ε(ω) are displayed as blue and red shaded areas, respectively. To some extent they overlap with the green shaded area representing the frequency-dependent ro-translational coupling εDJ(ω). The orange shaded areas correspond to the contribution of the dielectric conductivity ϑ0(ω). The experimental data (black triangles) are taken from Ref. 16.

At the lowest ion concentration cIL = 1.25M, the rotation of water molecules is responsible for more than 90% of the static dielectric constant Σ0 (see Fig. 3a). In contrast, at cIL = 3.77M the contributions of the water rotation, of the cationic rotation and of the dielectric conductivity ϑ0 are of comparable size as visible in Table II. Overall, the pure dielectric permittivity contribution of cations εCC seems to increase linearly with ΦIL as predicted by our geometric model and visible in Fig. 4. Both cross-terms, the rotational coupling between cations and water, εWC, as well as the ro-translational contribution εDJ go along with Φhyd in Eq. (20). This seems reasonable since the hydration water is the most important coupling partner for the ions. εWW is not displayed in Fig. 4 since the values are much higher compared to the other permittivity contributions with the exception of cIL = 3.77M. Moreover, since hydration and bulk water are contributing, the behavior of εWW when changing the IL concentration cannot be described by a function like Φbulk solely. Nevertheless, the overall trend is roughly decaying exponentially.

Table II.

Static values of various contributions to the generalized dielectric constant Σ0 = εCC + εWW + εWC + ϑ0 + 2εDJ. All ε-contributions emerge from rotational motions of the cations (C) and water (W). The anion is neglected here since it does not possess a permanent dipole. ϑ0 is the static dielectric conductivity and εDJ the static ro-translational coupling which can also be estimated by the predictions of Hubbard et al.30εHub) and Wolynes et al.34εWol)

ΦIL ε CC ε WW ε WC ϑ 0 2εDJ ΔεHub ΔεWol
0.000a 102.0
0.255 2.3 46.5 1.1 2.9 2.6 -9.2 -6.9
0.480 4.7 24.9 1.5 4.1 3.3 -12.7 -5.6
0.741 6.1 8.4 0.5 5.7 3.3 -10.7 -2.2
1.000b 8.2 4.6 -0.3
a

The value of pure TIP3P water model is reported in Ref. 48

b

The values of the pure [C4mim][BF4] is taken from Ref. 47.

Fig. 4.

Fig. 4

Static dielectric contributions to Σ0: The permittivity caused by cation-cation interactions εCC goes along with ΦIL whereas contributions between water and cations, εWC and εDJ follow Φhyd. In contrast, ΔεHub and ΔεWol are negative and do not show a clear trend. All static values can be found in Table II.

B. Electrolyte models

As a consequence of a decaying εWW, electrolyte theories such as the Debye-Hückel model or the dressed ion theory seem not appropriate to describe the depolarization since major contributions come from the depolarization of solvent molecules and are not due to pure inter-ionic interactions.37,54,55 Regarding the kinetic decrement, we noticed that, although the relaxation con-stant τD for the collective water rotation is increasing with increasing cIL (see supplementary material), the static electric conductivity σ(0) is simultaneously decreasing. This leads to a minimum of ΔεHub instead of the expected monotonic behavior (see Fig. 4). A monotonic behavior is recovered by ΔεWol due to the correlation of the vector connecting the ion with a water molecule rij and its dipole μW (c.f. Eq. (8)) but the trend goes in the wrong direction yielding a dielectric increase with increasing ion concentration. Furthermore, both values should be comparable to εDJ which has opposite sign and is slightly smaller. This is not only the case for [C4mim][BF4]/water mixtures but also for [C2mim][OTf]/water mixtures as well.18 Interestingly, the ro-translational contribution of pure ILs is negative.47

This discussion shows that kinetic depolarization models of Hubbard and Wolynes are not compatible with the dielectric decrement in case of our IL mixtures. The most obvious reason for this failure is the high concentration of ions cIL > 1M of the mixtures investigated here which are far from the very dilute concentrations assumed in both models. The interaction between the cations and anions with each other seems to be so strong in our case that the intuitive picture of the kinetic depolarization of water molecules around a single ion does not apply anymore. The hydration shells of the ions overlap to some extent due to the agglomeration of ions resulting in a non-homogeneous mixture which cannot be described by the continuum model of Hubbard. Moreover, ion aggregation seems common in our mixtures. We reported in Ref.57 mean residence times of cation-anion contacts of 64.0, 147.7 and 454.8 ps for cIL=1.25, 2.52 and 3.77 M, respectively. If charged or neutral clusters of ions move in one direction for this period of time, the rotation of neighboring water dipoles cannot be correlated with the electric current since cations and anions moving in the same direction prefer different water orientations. Consequently, the term in brackets in Eq. (8) will be oscillating around zero.

However, Levy et al.38 proposed a dipolar Poisson- Boltzmann theory concerning electrolyte solutions at higher ion concentrations. Here, the dielectric constant of water increases with increasing radial distance from a central ion reaching the bulk value for very long distances as visible in the inset of Fig. 1. However, the mean separation between two ions sets the maximum distance of water from any ion. Using spherical integration in Eq. (14) an average dielectric constant of water 〈ε(cIL)) in the proximity of ions can be computed. Since our molecular ionic liquids occupy a significant volume as pointed out in the Theory section, the overall dielectric constant Σ0 of the mixture may be estimated by

Σ0(1ΦIL)ϵ(cIL)+ΦILϵIL (21)

with εIL = 12.5 reported in Ref.47 The corresponding dependence of Σ0 on the volume occupancy ΦIL of the ionic liquid is depicted in Fig. 5 (green dots) and agrees fairly with the computed Σ0 (black crosses). So far, we used no fit parameters. Furthermore, this theory concurs nicely with our previous finding12 that the overall effective electric field reduces the interaction between cations and anions and not interstitial water molecules. In the previous publication we observed the effect of the water on the ions and in the present work the effect of the ions on the water. Our findings also agree with the results of Kobrak:63 The solute-solvent interactions are not governed by single ions, their steric structure or specific interactions with water. It is the ion density, which enters our considerations by their occupied volume ΦIL and by the upper limit of the spherical integration of ε(r), that yields 〈ε(cIL)〉.

Fig. 1.

Fig. 1

Dielectric constant 〈ε(cIL)〉 of water in the proximity of a point-size ion according to the dipolar Poisson-Boltzmann theory. The line at r=1.1Å corresponds to 〈ε(cIL)〉 = 40.2 = εhyd obtained from a spherical averaging (c.f. Eq. (14)) and indicated by the gray shaded area.

Fig. 5.

Fig. 5

Computational static dielectric constant Σ0 of [C4mim][BF4] / TIP3P mixtures and their agreement with various models discussed in the text.

C. Mixtures of polar solvents

In his review, Kaatze64 pointed out that strong local interactions lead to a failure of kinetic depolarization models and theories concerning bound water may be more applicable. For example, strong hydrogen bonds between the water molecules and the anions are known.12,1923,59 Consequently, one may think of two different water species in the electrolyte solution: hydration water which is strongly influenced by the IL ions and bulk water which is not. As mentioned in the Theory section (see Eq. (15) and (16)), the respective volume fractions Φhyd and Φbulk can be extrapolated from the amplitudes of MDW(0)MDW(t).

We now assume that the rotation of hydration water is different from bulk behavior due to the presence of the ions resulting in a different local permittivity εhyd. The overall dielectric constant of the electrolyte solution can be estimated by the average of the local dielectric constants of hydration water (εhyd), bulk water (εbulk = εW) and the ionic liquid (εIL) weighted by their volume occupancies

Σ0Φhydϵhyd+Φbulkϵbulk+ΦILϵIL. (22)

The only remaining fit parameter to match the static values of the generalized dielectric constant in Fig. 5 is the dielectric constant of the hydrated water molecules εhyd. Using a value of εhyd = 40.2 results in the pentagons in Fig. 5 which fit perfectly the computed Σ0-values. The hydration model relies on the computed Φhyd, Φbulk and ΦIL at the discrete given concentrations. In order to predict the volume occupancies at any concentration, they may be estimated by the geometric model in Eq. (20). The volume occupancy of the ionic liquid depends linearly on cIL, thus the exponential part of Eq. (20) represents Φbulk. Using this geometric model, the discrete values of the dielectric constant Σ0 (pentagons in Fig. 5) turn into a continous prediction shown as black dashed line.

The lower value of εhyd compared to εbulk seems reasonable since hydration water is expected to relax slower65, 66 and mutual position and orientation of water molecules are changed.12 This simple hydration picture assumes that only the water molecules are affected by the presence of the cations and anions, but the presence of the water molecules does not affect the rotational, εCC, and translational, ϑ0, contributions of the ions, since for instance εCC increases linearly with increasing ΦIL as shown by the red dashed line in Fig. 4. In principle, this simple hydration model differs not so much from the dipolar Poisson-Boltzmann approach: The hydration model varies the volume fraction as a function of the IL concentration (see Fig. 2) and keeps the local dielectric constants εhyd, εbulk and εIL constant. However, in the dipolar Poisson-Boltzmann model38 the water is not discrimated into hydrated and bulk water but the dielectric constant 〈ε(cIL)〉 varies continously as a function of the distance from the ion. Comparing both methods, the dielectric permittivity εhyd equals 〈ε(cIL)〉 for a distance of 1.1 Å depicted as gray area in the inset of Fig. 1. This distance is slightly less than the thickness of a complete hydration shell and may be reasonably explained by overlapping hydration shells in our IL mixtures.

The interpretation in terms of rotational depolarization is also endorsed by the success of the empirical logarithmic partitioning of Pujari et al.42 which was applied to aqueous mixtures of organic solvents and is depicted as red dash-dotted line in Fig. 5. They also suggested an additional correction term taking into account the polarity of the organic solvent

ln(Σ0)=(1ΦIL)ln(ϵW)+ΦILln(ϵIL)d(1ΦIL)ΦIL (23)

It turns out that our data results in a d-value of -0.08 which is comparable to methanol or ethanol. Interestingly, these alcohols share almost the same static dielectric constant of the pure [C4mim][BF4]. Hence, the aqueous IL mixtures investigated here seems to behave like common mixtures of polar solvents. Other models3941 describing these mixtures of polar solvents (see supplementary material for the corresponding figure) follow the overall trend of the decreasing dielectric constant but lack strong quantitative agreement. Moreover, all these models are intrinsically decaying with increasing ion concentration. However, experimental data on [C4mim][BF4] mixtures in dichloromethane56 and acetonitrile67 show a dielectric maximum as functions of mole fraction. This finding cannot be explained with those models but may be reproduced with a hydration model with εhyd > εbulk or two different hydration shells with different properties around cations and anions, respectively.

V. Conclusion

Although kinetic depolarization theories may be successful for very dilute electrolytes,3034 it seems not valid anymore in case of aqueous ionic liquid mixtures with moderate to high ion concentrations of cIL >1M. In fact, the simulated ro-translational coupling εDJ in this work turns out to have positive values due to strong interactions between the ions themselves and with water and consequently contradicts these models for dielectric decrement of electrolyte solutions.

Moreover, the εDJ contribution to the static dielectric constant shows a parabolic behavior as a function of the ion concentration which is similar to the prediction of a simple geometric model computing the volume occupancy Φhyd of water in the first solvation shell of the ions. However, not only the coupling between the translational motion of the ions and the rotation of the water molecules show the same behavior as Φhyd, but also the coupling between the rotation of the cations and the rotation of water. These findings point to a picture of rotational depolarization with strong dipolar interactions between the ions and the adjacent water molecules. Consequently, prediction models such as the hydration model or the dipolar Poisson-Boltzmann model38 prove more applicable.

In principle, this change from kinetic to rotational depolarization corresponds to a shift of the point of view on aqueous ionic liquids: due to the large volume occupied by the ions and the strong interaction with water and between the ions, these mixtures do not behave like classical electrolyte solutions at high salt concentrations anymore but have to be considered as mixtures of polar solvents.

Supplementary Material

Supplementary Material

VI. Acknowledgment

The computational work was performed at the Vienna Scientific Cluster of the University of Vienna, the Vienna University of Technology and the University of Natural Resources and Applied Life Science, Vienna. We thank for generous allocation of computer time. This work was supported by FWF Project No. P23494 and the European project FP7 331932. The authors would like to thank R. Buchner, H. Weingärtner and L.M. Varela for very helpful discussions.

See supplementary material at [URL will be inserted by AIP] for additional information on the fitting parameters of the corresponding correlation functions. Furthermore, additional information on empirical models of the dielectric constant for polar mixtures are displayed.

References

  • 1.Kohno Y, Ohno H. Chem Commun. 2012;48:7119. doi: 10.1039/c2cc31638b. [DOI] [PubMed] [Google Scholar]
  • 2.Endres F, Abedin SZE. Phys Chem Chem Phys. 2006;8:2101. doi: 10.1039/b600519p. [DOI] [PubMed] [Google Scholar]
  • 3.Baker GA, Baker SN. Aust J Chem. 2005;58:174. [Google Scholar]
  • 4.Seddon KR, Stark A, Torres M-J. Pure Appl Chem. 2000;72:2275 [Google Scholar]
  • 5.Vila J, Ginés P, Rilo E, Cabeza O, Varela LM. Fluid Phase Equilibria. 2006;247:32. [Google Scholar]
  • 6.Rilo E, Pico J, García-Garabal S, Varela LM, Cabeza O. Fluid Phase Equil. 2009;285:83. [Google Scholar]
  • 7.Varela LM, Carrete J, García M, Gallego LJ, Turmine M, Rilo E, Cabeza O. Fluid Phase Equil. 2010;298:280. [Google Scholar]
  • 8.Zhong X, Fan Z, Liu Z, Cao D. J Phys Chem B. 2012;116:3249. doi: 10.1021/jp3001543. [DOI] [PubMed] [Google Scholar]
  • 9.Leskiv M, Bernardes CES, Minas da Piedade ME, Canongia Lopes JN. J Phys Chem B. 2010;114:13179. doi: 10.1021/jp1067473. [DOI] [PubMed] [Google Scholar]
  • 10.Bernardes CES, Minas da Piedade ME, Canongia Lopes JN. J Phys Chem B. 2011;115:2067. doi: 10.1021/jp1113202. [DOI] [PubMed] [Google Scholar]
  • 11.Moreno M, Castiglione F, Mele A, Pasqui C, Raos G. J Phys Chem B. 2008;112:7826. doi: 10.1021/jp800383g. [DOI] [PubMed] [Google Scholar]
  • 12.Schröder C, Rudas T, Neumayr G, Benkner S, Steinhauser O. J Chem Phys. 2007;127:234503. doi: 10.1063/1.2805074. [DOI] [PubMed] [Google Scholar]
  • 13.Jiang W, Wang Y, Voth GA. J Phys Chem B. 2007;111:4812. doi: 10.1021/jp067142l. [DOI] [PubMed] [Google Scholar]
  • 14.Feng S, Voth GA. Fluid Phase Equil. 2010;294:148. [Google Scholar]
  • 15.Zhang X-X, Liang M, Hunger J, Buchner R, Maroncelli M. J Phys Chem B. 2013;117:15356. doi: 10.1021/jp4043528. [DOI] [PubMed] [Google Scholar]
  • 16.Schröder C, Hunger J, Stoppa A, Buchner R, Steinhauser O. J Chem Phys. 2008;129:184501. doi: 10.1063/1.3002563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Schröder C, Neumayr G, Steinhauser O. J Chem Phys. 2009;130:194503. doi: 10.1063/1.3127782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Schröder C, Steinhauser O. J Chem Phys. 2010;132:244109. doi: 10.1063/1.3432620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Cammarata L, Kazarian SG, Salter PA, Welton T. Phys Chem Chem Phys. 2001;3:5192 [Google Scholar]
  • 20.Köddermann T, Wertz C, Heintz A, Ludwig R. Angew Chem Int Ed. 2006;45:3697. doi: 10.1002/anie.200504471. [DOI] [PubMed] [Google Scholar]
  • 21.Spickermann C, Thar J, Lehmann SBC, Zahn S, Hunger J, Buchner R, Hunt PA, Welton T, Kirchner B. J Chem Phys. 2008;129:104505. doi: 10.1063/1.2974098. [DOI] [PubMed] [Google Scholar]
  • 22.Weingärtner H. Angew Chem Int Ed. 2008;47:654. doi: 10.1002/anie.200604951. [DOI] [PubMed] [Google Scholar]
  • 23.Buchner R. Pure Appl Chem. 2008;80:1239 [Google Scholar]
  • 24.Born M. Zeitschrift f Phys. 1920;1:221. [Google Scholar]
  • 25.Debye P, Hückel E. Physik Z. 1923;24:185. [Google Scholar]
  • 26.Hasted JB, Ritson DM, Collie CH. J Chem Phys. 1948;16:1. [Google Scholar]
  • 27.Haggis GH, Hasted JB, Buchanan TJ. J Chem Phys. 1952;20:1452 [Google Scholar]
  • 28.Stillinger FH, Jr, Lovett R. J Chem Phys. 1968;48:3858 [Google Scholar]
  • 29.Stillinger FH, Jr, Lovett R. J Chem Phys. 1969;49:1991 [Google Scholar]
  • 30.Hubbard J, Onsager L. J Chem Phys. 1977;67:4850 [Google Scholar]
  • 31.Hubbard JB. J Chem Phys. 1978;68:1649 [Google Scholar]
  • 32.Wolynes PG. J Chem Phys. 1978;68:473. [Google Scholar]
  • 33.Colonomos P, Wolynes PG. J Chem Phys. 1979;71:2644 [Google Scholar]
  • 34.Hubbard JB, Colonomos P, Wolynes PG. J Chem Phys. 1979;71:2652 [Google Scholar]
  • 35.Ennis J, Kjellander R, Mitchell DJ. J Chem Phys. 1995;102:975. [Google Scholar]
  • 36.Song X. J Chem Phys. 2009;131:044503. doi: 10.1063/1.3187147. [DOI] [PubMed] [Google Scholar]
  • 37.Xiao T, Song X. J Chem Phys. 2011;135:104104. doi: 10.1063/1.3632052. [DOI] [PubMed] [Google Scholar]
  • 38.Levy A, Andelman D, Orland H. Phys Rev Lett. 2012;108:227801. doi: 10.1103/PhysRevLett.108.227801. [DOI] [PubMed] [Google Scholar]
  • 39.Looyenga H. Physica. 1965;31:401. [Google Scholar]
  • 40.Böttcher CJF, Bordewijk P. Theory of electric polarization. Vol. 2. Elsevier; Amsterdam: 1978. [Google Scholar]
  • 41.Kraszewski A, Kulinksi S, Matuszewski M. J Appl Phys. 1976;47:1275 [Google Scholar]
  • 42.Pujari BR, Barik B, Behera B. Phys Chem Liq. 1998;36:105. [Google Scholar]
  • 43.Buchner R, Hölzl C, Stauber J, Barthel J. Phys Chem Chem Phys. 2002;4:2169 [Google Scholar]
  • 44.Koeberg M, Wu C-C, Kim D, Bonn M. Chem Phys Lett. 2007;439:60. [Google Scholar]
  • 45.Bešter-Rogač M, Hunger J, Stoppa A, Buchner R. J Chem Eng Data. 2011;56:1261 [Google Scholar]
  • 46.Bešter-Rogač M, Stoppa A, Hunger J, Hefter G, Buchner R. Phys Chem Chem Phys. 2011;13:17588. doi: 10.1039/c1cp21371g. [DOI] [PubMed] [Google Scholar]
  • 47.Schröder C, Haberler M, Steinhauser O. J Chem Phys. 2008;128:134501. doi: 10.1063/1.2868752. [DOI] [PubMed] [Google Scholar]
  • 48.Braun D, Boresch S, Steinhauser O. J Chem Phys. 2014;140:064107. doi: 10.1063/1.4864117. [DOI] [PubMed] [Google Scholar]
  • 49.Schröder C, Wakai C, Weingärtner H, Steinhauser O. J Chem Phys. 2007;126:084511. doi: 10.1063/1.2464057. [DOI] [PubMed] [Google Scholar]
  • 50.Schröder C, Steinhauser O. J Chem Phys. 2009;131:114504. doi: 10.1063/1.3220069. [DOI] [PubMed] [Google Scholar]
  • 51.Schröder C. J Chem Phys. 2011;135:024502. doi: 10.1063/1.3601750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Schröder C, Steinhauser O. J Chem Phys. 2010;133:154511. doi: 10.1063/1.3493689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Schröder C, Rudas T, Steinhauser O. J Chem Phys. 2006;125:244506. doi: 10.1063/1.2404674. [DOI] [PubMed] [Google Scholar]
  • 54.Kjellander R, Mitchell DJ. J Chem Phys. 1994;101:603. [Google Scholar]
  • 55.Hansen JP, McDonald IR. Theory of Simple Liquids. 3rd ed. Elsevier; 2006. [Google Scholar]
  • 56.Hunger J, Stoppa A, Buchner R, Hefter G. J Phys Chem B. 2008;112:12913. doi: 10.1021/jp8045627. [DOI] [PubMed] [Google Scholar]
  • 57.Neumayr G, Schröder C, Steinhauser O. J Chem Phys. 2009;131:174509. doi: 10.1063/1.3256003. [DOI] [PubMed] [Google Scholar]
  • 58.Bowers J, Butts CP, Martin PJ, Vergara-Gutierrez MC, Heenan RK. Langmuir. 2004;20:2191. doi: 10.1021/la035940m. [DOI] [PubMed] [Google Scholar]
  • 59.Singh T, Kumar A. J Phys Chem B. 2007;111:7843. doi: 10.1021/jp0726889. [DOI] [PubMed] [Google Scholar]
  • 60.Rollet A-L, Porion P, Vaultier M, Billard I, Deschamps M, Bessada C, Jouvensal L. J Phys Chem B. 2007;111:11888. doi: 10.1021/jp075378z. [DOI] [PubMed] [Google Scholar]
  • 61.Marcus Y, Hefter G. Chem Rev. 2006;106:4585. doi: 10.1021/cr040087x. [DOI] [PubMed] [Google Scholar]
  • 62.Fumino K, Wulff A, Ludwig R. Phys Chem Chem Phys. 2009;11:8790. doi: 10.1039/b905634c. [DOI] [PubMed] [Google Scholar]
  • 63.Kobrak MN. J Chem Phys. 2007;127:184507. doi: 10.1063/1.2790425. [DOI] [PubMed] [Google Scholar]
  • 64.Kaatze U. J Solution Chem. 1997;26:1049 [Google Scholar]
  • 65.Kropman MF, Bakker HJ. Science. 2001;291:2118. doi: 10.1126/science.1058190. [DOI] [PubMed] [Google Scholar]
  • 66.Omta AW, Kropman MF, Woutersen S, Bakker HJ. Science. 2003;301:347. doi: 10.1126/science.1084801. [DOI] [PubMed] [Google Scholar]
  • 67.Stoppa A, Hunger J, Hefter G, Buchner R. J Phys Chem B. 2012;116:7509. doi: 10.1021/jp3020673. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Material

RESOURCES