Modern Room Temperature Ionic Liquids, a Simple Guide to

Aug 5, 2015 - Piper , D. M.; Evans , T.; Leung , K.; Watkins , T.; Olson , J.; Kim , S. C.; Han , S. S.; Bhat , V.; Oh , K. H.; Buttry , D. A. Stable ...
0 downloads 0 Views 8MB Size
Feature Article pubs.acs.org/JPCB

Modern Room Temperature Ionic Liquids, a Simple Guide to Understanding Their Structure and How It May Relate to Dynamics Juan C. Araque, Jeevapani J. Hettige, and Claudio J. Margulis* Department of Chemistry, University of Iowa, Iowa City, Iowa 52242, United States ABSTRACT: Modern room temperature ionic liquids are structurally defined by symmetries on different length scales. Polar−apolar alternation defines their nanoscale structural heterogeneity, whereas positive−negative charge alternation defines short length scale order. Much progress has been made in the past few years as it pertains to the theoretical interpretation of X-ray scattering experiments for these liquids. Our group has contributed to the development of theoretical interpretation guidelines for the analysis of their structure function. Perhaps less well developed is our understanding of how transport and dynamics in general couple to the very unique structure of ionic liquids which are often dynamically and structurally heterogeneous. This article attempts to present our most current understanding of ionic liquid structure in general and its coupling to transport and dynamics in minimally technical terms for the benefit of the broadest audience.

1. INTRODUCTION The past 10 years have witnessed tremendous expansion in the design and synthesis of modern room-temperature ionic liquids (ILs). We now have task specific ionic liquids,1−11 ionic liquid drugs,12−16 ionic liquids for energy applications such as batteries and supercapacitors,17−31 those that capture carbon dioxide,3,32−38 and those that dissolve cellulose39−44 and proteins.45−49 In terms of their structure, interesting work has surfaced on protic ionic liquids,50−59 ILs that are highly fluorinated,59−63 ILs with polar tails,64−67 ILs with paramagnetic ions,68,69 those containing silicon,70−74 and some that are triphillic59,61−63 just to mention a few examples. Not all ionic liquids but a vast number of them can be categorized as having both charged components and also apolar components.75 The balance between charge and other polar or apolar components defines properties such as viscosity, melting points, solubility, hydrophilicity, and nanoscale structure. Neutron56,76−84 and X-ray66,78,79,83,85−107 scattering experiments as well as computer simulations63−65,67,85,87,96,103,108−132 have provided tremendous insight into the structure of ionic liquids. One of the areas in which our group has contributed is in deriving some basic rules for the analysis of X-ray data. For the most part, such rules should also apply to neutron scattering which when used in combination with partial deuteration becomes a most powerful experimental technique.76,77,84,133 Because ILs are ultimately molten salts, their underlying structure is governed to a large extent by charge alternation. At larger distances (lower q values), many ILs display what is commonly known as a prepeak or a first sharp diffraction peak. In systems where this is observed, the phenomenon arises because charge networks are necessarily separated from other charge networks by apolar components. As the complexity of ILs increases, other features beyond charge and polarity alternation © 2015 American Chemical Society

manifest including apolar-fluorinated alternation and the analysis of experiments becomes very complex without the guide of detailed simulations. Because ILs have structural features on different length scales, they are at the molecular level structurally heterogeneous.50,56,61,88,90,95,102,126,134−137 They are also known to be dynamically heterogeneous;138−142 chemical143,144 and spectroscopic139,145−149 signatures of such heterogeneity are ubiquitous in the literature. More recently, a different manifestation of heterogeneity has been observed when probing solute diffusion.150,151 Motion of charged solutes can be much slower than hydrodynamic predictions and that of neutral solutes much faster depending on the volume of the solute as compared to that of the solvent. Are anomalous solute mobility and solvent structural heterogeneity related? If so, what is the underlying mechanistic connection? In sections 2−5, we describe our most current understanding of IL structure, and in section 6, we discuss connections between structure and dynamics.

2. THE BASICS ABOUT IONIC LIQUID STRUCTURE It is easy to tell a generic ionic liquid from many non-hydrogenbonding conventional solvents by simple inspection of the structure function (S(q)) defined in terms of the coherent X-ray scattering intensity Icoh (eq 1). In eq 1, xi and f i(q) correspond to the mole fraction and X-ray form factor152 of atoms of type i. S(q) =

Icoh(q) − ∑i xi fi 2 (q) [∑i xi fi (q)]2

(1)

Received: June 9, 2015 Revised: July 20, 2015 Published: August 5, 2015 12727

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

lower q values. The one at the lowest q value, the prepeak, is indicative of polar−apolar alternation and has been linked to structural heterogeneity. The peak at intermediate q values, higher than the prepeak but lower than the adjacency peak, is the molten salt signature and indicates charge alternation. A missing prepeak is often indicative of small apolar components and a lack of polar−apolar alternation. A missing intermediate peak is likely due to cancellations because at room temperature charge alternation is always present. Charge and polarity alternation are not unique to the liquid phase and can be clearly identified in crystal structures. Examples of this are depicted in Figure 2. In fact, one can think of ILs as having lost most crystal symmetries except for charge and polarity alternation.112

Figure 1a displays computationally derived S(q) for an eclectic set of imidazolium, alkylammonium, 64,85,113 phospho-

3. PEAKS AND ANTIPEAKS IN PARTIAL STRUCTURE FUNCTIONS ARE SIMPLE INDICATORS OF ALTERNATION ON THE MOLECULAR LANDSCAPE The experimental S(q) carries all information about the liquid landscape, but it does so in a way that is difficult to dissect, since its nature is collective. In contrast, in computer simulations, the situation is the opposite as we construct the collective S(q) from specific atomic pair information. This is done either as a sum in real space that gets Fourier transformed (eq 2) S(q) = ρo ∑i ∑j xixj fi (q) f j (q) ∫

0

L /2

4πr 2(gij(r ) − 1)

sin qr W (r ) qr

dr

[∑i xi fi (q)]2 (2)

or directly in reciprocal space (eq 3). In eq 2, gij(r) represents the radial distribution function between atomic species of type i and j, and a Lorch function W(r)154,155 is often utilized to mitigate finite truncation effects at large r. Angle brackets in eq 3 correspond to the ensemble average of atomic densities in reciprocal space defined in eq 4 where, for atomic species of types i and j, Ni and Nj are the total number of atoms.156 Each of these computational approaches provides different physical insight that will be discussed below.

Figure 1. Computationally derived S(q) for (a) a diverse collection of ionic liquids64,109,111,113 and (b) conventional solvents. Perhaps with the exception of methanol that forms hydrogen bonds, the conventional solvents show a single feature in the region of S(q) relevant to intermolecular interactions. Instead the ionic liquids show three features or peaks. The peak at larger q values is associated with adjacency correlations between neighboring atoms. These correlations are intermolecular and intramolecular in nature. The intermediate peak is the signature of charge alternation. The peak at the smallest q value is caused by polarity alternation and is associated with structural heterogeneity. Absence of a charge alternation peak as in the case of [Im1,8][BF4] does not mean this symmetry is absent; instead, it denotes perfect cancellations of peaks and antipeaks at this particular q value (see section 3).

S(q) =

ρqi =

1 ∑i ∑j fi (q)f j (q)⎡⎣ N ⟨ρqi ρ−j q ⟩ − xiδi , j ⎤⎦

[∑i xi fi (q)]2

(3)

Ni

∑ exp(−iq·rα) α=1 Nj

ρ−j q

nium,111,114 and pyrrolidinium109,110 based ILs. Figure 1b shows instead S(q) for a selected set of conventional liquids. It is clear from the comparison that, perhaps with the exception of methanol which forms hydrogen bonds, these conventional solvents only have one peak at q values below 2.5 Å−1, whereas the ionic liquids commonly have three peaks or features. The only peak in the case of conventional solvents and the peak at larger q values in the case of the ionic liquids account for a myriad of different correlations between adjacent atoms. These can be both intermolecular and intramolecular in nature. In the case of ILs, two extra peaks or shoulders are commonly observed at

=

∑ exp(iq·rβ) β=1

(4)

To learn about the specific liquid motifs that give rise to different features in S(q), eq 2 can be conveniently partitioned into subcomponents. Many partitionings are possible that give rise to identical S(q), but we learn the most when such partitionings are wavenumber dependent.85,108−110 As an example, to study low q phenomena, it is most natural to partition S(q) into contributions from polar−polar, polar−apolar, and apolar−apolar interactions disregarding the positive or negative nature of the charge. 12728

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

Figure 2. Crystal structure of [Im10,1][PF6]153 showing (a) polarity alternation across Miller plane (0, 0, 1) and (b) charge alternation across Miller plane (1, 1, −1). These two symmetries, charge and polarity alternation, that are present in the crystalline state persist in the liquid phase and become the defining features of ionic liquid morphology (see ref 112).

Figure 3. On the left panel are idealized radial distribution functions g(r) for a liquid with patterns of alternation. The f1(r) function corresponds to the case of same-type objects, whereas f 2(r), to different-type objects. For example, f1(r) may represent positive−positive (++) correlations and f 2(r) may represent positive−negative (+−) correlations. Because the periodicity of the two functions is identical but f1(r) and f 2(r) are offset by half a period, their Fourier transforms (right-hand side panel) have a peak at identical q value but of different sign. Such peaks are always positive for same-type correlations and negative for different-type correlations. These are what we call peaks and antipeaks in the partial subcomponents of S(q). See ref 109 and the associated Supporting Information.

SPolar − Polar(q) = S a − a(q) + S c Head − c Head(q) + S c Head − a(q) + S a − c Head(q)

Instead, to study shorter range order, it is most appropriate to

S Apolar − Apolar(q) = S c Tail − c Tail(q)

positive, positive−negative, and negative−negative type con-

partition S(q) in a way that highlights and distinguishes positive− tributions, since such alternation is the phenomenon taking place

SPolar − Apolar(q) + S Apolar − Polar(q) = S c Head − c Tail(q) + S c Tail − c Head(q) + S c Tail − a(q) + S a − c Tail(q)

at length scales shorter than polarity alternation but longer than simple adjacency correlations. 12729

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B When alternations in the liquid phase are present and a judicious partitioning of S(q) for a given wavenumber region is chosen, what we have coined as peaks and antipeaks naturally manifest as the hallmark of such alternations in the partial subcomponents of S(q). The rationale for this is simple; seen from an arbitrary origin, alternating liquid components have the same periodicity but a phase offset. This leads to peaks at the same wavenumber but with different signs. Figure 3 demonstrates this for a perfectly idealized system for which “same-type” components have a periodic real space separation of 2π and are intercalated by “opposite-type” species. Parts a and b of Figure 4 show the same phenomenon when considering realistic polarity and charge alternations in an IL. In the polarity alternation region, same type corresponds to polar−polar and apolar−apolar correlations, whereas, in the charge alternation region, same type corresponds to positive−positive and negative−negative correlations. For the proverbial IL in which the anion is small and the cation has charged and apolar subcomponents, there is a simplest and most convenient way to analyze charge and polarity alternations in the liquid phase.110,113 The cation head−anion subcomponent of S(q) is a “natural variable” both in the charge alternation regime as well as in the polarity alternation regime. A plot of this function alone tells us most of what we need to know in both wavenumber ranges. This is easy to understand; cation heads and anions correspond to the same group when discussing polarity alternation, as they both form part of the polar network that alternates with apolar components of the liquid at low q values. Instead, cation heads and anions form part of opposite groups when considering charge alternation at intermediate q values. In Figure 4c, we see that same-type elements are on average periodically separated by a distance nλ (λ is large for polarity alternation and small for charge alternation), whereas differenttype elements are separated by a distance (n + 1/2)λ. At low q, because they are same-type liquid elements, cation heads and anions contribute a peak to the overall S(q) (because of the same reason, in the same q region, apolar components also contribute a peak, but polar−apolar components contribute an antipeak). At intermediate q values (larger than polarity alternation but smaller than simple adjacency correlations), cationic head−head correlations contribute a peak, anion−anion correlations contribute a peak, and cationic head−anion correlations contribute an antipeak to the overall S(q). Simply put, a plot of the cationic head−anion subcomponent of S(q) often shows a peak at the low q range where polarity alternation occurs and always shows an antipeak where charge alternation occurs. Examples of this for two different ILs of the pyrrolidinium series are depicted in Figure 5. No peak in the low q region means that the apolar subcomponent is not signif icant enough to allow for polar−apolar alternations and is often interpreted as indicative of an absence of polar−apolar structural heterogeneity ([Pyrr1,4][NTf2] in Figure 5). Whereas the feature corresponding to polarity alternation can be missing, we have never observed a situation in which the charge alternation antipeak is absent in a prototypical IL. Charge alternation is the signature of a room-temperature molten salt. This is written with the caveat that such an antipeak in the case of protic ILs can look very different in shape and wavenumber, as now it reflects hydrogen bonding. Interestingly, whereas peaks and antipeaks in the charge alternation region are prominent in partial subcomponents of S(q), it is often the case that the total S(q) only shows a small shoulder; in other cases, not even that. Such a situation is due to complete interference cancellation and not caused by the absence of charge alternation

Figure 4. (a) (+/+), (+/−), and (−/−) subcomponents of S(q) for (2ethoxyethoxy)ethyltriethylphosphonium bis(trifluoromethylsulfonyl)amide64 exhibit peaks and antipeaks associated with charge alternation. Both (+/+) and (−/−) are same-type correlations and therefore result in peak contributions, whereas (+/−) are different-type correlations and result in an antipeak. Notice how (−/−) correlations are of large intensity; this is quite common because anions are often the most important reporters of X-ray structure in ILs. (b) Polar−polar, polar− apolar, and apolar−apolar subcomponents of S(q) for tetradecyltrihexylphosphonium bis(trifluoromethylsulfonyl)amide114 exhibit peaks and antipeaks associated with polarity alternation. Because they are sametype interactions, polar−polar and apolar−apolar correlations appear as peaks, whereas polar−apolar correlations appear as an antipeak. The polar contribution is large compared to the apolar contribution; anions often heavily weight in the polar contribution. (c) Schematic representation depicting the origin of peaks and antipeaks. Same-type species are separated by nλ periods, but different-type species are separated by (n + 1/2) λ periods. The 1/2λ offset is the cause for antipeaks in correlations of different-type species.

in the liquid (a typical example is the case of [Im1,8][BF4] in Figure 1a). So far, we have discussed the benefits of using eq 2 to interpret IL structure. The route for this approach involves generating all possible radial distribution functions between atomic species and their properly weighted Fourier transformed sum. There are many advantages to this procedure but some limitations as well. Radial distribution functions are collective functions that report all spatial correlations between atomic classes (all carbon against all fluorine, for example). However, what if, for example, one is interested in teasing out the contribution to S(q) of a certain 12730

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

in Figure 6a, the prepeak anomalously increases with increasing temperature. Parts b and c of Figure 6 show that indeed the

Figure 5. For typical ionic liquids where cationic components have charged and apolar components and anions are small and symmetrically charged, a plot of the cationic head−anion subcomponent of S(q) highlights the most important features associated with liquid structure. Because the cationic head and the anion belong to same-type elements in the region associated with polarity alternation, a peak is expected from a plot of this subcomponent of S(q). Instead, because the cationic head and anion belong to opposite-type groups, in the case of charge alternation, an antipeak is expected at lower q value. The figure shows exactly this in the case of two pyrrolidinium based ILs.109 The system with a long alkyl tail shows both a polarity alternation peak at low q and a charge alternation antipeak at larger q. Instead, in the case of the cation with a shorter alkyl tail, polarity alternation is almost completely abolished but charge alternation is still present as an antipeak at about 0.8 Å−1.

specific liquid pattern involving only some cationic carbon atoms in tails and fluorinated anionic components? This is much more easily achieved with a sum over specific atomic contributions in reciprocal space as in eq 3. Reference 63 shows in detail how starting from eq 3 one can seek to identify atomic pair contributions that at a wavenumber of interest q contribute largely to S(q) and that at the same time are consistent with the condition that the atomic distance is d = 2π/q. Examples of the use of the approach discussed in ref 63 will appear in section 5.

Figure 6. (a) Total S(q), (b) polar−polar subcomponent of S(q), and (c) apolar−apolar subcomponent of S(q) for [P14,6,6,6][NTf2] at different temperatures. At the prepeak region which is indicative of polarity alternation and structural heterogeneity, the temperature dependence is anomalous. One would expect that higher temperatures would result in a less intense first sharp diffraction peak in the top panel, but the opposite is true. As expected, the increase in temperature coincides with a more disorganized apolar subcomponent but with a more organized polar component. Since anions are very important reporters of structure in ILs, the signal due to the polar component overwhelms that of the apolar part. A larger prepeak does not mean in this case that apolar domains are better defined; instead, it is a reflection of the predominant role of anions in these types of experiments (see ref 113).

4. CATIONS DEFINE THE LANDSCAPE BUT ANIONS REPORT Cations used in ILs are often more structurally rich and asymmetric than their anionic counterparts. This is by no means a general rule, but it applies to a large number of ILs. Cations are often the species that contribute both polar and apolar components, whereas anions are frequently smaller and of higher symmetry and charge is more evenly distributed. It is therefore reasonable to envision cation heads and tails as defining the landscape of polarity alternation. X-ray diffraction experiments probe electronic density which manifests in eq 2 via the atomic form factors. Because of their higher electronic density, anions tend to be better scatterers than cations. It is not uncommon for two systems that have similar pair correlation functions to have significant differences in S(q) depending on the identity of the anion.113 In other words, cations may determine the liquid landscape, but we only learn about them from the scattering of the less structurally rich anions. An example where this clearly manifests is in the case of the temperature dependence of the prepeak in long-tail phosphonium ILs.111,114 Since nanoscale order in the form of polarity alternation is expected to decrease with an increase in temperature, so is the prepeak intensity. In reality, as is shown

apolar contribution to S(q) diminishes as temperature increases; the loss of organization of the apolar component allows for a moderate increase in organization of the charged subcomponent. The apolar subcomponent disorder decreases the prepeak intensity, but the enhancement in positive−negative interactions results in an intensity increase. Why do these two opposite effects not cancel? The answer is simple: Because of their X-ray scattering form factors, anions tend to be the most prominent reporters of structure. This is why the enhancement to the purely polar component of the prepeak dominates what is seen in X-ray experiments. In this case, a more intense prepeak does not imply 12731

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

Figure 7. (a) Comparison of S(q) for phosphonium and ammonium isoelectronic analogues where the cation has either alkyl or ether containing tails. (b) Anionic first solvation shell around an ether-based cation shown in terms of a spatial density distribution. When comparing the ether substituted analogues with their isoelectronic alkyl based systems, we find that the prepeak indicative of polarity alternation and polar−apolar structural heterogeneity is completely abolished. This can be understood from part b where cationic tails curl and anions cap both ends of the cation. Anions close to the more polar tail occupy the space that apolar subcomponents would have taken in the case of the alkyl based cation. Because of this, polarity alternation is disrupted and the symmetry reflected as a prepeak in S(q) is completely absent from these ionic liquids with tails that are more polar (see ref 64).

the assignment of the prepeak on the center panel in Figure 8 as well as in the rendering in the bottom panel. Other features in S(q) highlight apolar-fluorinated cationic tail−anionic tail alternations and hydrogen bonding as well as adjacency correlations that are nonperiodic. To tease out detailed information on liquid patterns in complex systems like this one, the method based on Fourier transforms of pair distribution functions (eq 2) is less advantageous than the direct sum in reciprocal space in eq 3. In ref 63, we proposed a methodology to highlight patterns of atomic interactions in the liquid phase that not only contribute significantly at a given q value to the overall S(q) but that are also resonant with the Bragg condition d = 2π/q. Such a methodology has allowed us to assign all features in S(q), as highlighted in the top and center panels in Figure 8.

a better defined apolar domain. The opposite is true. The anomalous behavior is instead associated with the highly weighted role that anions play on scattering. In other words, if you want to learn about the cations, pay attention first to the anions.

5. BREAKING THE CONVENTIONAL PARADIGM. BEYOND APOLAR TAILS Ionic liquids do not have to be comprised of cations with polar heads and apolar tails combined with small symmetric anions. In fact, the paramount feature of ILs, charge alternation, can become hydrogen bonding in the case of protic ILs. Anions can have apolar or fluorinated tails, they can be large and asymmetric and cations can have tails that are polar or no tails at all. All of these possible changes and combinations dramatically alter the prototypical three-featured S(q) function. Figure 7a shows how when ethers replace alkyl tails charge alternation and adjacency correlations persist but polarity alternation characterized by a prepeak is completely vanished from S(q). Figure 7b shows that, when apolar tails are replaced by more polar ether containing tails, the morphology of said tails changes (tails curl) and anions efficiently solvate them. Since anions can solvate both the cationic charged head as well as the tail, the characteristic length scale corresponding to charged groups separated from other charged groups by apolar components disappears and so does the prepeak in S(q) at low q values.64 Among the most interesting systems that we have studied which break with the biphilic paradigm are protic ILs with large anionic fluorinated tails. 59 Some of these systems are continuously percolated by a network of hydrogen bonds. This network takes the form of continuous filaments decorated both by fluorinated anionic components and by alkyl cationic components (bottom of Figure 8). Several symmetries are present in the liquid phase. Because of charge alternation caused by hydrogen bonding along filaments, tails of cations and anions that decorateand are perpendicular tothe hydrogen bond filaments also alternate. This causes apolar-fluorinated alternation within each filament with interesting helical-like structures. The prepeak in this case highlights the typical separation between percolating hydrogen bond filaments spaced by decorating cationic and anionic tails. This is highlighted with

6. IONIC LIQUID STRUCTURE DEFINES SOLUTE DYNAMICS Do mechanisms of transport, anomalous diffusion, and dynamical heterogeneity have an underlying structural origin in ILs? Or are structural and dynamical heterogeneities fundamentally unrelated phenomena? A comparison between parts a and b of Figure 1 clearly highlights that many common ILs have at least two symmetries that are absent in most conventional solvents, one at shorter distances associated with charge alternation and the other at longer nanoscale distances (lowest q) linked to polar−apolar alternation. This lowest q feature associated with polarity alternation defines what is commonly associated with structural heterogeneity. Since the prepeak is indicative of nanoscale polarity alternations, one can envision regions of the liquid that are charge-dominated and others that are dominated by apolar interactions. It has been a long ongoing debate whether such structural heterogeneity may be related to other phenomena such as red edge effects commonly associated with dynamical heterogeneity or distributed dynamics in general.139 Recently, a set of intriguing articles157−168 discussing electron transfer reactivity at anomalous rates (too slow for charged species but too fast for neutral species) have surfaced in the literature. A comprehensive study150 on the mobility of neutral and charged species in comparison to hydrodynamic predictions indicates that when the molecular volume of the solute is small 12732

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

It is not difficult to see how stiff environments which are more polar and soft environments which are more apolar will also be linked to energy heterogeneities.169 Stiff or soft environments are associated with the distance to charge and should certainly affect the solvent contribution to the HOMO−LUMO gap for spectroscopic processes causing distributed dynamics and absorption wavelength dependent phenomena.139 The IL structure as well as the size and neutral or charged nature of the solute strongly couple to different mechanisms of diffusion. A pictorial explanation of the reason why Stokes−Einstein behavior fails for small solutes of different charge in ILs is presented in Figure 9. Figure 9a shows that, at 400 K, on a 600 ps time scale, a trajectory for methane in an ionic liquid with very modest apolar components is comprised of many cage and jump events linked to enhanced and diminished local friction. Cages and jumps have typical time scales much smaller than the overall trajectory which samples a significant volume. Instead, Figure 9b shows very different behavior in the case of ammonium where only a single jump event is observed on the same time scale and the overall volume sampled by the trajectory is much more modest. Both methane and ammonium have similar radii, and their diffusivity should be similar based on Stokes−Einstein predictions. We see that, as opposed to the case of methane, ammonium is not an innocent spectator of solvent environments but instead an integral part of the IL charge network. In general, the motion of ionic solutes is concerted with that of other ions in the liquid, apparent hydrodynamic radii become larger because of this, and the effective friction becomes much larger than could be predicted by the bulk viscosity. In general, liquid regions with local charge depletion are softer and more mobile than regions in which there is charge enhancement. However, the reader should understand that apolar tails are chemically bonded to charge and therefore it is not that the overall diffusion of apolar liquid components is faster. The softer nature of the more apolar liquid component plays an important dynamical role as it pertains to the dynamical coupling between neutral solute and solvent motion. In the case of neutral solutes, the fastest solute motion (solute jump) is associated with small variations in local density that are provided mostly by the soft apolar component. For example, apolar tails “flood” the space left behind by the diffusion of neutral solutes while also providing space for arrival at a new location. This is depicted in Figure 10a, where for a fixed distance of 1 Å the time dependent pair distribution function between the solute and different solvent components is shown. As the solute diffuses away, the vacated hole is promptly filled by cationic tails. Instead, charged liquid components only fill the hole on much longer time scales as memory of the presence of an apolar solute is lost. The dynamical coupling between charged solutes and solvent is dominated by charge−charge interactions, as the fastest motion is associated with the shedding of a significant portion of the solvation shell, the transport of the partial solvate, and the reassembly of a full solvation environment that is an integral part of the charge network.151 Figure 10b shows that the space left behind by small positively charged solutes is initially flooded by anions. These anions, which are associated with the solute, swipe the space it leaves behind as they follow its motion. Other liquid components only fill the hole on much longer time scales associated with randomization.

Figure 8. Assignment of structural features for BAOF IL (butylammonium pentadecafluorooctanoate). The top panel shows computationally derived S(q) and its assignment based on eq 3 and methodology described in ref 63. The prepeak is due to polar-fluorinated tail alternations, an intermediate peak due to alkyl cationic tail-fluorinated anionic tail alternations, and the peak at around 1.25 Å−1 due to hydrogen bonding associated with charge alternation. The bottom panel shows that this liquid is continuously percolated by a network of hydrogen bonded filaments which are decorated by tails. Because fluorinated tails are attached to negative charge and alkylated tails attached to positive charge that alternate, within each of the filaments the tails also alternate in a quasi-helical shape. The prepeak corresponds to the typical separation between adjacent filaments. Filaments are shown in yellow. Figure reprinted in part with permission from ref 63. Copyright 2014 American Chemical Society.

compared to that of the solvent deviations from Stokes−Einstein behavior are very large in ILs, more so than in other solvents.150 A recent study from our group151 has shed some light on the relation between anomalous dynamics and structure. A solute of comparable size to the solvent or smaller will experience local friction that is very different in different liquid locations. For example, a neutral gas molecule will be caged in regions of excess local charge and its mobility will be enhanced when in regions of charge depletion. In other words, a small neutral solute will be an innocent spectator of high and low electrostriction regions associated with stiff and soft environments that slow or enhance its mobility. In the case of small neutral solutes, this is at the origin of solute dynamical heterogeneity; structure defines dynamics. 12733

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

Figure 9. Comparison of 600 ps trajectories for (a) methane and (b) ammonium in [Pyrr1,4][NTf2] at 400 K. Part a shows that, in the case of methane, on this time scale, a trajectory is composed of many cage regimes separated by jumps. The overall volume visited by the probe is significant. Instead, part b shows that, on the same time scale, ammonium which has a very similar molecular volume undergoes a single jump event and visits a much smaller configuration space volume. Such disparity in diffusive behavior is common in ILs when the size of the tracer is smaller than the average size of the solvent. In such cases, diffusion predictions based on the bulk viscosity can fail by orders of magnitude. When a solute is small, its charge and volume become key components that define mechanisms of transport. Figure reprinted in part with permission from ref 151. Copyright 2015 American Chemical Society.

Figure 10. As a function of time, the probability of finding a solvent component (head, tail, or anion) at 1 Å from the position the solute held at time zero for (a) methane and (b) ammonium. In each case, color coded is a solute trajectory segment and a solvent interaction that qualitatively helps visualize a particular mechanism of transport. In the case of methane, we see from gmethane−tail(r − r0 = 1 Å, t − t0) that apolar tail components promptly fill the space left behind by the diffusing solute. This can also be qualitatively seen from the color coded trajectory, as a cationic tail initially opens up for solute passage and later closes filling the void left by the solute. In the case of ammonium, it is the anionic component that most promptly fills the space left by the diffusing cation. This can be seen quantitatively from gammonium−anion(r − r0 = 1 Å, t − t0) and qualitatively from the color coded trajectory coupled with one selected anion. Both parts a and b show that other liquid components only fill the hole left by the solute on a time compatible with randomization.151

12734

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B Biographies

7. CONCLUSIONS We see conventional solvents as often willing to forego most symmetries present in the solid state. Ionic liquids do too, except for polarity alternation and charge alternation. These two symmetries present both in the crystal and liquid phases are their hallmark and make them somewhat different from other liquids. When trying to interpret the structure of ILs via X-ray scattering, we learn that not all peaks are built the same way. Because polarity and charge symmetry are to an extent periodic, they are characterized by peaks and antipeaks in the partial subcomponents of S(q); instead, adjacency correlations are nonperiodic. From an experimental perspective, the analysis of X-ray liquid structure is difficult, as one is left to attempt an inversion of a single S(q) function to derive many pair correlations that interfere and often cancel out. The analysis is much facilitated when molecular simulations are available. A q-dependent partition of the simulated S(q) enormously simplifies the analysis, but in some cases, techniques involving direct sums in reciprocal space prove more advantageous particularly when trying to identify complex liquid structural patterns. Charge and polarity alternation not only dictate liquid structure but also affect local dynamics. If a solute is small, the friction felt in charge enhanced (stiff) and charge depleted (soft) regions can be very different. It is reasonable to expect soft and stiff environments (i.e., structure) to correlate with solvation environments of different local charge density which may be related to distributed chemical dynamics. This energy heterogeneity was recently highlighted by Hu and co-workers as correlated with dynamical heterogeneity of the type that gives rise to excitation wavelength dependent spectroscopy.169 From a transport perspective, small neutral solutes are to some extent innocent spectators of different stiff and soft frictional environments, whereas charged solutes become intrinsic components of the charge network. The motion of small charged solutes can be very slow, since their effective volume becomes very large. Instead, motion of small neutral solutes can be very fast. Because of this, solutes of similar volume can have vastly different diffusivities and significant deviation from Stokes− Einstein behavior can be expected. There is also strong coupling between solute motion and solvent motion. When small neutral solutes undergo fast motional jumps, it is the soft apolar component of the IL that fills the void left behind. By time reversal symmetry, it is the soft apolar components of the IL that vacate the space needed for fast neutral solute diffusion. Instead, a small cationic solute will see the space it leaves behind promptly filled by anions that swipe the space as they follow its motion. In other words, mechanisms of transport can show strong solute size and charge dependence. It would appear that ionic liquid structure and dynamics are strongly intertwined and one needs to understand one to fathom the other.



Juan Carlos Araque earned his B.S. degree in Chemical Engineering from the University of the Andes in Venezuela, and then worked for 2 years as a R&D engineer at PDVSA’s Petroleum Research Center. After completing his Ph.D. in Chemical Engineering at Rice University he did post-doctoral work at Cornell University. He is currently a post-doctoral scholar at the University of Iowa working on ionic liquids with emphasis on the relation between their structural and dynamical properties.

Jeevapani Hettige earned her B.S. in Chemistry at the University of Colombo, Sri Lanka, in 2008. Since 2010, she has been a Ph.D. student under the mentorship of Professor Claudio J. Margulis at the Department of Chemistry of the University of Iowa. Her research interests are mainly focused on investigating the nanoscale structure of ionic liquids using molecular dynamics simulations.

AUTHOR INFORMATION

Corresponding Author

Claudio J. Margulis is a professor of chemistry at the University of Iowa. He received his Licenciado en Ciencias Quı ́micas degree from the University of Buenos Aires; he later obtained his Ph.D. from Boston University working with David Coker. He did postdoctoral work at

*E-mail: [email protected]. Notes

The authors declare no competing financial interest. 12735

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

(15) Stoimenovski, J.; MacFarlane, D.; Bica, K.; Rogers, R. Crystalline vs. Ionic Liquid Salt Forms of Active Pharmaceutical Ingredients: A Position Paper. Pharm. Res. 2010, 27, 521−526. (16) Moniruzzaman, M.; Tamura, M.; Tahara, Y.; Kamiya, N.; Goto, M. Ionic Liquid-in-Oil Microemulsion as a Potential Carrier of Sparingly Soluble Drug: Characterization and Cytotoxicity Evaluation. Int. J. Pharm. 2010, 400, 243−250. (17) Garcia, B.; Lavallee, S.; Perron, G.; Michot, C.; Armand, M. Room Temperature Molten Salts as Lithium Battery Electrolyte. Electrochim. Acta 2004, 49, 4583−4588. (18) Lewandowski, A.; Świderska Mocek, A. Ionic Liquids as Electrolytes for Li-ion Batteries-An Overview of Electrochemical Studies. J. Power Sources 2009, 194, 601−609. (19) Diaw, M.; Chagnes, A.; Carré, B.; Willmann, P.; Lemordant, D. Mixed Ionic Liquid as Electrolyte for Lithium Batteries. J. Power Sources 2005, 146, 682−684. (20) Xu, K. Nonaqueous Liquid Electrolytes for Lithium-Based Rechargeable Batteries. Chem. Rev. 2004, 104, 4303−4418. (21) Sato, T.; Masuda, G.; Takagi, K. Electrochemical Properties of Novel Ionic Liquids for Electric Double Layer Capacitor Applications. Electrochim. Acta 2004, 49, 3603−3611. (22) Simon, P.; Gogotsi, Y. Materials for Electrochemical Capacitors. Nat. Mater. 2008, 7, 845−854. (23) Frackowiak, E.; Lota, G.; Pernak, J. Room-temperature Phosphonium Ionic Liquids for Supercapacitor Application. Appl. Phys. Lett. 2005, 86, 164104. (24) McEwen, A. B.; Ngo, H. L.; LeCompte, K.; Goldman, J. L. Electrochemical Properties of Imidazolium Salt Electrolytes for Electrochemical Capacitor Applications. J. Electrochem. Soc. 1999, 146, 1687−1695. (25) Hagiwara, R.; Lee, J. S. Ionic Liquids for Electrochemical Devices. Electrochemistry 2007, 75, 23−34. (26) Lin, M.; Gong, M.; Lu, B.; Wu, Y.; Wang, D.; Guan, M.; Angell, M.; Chen, C.; Yang, J.; Hwang, B.; et al. An Ultrafast Rechargeable Aluminium-ion Battery. Nature 2015, 520, 324−328. (27) Girard, G. M. A.; Hilder, M.; Zhu, H.; Nucciarone, D.; Whitbread, K.; Zavorine, S.; Moser, M.; Forsyth, M.; MacFarlane, D. R.; Howlett, P. C. Electrochemical and Physicochemical Properties of Small Phosphonium Cation Ionic Liquid Electrolytes with High Lithium Salt Content. Phys. Chem. Chem. Phys. 2015, 17, 8706−8713. (28) Piper, D. M.; Evans, T.; Leung, K.; Watkins, T.; Olson, J.; Kim, S. C.; Han, S. S.; Bhat, V.; Oh, K. H.; Buttry, D. A.; et al. Stable SiliconIonic Liquid Interface for Next-Generation Lithium-Ion Batteries. Nat. Commun. 2015, 6, 6230. (29) Matsui, Y.; Yamagata, M.; Murakami, S.; Saito, Y.; Higashizaki, T.; Ishiko, E.; Kono, M.; Ishikawa, M. Design of an Electrolyte Composition for Stable and Rapid Charging-Discharging of a Graphite Negative Electrode in a bis(fluorosulfonyl)imide Based Ionic Liquid. J. Power Sources 2015, 279, 766−773. (30) Park, J.; Ueno, K.; Tachikawa, N.; Dokko, K.; Watanabe, M. Ionic Liquid Electrolytes for Lithium-Sulfur Batteries. J. Phys. Chem. C 2013, 117, 20531−20541. (31) Zhang, C.; Yamazaki, A.; Murai, J.; Park, J.; Mandai, T.; Ueno, K.; Dokko, K.; Watanabe, M. Chelate Effects in Glyme/Lithium bis(trifluoromethanesulfonyl)amide Solvate Ionic Liquids, Part 2: Importance of Solvate-Structure Stability for Electrolytes of Lithium Batteries. J. Phys. Chem. C 2014, 118, 17362−17373. (32) Huang, X.; Margulis, C. J.; Li, Y.; Berne, B. J. Why Is the Partial Molar Volume of CO2 So Small When Dissolved in a Room Temperature Ionic Liquid? Structure and Dynamics of CO2 Dissolved in [Bmim+] [PF−6 ]. J. Am. Chem. Soc. 2005, 127, 17842−17851. (33) Pérez-Salado Kamps, A.; Tuma, D.; Xia, J.; Maurer, G. Solubility of CO2 in the Ionic Liquid [bmim][PF6]. J. Chem. Eng. Data 2003, 48, 746−749. (34) Cadena, C.; Anthony, J. L.; Shah, J. K.; Morrow, T. I.; Brennecke, J. F.; Maginn, E. J. Why Is CO2 So Soluble in Imidazolium-Based Ionic Liquids? J. Am. Chem. Soc. 2004, 126, 5300−5308. (35) Camper, D.; Bara, J. E.; Gin, D. L.; Noble, R. D. RoomTemperature Ionic Liquid-Amine Solutions: Tunable Solvents for

Columbia University with Bruce J. Berne. His research is theoretical and computational in nature with a strong focus on the structure and dynamics of ionic liquids.



ACKNOWLEDGMENTS This article provides an overview of selected work done in the Margulis group in the past five years. Some of this work was carried out by current group members, but a significant portion was also carried out by prior group members. In particular, Harsha V. R. Annapureddy and Hemant Kashyap were instrumental to many of the developments here presented. Almost all of these works were done in collaboration with the groups of Edward W. Castner Jr. at Rutgers University, James Wishart at BNL, and Mark Maroncelli at Penn State University to whom we are greatly indebted. This work was supported by the US Department of Energy under grant DE-SC0008644 and by the US National Science Foundation under grant CHE-1362129.



REFERENCES

(1) Giernoth, R. Task-Specific Ionic Liquids. Angew. Chem., Int. Ed. 2010, 49, 2834−2839. (2) Visser, A. E.; Swatloski, R. P.; Reichert, W. M.; Mayton, R.; Sheff, S.; Wierzbicki, A.; Davis, J. H., Jr.; Rogers, R. D. Task-Specific Ionic Liquids for the Extraction of Metal Ions from Aqueous Solutions. Chem. Commun. 2001, 135−136. (3) Bates, E. D.; Mayton, R. D.; Ntai, I.; Davis, J. H. CO2 Capture by a Task-Specific Ionic Liquid. J. Am. Chem. Soc. 2002, 124, 926−927. (4) Visser, A. E.; Swatloski, R. P.; Reichert, W. M.; Mayton, R.; Sheff, S.; Wierzbicki, A.; Davis, J. H.; Rogers, R. D. Task-Specific Ionic Liquids Incorporating Novel Cations for the Coordination and Extraction of Hg2+ and Cd2+: Synthesis, Characterization, and Extraction Studies. Environ. Sci. Technol. 2002, 36, 2523−2529. (5) Lee, S. Functionalized Imidazolium Salts for Task-Specific Ionic Liquids and Their Applications. Chem. Commun. 2006, 1049−1063. (6) Gui, J.; Deng, Y.; Hu, Z.; Sun, Z. A Novel Task-Specific Ionic Liquid for Beckmann Rearrangement: A Simple and Effective Way for Product Separation. Tetrahedron Lett. 2004, 45, 2681−2683. (7) Zhang, Z.; Xie, Y.; Li, W.; Hu, S.; Song, J.; Jiang, T.; Han, B. Hydrogenation of Carbon Dioxide is Promoted by a Task-Specific Ionic Liquid. Angew. Chem., Int. Ed. 2008, 47, 1127−1129. (8) Nockemann, P.; Thijs, B.; Parac-Vogt, T. N.; Van Hecke, K.; Van Meervelt, L.; Tinant, B.; Hartenbach, I.; Schleid, T.; Ngan, V. T.; Nguyen, M. T.; et al. Carboxyl-Functionalized Task-Specific Ionic Liquids for Solubilizing Metal Oxides. Inorg. Chem. 2008, 47, 9987− 9999. (9) Nockemann, P.; Thijs, B.; Pittois, S.; Thoen, J.; Glorieux, C.; Van Hecke, K.; Van Meervelt, L.; Kirchner, B.; Binnemans, K. Task-Specific Ionic Liquid for Solubilizing Metal Oxides. J. Phys. Chem. B 2006, 110, 20978−20992. (10) Sawant, A.; Raut, D.; Darvatkar, N.; Salunkhe, M. Recent Developments of Task-Specific Ionic Liquids in Organic Synthesis. Green Chem. Lett. Rev. 2011, 4, 41−54. (11) Tang, S.; Baker, G. A.; Zhao, H. Ether- and AlcoholFunctionalized Task-Specific Ionic Liquids: Attractive Properties and Applications. Chem. Soc. Rev. 2012, 41, 4030−4066. (12) Sahbaz, Y.; Williams, H. D.; Nguyen, T.; Saunders, J.; Ford, L.; Charman, S. A.; Scammells, P. J.; Porter, C. J. H. Transformation of Poorly Water-Soluble Drugs into Lipophilic Ionic Liquids Enhances Oral Drug Exposure from Lipid Based Formulations. Mol. Pharmaceutics 2015, 12, 1980−1991. (13) Hough, W. L.; Smiglak, M.; Rodriguez, H.; Swatloski, R. P.; Spear, S. K.; Daly, D. T.; Pernak, J.; Grisel, J. E.; Carliss, R. D.; Soutullo, M. D.; et al. The Third Evolution of Ionic Liquids: Active Pharmaceutical Ingredients. New J. Chem. 2007, 31, 1429−1436. (14) Hart, M. L.; Do, D. P.; Ansari, R. A.; Rizvi, S. A. Brief Overview of Various Approaches to Enhance Drug Solubility. J. Dev. Drugs 2013, 2, 1000110−1000116. 12736

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B Efficient and Reversible Capture of CO2. Ind. Eng. Chem. Res. 2008, 47, 8496−8498. (36) Karadas, F.; Atilhan, M.; Aparicio, S. Review on the Use of Ionic Liquids (ILs) as Alternative Fluids for CO2 Capture and Natural Gas Sweetening. Energy Fuels 2010, 24, 5817−5828. (37) Gurkan, B. E.; de la Fuente, J. C.; Mindrup, E. M.; Ficke, L. E.; Goodrich, B. F.; Price, E. A.; Schneider, W. F.; Brennecke, J. F. Equimolar CO2 Absorption by Anion-Functionalized Ionic Liquids. J. Am. Chem. Soc. 2010, 132, 2116−2117. (38) Hazelbaker, E. D.; Budhathoki, S.; Wang, H.; Shah, J.; Maginn, E. J.; Vasenkov, S. Relationship between Diffusion and Chemical Exchange in Mixtures of Carbon Dioxide and an Amine-Functionalized Ionic Liquid by High Field NMR and Kinetic Monte Carlo Simulations. J. Phys. Chem. Lett. 2014, 5, 1766−1770. (39) Swatloski, R. P.; Spear, S. K.; Holbrey, J. D.; Rogers, R. D. Dissolution of Cellose with Ionic Liquids. J. Am. Chem. Soc. 2002, 124, 4974−4975. (40) Zhu, S.; Wu, Y.; Chen, Q.; Yu, Z.; Wang, C.; Jin, S.; Ding, Y.; Wu, G. Dissolution of Cellulose with Ionic Liquids and its Application: A Mini-Review. Green Chem. 2006, 8, 325−327. (41) Fukaya, Y.; Hayashi, K.; Wada, M.; Ohno, H. Cellulose Dissolution with Polar Ionic Liquids under Mild Conditions: Required Factors for Anions. Green Chem. 2008, 10, 44−46. (42) Wang, H.; Gurau, G.; Rogers, R. D. Ionic Liquid Processing of Cellulose. Chem. Soc. Rev. 2012, 41, 1519−1537. (43) Zhang, H.; Wu, J.; Zhang, J.; He, J. 1-allyl-3-methylimidazolium chloride Room Temperature Ionic Liquid: A New and Powerful Nonderivatizing Solvent for Cellulose. Macromolecules 2005, 38, 8272− 8277. (44) Socha, A. M.; Parthasarathi, R.; Shi, J.; Pattathil, S.; Whyte, D.; Bergeron, M.; George, A.; Tran, K.; Stavila, V.; Venkatachalam, S.; et al. Efficient Biomass Pretreatment Using Ionic Liquids Derived from Lignin and Hemicellulose. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, E3587−E3595. (45) Fujita, K.; MacFarlane, D. R.; Forsyth, M. Protein Solubilising and Stabilising Ionic Liquids. Chem. Commun. 2005, 4804−4806. (46) Pei, Y.; Wang, J.; Wu, K.; Xuan, X.; Lu, X. Ionic Liquid-Based Aqueous Two-Phase Extraction of Selected Proteins. Sep. Purif. Technol. 2009, 64, 288−295. (47) Chen, X.; Liu, J.; Wang, J. Ionic Liquids in the Assay of Proteins. Anal. Methods 2010, 2, 1222−1226. (48) Gorke, J.; Srienc, F.; Kazlauskas, R. Toward Advanced Ionic Liquids. Polar, Enzyme-Friendly Solvents for Biocatalysis. Biotechnol. Bioprocess Eng. 2010, 15, 40−53. (49) Taha, M.; Almeida, M. R.; Silva, F. A.; Domingues, P.; Ventura, S. P. M.; Coutinho, J. A. P.; Freire, M. G. Novel Biocompatible and Selfbuffering Ionic Liquids for Biopharmaceutical Applications. Chem. - Eur. J. 2015, 21, 4781−4788. (50) Greaves, T. L.; Kennedy, D. F.; Mudie, S. T.; Drummond, C. J. Diversity Observed in the Nanostructure of Protic Ionic Liquids. J. Phys. Chem. B 2010, 114, 10022−10031. (51) Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Amphiphilicity Determines Nanostructure in Protic Ionic Liquids. Phys. Chem. Chem. Phys. 2011, 13, 3237−3247. (52) Greaves, T. L.; Kennedy, D. F.; Weerawardena, A.; Tse, N. M. K.; Kirby, N.; Drummond, C. J. Nanostructured Protic Ionic Liquids Retain Nanoscale Features in Aqueous Solution While Precursor Bronsted Acids and Bases Exhibit Different Behavior. J. Phys. Chem. B 2011, 115, 2055−2066. (53) Greaves, T. L.; Kennedy, D. F.; Kirby, N.; Drummond, C. J. Nanostructure Changes in Protic Ionic Liquids (PILs) Through Adding Solutes and Mixing PILs. Phys. Chem. Chem. Phys. 2011, 13, 13501− 13509. (54) Greaves, T. L.; Weerawardena, A.; Fong, C.; Drummond, C. J. Many Protic Ionic Liquids Mediate Hydrocarbon-Solvent Interactions and Promote Amphiphile Self-Assembly. Langmuir 2007, 23, 402−404. (55) Zahn, S.; Thar, J.; Kirchner, B. Structure and Dynamics of the Protic Ionic Liquid monomethylammonium nitrate ([CH3NH3]

[NO3]) from Ab Initio Molecular Dynamics Simulations. J. Chem. Phys. 2010, 132, 124506. (56) Atkin, R.; Warr, G. G. The Smallest Amphiphiles: Nanostructure in Protic Room-Temperature Ionic Liquids with Short Alkyl Groups. J. Phys. Chem. B 2008, 112, 4164−4166. (57) Gontrani, L.; Bodo, E.; Triolo, A.; Leonelli, F.; D’Angelo, P.; Migliorati, V.; Caminiti, R. The Interpretation of Diffraction Patterns of Two Prototypical Protic Ionic Liquids: a Challenging Task for Classical Molecular Dynamics Simulations. J. Phys. Chem. B 2012, 116, 13024− 13032. (58) Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Effect of Cation Alkyl Chain Length and Anion Type on Protic Ionic Liquid Nanostructure. J. Phys. Chem. C 2014, 118, 13998−14008. (59) Shen, Y.; Kennedy, D. F.; Greaves, T. L.; Weerawardena, A.; Mulder, R. J.; Kirby, N.; Song, G.; Drummond, C. J. Protic Ionic Liquids with Fluorous Anions: Physicochemical Properties and Self-assembly Nanostructure. Phys. Chem. Chem. Phys. 2012, 14, 7981−7992. (60) Greaves, T. L.; Kennedy, D. F.; Shen, Y.; Weerawardena, A.; Hawley, A.; Song, G.; Drummond, C. J. Fluorous Protic Ionic Liquid Exhibits a Series of Lyotropic Liquid Crystalline Mesophases Upon Water Addition. J. Mol. Liq. 2015, DOI: 10.1016/j.molliq.2015.03.037. (61) Russina, O.; Lo Celso, F.; Di Michiel, M.; Passerini, S.; Appetecchi, G. B.; Castiglione, F.; Mele, A.; Caminiti, R.; Triolo, A. Mesoscopic Structural Organization in Triphilic Room Temperature Ionic Liquids. Faraday Discuss. 2013, 167, 499−513. (62) Pereiro, A. B.; Pastoriza-Gallego, M. J.; Shimizu, K.; Marrucho, I. M.; Lopes, J. N. C.; Piñeiro, M. M.; Rebelo, L. P. N. On the Formation of a Third, Nanostructured Domain in Ionic Liquids. J. Phys. Chem. B 2013, 117, 10826−10833. (63) Hettige, J. J.; Araque, J. C.; Margulis, C. J. Bicontinuity and Multiple Length Scale Ordering in Triphilic Hydrogen-Bonding Ionic Liquids. J. Phys. Chem. B 2014, 118, 12706−12716. (64) Kashyap, H. K.; Santos, C. S.; Daly, R. P.; Hettige, J. J.; Murthy, N. S.; Shirota, H.; Castner, E. W., Jr.; Margulis, C. J. How Does the Ionic Liquid Organizational Landscape Change when Nonpolar Cationic Alkyl Groups Are Replaced by Polar Isoelectronic Diethers? J. Phys. Chem. B 2013, 117, 1130−1135. (65) Shimizu, K.; Bernardes, C. E. S.; Triolo, A.; Canongia Lopes, J. N. Nano-Segregation in Ionic Liquids: Scorpions and Vanishing Chains. Phys. Chem. Chem. Phys. 2013, 15, 16256−16262. (66) Triolo, A.; Russina, O.; Caminiti, R.; Shirota, H.; Lee, H. Y.; Santos, C. S.; Murthy, N. S.; Castner, E. W., Jr. Comparing Intermediate Range Order for Alkyl- vs. Ether-Substituted Cations in Ionic Liquids. Chem. Commun. 2012, 48, 4959−4961. (67) Freitas, A. A.; Shimizu, K.; Canongia Lopes, J. N. Complex Structure of Ionic Liquids. Molecular Dynamics Studies with Different Cation-Anion Combinations. J. Chem. Eng. Data 2014, 59, 3120−3129. (68) Peppel, T.; Kockerling, M.; Geppert-Rybczynska, M.; Ralys, R. V.; Lehmann, J. K.; Verevkin, S. P.; Heintz, A. Low-Viscosity Paramagnetic Ionic Liquids with Doubly Charged [Co(NCS)4]2− Ions. Angew. Chem., Int. Ed. 2010, 49, 7116−7119. (69) Bernardes, C. E. S.; Mochida, T.; Canongia Lopes, J. N. Modeling the Structure and Thermodynamics of Ferrocenium-Based Ionic Liquids. Phys. Chem. Chem. Phys. 2015, 17, 10200−10208. (70) Li, P.; Wang, W.; Du, Z.; Wang, G.; Li, E.; Li, X. Adsorption and Aggregation Behavior of Surface Active Trisiloxane Room-Temperature Ionic Liquids. Colloids Surf., A 2014, 450, 52−58. (71) Du, Z.; Li, E.; Cao, Y.; Li, X.; Wang, G. Synthesis of TrisiloxaneTailed Surface Active Ionic Liquids and Their Aggregation Behavior in Aqueous Solution. Colloids Surf., A 2014, 441, 744−751. (72) Shirota, H.; Castner, E. W. Why Are Viscosities Lower for Ionic Liquids with -CH2Si(CH3)3 vs -CH2C(CH3)3 Substitutions on the Imidazolium Cations? J. Phys. Chem. B 2005, 109, 21576−21585. (73) Castner, E. W., Jr.; Wishart, J. F.; Shirota, H. Intermolecular Dynamics, Interactions, and Solvation in Ionic Liquids. Acc. Chem. Res. 2007, 40, 1217−1227. (74) Niedermeyer, H.; Ab Rani, M. A.; Lickiss, P. D.; Hallett, J. P.; Welton, T.; White, A. J. P.; Hunt, P. A. Understanding Siloxane 12737

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B Functionalised Ionic Liquids. Phys. Chem. Chem. Phys. 2010, 12, 2018− 2029. (75) Hayes, R.; Warr, G. G.; Atkin, R. Structure and Nanostructure in Ionic Liquids. Chem. Rev. 2015, 115, 6357−6426. (76) Kofu, M.; Nagao, M.; Ueki, T.; Kitazawa, Y.; Nakamura, Y.; Sawamura, S.; Watanabe, M.; Yamamuro, O. Heterogeneous Slow Dynamics of Imidazolium-Based Ionic Liquids Studied by Neutron Spin Echo. J. Phys. Chem. B 2013, 117, 2773−2781. (77) Hardacre, C.; Holbrey, J. D.; Mullan, C. L.; Youngs, T. G. A.; Bowron, D. T. Small Angle Neutron Scattering from 1-alkyl-3methylimidazolium hexafluorophosphate Ionic Liquids ([Cnmim][PF6], N = 4, 6, and 8). J. Chem. Phys. 2010, 133, 74510−74517. (78) Hardacre, C.; Holbrey, J. D.; Nieuwenhuyzen, M.; Youngs, T. G. A. Structure and Solvation in Ionic Liquids. Acc. Chem. Res. 2007, 40, 1146−1155. (79) Macchiagodena, M.; Gontrani, L.; Ramondo, F.; Triolo, A.; Caminiti, R. Liquid Structure of 1-alkyl-3-methylimidazolium hexafluorophosphates by Wide Angle X-ray and Neutron Scattering and Molecular Dynamics. J. Chem. Phys. 2011, 134, 114521−114535. (80) Triolo, A.; Russina, O.; Arrighi, V.; Juranyi, F.; Janssen, S.; Gordon, C. Quasielastic Neutron Scattering Characterization of the Relaxation Processes in a Room Temperature Ionic Liquid. J. Chem. Phys. 2003, 119, 8549−8557. (81) Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Pronounced SpongeLike Nanostructure in propylammonium nitrate. Phys. Chem. Chem. Phys. 2011, 13, 13544−13551. (82) Hardacre, C.; Holbrey, J. D.; McMath, S. E. J.; Bowron, D. T.; Soper, A. K. Structure of Molten 1,3-dimethylimidazolium chloride Using Neutron Diffraction. J. Chem. Phys. 2003, 118, 273−278. (83) Fujii, K.; Kanzaki, R.; Takamuku, T.; Kameda, Y.; Kohara, S.; Kanakubo, M.; Shibayama, M.; ichi Ishiguro, S.; Umebayashi, Y. Experimental Evidences for Molecular Origin of Low-q Peak in Neutron/X-ray Scattering of 1-alkyl-3-methylimidazolium bis(trifluoromethanesulfonyl)amide Ionic Liquids. J. Chem. Phys. 2011, 135, 244502. (84) Burankova, T.; Hempelmann, R.; Wildes, A.; Embs, J. P. Collective Ion Diffusion and Localized Single Particle Dynamics in Pyridinium-Based Ionic Liquids. J. Phys. Chem. B 2014, 118, 14452− 14460. (85) Santos, C. S.; Annapureddy, H. V. R.; Murthy, N. S.; Kashyap, H. K.; Castner, E. W., Jr.; Margulis, C. J. Temperature-Dependent Structure of methyltributylammonium bis(trifluoromethylsulfonyl)amide: X-Ray Scattering and Simulations. J. Chem. Phys. 2011, 134, 064501. (86) Santos, C. S.; Murthy, N. S.; Baker, G. A.; Castner, E. W., Jr. Communication: X-ray Scattering from Ionic Liquids with pyrrolidinium Cations. J. Chem. Phys. 2011, 134, 121101. (87) Macchiagodena, M.; Ramondo, F.; Triolo, A.; Gontrani, L.; Caminiti, R. Liquid Structure of 1-ethyl-3-methylimidazolium Alkyl Sulfates by X-ray Scattering and Molecular Dynamics. J. Phys. Chem. B 2012, 116, 13448−13458. (88) Russina, O.; Triolo, A.; Gontrani, L.; Caminiti, R. Mesoscopic Structural Heterogeneities in Room-Temperature Ionic Liquids. J. Phys. Chem. Lett. 2012, 3, 27−33. (89) Russina, O.; Triolo, A.; Gontrani, L.; Caminiti, R.; Xiao, D.; Hines, L. G., Jr.; Bartsch, R. A.; Quitevis, E. L.; Plechkova, N.; Seddon, K. R. Morphology and Intermolecular Dynamics of 1-alkyl-3-methylimidazolium bis(trifluoromethane)sulfonylamide Ionic Liquids: Structural and Dynamic Evidence of Nanoscale Segregation. J. Phys.: Condens. Matter 2009, 21, 424121. (90) Russina, O.; Triolo, A. New Experimental Evidence Supporting the Mesoscopic Segregation Model in Room Temperature Ionic Liquids. Faraday Discuss. 2012, 154, 97−109. (91) Russina, O.; Gontrani, L.; Fazio, B.; Lombardo, D.; Triolo, A.; Caminiti, R. Selected Chemical-Physical Properties and Structural Heterogeneities in 1-ethyl-3-methylimidazolium alkyl-sulfate Room Temperature Ionic Liquids. Chem. Phys. Lett. 2010, 493, 259−262. (92) Bodo, E.; Gontrani, L.; Caminiti, R.; Plechkova, N. V.; Seddon, K. R.; Triolo, A. Structural Properties of 1-alkyl-3-methylimidazolium bis(trifluoromethyl)sulfonylamide Ionic Liquids: X-ray Diffraction Data

and Molecular Dynamics Simulations. J. Phys. Chem. B 2010, 114, 16398−16407. (93) Triolo, A.; Russina, O.; Fazio, B.; Appetecchi, G. B.; Carewska, M.; Passerini, S. Nanoscale Organization in Piperidinium-Based Room Temperature Ionic Liquids. J. Chem. Phys. 2009, 130, 164521. (94) Triolo, A.; Russina, O.; Fazio, B.; Triolo, R.; Di Cola, E. Morphology of 1-alkyl-3-methylimidazolium hexafluorophosphate Room Temperature Ionic Liquids. Chem. Phys. Lett. 2008, 457, 362− 365. (95) Triolo, A.; Russina, O.; Bleif, H.; Di Cola, E. Nanoscale Segregation in Room Temperature Ionic Liquids. J. Phys. Chem. B 2007, 111, 4641−4644. (96) Gontrani, L.; Russina, O.; Celso, F. L.; Caminiti, R.; Annat, G.; Triolo, A. Liquid Structure of Trihexyltetradecylphosphonium Chloride at Ambient Temperature: An X-ray Scattering and Simulation Study. J. Phys. Chem. B 2009, 113, 9235−9240. (97) Fukuda, S.; Takeuchi, M.; Fujii, K.; Kanzaki, R.; Takamuku, T.; Chiba, K.; Yamamoto, H.; Umebayashi, Y.; Ishiguro, S. Liquid Structure of N-butyl-N-methylpyrrolidinium bis-(trifluoromethanesulfonyl)amide Ionic Liquid Studied by Large Angle X-ray Scattering and Molecular Dynamics Simulations. J. Mol. Liq. 2008, 143, 2−7. (98) Rocha, M. A. A.; Neves, C. M. S. S.; Freire, M. G.; Russina, O.; Triolo, A.; Coutinho, J. A. P.; Santos, L. M. N. B. F. Alkylimidazolium Based Ionic Liquids: Impact of Cation Symmetry on Their Nanoscale Structural Organization. J. Phys. Chem. B 2013, 117, 10889−10897. (99) Bradley, A. E.; Hardacre, C.; Holbrey, J. D.; Johnston, S.; McMath, S. E. J.; Nieuwenhuyzen, M. Small-Angle X-ray Scattering Studies of Liquid Crystalline 1-alkyl-3-methylimidazolium Salts. Chem. Mater. 2002, 14, 629−635. (100) Xiao, D.; Hines, L. G., Jr.; Li, S.; Bartsch, R. A.; Quitevis, E. L.; Russina, O.; Triolo, A. Effect of Cation Symmetry and Alkyl Chain Length on the Structure and Intermolecular Dynamics of 1,3dialkylimidazolium bis(trifluoromethanesulfonyl)amide Ionic Liquids. J. Phys. Chem. B 2009, 113, 6426−6433. (101) Fujii, K.; Mitsugi, T.; Takamuku, T.; Yanaguchi, T.; Umebayashi, Y.; Ishiguro, S. Effect of Methylation at the C2 Position of Imidazolium on the Structure of Ionic Liquids Revealed by Large Angle X-ray Scattering Experiments and MD Simulations. Chem. Lett. 2009, 38, 340−341. (102) Aoun, B.; Goldbach, A.; González, M. A.; Kohara, S.; Price, D. L.; Saboungi, M. Nanoscale Heterogeneity in alkyl-methylimidazolium bromide Ionic Liquids. J. Chem. Phys. 2011, 134, 104509. (103) Li, S.; Bañuelos, J. L.; Guo, J.; Anovitz, L.; Rother, G.; Shaw, R. W.; Hillesheim, P. C.; Dai, S.; Baker, G. A.; Cummings, P. T. Alkyl Chain Length and Temperature Effects on Structural Properties of Pyrrolidinium-Based Ionic Liquids: A Combined Atomistic Simulation and Small-Angle X-ray Scattering Study. J. Phys. Chem. Lett. 2012, 3, 125−130. (104) Pott, T.; Méléard, P. New Insight into the Nanostructure of Ionic Liquids: A Small Angle X-ray Scattering (SAXS) Study on Liquid trialkyl-methyl-ammonium bis(trifluoromethanesulfonyl)amides and Their Mixtures. Phys. Chem. Chem. Phys. 2009, 11, 5469−5475. (105) Zheng, W.; Mohammed, A.; Hines, L. G.; Xiao, D.; Martinez, O. J.; Bartsch, R. A.; Simon, S. L.; Russina, O.; Triolo, A.; Quitevis, E. L. Effect of Cation Symmetry on the Morphology and Physicochemical Properties of Imidazolium Ionic Liquids. J. Phys. Chem. B 2011, 115, 6572−6584. (106) Fujii, K.; Seki, S.; Fukuda, S.; Takamuku, T.; Kohara, S.; Kameda, Y.; Umebayashi, Y.; Ishiguro, S. Liquid Structure and Conformation of a Low-Viscosity Ionic Liquid, N-methyl-N-propyl-pyrrolidinium bis(fluorosulfonyl)imide Studied by High-Energy X-ray Scattering. J. Mol. Liq. 2008, 143, 64−69. (107) Nemoto, F.; Kofu, M.; Yamamuro, O. Thermal and Structural Studies of Imidazolium-Based Ionic Liquids with and without LiquidCrystalline Phases: The Origin of Nanostructure. J. Phys. Chem. B 2015, 119, 5028−5034. (108) Kashyap, H. K.; Margulis, C. J. Theoretical Deconstruction of the X-ray Structure Function Exposes Polarity Alternations in Room Temperature Ionic Liquids. ECS Trans. 2013, 50, 301−307. 12738

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B

Pyrrolidinium-Based Ionic Liquids: A Molecular Simulation Study. J. Phys. Chem. B 2014, 118, 731−742. (129) Ji, Y.; Shi, R.; Wang, Y.; Saielli, G. Effect of the Chain Length on the Structure of Ionic Liquids: from Spatial Heterogeneity to Ionic Liquid Crystals. J. Phys. Chem. B 2013, 117, 1104−1109. (130) Wang, Y.; Jiang, W.; Yan, T.; Voth, G. A. Understanding Ionic Liquids through Atomistic and Coarse-Grained Molecular Dynamics Simulations. Acc. Chem. Res. 2007, 40, 1193−1199. (131) Shimizu, K.; Bernardes, C. E. S.; Canongia Lopes, J. N. Structure and Aggregation in the 1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide Ionic Liquid Homologous Series. J. Phys. Chem. B 2014, 118, 567−576. (132) Shimizu, K.; Canongia Lopes, J. N. Probing the Structural Features of the 1-alkyl-3-methylimidazolium hexafluorophosphate Ionic Liquid Series Using Molecular Dynamics Simulations. J. Mol. Liq. 2015, DOI: 10.1016/j.molliq.2015.04.014. (133) Yamamuro, O.; Yamada, T.; Kofu, M.; Nakakoshi, M.; Nagao, M. Hierarchical Structure and Dynamics of an Ionic Liquid 1-octyl-3methylimidazolium chloride. J. Chem. Phys. 2011, 135, 054508. (134) Pádua, A. A. H.; Costa Gomes, M. F.; Canongia Lopes, J. N. A. Molecular Solutes in Ionic Liquids: A Structural Perspective. Acc. Chem. Res. 2007, 40, 1087−1096. (135) Wang, Y.; Voth, G. A. Tail Aggregation and Domain Diffusion in Ionic Liquids. J. Phys. Chem. B 2006, 110, 18601−18608. (136) Canongia Lopes, J. N. A.; Pádua, A. A. H. Nanostructural Organization in Ionic Liquids. J. Phys. Chem. B 2006, 110, 3330−3335. (137) Ren, Z.; Brinzer, T.; Dutta, S.; Garrett-Roe, S. Thiocyanate as a Local Probe of Ultrafast Structure and Dynamics in Imidazolium-Based Ionic Liquids: Water-Induced Heterogeneity and Cation-Induced Ion Pairing. J. Phys. Chem. B 2015, 119, 4699−4712. (138) Del Popolo, M. G.; Voth, G. A. On the Structure and Dynamics of Ionic Liquids. J. Phys. Chem. B 2004, 108, 1744−1752. (139) Hu, Z. H.; Margulis, C. J. Heterogeneity in a Room-Temperature Ionic Liquid: Persistent Local Environments and the Red-Edge Effect. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 831−836. (140) Habasaki, J.; Ngai, K. L. Heterogeneous Dynamics of Ionic Liquids from Molecular Dynamics Simulations. J. Chem. Phys. 2008, 129, 194501. (141) Kim, D.; Jeong, D.; Jung, Y. Dynamic Propensity as an Indicator of Heterogeneity in Room-Temperature Ionic Liquids. Phys. Chem. Chem. Phys. 2014, 16, 19712−19719. (142) Bardak, F.; Rajian, J. R.; Son, P.; Quitevis, E. L. Heterogeneous Dynamics in Ionic Liquids at the Glass Transition: Fluorescence Recovery after Photobleaching Measurements of Probe Rotational Motion from Tg-6 k To Tg+4 k. J. Non-Cryst. Solids 2015, 407, 324−332. (143) Belding, S. R.; Rees, N. V.; Aldous, L.; Hardacre, C.; Compton, R. G. Behavior of the Heterogeneous Electron-Transfer Rate Constants of Arenes and Substituted Anthracenes in Room-Temperature Ionic Liquids. J. Phys. Chem. C 2008, 112, 1650−1657. (144) Nagasawa, Y.; Miyasaka, H. Ultrafast Solvation Dynamics and Charge Transfer Reactions in Room Temperature Ionic Liquids. Phys. Chem. Chem. Phys. 2014, 16, 13008−13026. (145) Mandal, P. K.; Sarkar, M.; Samanta, A. Excitation-WavelengthDependent Fluorescence Behavior of Some Dipolar Molecules in Room-Temperature Ionic Liquids. J. Phys. Chem. A 2004, 108, 9048− 9053. (146) Paul, A.; Mandal, P. K.; Samanta, A. On the Optical Properties of the Imidazolium Ionic Liquids. J. Phys. Chem. B 2005, 109, 9148−9153. (147) Hu, Z.; Margulis, C. J. Room-Temperature Ionic Liquids: Slow Dynamics, Viscosity, and the Red Edge Effect. Acc. Chem. Res. 2007, 40, 1097−1105. (148) Jin, H.; Li, X.; Maroncelli, M. Heterogeneous Solute Dynamics in Room Temperature Ionic Liquids. J. Phys. Chem. B 2007, 111, 13473−13478. (149) Annapureddy, H. V. R.; Margulis, C. J. Controlling the Outcome of Electron Transfer Reactions in Ionic Liquids. J. Phys. Chem. B 2009, 113, 12005−12012.

(109) Kashyap, H. K.; Hettige, J. J.; Annapureddy, H. V. R.; Margulis, C. J. SAXS Anti-Peaks Reveal the Length-Scales of Dual PositiveNegative and Polar-Apolar Ordering in Room-Temperature Ionic Liquids. Chem. Commun. 2012, 48, 5103−5105. (110) Kashyap, H. K.; Santos, C. S.; Murthy, N. S.; Hettige, J. J.; Kerr, K.; Ramati, S.; Gwon, J.; Gohdo, M.; Lall-Ramnarine, S. I.; Wishart, J. F.; et al. St r ucture of 1-alkyl-1-methylpyrrolidinium b is(trifluoromethylsulfonyl)amide Ionic Liquids with Linear, Branched, and Cyclic Alkyl Groups. J. Phys. Chem. B 2013, 117, 15328−15337. (111) Kashyap, H. K.; Santos, C. S.; Annapureddy, H. V. R.; Murthy, N. S.; Margulis, C. J.; Castner, E. W., Jr. Temperature-Dependent Structure of Ionic Liquids: X-ray Scattering and Simulations. Faraday Discuss. 2012, 154, 133−143. (112) Annapureddy, H. V. R.; Kashyap, H. K.; De Biase, P. M.; Margulis, C. J. What is the Origin of the Prepeak in the X-ray Scattering of Imidazolium-Based Room-Temperature Ionic Liquids? J. Phys. Chem. B 2010, 114, 16838−16846. (113) Hettige, J. J.; Kashyap, H. K.; Annapureddy, H. V. R.; Margulis, C. J. Anions, the Reporters of Structure in Ionic Liquids. J. Phys. Chem. Lett. 2013, 4, 105−110. (114) Hettige, J. J.; Kashyap, H. K.; Margulis, C. J. Communication: Anomalous Temperature Dependence of the Intermediate Range Order in Phosphonium Ionic Liquids. J. Chem. Phys. 2014, 140, 111102. (115) Bhargava, B.; Klein, M.; Balasubramanian, S. Structural Correlations and Charge Ordering in a Room-Temperature Ionic Liquid. ChemPhysChem 2008, 9, 67−70. (116) Bhargava, B. L.; Devane, R.; Klein, M. L.; Balasubramanian, S. Nanoscale Organization in Room Temperature Ionic Liquids: A Coarse Grained Molecular Dynamics Simulation Study. Soft Matter 2007, 3, 1395−1400. (117) Li, S.; Feng, G.; Bañuelos, J. L.; Rother, G.; Fulvio, P. F.; Dai, S.; Cummings, P. T. Distinctive Nanoscale Organization of Dicationic versus Monocationic Ionic Liquids. J. Phys. Chem. C 2013, 117, 18251− 18257. (118) Shimizu, K.; Pádua, A. A. H.; Canongia Lopes, J. N. Nanostructure of trialkylmethylammonium bistriflamide Ionic Liquids Studied by Molecular Dynamics. J. Phys. Chem. B 2010, 114, 15635− 15641. (119) Maginn, E. J. Molecular Simulation of Ionic Liquids: Current Status and Future Opportunities. J. Phys.: Condens. Matter 2009, 21, 373101. (120) Urahata, S. M.; Ribeiro, M. C. C. Structure of Ionic Liquids of 1alkyl-3-methylimidazolium Cations: A Systematic Computer Simulation Study. J. Chem. Phys. 2004, 120, 1855−1863. (121) Canongia Lopes, J. N. A.; Pádua, A. A. H. Nanostructural Organization in Ionic Liquids. J. Phys. Chem. B 2006, 110, 3330−3335. (122) Canongia Lopes, J. N.; Shimizu, K.; Pádua, A. A. H.; Umebayashi, Y.; Fukuda, S.; Fujii, K.; Ishiguro, S. A Tale of Two Ions: The Conformational Landscapes of bis(trifluoromethanesulfonyl)amide and N,N-dialkylpyrrolidinium. J. Phys. Chem. B 2008, 112, 1465−1472. (123) 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. (124) Siqueira, L. J. A.; Ribeiro, M. C. C. Charge Ordering and Intermediate Range Order in Ammonium Ionic Liquids. J. Chem. Phys. 2011, 135, 204506. (125) Wang, Y. T.; Izvekov, S.; Yan, T. Y.; Voth, G. A. multiscale Coarse-Graining of Ionic Liquids. J. Phys. Chem. B 2006, 110, 3564− 3575. (126) Wang, Y.; Voth, G. A. Unique Spatial Heterogeneity in Ionic Liquids. J. Am. Chem. Soc. 2005, 127, 12192−12193. (127) Borodin, O.; Smith, G. D. Structure and Dynamics of N-methylN-propylpyrrolidinium bis(trifluoromethanesulfonyl)imide Ionic Liquid from Molecular Dynamics Simulations. J. Phys. Chem. B 2006, 110, 11481−11490. (128) Paredes, X.; Fernández, J.; Pádua, A. A. H.; Malfreyt, P.; Malberg, F.; Kirchner, B.; Pensado, A. S. Bulk and Liquid-Vapor Interface of 12739

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740

Feature Article

The Journal of Physical Chemistry B (150) Kaintz, A.; Baker, G.; Benesi, A.; Maroncelli, M. Solute Diffusion in Ionic Liquids, NMR Measurements and Comparisons to Conventional Solvents. J. Phys. Chem. B 2013, 117, 11697−11708. (151) Araque, J. C.; Yadav, S. K.; Shadeck, M.; Maroncelli, M.; Margulis, C. J. How is Diffusion of Neutral and Charged Tracers Related to the Structure and Dynamics of a Room-Temperature Ionic Liquid? Large Deviations from Stokes-Einstein Behavior Explained. J. Phys. Chem. B 2015, 119, 7015−7029. (152) Hahn, T.; Shmueli, U.; Wilson, A. J. C.; Prince, E. International Tables for Crystallography; Prince, E., Ed.; International Union of Crystallography: Chester, England, 2006; Vol. C. (153) Gordon, C. M.; Holbrey, J. D.; Kennedy, A. R.; Seddon, K. R. Ionic Liquid Crystals: Hexafluorophosphate Salts. J. Mater. Chem. 1998, 8, 2627−2636. (154) Lorch, E. Neutron Diffraction by Germania Silica and RadiationDamaged Silica Glasses. J. Phys. C: Solid State Phys. 1969, 2, 229−237. (155) Du, J.; Benmore, C. J.; Corrales, R.; Hart, R. T.; Weber, J. K. R. A Molecular Dynamics Simulation Interpretation of Neutron and X-ray Diffraction Measurements on Single Phase Y(2)o(3)-al(2)o(3) Glasses. J. Phys.: Condens. Matter 2009, 21, 205102. (156) Hansen, J.-P.; McDonald, I. R. Theory of Simple Liquids; Elsevier Academic Press: Burlington, MA, 3rd ed., 2006. (157) McLean, A.; Muldoon, M.; Gordon, C.; Dunkin, I. Bimolecular Rate Constants for Diffusion in Ionic Liquids. Chem. Commun. 2002, 1880−1881. (158) Skrzypczak, A.; Neta, P. Diffusion-Controlled Electron-Transfer Reactions in Ionic Liquids. J. Phys. Chem. A 2003, 107, 7800−7803. (159) Grodkowski, J.; Neta, P.; Wishart, J. F. Pulse Radiolysis Study of the Reactions of Hydrogen Atoms in the Ionic Liquid methyltributylammonium bis[(trifluoromethyl)sulfonyl]imide. J. Phys. Chem. A 2003, 107, 9794−9799. (160) Vieira, R. C.; Falvey, D. E. Photoinduced Electron-Transfer Reactions in Two Room-Temperature Ionic Liquids: 1-butyl-3methylimidazolium hexafluorophosphate and 1-octyl-3-methylimidazolium hexafluorophosphate. J. Phys. Chem. B 2007, 111, 5023−5029. (161) Paul, A.; Samanta, A. Photoinduced Electron Transfer Reaction in Room Temperature Ionic Liquids: A Combined Laser Flash Photolysis and Fluorescence Study. J. Phys. Chem. B 2007, 111, 1957−1962. (162) Sarkar, S.; Pramanik, R.; Seth, D.; Setua, P.; Sarkar, N. Photoinduced Electron Transfer (PET) from N,N-dimethylaniline to 7amino coumarin Dyes in a Room Temperature Ionic Liquid (RTIL): Slowing Down of Electron Transfer Rate Compared to Conventional Solvent. Chem. Phys. Lett. 2009, 477, 102−108. (163) Hallett, J. P.; Welton, T. Room-Temperature Ionic Liquids: Solvents for Synthesis and Catalysis. 2. Chem. Rev. 2011, 111, 3508− 3576. (164) Liang, M.; Kaintz, A.; Baker, G. A.; Maroncelli, M. Bimolecular Electron Transfer in Ionic Liquids: Are Reaction Rates Anomalously High? J. Phys. Chem. B 2012, 116, 1370−1384. (165) Gordon, C.; McLean, A. Photoelectron Transfer from ExcitedState ruthenium(ii) tris(bipyridyl) to methylviologen in an Ionic Liquid. Chem. Commun. 2000, 1395−1396. (166) Grodkowski, J.; Neta, P. Formation and Reaction of Br•− 2 Radicals in the Ionic Liquid methyltributylammonium bis(trifluoromethylsulfonyl)imide and in Other Solvents. J. Phys. Chem. A 2002, 106, 11130−11134. (167) Wishart, J. F.; Neta, P. Spectrum and Reactivity of the Solvated Electron in the Ionic Liquid methyltributylammonium bis(trifluoromethylsulfonyl)imide. J. Phys. Chem. B 2003, 107, 7261−7267. (168) Takahashi, K.; Sakai, S.; Tezuka, H.; Hiejima, Y.; Katsumura, Y.; Watanabe, M. Reaction between Diiodide Anion Radicals in Ionic Liquids. J. Phys. Chem. B 2007, 111, 4807−4811. (169) Zhao, Y.; Hu, Z. Structural Origin of Energetic Heterogeneity in Ionic Liquids. Chem. Commun. 2012, 48, 2231−2233.

12740

DOI: 10.1021/acs.jpcb.5b05506 J. Phys. Chem. B 2015, 119, 12727−12740