Paschoal Et Al. 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 60

jwp00 | ACSJCA | JCA10.0.1465/W Unicode | research.3f (R3.6.i12 HF01:4457 | 2.

0 alpha 39) 2016/10/28 09:46:00 | PROD-JCA1 | rq_6805775 | 1/03/2017 09:38:49 | 60 | JCA-DEFAULT

Review

pubs.acs.org/CR

1 Vibrational Spectroscopy of Ionic Liquids


2 Vitor H. Paschoal, Luiz F. O. Faria, and Mauro C. C. Ribeiro*
3 Laboratório de Espectroscopia Molecular, Departamento de Química Fundamental, Instituto de Química, Universidade de São Paulo,
4 Av. Prof. Lineu Prestes 748, São Paulo 05508-000, Brazil
5 *
S Supporting Information

6 ABSTRACT: Vibrational spectroscopy has continued use as a powerful tool to


7 characterize ionic liquids since the literature on room temperature molten salts
8 experienced the rapid increase in number of publications in the 1990’s. In the past years,
9 infrared (IR) and Raman spectroscopies have provided insights on ionic interactions and
10 the resulting liquid structure in ionic liquids. A large body of information is now available
11 concerning vibrational spectra of ionic liquids made of many different combinations of
12 anions and cations, but reviews on this literature are scarce. This review is an attempt at
13 filling this gap. Some basic care needed while recording IR or Raman spectra of ionic
14 liquids is explained. We have reviewed the conceptual basis of theoretical frameworks
15 which have been used to interpret vibrational spectra of ionic liquids, helping the reader
16 to distinguish the scope of application of different methods of calculation. Vibrational
17 frequencies observed in IR and Raman spectra of ionic liquids based on different anions
18 and cations are discussed and eventual disagreements between different sources are critically reviewed. The aim is that the reader
19 can use this information while assigning vibrational spectra of an ionic liquid containing another particular combination of anions
20 and cations. Different applications of IR and Raman spectroscopies are given for both pure ionic liquids and solutions. Further
21 issues addressed in this review are the intermolecular vibrations that are more directly probed by the low-frequency range of IR
22 and Raman spectra and the applications of vibrational spectroscopy in studying phase transitions of ionic liquids.

23 CONTENTS Corresponding Author AS 53


ORCID AS 54
25 1. Introduction A Notes AS 55
26 2. Experimental Details C Biographies AS 56
27 3. Computational Details E Acknowledgments AS 57
28 4. Vibrational Spectroscopy of Pure Ionic Liquids in References AS 58
29 the Mid-Frequency Range H
30 4.1. Vibrational Frequencies of Anions H
31 4.1.1. Small Symmetric Anions H
32 4.1.2. More Complex Fluorinated Anions K 1. INTRODUCTION
33 4.1.3. Alkylsulfates and Hydrogen Sulfate N Vibrational and NMR spectroscopies are fundamental tools to 59
34 4.1.4. Carboxilates O characterize ionic liquids. The vibrational spectroscopy 60
35 4.2. Vibrational Frequencies of Cations Q techniques of infrared (IR) and Raman permeate the literature 61
36 4.2.1. Imidazolium Q on essentially all of the actual or potential applications 62
37 4.2.2. Pyridinium U envisaged for ionic liquids. IR and Raman spectroscopies 63
38 4.2.3. Pyrrolidinium U have provided deep insights on the nature of ionic interactions, 64
39 4.2.4. Piperidinium V the role played by anion−cation hydrogen bonds, molecular 65
40 4.2.5. Ammonium W conformations, and their modifications as pressure and 66
41 4.2.6. Other Cations Y temperature is varied in the normal liquid phase, during 67
42 4.3. Applications Y phase transition to crystalline or amorphous (glassy) solid 68
43 5. Vibrational Spectroscopy of Pure Ionic Liquids in phases, after vaporization, etc. The user of a vibrational 69
44 the Low-Frequency Range AC spectroscopy technique needs to have in hand reliable 70
45 6. Vibrational Spectroscopy for Studying Ionic interpretation of vibrational spectra, in particular, assignment 71
46 Liquid Phase Transitions AH of experimental frequencies to vibrational motions of the 72
47 7. Vibrational Spectroscopy of Ionic Liquid Solu- common ions of ionic liquids. This information is usually 73
48 tions AL
49 8. Concluding Remarks AR
Special Issue: Ionic Liquids
50 Associated Content AS
51 Supporting Information AS Received: July 15, 2016
52 Author Information AS

© XXXX American Chemical Society A DOI: 10.1021/acs.chemrev.6b00461


Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

74 available from studies which are more dedicated to theoretical


75 and computational issues related to vibrational spectroscopy.
76 Quantum chemistry methods are now routinely used to
77 calculate vibrational frequencies to be compared to exper-
78 imental data. However, strong ionic interactions may imply that
79 comparison between experimental and calculated frequencies
80 by a relatively fast ab initio calculation is far from
81 straightforward. In fact, scrutinizing the large body of literature
82 on vibrational spectroscopy of ionic liquids, eventually there is
83 no full agreement between different authors for the same
84 system. Other schemes have been considered for assigning
85 vibrational spectra of ionic liquids (e.g., taking advantage of
86 molecular dynamics simulations of liquids). On the other hand,
87 even though vibrational spectroscopy has a long history in
88 studying high temperature molten salts and some ions are
Figure 1. Molecular structures of some cations and anions whose
89 common species for both high and room temperature molten vibrational spectra are discussed in this review. The Ri, Rj, Rk, and Rl
90 salts (e.g., relatively simple polyatomic anions), sometimes indicate a hydrogen atom or an alkyl chain group. We adopt the usual
91 previous knowledge on vibrational spectroscopy of molten salts notation of cations throughout this work (e.g., 1-alkyl-3-methylimida-
92 is not fully appreciated within the context of ionic liquids. zolium, [CnC1im]+, where n is the number of carbon atoms in the alkyl
93 Moreover, vibrational spectroscopy is well appropriate for chain, and 1-alkyl-2,3-dimethylimidazolium, [CnC1C1im]+). Deriva-
94 studying intermolecular interactions not only in pure ionic tives of ammonium cations are also indicated by the length of the alkyl
95 liquids but also in ionic liquids mixtures, solutions of molecular chains (e.g., [C4C1C1C1N]+ means butyl-trimethylammonium and
96 or ionic solutes in ionic liquids, gas absorption, etc. [C3NH3]+ is the propylammonium cation which forms protic ionic
97 This review addresses the above-mentioned issues and liquids). Similar notation follows for derivatives of pyridinium,
pyrrolidinium, and piperidinium and usual notation for the anions.
98 others, aiming to be useful for users who employ IR and
99 Raman spectroscopies combined with other techniques while
100 investigating ionic liquids and also for those who are involved been especially obtained for the purpose of this review. Raman 137
101 in assignments of vibrational spectra of ionic liquids. spectroscopy of high temperature molten salts demand 138
102 Papatheodorou et al.1,2 published important reviews on nontrivial experimental skills for building furnaces and sample 139
103 Raman spectroscopy of high temperature molten salts, but containers for handling corrosive and air-sensitive melts.2 In 140
104 previous reviews on vibrational spectroscopy of ionic liquids are contrast, IR and Raman measurements of ionic liquids are more 141
105 scarce. In 2007, Berg3 published a review on Raman easily carried out since the liquid sample may be simply 142
106 spectroscopy and ab initio calculations of ionic liquids. Berg’s accommodated in quartz tubes. Nevertheless, some warnings 143
107 review focused on the spectroscopic signatures of molecular on the experimental side are timely. All of IR and Raman 144
108 conformations achieved by the ions, mainly 1-alkyl-3- spectra recorded in this work are available for the reader in 145
109 methylimidazolium and N,N-dialkylpyrrolidinium cations, and TXT files. The usefulness of vibrational spectroscopy is heavily 146
110 the bis(trifluoromethanesulfonyl)imide anion, [NTf2]−. These linked to the calculation of vibrational frequencies, so that 147
111 Raman spectroscopic studies have been reviewed more recently section 3 reviews the methods which have been applied for 148
112 by Saha et al.4 The more specific issue of vibrational calculating vibrational spectra of ionic liquids. These methods 149
113 spectroscopy of ionic liquid surfaces using linear and nonlinear include classical normal coordinate analysis, ab initio calculation 150
114 techniques has also been reviewed recently.5 In this review, we for an isolated ion or for a cluster of ions, and classical or ab 151
115 will address both IR and Raman studies encompassing a wide initio molecular dynamics simulations of liquids. The scope of 152
116 group of cations and anions commonly used in forming ionic section 3 is not technical details of all of these methods; instead 153
f1 117 liquids. Figure 1 shows molecular structures of several cations the section focuses on the conceptual basis of them, helping the 154
118 and anions whose vibrational spectra will be discussed in this reader to distinguish the assumptions and the need for different 155
119 review, together with notation used throughout this work. methods of calculation. Section 4 reviews vibrational spectros- 156
120 This review addresses many issues in which vibrational copy studies of pure ionic liquids, being the longest section of 157
121 spectroscopy has given fundamental contributions for our this work. This section discusses the assignment of vibrational 158
122 current understanding about interactions and structure of ionic frequencies of the most common anions (section 4.1) and 159
123 liquids along the last two decades. The review focusesd on cations (section 4.2) which form ionic liquids. These two 160
124 linear IR and Raman spectroscopies, being beyond the scope of sections are already plenty of applications of vibrational 161
125 techniques such as time-resolved IR spectroscopy and coherent spectroscopy in studying pure ionic liquids, but further 162
126 anti-Stokes Raman scattering (CARS) which have recently been applications are given in section 4.3. It is our hope that the 163
127 applied for investigating ionic liquids.6−11 On the other hand, vibrational assignments discussed thoroughly in this section 164
128 we will address far-infrared (FIR) and low-frequency Raman helps the reader to interpret vibrational spectra when working 165
129 spectroscopy studies of ionic liquids. Thus, results obtained with a given ionic liquid based on a different combination of 166
130 from optical Kerr effect (OKE) spectroscopy, being the anions and cations. We separated in section 5 the discussion of 167
131 counterpart of low-frequency Raman spectroscopy, will be the low-frequency range, as this range directly manifests the 168
132 mentioned, even though OKE is a time-resolved spectroscopy liquid structure and intermolecular dynamics of ionic liquids. 169
133 technique, because it has been the most common technique Studies concerning the low-frequency range of vibrational 170
134 probing the low-frequency vibrations of ionic liquids. spectra of ionic liquids have been developed along two different 171
135 The review is organized as follows. We provide experimental research lines, one by workers using far-infrared (FIR) and 172
136 details in section 2 since most of spectra shown below have other by workers using low-frequency Raman spectroscopy, in 173

B DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

174 particular OKE spectroscopy. Section 5 is an attempt at putting transmission and ATR infrared spectroscopy has been used by 216
175 into a proper perspective the results from FIR and low- Burba et al.20−22 as a methodology to infer about charge 217
176 frequency Raman spectroscopy that have been obtained along organization in the series [CnC1im][CF3SO3], n = 2−8. They 218
177 the past decade. The accumulated knowledge on the nature of considered the intense IR band of the anion symmetric 219
178 vibrational motions and characteristic frequencies in the normal stretching mode, νs(SO3), with maximum at 1031 cm−1. The 220
179 liquid phase discussed in sections 4 and 5 is then applied in approach is based on a first estimate of the dipole moment 221
180 section 6 for studying phase transitions, in particular derivative from a transmission measurement according to the 222
181 crystallization and glass transition experienced by ionic liquids dipolar coupling theory, which assumes a quasilattice 223
182 under low temperature or high pressure. In section 7, we move organization of the ions. A second estimate of the dipole 224
183 to ionic liquids solutions. Since ionic liquids as solvents moment derivative is obtained from optical constants obtained 225
184 encompass a very large range of applications, from gas from an ATR measurement, which is not based on the 226
185 molecules to cellulose, and in each of these areas vibrational quasillatice model. The ratio between these two values of dipole 227
186 spectroscopy has given its contribution, section 7 gives some moment derivatives indicates the degree of quasilattice 228
187 representative examples of the capability of vibrational structure in the ionic liquid. Burba et al.21,22 showed that 229
188 spectroscopy in shedding light on solute−ionic liquid such charge ordering decreases with increasing length of the 230
189 interactions. Section 8 closes the review with some concluding alkyl chain in [CnC1im][CF3SO3]. It should be noted, however, 231
190 remarks. that the method also relies on assigning the asymmetric shape 232
of the νs(SO3) IR band as the result of longitudinal optic− 233
2. EXPERIMENTAL DETAILS transverse optic (LO-TO) splitting23,24 to be considered within 234

191 IR and Raman spectra of several ionic liquids were recorded for the dipolar coupling theory.20−22 All of the IR spectra recorded 235

192 specific purposes of this review. The spectra are available as in this work will be reported in transmittance, except Figures 2 236

193 TXT files. The ionic liquids used in this work were purchased and 5 where IR spectra are shown in absorbance. 237

194 from different suppliers (e.g., Iolitec, Solvionic, Aldrich, and Raman spectra were obtained with a Horiba-Jobin-Yvon 238

195 Merck). Most of the purchased ionic liquids have purity T64000 triple monochromator spectrometer equipped with 239
196 superior to 98%, and they were used without further CCD. The need of triple or double monochromator, or a single 240
197 purification except for the drying process as discussed below. monochromator spectrometer with notch and bandpass filters 241
198 Fourier transform IR spectra were recorded with a Bruker suitable to reduce the Rayleigh scattering line, is particularly 242
199 Alpha equipment with a DTGS detector and KBr optics. IR important for the low-frequency range, 5 < ω < 100 cm−1, to be 243
200 spectra measured by transmission were obtained from thin discussed in section 5. Raman spectra were obtained in 180° 244
201 films of liquid samples between KRS-5 windows. In the case of scattering geometry. There was no selection of polarization of 245
202 solid samples, we used the attenuated total reflection (ATR) scattered light for most of the Raman spectra shown in this 246
203 Platinum accessory (Bruker) with diamond crystal and a single review, otherwise some polarized and depolarized Raman 247
204 reflection. Spectral resolution was 2 cm−1. Optical effects can spectra will be shown as mentioned in text. The excitation line 248
205 result in differences between transmission and reflection used was the 647.1 nm line of a mixed argon−krypton 249
206 measurements.12−15 ATR is a technique appropriate for Coherent laser, typically with 200 mW of output power. 250
207 quantitative IR spectroscopy since it allows for better control Fluorescence background is a known issue in Raman 251
208 of sample size and thickness.16,17 Moreover, ATR Fourier spectroscopy of (high temperature) molten salts.2 Even though 252
209 transform IR spectrum can be used to obtain optical most ionic liquids are colorless, we use a red laser line to excite 253
210 constants12−14,16,18,19 (i.e., frequency-dependent refractive Raman spectra in order to reduce eventual fluorescence 254
211 index and extinction coefficient), as shown by Buffeteau et background. For instance, Figure 3 shows Raman spectra 255 f3
f2 212 al.15 for different ionic liquids. Figure 2 illustrates differences in obtained with excitations at 514.5 and 647.1 nm for the same 256
213 relative intensities and vibrational frequencies that can be found sample of a colorless ionic liquid, [C4C1im][PF6]. The same 257
214 between transmission and ATR measurements of IR spectra of laser power was used to obtain the spectra of Figure 3, and no 258
215 a given ionic liquid, [C4C1im][CF3SO3]. A combination of baseline correction was done, so that it is clear that fluorescence 259
background is strongly reduced when using the 647.1 nm laser 260
line. Whenever fluorescence precluded obtaining suitable 261

Figure 2. IR spectra of [C4C1im][CF3SO3] obtained in transmittance


(black) and ATR (red). The IR spectra have their intensities Figure 3. Raman spectra of [C4C1im][PF6] obtained using the 514.5
normalized by the most intense band for comparison purpose. nm (black) and the 647.1 nm (red) laser line.

C DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

262 Raman spectrum with the 647.1 nm laser line, we used a FT- All the samples used in this work were submitted to a drying 277
263 Raman Bruker RFS-100 spectrometer with a 1064 nm exciting process under high vacuum (10−5 mbar) at ∼60 °C for 48 h 278
264 radiation (Nd:YAG laser Coherent Compass 1064−500N) and prior the analysis. The samples were then manipulated inside a 279
265 Germanium detector. drybox with argon atmosphere. It has been shown that small 280
266 A relatively stringent spectral resolution is eventually amount of water can have significant effect on ionic liquids 281
267 required since the complex molecular structures of the ions properties.25−28 Karl Fischer coulometric titration is the usual 282
f4 268 result in overlapping of bands. Figure 4 illustrates the effect of method to quantify water, but IR spectroscopy has been applied 283
to measure water content in ionic liquids. Andanson et al.27 284
used the O−H stretching mode of water in the range 3400− 285
3800 cm−1, and Fadeeva et al.28 used a combination band of 286
water at ∼5250 cm−1 to measure the water content in ionic 287
liquids. Cammarata et al.29 showed by using IR spectroscopy 288
that interactions between water and the ions are dominated by 289
anion−water interactions in the case of nonprotic ionic liquids 290
based on imidazolium cations. In the case of a mixture of 291
liquids, it is worth noting that accurate analysis of band 292
intensity should take into account the fact that the effective 293
path length eventually changes in ATR-IR measurements 294
because the refractive index of the sample depends on the 295
concentration of the solution.30 Figure 5 (panels A−C) 296 f5
illustrate the evolution of the drying process as monitored by 297
IR spectroscopy for three ionic liquids with the same 298
[C4C1im]+ cation but anions with distinct coordination 299
Figure 4. Raman spectra of [C4C1im][NTf2] in the range of SO2 strength or basicity, [NTf2]−, [CF3SO3]−, and [CH3COO]−. 300
wagging mode recorded with different spectrometer resolutions. The Karl Fischer analyses of [C4C1im][NTf2], [C4C1im]- 301
Raman spectrum obtained with 0.5 cm−1 of spectral resolution is [CF3SO3], and [C4C1im][CH3COO] indicated water concen- 302
shown with the intensity multiplied by a factor of 20.
tration of 284, 835, and 22353 ppm, respectively, for the 303
samples as taken straight from their flask prior to any drying 304
attempt. Water content dropped to 214, 298, and 18795 ppm 305
269 spectral resolution in the Raman spectrum of [C4C1im][NTf2] after 48 h under high-vacuum at room temperature, and when 306
270 within the spectral range of the anion SO2 wagging mode. This the sample was simultaneously warmed to ∼60 °C, the water 307
271 figure exhibits a single broad band at ∼400 cm−1 when using content achieved 45, 118, and 8580 ppm, respectively. This 308
272 poor spectral resolution, but the band is resolved into two trend is manifested in the IR spectra of Figure 5 (panels A−C) 309
273 peaks at 397 and 404 cm−1 when using better spectral by the comparison between relative intensities of water bands 310
274 resolution. In this work, Raman spectra were obtained with (3400−3800 cm−1) and ionic liquid bands (2800−3200 cm−1). 311
275 spectral resolution of 2 cm−1, taking into consideration The drying protocol using high vacuum at room temperature 312
276 compromise on good signal-to-noise ratio. seems enough for a less viscous ionic liquid and less 313

Figure 5. IR spectra of (A) [C4C1im][NTf2], (B) [C4C1im][CF3SO3], and (C) [C4C1im][CH3COO] before and after the drying processes showing
the spectral range of water bands (3400−3800 cm−1). The asterisk in each panel marks the ionic liquid band used to normalize IR intensities within
the spectral range shown in the figure. (C) also compares transmission (red) and ATR (green) measurements of IR spectra of the same sample of
[C4C1im][CH3COO] after the drying process. (D) shows the molar absorption coefficient Em (black, scale at right), the real part of the refractive
index n (green, scale at left), and the imaginary part of the dielectric constant ε″ (red, scale at left) of liquid H2O as provided by Bertie and Lan.31

D DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

314 coordinating anion such as [C4C1im][NTf2]. Otherwise heating with the 647.1 nm laser line focused into the sample by a 20× 376
315 is needed to reduce the viscosity and facilitate the drying Leica objective. We used a DAC from Almax EasyLab, model 377
316 process for more viscous ionic liquids and more strongly Diacell LeverDAC-Maxi, having a diamond culet size of 500 378
317 coordinating anions, [C4C1im][CF3SO3] and [C4C1im]- μm. The Boehler microDriller (Almax EasyLab) was used to 379
318 [CH3COO].29 It is worth mentioning that increase of drill a 250 μm hole in a stainless steel gasket (10 mm diameter, 380
319 temperature implies a certain degree of decomposition as 250 μm thick) preindented to ∼150 μm. Pressure calibration 381
320 discussed by Gurau et al.32 for [C4C1im][CH3COO]. Even in a has been done by the usual method of measuring the shift of 382
321 situation where the minor amount of impurity is generated, this the fluorescence line of ruby.38,39 Figure 6 illustrates the 383 f6
322 heating effect may have a direct consequence in Raman
323 spectroscopy because of eventual increase in fluorescence
324 background. The IR spectra reported by Andanson et al.27 for
325 [C4C1im][NTf2] and [C4C1im][CF3SO3] and by Thomas et
326 al.33 and Marekha et al.34 for [C4C1im][CH3COO] exhibit less
327 intense water bands in comparison with the spectra shown in
328 Figure 5. However, it should be noted that these authors
329 reported IR spectra obtained by ATR while we report
330 transmission IR spectra. In fact, Figure 5C compares trans-
331 mission and ATR spectra of [C4C1im][CH3COO], the latter
332 exhibiting lower intensity of water bands in comparison with
333 the former, being the spectra normalized by an ionic liquid
334 band. The strong variation of the real part of the refractive
335 index, n(ω), within the range of wavenumbers ω of an IR
336 absorption (i.e., the so-called anomalous dispersion as
Figure 6. Pressure dependence of the fluorescence emission spectra of
337 illustrated in Figure 5D for liquid water (green line),31 might
ruby in the ionic liquid [Pyr14][NTf2]. Spectra are normalized by the
338 imply significant differences between ATR and transmission most intense peak. The inset shows a photograph of a DAC chamber
339 measurements of IR spectra.12−14 The molar absorption containing the ionic liquid sample and ruby spheres used for pressure
340 coefficient Em, given by Em = A10/Cd, where d is the path calibration.
341 length, C is the molar concentration, and A10 is the decadic
342 absorbance following the Beer’s Law, is related to the imaginary
4πωk(ω) pressure-dependent fluorescence emission spectra of ruby in 384
part of the refractive index, C = 2.303Em(ω).18,19,35 On
343 [Pyr14][NTf2]. The inset is a photograph of the gasket hole in 385
344 the other hand, Stuchebryukov and Rudoy12 showed that an between the diamonds of the DAC containing the liquid 386
345 ATR spectrum can be considered as the spectrum of the sample and the ruby spheres used for pressure calibration. In 387
346 imaginary part of the dielectric constant, ε″(ω). Since the fact, Faria et al.40 showed that the characteristic [NTf2]− Raman 388
347 complex dielectric constant, ε̂ = ε′ + iε″, and refractive index, n̂ band at 741 cm−1 can be used for pressure calibration. The 389
348 = n + ik, are related by the fundamental equation ε̂ = n̂2, then pressure induced frequency shift of this vibrational mode of 390
349 ε′(ω) = n2(ω) − k2(ω) and ε″(ω) = 2n(ω)k(ω). As a [NTf2]− exhibits a linear variation, ca. 4.2 cm−1/GPa, 391
350 consequence of the anomalous dispersion of n(ω) around an IR depending on the ionic liquid for pressures up to ∼2.5 GPa, 392
351 band, the frequency of the ε″(ω) spectrum is shifted toward the so that within this pressure range the ionic liquid can be used as 393
352 low-frequency side of the band. This is clearly seen in the pressure-transmitting medium and pressure marker. 394
353 comparison between ATR and transmission IR spectra of
354 [C4C1im][CF3SO3] shown in Figure 2. Relative intensities of 3. COMPUTATIONAL DETAILS
355 IR bands in ATR and transmission measurements are also Polyatomic ions usually involved in forming ionic liquids range 395
356 affected by the different dependence on the optical constants. from highly symmetric small ions to complex flexible organic 396
357 This is evident in Figure 5D, which shows Em(ω) and ε″(ω) ions, so that different theoretical methods have been used for 397
358 spectra provided by Bertie and Lan31 from an investigation of the calculation of vibrational frequencies and normal mode 398
359 the optical constants of liquid water. It is clear from Figure 5D assignment. These include classical normal coordinates analysis 399
360 the difference of intensities in the O−H stretching region for relying on an assumed intramolecular force field, ab initio 400
361 each spectrum. Therefore, the word of caution is that quantum chemistry calculations at different levels of theory for 401
362 transmission IR spectrum might indicate that an apparent an isolated ion, ion pair or cluster of ions, and molecular 402
363 “dry” sample of ionic liquid still contains a significant amount of dynamics simulation (classical or ab initio) to calculate an 403
364 water, depending on the ionic liquid. appropriate time correlation function and then its Fourier 404
365 Raman spectra as a function of temperature were obtained in transform to result in the vibrational spectrum. Different 405
366 this work for some ionic liquids. We used an OptistatDN methods are chosen not merely on the basis of molecular 406
367 cryostat (Oxford Instruments) filled with liquid nitrogen, thus structure complexity but also on the level of analysis one wishes 407
368 allowing for the achievement of temperatures as low as 77 K at and approximations allowed for. All of these theoretical 408
369 which ionic liquids are in crystalline or glassy phases. Raman methods have been used for assigning vibrational spectra of 409
370 spectra were also obtained for some ionic liquids under the most common ions to be discussed in section 4, so that it is 410
371 pressure within the GPa range by using a diamond anvil cell useful to provide here a brief account of these methods for 411
372 (DAC).36,37 Spectra as a function of pressure at room future reference along the review. 412
373 temperature were obtained with the already mentioned Classical normal coordinate analysis is based on a ball-and- 413
374 Horiba-Jobin-Yvon T64000 spectrometer having a coupled spring model for the nuclei vibrations with no consideration of 414
375 Olympus BX41 microscope. The Raman spectra were excited electronic degrees of freedom.41,42 Taking the [SCN]− anion as 415

E DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

416 an example, a quadratic force field proposed by Baddiel and There are some accounts in the literature of ionic liquids 473
417 Janz43 involves internal coordinates for stretching of C−N and concerning the performance of quantum chemistry methods for 474
418 C−S bonds, ΔrCN and ΔrCS, and bending of the SCN angle, calculating different properties (e.g., anion−cation binding 475
419 Δθ, energy44,45 or gas−ion interactions for studying gas capture by 476
ionic liquids).46 On the other hand, there are no benchmark 477
2 2
2V = k CNΔrCN + k CSΔrCS + k CN,CSΔrCNΔrCS + kθ Δθ 2 studies for ab initio calculations of vibrational spectra of ionic 478
420 (1) liquids, although some works provide specific comparisons of 479
vibrational frequencies calculated with different levels of 480
421 with corresponding force constants kCN, kCS, and kθ, and theory.47−49 Density functional theory (DFT) is one of the 481
422 interaction constant kCN,CS. Force constants are second most widely used level of theory for the calculation of 482
423 derivatives of the potential energy function at the equilibrium vibrational frequencies of ionic liquids since it provides a 483
424 configuration, and they are optimized in order to reproduce the satisfactory compromise between accuracy and computational 484
425 experimental vibrational frequencies. A normal mode Qα is a time. Overall, the DFT/B3LYP with 6-311++G(d,p) basis set 485
426 coordinate in which all atoms execute harmonic vibrations with gives good results without being computationally expensive. 486
427 the same frequency leading to a Hamiltonian function, H = T + Second-order Møller−Plesset perturbation theory (MP2) has 487
428 V, where T is the kinetic energy, as a collection of independent been also extensively used in the context of vibrational 488
429 oscillators without cross-terms between different coordinates: spectroscopy of ionic liquids. We used these methods to 489

3N calculate vibrational frequencies for specific purposes of this 490


1 ̇2 review.
H= ∑ (Q + λαQ α2) 491

430 α=1
2 α (2) The ultimate aim of calculating vibrational frequencies is the 492
assignment of vibrational motions to observed frequencies. 493
431 where the eigenvalue λα of mode α is related to the square of its Most of the works dealing with calculation of vibrational 494
432 vibrational frequency, λα = (2πνα)2, and the coordinates for spectra of ionic liquids have interpreted the normal mode 495
433 convenience are weighted by the square root of atomic masses. composition by direct visualization of atomic displacements on 496
434 The number of nuclei is N, and the number of actual vibrational the computer screen. This is a helpful and easy way of assigning 497
435 motions, excluding rotational and translational motions of zero the vibrations, although not a quantitative one. Furthermore, it 498
436 frequency, is 3N − 6 in general or 3N − 5 for linear molecules. might be misleading because large displacement of hydrogen 499
437 If this classical Hamiltonian function is used to build the atoms eventually is only a matter of the small mass of hydrogen 500
438 Hamiltonian operator in the quantum chemistry treatment, the but not implying large contribution to the normal mode energy. 501
439 set of independent harmonic oscillators implies that the total Potential energy distribution (PED) in the classical normal 502
440 vibrational energy is the sum of the well-known textbook result coordinate analysis is accomplished by evaluating the percent 503
441 for each normal mode, Eα = hνα (vα + 1/2), where h is the weight of each symmetry coordinate Sj contributing to the 504
442 Planck constant, vα is the vibrational quantum number, vα = 0, potential energy of the normal mode Qα. The PED can also be 505
443 1, 2···, and να is the same as the classical vibrational frequency. written in terms of internal coordinates, an approach which is 506
444 Each Qα is a linear combination of the original coordinates of chemically appealing as one understands the vibrations in terms 507
445 displacements for all of the atoms of the molecule. The linear of changes in bond lengths and angles of valence, out-of-plane, 508
446 combination can be done in terms of the 3N atomic Cartesian and torsional. Jamróz50 made available the computer code 509
447 coordinates, in terms of internal coordinates of bond-length VEDA (vibrational energy distribution analysis) which takes the 510
448 and angle variations, or in terms of symmetry coordinates, Sj. In output of the commonly used Gaussian51 package of quantum 511
449 the so-called method of GF matrixes of Wilson, Decius, and chemistry. The matrixes containing atomic displacements and 512
450 Cross,41 the symmetry coordinate is a new coordinate system force constants in terms of Cartesian coordinates are available 513
451 mixing the internal coordinates, so that each Sj belongs to a along the calculation of vibrational frequencies with the 514
452 given symmetry species of the point group of the molecule. The Gaussian program. The VEDA program then uses the 515
453 advantage of proposing symmetry coordinates relies on the fact molecular structure, automatically sets internal coordinates, 516
454 that only Sj of the same symmetry species will mix together in and evaluates how much each one contributes to the energy of 517
455 the composition of a given normal mode Qα. Furthermore, a given normal mode. We show in Figure 23 atomic 518
456 defining symmetry coordinates makes easier the resolution of displacements for two normal modes of [C 4 C 1 C 1 im] + 519
457 the secular equation for obtaining the eigenvalues because it calculated by the Gaussian program that illustrate the large 520
458 factors into diagonal blocks for each symmetry species. amplitude of motions of hydrogen atoms, although PED 521
459 In the context of ionic liquids, the classical normal coordinate calculated by the VEDA program indicate they contribute little 522
460 analysis is more appropriate for simple ions, typically highly to the normal modes energies. 523
461 symmetric anions. Most of vibrational frequency calculations of Harmonic vibrational frequencies are obtained in these ab 524
462 ionic liquids referred to in section 4 are based instead on initio quantum chemistry calculations. No negative eigenvalue 525
463 quantum chemistry methods. The electronic structure problem (i.e., no imaginary frequency) is obtained as long as the 526
464 of the molecule is solved within a given approximation for the minimum energy molecular structure has been correctly 527
465 electronic wave function, and the minimum energy config- identified. However, theoretical harmonic frequencies are 528
466 uration of nuclei is found. For this configuration, one obtains usually higher than experimentally observed, so that scaling 529
467 the 3N × 3N Hessian matrix (i.e., the matrix containing second factors multiplying the frequencies are needed to bring 530
468 derivatives of potential energy in terms of the nuclei Cartesian calculations into agreement with the experiment. Taking 1- 531
469 coordinates). Diagonalization of the Hessian matrix gives the alkyl-3-methylimidazolim cations as examples, Talaty et al.52 532
470 eigenvalues λα along the diagonal, and the eigenvector and Heimer et al.53 found for [CnC1im][PF6] and [CnC1im]- 533
471 corresponding to each frequency gives the composition of Qα [BF4], n = 2, 3, and 4, it was necessary to multiply harmonic 534
472 in terms of atomic displacements. frequencies calculated by the DFT/B3LYP level of theory by 535

F DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

536 scaling factors within the range 0.963−0.967. Katsyuba et al.54 Raman spectroscopy (see section 5). The INM analysis of 599
537 proposed sets of different scaling factors within the range of molten ZnCl2 was useful to disentangle collective soundlike 600
538 0.889−1.0144 for each kind of normal mode (stretch, bend, modes and localized vibrations of a molecular-like structure of 601
539 torsion, and out of plane) of the [C2C1im]+ cation. In order to ZnCl4 tetrahedral.67,68 In the case of BeCl2 and NaF-AlF3 602
540 go beyond the calculation of harmonic frequencies and to avoid mixtures, INM of frequencies up to ∼1000 cm−1 are obtained 603
541 empirical scaling factors, anharmonicity should be included. because of relatively stable [BeCl2]n oligomers69 and AlFn(n−3)− 604
542 The concept of independent normal coordinates is no longer polyhedrals.70 605
543 rigorously valid when anharmonicity is considered because Dynamical effects are not included in a quantum chemistry 606
544 there appears additional potential terms in the Hamiltonian calculation of vibrational spectrum once it is carried out in a 607
545 function of eq 2 mixing together different coordinates, minimum energy configuration of an isolated ion, an ionic pair, 608
546 ∑α,β,γgαβγQαQβQγ + ∑α,β,γ,δhαβγδQαQβQγQδ + ..., where gαβγ or a cluster of ions. In contrast, the theoretical framework of 609
547 and hαβγδ are anharmonic constants corresponding, respectively, time correlation function71−74 is appropriated for calculating 610
548 to third and fourth derivatives of the potential energy.42 vibrational spectra by MD simulations of liquids. A time 611
549 Nevertheless, one still thinks in terms of normal modes because correlation function measures how a system property at given 612
550 anharmonic effects are taken into account as perturbations on time, A(t), correlates with another property at a previous time, 613
551 the states and energies of the harmonic model according to B(0). In an autocorrelation function, CA(t) = < A(0)·A(t)>, 614
552 perturbation theory of quantum chemistry. The harmonic where <···> indicates an average according to statistical 615
553 frequency of a given normal mode continues being related to mechanics, CA(t) starts from < |A|2> at time zero and reaches 616
554 the force constant as in the harmonic case, but the correction the long time value < |A|>2. Figure 7 illustrates the behavior of a 617 f7
555 on energy levels mixes gαβγ and hαβγδ in complicated expressions time correlation function for the atomic velocities, Cv(t) = < 618
556 depending on the order of the expansion of the potential vi(0)·vi(t)>, calculated by MD simulation of [C2C1im][NTf2]. 619
557 energy function and the order of perturbation theory.42,55,56 The intra- and intermolecular potential function we used in this 620
558 Barone et al.57 made available a code including cubic and simulation is the CL&P model proposed by Lopes and Padua 621
559 quartic anharmonicity constants at second-order perturbation
560 theory (VPT2) to correct harmonic vibrational frequencies
561 calculated by the Gaussian program. Anharmonic frequencies
562 have been indeed calculated by the method of Barone for ionic
563 liquids based on 1-alkyl-3-methylimidazolium cations.58 The
564 vibrational frequency shift when anharmonicity is included in
565 the calculation of an isolated 1-alkyl-3-methylimidazolium
566 cation might be comparable in magnitude to the effect of
567 considering a cation−anion pair, in particular for those
568 vibrations involving C−H stretching motions. Including
569 anharmonicity not only downshifts vibrational frequencies but
570 also allows for addressing important effects on the experimental
571 spectra (e.g., Fermi resonance).42 This has been found
572 particularly important in the high-frequency range of vibrational
573 spectra of 1-alkyl-3-methylimidazolium cations because this
574 spectral range is prone to Fermi resonance between overtones
575 or combination bands of ring vibrations with the C−H
576 stretching modes.58,59
577 The harmonic model of expanding the potential energy
578 function to quadratic terms in coordinates has also been
579 proposed for an assembly of molecules forming a liquid. If
580 liquid phase configurations are generated along a computer
581 simulation, either by Monte Carlo (MC) or molecular
582 dynamics (MD),60 the Hessian matrix can be evaluated for
583 each configuration. Eigenvalues and eigenvectors resulting from
584 the diagonalization of the Hessian matrix give frequencies and
585 composition of normal coordinates for an instantaneous
586 configuration of the liquid, so that these are called
587 instantaneous normal modes (INM).61,62 Part of the INM
588 has imaginary frequency because a liquid configuration of
589 thousands of particles is not a minimum energy configuration. Figure 7. Upper panel: power spectra, P(ω), calculated according to
590 Nevertheless, the fraction of imaginary frequency INM also eq 3 by classical MD simulation of [C2C1im][NTf2] at 400 K and
591 carries physical information; for instance, it has been related to density 1.05 g cm−3. The MD simulation considered 500 ion pairs and
592 ionic diffusion coefficients.63−65 The INM method was applied the CL&P force field.75 The total P(ω) (black line) has been split into
contributions of atoms belonging to anions (red line) and cations
593 in MD simulations of (high temperature) molten salts (e.g.
(green line). Intensity was normalized by the most intense band of
594 ZnCl2,66−68 BeCl2,69 and cryolitic NaF-AlF3 mixtures)70 but each P(ω). The inset highlights the low-frequency range of the total
595 not yet for (room temperature) ionic liquids. In atomic molten P(ω). Bottom panel: the normalized time correlation functions of
596 salts, the positive frequency INM covering the low-frequency velocity, Cv(t), obtained from the Fourier transform of the
597 range are direct probes of the intermolecular dynamics as corresponding power spectrum. The inset shows the total Cv(t) in a
598 experimentally accessible by far-infrared and low-frequency wider time range.

G DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

622 for MD simulations of ionic liquids.75 Figure 7 shows Cv(t), M(t) may be approximated at first order as the sum of 679
623 including all of the atoms i, and also Cv(t) calculated separately individual molecular dipole moments, although induced 680
624 for atoms belonging either to cations or anions. The Cv(t) is interaction effects imply that M(t) is a collective property.86 681
625 shown normalized by its initial value, decaying to the average An expression analogous to eq 4 follows for the Raman 682
626 value of velocity which is zero. Figure 7 also shows the so-called spectrum as the Fourier transform of the time correlation 683
627 power spectrum, P(ω) (i.e., the density of vibrational states). function of polarizability fluctuation. The alternative method of 684
628 Given any time-dependent property A(t), its power spectrum eq 3 also applies to the calculation of IR or Raman spectrum 685
629 PA(ω) can be obtained as the Fourier transform of the from the time-dependence of dipole moment or polarizability. 686
630 corresponding time correlation function, CA(t) = < A(0)·A(t)>. Vibrational spectra resulting from time correlation functions of 687
631 An alternative method to calculate PA(ω) more efficiently dipole moment or polarizability fluctuations calculated by ab 688
632 follows from the spectral density of A(t):74,76−78 initio MD simulations show reasonable agreement to 689
experimental spectra of ionic liquids.33 This approach for 690

calculating vibrational spectra of liquids has been implemented 691
PA(ω) = | ∫ A(t ) e−iωt |2 in the TRAVIS (Trajectory Analyzer and VISualizer) package,87 692
633 −∞ (3) which includes several routines for analyzing trajectories 693
generated by computer simulations. 694
634 Then, the CA(t) can be obtained by Fourier transforming the The time correlation function approach for vibrational 695
635 PA(ω). spectroscopy allows for a direct comparison between simulation 696
636 The Cv(t) shown in the inset of the bottom panel of Figure 7 and experiment in terms of peak positions, band shapes, and 697
637 exhibits an overall damped oscillatory decay within the relative intensities, but the ultimate goal of the normal mode 698
638 picosecond time range, arising from rattling dynamics of ions assignment is not yet reached. One possibility is to take a few 699
639 within a temporary cage made of neighboring ions. This liquid configurations generated by the MD simulation for 700
640 intermolecular dynamics is manifested in the P(ω) within the energy minimization using, for instance, the conjugate gradient 701
641 frequency range below ca. 100 cm−1 (see inset in the top panel method.88,89 In contrast to the INM analysis discussed above, 702
642 of Figure 7), which is the range accessible by far-IR and low- this approach calculates the Hessian matrix and performs the 703
643 frequency Raman spectroscopy. The CL&P model considers normal-mode analysis for those quenched configurations. In the 704
644 flexible ions, so that the very fast oscillations in Cv(t) (main time correlation function approach, however, the actual nature 705
645 figure at the bottom panel of Figure 7) arise from the of the molecular vibration responsible for a given band should 706
646 intramolecular vibrations. Accordingly, P(ω) shows high- be retrieved from the spectrum obtained from the Fourier 707
647 frequency peaks assigned to cation and anion intramolecular transform. A generalized normal coordinates approach90,91 has 708
648 vibrations. been applied for an anharmonic Hamiltonian as it is the case in 709
649 Vibrational frequencies in P(ω) manifest condensed phase ab initio MD simulations of ionic liquids. The approach is 710
650 and anharmonicity effects, as long as an anharmonic intra- based on a generalization of <vi(0)·vi(t)>, which is a single 711
651 molecular potential function is included in the model. It is particle time correlation function as it involves the property of a 712
652 worth noting, however, that the CL&P model75 considers given particle i. The generalization accounts for calculating the 713
653 harmonic terms for stretching and bending motions. On the collective counterpart of mass-weighted velocities, <mi1/2vi(0)· 714
654 other hand, it is well-known that proper coupling between mj1/2vj(t)>, whose Fourier transform gives a power spectrum 715
655 intra- and intermolecular degrees of freedom, and the Pij(ω), including cross-correlations between different particles. 716
656 consequent vibrational frequency shift and vibrational relaxa- The matrix relating Cartesian to the generalized normal 717
657 tion in the liquid with respect to the gas phase, is heavily coordinates is obtained according to a recipe which minimizes 718
658 dependent on anharmonic terms in the intramolecular off-diagonal terms of Pij(ω) for all of the frequencies. The 719
659 potential.79−81 Calculations of P(ω) have been done by ab procedure brings the collective power spectrum as close as 720
660 initio MD simulations,82,78 which do not rely on an empirical possible to a diagonal form leading to a representation of 721
661 parametrized force field, instead the electronic structure is vibrations in terms of normal modes. This approach for 722
662 solved along the simulation run giving the forces that move the calculating normal modes from computer simulation includes 723
663 nuclei to the new configuration. Therefore, anharmonicity of anharmonicity and condensed phase effects, and it has been 724
664 vibrations are taken into account when P(ω) of ionic liquids are applied to assign vibrational spectra of [C2C1im][CH3COO] 725
665 calculated by ab initio MD simulations. and its mixture with CO2 and water.33 726
666 The power spectrum shown in Figure 7 is not a theoretical
667 IR or Raman spectrum.78 P(ω) exhibits all of vibrations 4. VIBRATIONAL SPECTROSCOPY OF PURE IONIC
668 included in the model, and it has to be weighted by how much LIQUIDS IN THE MID-FREQUENCY RANGE 727
669 the intra- and intermolecular dynamics fluctuate the electric
670 dipole moment or the polarizability in order to represent the IR 4.1. Vibrational Frequencies of Anions
671 or the Raman spectrum, respectively. In other words, IR and 4.1.1. Small Symmetric Anions. Vibrational spectroscopy 728
672 Raman activities are determined by the coupling between is of long usage for studying (high temperature) molten salts of 729
673 mechanical vibrations and electronic molecular properties. It is polyatomic inorganic anions, some of which are common 730
674 a formal result of nonequilibrium statistical mechanics that the species in ionic liquids. Reliable assignment of vibrational 731
675 IR spectrum, IIR(ω), is proportional to the Fourier transform of frequencies is needed in order to use IR and Raman 732
676 the time correlation function of fluctuations of the electric spectroscopies as useful tools for unravelling molecular 733
677 dipole moment of the whole system:74,77,83−85 conformations, intermolecular interactions, hydrogen bonds, 734


etc. The highly symmetric structures of inorganic anions allow 735

678
IIR (ω) ∝ ∫−∞ ⟨M(0) ·M(t )⟩eiωt dt
(4)
for more straightforward assignment of vibrational frequencies
in comparison with the ionic liquids forming organic cations.
736
737

H DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 1. Fundamental Frequencies of Some Highly Symmetric Anions Commonly Used in Ionic Liquidsa
[NO3]− [BF4]− [PF6]−
IR Raman D3h IR Raman Td IR Raman Oh
1346 vs νas (E′) 1062 vs νas (F2) 843 vs 864 wb νas (F1u)
1041 wb 1041 vs νs (A1′) 764 mb 764 vs νs (A1) 741 mb 740 vs νs (A1g)
830 m γ (A2″) 522 m 521 w δ (F2) 565 wb 568 m ν (Eg)
708 w 706 m δ (E′) 352 w δ (E) 558 s 560 wb δ (F1u)
470 wb 471 m δ (F2g)
[SCN]− [N(CN)2]−
IR Raman C∞v IR Raman C2v
2056 vs 2054 s ν(CN) (Σ ) +
2192 s 2192 s νs(CN) (A1)
737 w 738 m ν(CS) (Σ+) 2133 vs 2133 w νas(CN) (B2)
471 w δ (Π) 1309 s νas(N−C) (B2)
904 w 904 w νs(N−C) (A1)
666 m δ(CNC) (A1)
524 wb γs(N−CN) (A2)
508 w γas(N−CN) (B1)
496 w δas(N−CN) (B2)
[C(CN)3]− [B(CN)4]−
IR Raman D3h IR Raman Td
2209 vwb 2209 vs νs(CN) (A1′) 2223 vs νs(CN) (A1)
2165 vs 2163 s νas(CN) (E′) 2223 m νas(CN) (F2)
1257 m 1257 m δ((CN)C(CN)) + ν(C−CN) (E′) 938 s ν(B−C) (F2)
647 wb 646 m νs(C−C) (A1′) 521 w δ(BCN) (E)
563 s 565 wb δ(CCN) (A2″) 496 m δ(BCN) (F2)
483 w δ(CCN) (E″) 483 m ν(B(CN)4) (A1)
a
Values (cm−1) correspond to frequencies observed in IR and Raman spectra at room temperature for ionic liquids with the [C4C1im]+ cation for
[NO3]−, [BF4]−, and [PF6]− anions and the [C2C1im]+ cation for the cyanate anions. ν, stretch; δ, bend; γ, out-of-plane; s, symmetric; as,
antisymmetric; w, weak; m, medium; s, strong; v, very. bInactive according to the symmetry of isolated species.

Figure 8. IR (red, transmittance scale at right) and Raman (black) spectra of [C4C1im][BF4] (left panel) and [C4C1im][PF6] (right panel) at room
temperature in the range of the totally symmetric stretching mode of the anion.

738 There is accumulated knowledge on how vibrational molten salts based on alkali cations. For instance, the 755
739 frequencies of anions depend on cation charge and size, vibrational frequency of the totally symmetric stretching 756
740 solid−liquid phase transition, temperature, and dilution in mode of the nitrate anion, νs(NO3), in molten alkali nitrates 757
741 solvents of different dielectric constants. Vibrational frequency (at T ∼ 400 °C) exhibits a significant downward shift from 758
742 of stretching motion of simple anion decreases when the cation 1067 to 1043 cm−1 when the counterion is changed along the 759
743 is replaced by another with lower polarizing power (i.e., the sequence of Li+, Na+, K+, Rb+, and Cs+.93 Accordingly, νs(NO3) 760
744 ratio between charge and ionic radius of the cation).92 In terms is observed at 1041 cm−1 in molten [C4C1im][NO3] at 313 K. 761
745 of mechanical models of balls and springs,80,79 vibrational Analogous effect of cation polarizing power is found in the 762
746 frequency shift in the condensed phase is a subtle balance CN stretching mode of [SCN]−. In molten LiSCN, NaSCN, 763
747 between attractive and repulsive intermolecular forces con- and KSCN, ν(CN) follows the trend 2083, 2074, and 2068 764
748 tributing, respectively, to negative and positive shift in cm−1.43 Arguing beyond mechanical springs−and−balls models, 765
749 comparison with the vibrational frequency of the free species. Chabanel et al.94 claimed that strongly polarizing cations 766
750 The very fact that anion vibrational frequencies increase with stabilize the anion σ orbitals. This effect of strengthening the 767
751 strength of interaction with the cation points out the role CN bond is partially counterbalanced by the effect of resonance 768
752 played by short-range repulsive forces on the probe oscillator.92 structures, −S−CN ↔ [SCN]−, that softens the CN 769
753 As ionic liquids are usually made of bulky organic cations, the bond and becomes predominant as larger is the cation. 770
754 anion stretching mode is expected at lower frequency than Accordingly, the ν(CN) mode is observed at 2054 cm−1 in 771

I DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

772 the ionic liquid [C2C1im][SCN]. Since a vast literature is dipole or polarizability adds further mechanism of fluctuation of 835
773 available reporting vibrational frequencies for simple anions electrical properties beyond vibrational and rotational dynamics 836
t1 774 with different counterions in solid or liquid phases,95 Table 1 of the probe molecule.86 Being (o)μi(t) and (o)αi(t), respectively, 837
775 gives for reference purpose the values actually observed for dipole and polarizability of the isolated molecule i, where time 838
776 some highly symmetrical anions in ionic liquids based on the dependence comes from vibration and rotation of the molecule, 839
777 [C2C1im]+ or [C4C1im]+ cations. the actual dipole and polarizability of the molecule in the liquid 840
778 Symmetry considerations are more easily applied for have additional contributions, (I)μi(t) and (I)αi(t), induced by 841
779 vibrations of relatively rigid inorganic anions than for the interactions with neighbor molecules. Let Qi be an IR inactive 842
780 complex cations of ionic liquids. Vibrational spectroscopy was a normal mode of molecule i (i.e., the transition dipole ∂(o)μi/∂Qi 843
781 powerful tool to unravel the presence of complex anionic is zero for the isolated molecule) but Raman active (i.e., 844
782 species in the early period of highly hygroscopic ionic liquids nonzero ∂(o)αi/∂Qi). If the molecule i is surrounded by dipolar 845
783 based on mixing AlCl3 and organic cations chloride. Using molecules j, the dipole-induced dipole (DID) mechanism at 846
784 Raman spectroscopy, Takahashi et al.96 showed that in the this order is enough to account for an interaction-induced 847
785 basic solution of AlCl3/[C2C1im]Cl, x(AlCl3) = 0.54, the contribution to the dipole of i, (I)μi = Σj≠i (∂(o)αi/∂Qi)· 848
786 prevailing anionic species is [AlCl4]− belonging to the Td point T(rij)·(o)μj, where rij is the distance vector between the 849
787 group, whereas in acidic solution, x(AlCl3) = 0.67 and the molecules, and the dipole−dipole tensor is T(r ij ) = 850
788 prevailing species is [Al2Cl7]− belonging to the C2 point group (4πεo)−1(3rij·rij − rij2·I)r−5, where I is the unitary matrix. In 851
789 (acidic melts have AlCl3:organic salt mole ratio greater than 1). general, the total dipole and polarizability must be obtained in a 852
790 Grondin et al.58 suggested that the octahedral symmetry of self-consistent way because the electric properties of all of the 853
791 the [PF6]− anion is perturbed in the ionic liquid [C4C1im]- molecules contain permanent and interaction-induced parts.86 854
792 [PF6] since the totally symmetric stretching mode (A1g), which In the case of Raman spectra of ionic systems, Madden et 855
793 is IR inactive on the basis of the Oh point group of the isolated al.100,101 showed that other mechanisms might contribute to the 856
794 [PF6]−, is actually observed in the IR spectrum. Analogous fluctuating polarizability besides the DID mechanism (e.g., 857
795 effect has been found by Katsyuba et al.,54 who observed the distortion of a given ion by the Coulomb field of other ions and 858
796 [BF4]− band corresponding to the νs(A1) mode in the IR short-range overlap interactions). 859
797 spectrum of [C2C1im][BF4]. These findings are shown in Liquid carbon disulfide is a well-known example of 860
f8 798 Figure 8, which compares IR and Raman spectra of interaction−induced effect contributing to vibrational spec- 861
799 [C4C1im][BF4] and [C4C1im][PF6] in the range where the tra:102−104 the doubly degenerate bending and the antisym- 862
800 anion totally symmetric mode is observed. Environmental effect metric stretching modes, which are Raman inactive for an 863
801 on the anion symmetry is an issue in vibrational spectroscopy isolated CS2 molecule, become active in the Raman spectrum of 864
802 of molten salts,92 nitrate certainly being the most investigated the liquid phase. Additional signature of interaction−induced 865
803 anion.93,97,98 Experimental evidence of perturbation on the D3h effect is an exponential long tail, e−ω/Δ, where Δ is a parameter, 866
804 symmetry of [NO3]− is the IR band of the νs(A1′) or the in a Raman band corresponding to a normal mode which is 867
805 Raman band of the γ(A2″) mode and split of νas(E′) because of allowed by symmetry of the isolated molecule.102,103 Depolar- 868
806 the degeneracy lift. The split of νas(E′) is seen in molten LiNO3 ized Raman spectra have been interpreted by assuming there is 869
807 but not in molten NaNO3, KNO3, RbNO3, and CsNO3. These time separation between fast rattling dynamics of molecules 870
808 condensed phase effects in the [NO3]− vibrations might be within the cage of neighbors, resulting in interaction-induced 871
809 understood on the basis of relatively strong ion-pairing leading effect as intermolecular distances are changed, and slow 872
810 to C2v symmetry when the cation has high polarizing power. reorientational dynamics of the molecule as a whole. The 873
811 For less polarizing cations, one considers that the D3h symmetry short-time rattling dynamics contributes to the high-frequency 874
812 of [NO3]− is retained, but the environmental effects are enough tail of the band, whereas the relatively slow reorientational 875
813 to breakdown selection rules for the isolated nitrate. dynamics contributes to the center of the band. Unfortunately, 876
814 Accordingly, Table 1 indicates that the νs(A1′) mode is there are overlaps of bands in vibrational spectra of ionic 877
815 observed as a weak band in the IR spectrum of [C4C1im]- liquids, being difficult to address whether the high-frequency 878
816 [NO3]. In the case of [C4C1im][PF6], even though νs(A1g) is range of a Raman band exhibits exponential shape. A particular 879
817 observed in the IR spectrum, no lifting of degeneracy of anion band that could be considered in this respect is the stretching 880
818 bands has been observed in the IR and Raman spectra.58 mode ν(CN) of [SCN]− because it appears in a spectral range 881
819 Therefore, Grondin et al.58 proposed that [PF6]− preserves a free of overlaps. Figure 9 shows the depolarized ν(CN) Raman 882 f9
820 quasi-octahedral symmetry in the ionic liquid. We found the band of [C2C1im][SCN] at room temperature. The ν(CN) 883
821 νas(F2) mode of [BF4]− as a broad band with maximum at 1062 Raman band indeed exhibits a long tail which extends far from 884
822 cm−1 in the IR spectrum of [C4C1im][BF4]. In contrast, the band-center. The inset of Figure 9 makes clear that the tail 885
823 Holomb et al.99 found that the νas(F2) mode splits into three exhibits the exponential shape, strongly suggesting that 886
824 peaks (1015, 1033, and 1045 cm−1) in the IR spectrum of interaction-induced mechanisms should not be ruled out. It 887
825 [C4C1im][BF4]. We found in the Raman spectrum of would be interesting for future work attempts to disentangle 888
826 [C4C1im][BF4] a band at 493 cm−1, which may be the result symmetry reduction and interaction-induced effects in vibra- 889
827 of splitting of the F2 bending mode of [BF4]−. This assignment tional spectroscopy of ionic liquids. 890
828 is supported by ab initio calculations performed in this work Classical intramolecular force fields for computing vibrational 891
829 (MP2 level of theory, aug-cc-pVDZ basis set) for a [BF4]− frequencies are available for some ionic liquid forming anions 892
830 anion under the perturbation of an external positive charge. (e.g., [NO3]−,98 [SCN]−,43 and [C(CN)3]−).105 On the other 893
831 Symmetry reduction is not the only condensed phase effect hand, the complex molecular structures of the ions typically 894
832 leading to IR or Raman activity of an otherwise inactive involved in ionic liquids prompt for quantum chemistry as the 895
833 vibration according to the point group of the isolated molecule. more appropriate approach to calculate vibrational frequencies. 896
834 The so-called interaction-induced contribution to the molecular Hipps and Aplin105 accounted for the vibrations of [C(CN)3]− 897

J DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

opposite is true for [N(CN)2]− and [C(CN)3]−. The IR, 927


polarized, and depolarized Raman spectra shown in Figure 10 928
for [C2C1im][N(CN)2] and [C2C1im][C(CN)3] in the spectral 929
range covering νas(CN) and νs(CN) modes support the 930
assignment. Crystallographic analysis108 of Li[N(CN)2] 931
showed that the N−C bond length (131 pm) is in between 932
typical values for single and double bonds, whereas the CN 933
bond length (116 pm) is in between double and triple bonds. 934
Therefore, Reckeweg et al.108 concluded that the electronic 935
structure is better represented as the resonance hybrid −N 936
CN−CN ↔ NC−NCN−, rather than NC− 937
N−−CN. On the basis of this resonance, it is reasonable that 938
a vibration will be of lower energy when the two CN moieties 939
oscillate antisymmetrically. The IR spectrum of [C2C1im][N- 940
(CN)2] exhibits another band at 2227 cm−1 (just out of the 941
Figure 9. Depolarized Raman spectrum of the ionic liquid [C2C1im]-
[SCN] at room temperature in the range of the ν(CN) normal mode. spectral window shown in Figure 10), which has been 942

Intensity has been normalized by the maximum of the band. The inset assigned109 to a combination band of νs(N−C) and νas(N− 943
shows the logarithm of the high frequency side of the band, where the C) modes whose frequency is shifted and intensity is enhanced 944
red line is a linear fit highlighting the exponential tail, e−ω/Δ, with Δ = because of the Fermi resonance with the fundamental of the 945
36.7 cm−1. νas(CN) mode. In the case of [B(CN)4]−, the Raman active 946
νs(CN) and the IR active νas(CN) are observed at the same 947
898 in the potassium salt with a quadratic force field with force frequency.110 948
899 constants for internal coordinates of bond displacements and 4.1.2. More Complex Fluorinated Anions. The bis- 949
900 angles based on ab initio calculation. The carbon−carbon (trifluoromethanesulfonyl)imide, [NTf2]−, is one of the most 950
901 distance and force constant obtained for [C(CN)3]− are similar popular anions, resulting in low melting point salts when 951
902 to values for benzene, indicating significant resonance combined with the majority of common organic cations. 952
903 stabilization in [C(CN)3]−.105 We have found bands at 1231 Assignment of vibrational frequencies of [NTf2]− has been an 953
904 and 1230 cm−1 in IR and Raman spectra, respectively, of issue in the literature because the lithium salt of [NTf2]− is 954
905 [C2C1im][C(CN)3]. In accordance with the assignment commonly used in polymer electrolytes. Rey et al.111 provided a 955
906 suggested by Hipps and Aplin for simpler [C(CN)3 ] − detailed analysis of the [NTf2]− normal modes on the basis of 956
907 salts,105 this is a combination band intensified by Fermi quantum chemistry calculations for its C2 conformer. In a 957
908 resonance with the fundamental E′ mode observed at 1257
subsequent IR and Raman spectroscopic study, Herstedt et
cm−1 in IR and Raman spectra of [C2C1im][C(CN)3]. The
958
909
al.112 were able to distinguish bands characteristic of C2 and C1 959
910 occurrence of this Fermi resonance was indeed confirmed by
conformers of [NTf2]−, commonly called transoid and cisoid,
911 anharmonic ab initio calculations of an isolated [C(CN)3]− 960

912 anion performed in this work (MP2/VPT2 level of theory, aug- respectively, in solutions of Li[NTf2] in ethers CH3O- 961

913 cc-pVDZ basis set). The calculation showed that the 1230 cm−1 (CH2CH2O)nCH3, where n = 1, 2, 3, and 4. A complete 962

914 band is the combination of A1′ δ(CCN) and E′ νs(C−CN) table of experimental versus calculated vibrational frequencies, 963

915 modes calculated at 606 and 641 cm−1, respectively. internal force constants, and potential energy distribution of 964

916 The vibrational frequency of the totally symmetric CN [NTf2]− normal modes can be found in these papers.111,112 965

917 stretching mode decreases along the sequence [B(CN)4]−, Some of the bands assigned to [NTf2]− vibrations actually 966
918 [C(CN)3]−, and [N(CN)2]− as the electronic delocalization observed in IR and Raman spectra of a typical ionic liquid, 967
919 increases by 2223, 2209, and 2192 cm−1, respectively, in ionic [C2C1im][NTf2], are marked in Figure 11. Tables containing 968 f11
f10 920 liquids with the same [C2C1im]+ cation (see Figure 10).106 In the [NTf2]− vibrational frequencies are available in many 969
921 an IR spectroscopy study of solid [NH4][N(CN)2], Sprague et papers,111−123 but values strongly depend on the counterion, 970
922 al.107 assigned the nontotally symmetric mode νas(CN) at physical state, and solvent. Thus, Table 2 lists the actual values 971 t2
923 higher frequency than the totally symmetric mode νs(CN). It observed in [C2C1im][NTf2] at room temperature. The 972
924 is a rule of thumb in vibrational spectroscopy assigning the eigenvectors of [NTf2]− normal modes exhibit a complex 973
925 nontotally symmetric stretching at higher frequency than the pattern involving displacements of many atoms,112 so that the 974
926 totally symmetric stretching of a given moiety. However, the assignment in Table 2 is only a simplification of the potential 975

Figure 10. IR (red, transmittance scale at right) and Raman spectra (black; full line, polarized; dashed line, depolarized) of ionic liquids
[C2C1im][N(CN)2] (left), [C2C1im][C(CN)3] (middle), and [C2C1im][B(CN)4] (right) at room temperature.

K DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

cisoid [NTf2]− conformers in the normal liquid phase of ionic 985


liquids. Ionic liquids in which [NTf2]− conformers have been 986
found by vibrational spectroscopy include systems based on 987
derivatives of imidazolium,119,3,124,125 pyrrolidinium126,127 988
(either protic or nonprotic derivatives of these cations),116,118 989
piperidinium,120 and ammonium.122,123,128 Quantum chemistry 990
calculations indicate that the spectral ranges 260−370 cm−1 and 991
620−660 cm−1 are well-suited for finding bands that character- 992
ize the transoid or the cisoid conformation. These spectral 993
ranges of the Raman spectrum of [C2C1im][NTf2] are 994
highlighted in Figure 12. The antisymmetric SO2 out-of-plane 995 f12
bending at 623 cm−1 characterizes the transoid conformer, 996
whereas the S−N−S bending at 650 cm−1 characterizes the 997
cisoid conformer. These are very weak Raman bands so that the 998
Figure 11. IR (red, transmittance scale at right) and Raman spectra
(black) of [C2C1im][NTf2] at room temperature. Some bands
260−370 cm−1 range is more appropriated for identifying the 999

assigned to the [NTf2]− normal modes are indicated by arrows. [NTf2]− conformation. In particular, the SO2 rocking at 326 1000
cm−1 is characteristic of the cisoid conformer. The quantum 1001
chemistry assignments marked in Figure 12 are supported by 1002
976 energy distribution among the different internal coordinates remarkable spectral changes eventually observed in crystalline 1003
977 given in ref 111. phases as bands belonging to a given conformer might be 1004
978 It is clear from Figure 11 that the Raman spectrum of absent. Following phase transitions of ionic liquids by 1005
979 [C2C1im][NTf2] is dominated by anion bands because of large vibrational spectroscopy will be discussed in section 6. The 1006
980 polarizability fluctuation resulting from anion vibrations. The most intense Raman band at 741 cm−1, corresponding to a 1007
981 remaining bands assigned to the cation are easily identified by normal mode in which the [NTf2]− anion breathes as a whole, 1008
982 comparison with the spectrum of an atomic anion counterpart exhibits an asymmetric band shape that has been also assigned 1009
983 (e.g., [C2C1im]Cl). It has been a successful application of to transoid and cisoid conformers, giving two components with 1010
984 Raman spectroscopy unveiling the coexistence of transoid and a small difference of ∼3 cm−1.116 On the other hand, it will be 1011

Table 2. Vibrational Frequencies of Some Fluorinated Anions Commonly Used in Ionic Liquidsa
[NTf2]− [N(SO2F)2]− [CF3SO3]−
IR Raman IR Raman IR Raman
120 t(CF3) 291 τ(SO2F) 210 ρ(CF3)
165 326 δop(SO2F) 312
278 ρ(CF3) 359 δop(SO2F) 348 ρ(SO3)
297 454 456 δsci(SOF) + δ(OSNSO) 518 518 δas(SO3)
313 ρ(SO2) 482 482 δsci(SOF) + δop(SO2F) 573 573 δas(CF3)
326 ρ(SO2) 524 524 δsci(SO2) + ν(SF) 638 640 δs(SO3)
340 τ(SO2) 572 568 δip(O2SNSO2) + ν(SF) + δip(SO2F) 755 755 δs(CF3)
351 τ(SO2) 725 727 δsci(SO2) + ν(SF) + δsci(SNS) 1031 1034 νs(SO3)
397 ω(SO2) 831 831 ν(SF) + ρ(SO2) + νs(SNS) 1162 1167 νas(CF3)
408 404 ω(SO2) 1218 1217 νs(SO2) 1225 1226 νs(SO3)
514 1365 1361 νas(SO2) 1261 1259 νas(SO3)
551 δs(SO2) 1383 1380 νas(SO2) 1281 1281 νas(SO3)
571 571 δa(CF3) 1388 νas(SO2)
602 590 δa,ip(SO2)
619 623 δa,op(SO2)
651 650 δ(SNS)
741 741 νs(SNS)
762 764 νs(SNS)
790 796 ν(CS)
1058
1139 1136 νs,ip(SO2)
1195
1231
1243 νs(CF3)
1333 1337 νa,op(SO2)
1352 1353

a
Values (cm−1) correspond to frequencies observed in IR and Raman spectra at room temperature for ionic liquids with the [C2C1im]+ cation for
[NTf2]− and [N(SO2F)2]− anions and the [C4C1im]+ cation for [CF3SO3]−. ν, stretch; δ, bend; δsci, scissoring; ω, wagging; τ, twisting; ρ, rocking; t,
torsion; s, symmetric; as, antisymmetric; ip, in plane; and op, out-of-plane.

L DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 12. Raman spectrum of [C2C1im][NTf2] in the spectral ranges covering the bands characteristics of [NTf2]− in transoid (marked #) and
cisoid (marked *) conformations. Optimized structures of transoid and cisoid conformers of [NTf2]− are shown.

1012 discussed that the 741 cm−1 band is very sensitive to bands of [N(SO2F)2]− in the range of 250−400 cm−1 are 1029
1013 coordination of [NTf2]− to small cations of high polarizing asymmetric because of the presence of both the [N(SO2F)2]− 1030
1014 power (see section 7). conformers. The bands seen in Figure 14 at 291, 326, and 359 1031 f14
1015 The closely related bis(fluorosulfonyl)imide anion, [N-
cm−1 in the Raman spectrum of [C2C1im][N(SO2F)2]
1016 (SO2F)2]−, also exhibits equilibrium between C1 and C2 1032

1017 conformers in ionic liquids. Fujii et al.129 discussed the correspond to SO2F out-of-plane bending for cisoid conformer 1033

1018 signature of conformational equilibrium of [N(SO2F)2]− in


1019 vibrational spectra of ionic liquids containing the cations 1-
1020 ethyl-3-methylimidazolium129 and N-methyl-N-propyl-pyrroli-
f13 1021 dium.130 Figure 13 shows vibrational spectra of [C2C1im][N-

Figure 13. IR (red, transmittance scale at right) and Raman spectra


(black) of [C2C1im][N(SO2F)2] at room temperature. Some bands
assigned to [N(SO2F)2]− normal modes are indicated by arrows.

1022 (SO2F)2] with the anion bands indicated by arrows. Assign-


1023 ment of [N(SO2F)2]− normal modes has not been provided in Figure 14. Raman spectrum of [C2C1im][N(SO2F)2] at room
1024 refs 129 and 130, but it is available in the work of Matsumoto et temperature. Vibrational frequencies and relative intensities of
1025 al.131 concerning polymorphism in crystals of Na+, K+, and Cs+ Raman bands calculated by the DFT/B3LYP level of theory are
1026 salts of [N(SO2F)2]−. indicated for [N(SO2F)2]− at C1 (cisoid, blue lines) and C2 (transoid,
1027 On the basis of quantum chemistry calculations at the DFT/ red lines) conformation. Optimized structures of transoid and cisoid
1028 B3LYP level of theory, Fujii et al.129,130 proposed that Raman conformers of [N(SO2F)2]− are shown.

M DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1034 or SO2F twisting for transoid conformer, SO2 rocking for cisoid
1035 conformer or SO2F out-of-plane bending for transoid con-
1036 former, and SO2F in-plane bending for cisoid conformer or
1037 SO2F out-of-plane bending for transoid conformer, respec-
1038 tively.131 The calculated frequencies for each [N(SO2F)2]−
1039 conformer are relatively close, so that the more strong
1040 experimental evidence in favor of conformational equilibrium
1041 in the ionic liquid is the temperature dependence of the spectral
1042 pattern.129,130,132 The relative intensities of components
1043 belonging to transoid conformer decrease with increasing
1044 temperature providing support to the calculations that transoid
1045 is the lower energy conformer between them.129,130
1046 Giffin et al.133 prepared pyrrolidinium ionic liquids with an
1047 anion whose molecular structure lies in between [NTf2]− and
1048 [N(SO2F)2]−, namely, (fluorosulfonyl) Figure 15. IR (red, transmittance scale at right) and Raman spectra of
1049 (trifluoromethanesulfonyl)imide, [N(SO2F) (CF3SO2)]−. The [C4C1im][CF3SO3] (black) at room temperature. Some bands
1050 most intense Raman band of [N(SO2F) (CF3SO2)]− at 730 assigned to the [CF3SO3]− normal modes are indicated by arrows.
1051 cm−1 is the counterpart of the characteristic band due to the
1052 expansion and contraction modes of [NTf2]−. Giffin et al.133 phase spectrum shown in Figure 15. This allowed a fine 1097
1053 emphasized that this band of [N(SO2F) (CF3SO2)]− is broader comparison between experimental and calculated frequencies at 1098
1054 than analogous band of [NTf2]− proper to distribution of three the DFT/B3LYP level of theory for two different arrangements 1099
1055 rotamers of [N(SO2F) (CF3SO2)]−. The authors of ref 133 of the [C2C1im]+−[CF3SO3]− pair.145 The experimental 1100
1056 provided a table of experimental versus calculated vibrational matrix-isolated IR spectrum of [C2C1im][CF3SO3] was more 1101
1057 frequencies and normal mode assignment for different consistent with a local arrangement in which the anion SO3 1102
1058 conformers of [N(SO2F) (CF3SO2)]−, but overlap of bands group points toward the cation C2 atom with five anion−cation 1103
1059 in the spectral range 280−400 cm−1 implies that distinguishing hydrogen bonds.145 1104
1060 the presence of [N(SO2F) (CF3SO2)]− conformers by Raman The vibrations of tris(pentafluoroethyl)-trifluorophosphate, 1105
1061 spectroscopy is a more challenging task than [NTf 2 ] − [FAP]−, in the ionic liquid [C2C1im][FAP] have been 1106
1062 conformers. discussed in an IR and Raman study by Mao and Damodaran146 1107
1063 The trifluoromethanesulfonate (triflate) anion, [CF3SO3]−, and in an IR study by Voroshylova et al.147 These two works 1108
1064 was the subject of many vibrational spectroscopy studies provide a full list of observed frequencies of [C2C1im][FAP] 1109
1065 because it has been extensively used in polymer electro- and assignments based on the potential energy distribution of 1110
1066 lytes.134−137 One recurrent issue in these works is to unveil, normal modes calculated for the ionic pair at the DFT/B3LYP 1111
1067 from frequency shift and band split, ionic pairing between level of theory. Concerning those bands belonging to [FAP]− 1112
1068 [CF3SO3]− and alkali metal cations. Here again vibrational normal modes, there are some disagreement on assignments 1113
1069 frequencies of some normal modes are very dependent on the proposed in these works. For example, the intense IR band at 1114
1070 polarizing power of the cation. For instance, δs(CF3) and 1209 cm−1 was assigned to ν(CC) by Mao and Damodaran,146 1115
1071 νs(SO3) are observed at 755 and 1034 cm−1, respectively, in the but to νas(CF2) by Voroshylova et al.147 Two isomers of 1116
1072 Raman spectrum of [C4C1im][CF3SO3]. These frequencies are [FAP]− are possible, meridional and facial, each one with 1117
1073 close to values of “free” [CF3SO3]− in 2-methyltetrahydrofuran several conformers. The calculations of Voroshylova et al.147 1118
1074 solution,138 whereas these modes are observed at 767 and 1053 indicated IR bands at ∼800 and ∼700 cm−1 as characteristic 1119
1075 cm−1 for [CF3SO3]− in aggregates with the strongly polarizing features of meridional and facial isomers, respectively, and 1120
1076 Li+ cation. In a quantum chemistry investigation of [CF3SO3]− mixture of anion conformers in [C2C1im][FAP]. 1121
1077 coordinated to Li+, Gejji et al.139 calculated the potential energy 4.1.3. Alkylsulfates and Hydrogen Sulfate. Few works 1122
1078 distribution of normal modes among the internal coordinates of have discussed vibrational spectra of pure ionic liquids 1123
1079 [CF3SO3]−. Huang et al.138 compared the nature of the normal containing alkylsulfate anions, [R−O−SO3]−, in spite of the 1124
1080 modes of free and lithium-coordinated [CF3SO3]−. Potential relevance of this class of anion forming ionic liquids. IR and 1125
1081 energy distribution indicates less delocalized normal modes for Raman spectra of 1-ethyl-3-methylimidazolium ethylsulfate, 1126
1082 [CF3SO3]− than [NTf2]− or [N(SO2F)2]−, as one would expect [C2C1im][C2SO4], have been first discussed by Kiefer et al.,115 1127
1083 from the electronic structure of [CF3SO3]−. Since the and in a subsequent paper by Dhumal et al.148 The IR 1128
1084 [CF3SO3]− vibrational frequencies are very sensitive to the frequencies listed in these two works agree with each other; 1129
1085 local environment experienced by the anion,140−143 we show in however, there is systematic mismatching between them for the 1130
f15 1086 Figure 15 IR and Raman spectra of a common [CF3SO3]− Raman frequencies of [C2C1im][C2SO4]. For instance, the 1131
1087 based ionic liquid. Vibrational frequencies belonging to most intense Raman band of [C2SO4]− is reported at 1072 1132
1088 [CF3SO3]− normal modes actually observed in spectra of cm−1 in ref 115, whereas it is reported at 1060 cm−1 in ref 148. 1133
1089 [C4C1im][CF3SO3] are listed in Table 2. Assignments are The latter is most probably the correct value, being close to 1134
1090 available from the previous works on polymer electrolytes and 1062 cm−1 as found in aqueous solution of Na[C2SO4].149 In 1135
1091 reconsidered by Schwenzer et al.144 within the context of ionic light of disagreement of reported frequencies for [C2SO4]−, 1136
1092 liquids. A distinctive feature of a study published by Akai et Figure 16 shows IR and Raman spectra obtained in this work 1137 f16
1093 al.145 was recording IR spectrum of the ionic pair after for [C2C1im][C2SO4] at room temperature. Raman frequencies 1138
1094 evaporating [C2C1im][CF3SO3] and trapping in a cryogenic of [C2SO4]− we obtained agree with values given in ref 148. 1139
1095 neon matrix. Of course the bands in the IR spectrum of matrix- Dhumal et al.148 calculated vibrational frequencies at the 1140
1096 isolated ionic pair become much sharper than the normal liquid DFT/B3LYP level of theory for isolated ions and the ionic pair. 1141

N DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

context of ionic liquids, Kiefer and Pye155 used IR and Raman 1172
spectroscopies and quantum chemistry calculations to charac- 1173
terize the conformations of ions in [C6C1im][HSO4], 1174
concluding that [HSO4]− occurs in the trans conformation. 1175
Kiefer and Pye155 found a group of bands which they assigned 1176
to sulfuric acid in the IR (903, 1158, and 2963 cm−1) and 1177
Raman (959, 1157, and 1364 cm−1) spectra of [C6C1im]- 1178
[HSO4]. They did not conclude whether the sulfuric acid 1179
originated from proton transfer, like the case of 1-alkyl-3- 1180
methylimidazolium acetate, or from a residual of synthesis. On 1181
the other hand, sulfuric acid bands have not been found either 1182
in the IR spectrum of [C4C1im][HSO4] reported by Schwenzer 1183
et al.144 or in the Raman spectra of [C2C1im][HSO4] and 1184
[C4C1im][HSO4] reported by Ribeiro.156 The [HSO4]− Raman 1185
Figure 16. IR (red, transmittance scale at right) and Raman (black) bands observed in the ionic liquids at 416 [δas(SO3)], 1186
spectra of [C2C1im][C2SO4] at room temperature. Some bands 581[δs(SO3)], 845 [ν(S−OH)], and 1046 cm−1 [νs(S 1187
assigned to [C2SO4]− normal modes are indicated by arrows.
O)]156 nicely match the corresponding bands in the (high 1188
temperature) molten salt KHSO4.151 Another νs(SO) band 1189

1142 Assignment of vibrational frequencies on the basis of at 1010 cm−1 has also been found, whose intensity increases as 1190

1143 comparison between calculated and experimental data is temperature decreases, and assigned to [HSO4]−, engaged in 1191

1144 particularly difficult for this system because strong anion− chains of hydrogen-bonded anions.156 Signature of anion− 1192

1145 cation interaction imply significant frequency shifts. For anion hydrogen bonding has been found in vibrational spectra 1193

1146 instance, the most intense and sharp Raman band at ∼1060 of alkali bisulfate crystals.157−160 It had been proposed decades 1194

1147 cm−1, which is assigned to S−O symmetric stretching of ago that simple [HSO4]− molten salts are very viscous because 1195

1148 [C2SO4]− in ref 149, is assigned to C−O stretching in ref 148. structures of hydrogen-bonded anions existing in the crystalline 1196

1149 In this work, we calculated vibrational frequencies of an ionic phase remain in the liquid phase just above the melting 1197

1150 pair made of [C2SO4]− and tetramethylammonium cation at temperature.161 Therefore, it was proposed that the very high 1198
1151 the DFT/B3LYP level of theory with a basis set 6-311+ viscosity of [HSO4]− ionic liquids in comparison with 1199
1152 +G(d,p). Our results are inline with the assignment of Dhumal alkylsulfates for a given 1-alkyl-3-methylimidazolium cation is 1200
1153 et al.,148 that is, the Raman bands observed at 1061.5 and 959 due to anion−anion, rather than anion−cation, hydrogen 1201
1154 cm−1 in Figure 16 are assigned to C−O and S−O stretching bonding.156 1202
1155 modes, respectively. We also found that the vibrational 4.1.4. Carboxilates. Vibrational spectroscopy studies of 1203
1156 frequencies of the main bands of methylsulfate, [C1SO4]−, are carboxylate-based ionic liquids are heavily linked to applications 1204
1157 essentially the same in the ionic liquid [C4C1im][C1SO4] in gas absorption to be discussed in section 7. Here we focus on 1205
1158 (spectra not shown) since differences between [C2C1im]- vibrational frequencies of acetate in pure ionic liquids based on 1206
1159 [C2SO4] and [C4C1im][C1SO4] spectra arise from different 1-alkyl-3-methylimidazolium cations. IR and Raman spectra of 1207
1160 alkyl chain lengths. Vibrational frequencies of stretching modes [C2C1im][CH3COO] and [C4C1im][CH3COO] have been 1208
1161 of C−H bonds of alkylsulfate anions overlap the same high- discussed by Thomas et al.33 and Cabaço et al.,162 respectively. 1209
1162 frequency range of corresponding vibrations of 1-alkyl-3- Raman frequencies of [CH3COO]− reported in these works are 1210
1163 methylimidazolium cations discussed in the next section. listed in Table 3 showing some inconsistencies between them. 1211 t3
1164 At first sight, one would expect more simple spectra for ionic Ito and Bernstein163 discussed IR and Raman spectra of 1212
1165 liquids based on hydrogen sulfate (or bisulfate), [HSO4]−, since aqueous solutions of formate, oxalate, and acetate; the 1213
1166 it could be considered the first of the [R−O−SO3]− series. vibrational frequencies of the latter are given in Table 3 for 1214
1167 However, hydrogen bonding implies nontrivial features in comparison purposes. Assignment of νs(COO) and νas(COO) 1215
1168 vibrational spectra of molten bisulfates. Bisulfate belongs to the modes in ionic liquids [CnC1im][CH3COO] is cumbersome 1216
1169 class of relative simple polyatomic anions with a large number because of overlaps with bands of imidazolium ring modes in 1217
1170 of spectroscopic studies concerning alkali cation molten the same spectral range. A large number of works address 1218
1171 salts150−152 or aqueous solution.153,154 Within the more recent vibrational spectra of acetate in alkali salts and metal 1219

Table 3. Fundamental Frequencies (cm−1) and Assignments Reported in the Literature for Raman Spectra of the Acetate Anion
in Ionic Liquids. Raman Frequencies of Na[CH3COO] Aqueous Solution are Given for Comparison Purposesa
Thomas et al.33 [C2C1im][CH3COO] Cabaço et al.162 [C4C1im][CH3COO] Ito and Bernstein163 Na[CH3COO](aq)
454 ρ(COO) 471 ρip(COO), B1
635 δ(OCO)+ν(CC) 637 δ(OCO) 650 δ(OCO), A1
899 δ(OCO)+ν(CC) 902 ν(CC) 926 ν(CC), A1
1334 νs(COO) 1326 δ(CH3) 1344 δ(CH3), A1
1382 νs(COO) 1413 νs(COO), A1
1567 νas(COO) 1582 νas(COO) 1556 νas(COO), B1
2917 νs(CH3) 2936 νs(CH3), A1

a
For comparison purposes, corresponding Raman frequencies of Na[CH3COO] aqueous solution are given with their assignment and symmetry. ν,
stretch; δ, bend; ρ, rocking; s, symmetric; as, antisymmetric; ip, in plane.

O DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1220 complexes.163−166 However, direct comparison with the Table 4. Assignment According Different Authors for Some
1221 frequencies observed in ionic liquids is not straightforward Characteristic Imidazolium Ring Vibrations Numbered as
1222 since [CH3COO]− frequencies are very sensitive to coordina- Figure 18a
1223 tion. Furthermore, theoretical calculations for [CnC1im]+ and
#
1224 [CH3COO]− in the gas phase might lead to proton transfer and IR Raman assignment
1225 formation of the N-heterocyclic carbene. This is regarded as an 1 1021 1021 ref 58: ν(CNring) + δ(CNCring) + δ(CH)
1226 important step along the process of CO2 absorption, but the ref 53: νring,ip,s
1227 ionic environment in the liquid phase stabilizes the ions.167 The ref 54: breathing + ν (N−CH2) + ν (N−CH3)
1228 theoretical tool employed by Thomas et al.33 was ab initio 2 1337 1337 ref 58: νip(C(2)H3N) + νip(CNring) + w(CH2)
1229 molecular dynamics simulation of 36 pairs of [C2C1im]+− ref 53: ν[CN CH3(N)] + ν[CN CH2(N)] + νring,ip,s
1230 [CH3COO]− ions, whereas Cabaço et al.162 performed DFT/ ref 54: breathing + ν(N−CH2) + ν(N−CH3)
1231 B3LYP calculations of a cluster made of two pairs of 3 1378 1385 ref 58: νas(C(2,4)N) + δ(CH) + δs(CH3)
1232 [C4C1im]+−[CH3COO]− ions. In order to identify the ref 53: δs[HCH CH3(N)] + ν[CN CH2(N)] + νring,ip,s
1233 [CH3COO]− bands, it is useful to compare vibrational spectra ref 54: νas(C(2)N(1)C(5)) + δs(CH3)
1234 of acetate-based and other ionic liquids with the same cation. 4 1416 ref 58: νas(C(2,5)N) + δ(CH2) + δ(C(2,4,5)H)+
ν(C(2)N)
1235 Cabaço et al.152 compared IR and Raman spectra of ref 53: νring,ip,as + ν[CN CH2(N)]+ δs[HCH butyl]
1236 [C4C1im][CH3COO], [C4C1im][BF4], and [C4C1im][PF6]. ref 54: νring + δ(CH2)
f17 1237 In Figure 17, we provide a comparison of IR and Raman spectra 5 1568 1564 ref 58.:ν(C(2)N) + δ(CH) + ν(CCring)
ref 53: νring,ip,as + ν[CN CH2(N)] + ν[CN CH3(N)]
ref54: νas(N(1)C(2)N(3)) + r(C(2)−H)
a
v, stretching; δ, in plane bending; w, wagging; r, rocking; s,
symmetric; as, antisymmetric; ip, in-phase.

three different carboxylate anions (formate, propanoate, and 1262


butanoate) and a common cation (choline). The values of Δ in 1263
formate and propanoate ionic liquids are higher, whereas it is 1264
slightly smaller in butanoate than corresponding values in the 1265
sodium salts. The work of Tanzi et al.170 provided vibrational 1266
frequencies of these carboxylate anions and assignments on the 1267
basis of DFT calculations for isolated ions or ionic pairs and by 1268
ab initio MD simulations of clusters of ions. The sequence of 1269
calculations for isolated ions, ionic pairs, and clusters highlight 1270
Figure 17. IR (red) and Raman (black) spectra of [C4C1im]- the spectral signature of strong hydrogen bonds between 1271
[CH3COO] at room temperature. For comparison purposes, IR carboxylate anions and the choline cation. 1272
(green) and Raman (blue) spectra of [C4C1im]Br in the liquid phase
The Raman band shape of the ν(CC) mode is also a probe of 1273
are shown. Arrows indicate bands assigned to [CH3COO]− normal
modes. the coordinating structure of acetate in solution.171 Cabaço et 1274
al.162 considered the asymmetry of the ν(CC) Raman band of 1275
[C4C1im][CH3COO] at 902 cm−1, which exhibits a high 1276
1238 of [C4C1im]Br and [C4C1im][CH3COO], where bands that frequency tail due to a low-intensity component at 909 cm−1, as 1277
1239 Cabaço et al.162 assigned to [CH3COO]− are indicated by an indication of a fraction of anions in the bidentate 1278
1240 arrows. The Raman bands at 1334 and 1567 cm−1 assigned by coordination. A jointed neutron diffraction and MD simulation 1279
1241 Thomas et al.33 to the acetate anion actually belongs to study of [C2C1im][CH3COO] by Bowron et al.172 indeed 1280
1242 imidazolium ring modes as discussed in the next section (see suggested dominance of unidentate over bidentate coordination 1281
t4 1243 Table 4). Therefore, we follow Cabaço et al.162 assignment of of [CH3COO]− to the hydrogen atoms bounded to the 1282
1244 [CH3COO]− bands. However, it is worth mentioning that we imidazolium ring of [C2C1im]+. It will be discussed in the next 1283
1245 found the most intense [CH3COO]− Raman band at 911 cm−1 section that an anion hydrogen-bonded to 1-alkyl-3-methyl- 1284
1246 in the Raman spectrum of [C4C1im][CH3COO]. imidazolium cations has significant effect on vibrational 1285
1247 It is worth noting the difference Δ = νas(COO) − νs(COO) frequencies of stretching of C−H ring bonds of the cation. 1286
1248 = 200 cm−1 in the ionic liquid [C4C1im][CH3COO]. In the The Raman band at 2917 cm−1 assigned to the νs(CH3) mode 1287
1249 context of coordination chemistry, the magnitude of Δ has of [CH3COO]− (see Table 3) overlaps the spectral range of 1288
1250 been used as a signature of the coordinating structure of acetate C−H stretching modes of the butyl chain of [C4C1im]+ (see 1289
1251 to cations, 164−166,168 taking the value of Δ in Na- Figure 19 below). 1290
1252 [CH3COO]163,169 as a reference of purely ionic interaction. Vibrations of the trifluoroacetate anion have been discussed 1291
1253 In the case of unidentate coordination (i.e., only one oxygen by Cabaço et al.173 These authors provided a list of vibrational 1292
1254 atom of [CH3COO]− coordinating the metal cation), the frequencies and assignments for [CF3COO]− in [C4C1im]- 1293
1255 equivalence of oxygens is removed, implying large Δ, typically [CF3COO] together with corresponding values for the simple 1294
1256 Δ > 200 cm−1. In contrast, bidentate coordination (i.e., salt Na[CF3COO] in aqueous solution.174 Some effects of 1295
1257 chelating structure) implies small Δ, typically Δ < 150 cm−1.165 replacing H by F atoms in going from [CH3COO]− to 1296
1258 The value of Δ in [C4C1im][CH3COO] suggests dominance of [CF3COO]− are worth noting on the most important 1297
1259 unidentate interaction of acetate to imidazolium cation. The vibrations of the anion. In the case of [C4C1im][CF3COO], 1298
1260 magnitude of Δ has also been commented by Tanzi et al.170 in νas(COO) and νs(COO) are observed at 1691 and 1406 cm−1, 1299
1261 an IR and Raman investigation carried out for ionic liquids with respectively.173 The difference between them (Δ = 285 cm−1), 1300

P DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1301 being significantly larger than in Na[CF3COO] (Δ = 246


1302 cm−1),165 is also an indication of unidentate interaction
1303 between [CF3COO]− and the imidazolium cation. In contrast
1304 to the asymmetric Raman band shape of ν(CC) of
1305 [CH3COO]−, the ν(CC) mode of [CF3COO]− gives a single
1306 inhomogeneously broadened Gaussian profile at 824 cm−1 in
1307 the Raman spectrum of [C4C1im][CF3COO].173 Another
1308 worth mentioning difference between [C4C1im][CH3COO]
1309 and [C4C1im][CF3COO] is the Raman band at 1203 cm−1 in
1310 the latter due to the CF3 stretching mode.173
4.2. Vibrational Frequencies of Cations
1311 4.2.1. Imidazolium. Vibrations of 1-alkyl-3-methylimidazo-
1312 lium cations have been discussed in more detail than other
1313 ionic liquid forming cations. Assignment of vibrational Figure 18. IR and Raman spectra in the 1000−1600 cm−1 range of
1314 frequencies has been done with the help of quantum chemistry [C4C1im]Br (green and blue) and [C6C1im]Br (red and black) in
calculations for the isolated cation or ion pairs. Berg3 used MP2 liquid phase at room temperature. Table 4 gives the assignment of
1315
Raman bands marked 1−5.
1316 level of theory to calculate vibrational frequencies of [C4C1im]+
1317 with two different conformations of the butyl chain. A well- of the imidazolium ring. Figure 19 shows for [C4C1im][PF6] 1362 f19
1318 suited level of theory to calculate vibrational frequencies is and [C4C1im][CF3SO3] such separation of two groups of 1363
1319 DFT/B3LYP with a typical basis set, say 6-31+G(d,p). DFT/
1320 B3LYP has been used by Dhumal et al.119 for isolated ions and
1321 ionic pairs of [C2C1im]+ and [Tf2N]−, by Talaty et al.52 for
1322 ionic pairs of [CnC1im][PF6] and by Heimer et al.53 for
1323 [CnC1im][BF4], n = 2, 3, and 4, and by Katsyuba et al.54 for
1324 [C2C1im][BF4]. Once harmonic frequencies follow from these
1325 standard calculations, scaling factors are needed for agreement
1326 between calculated and experimental data. Grondin et al.58
1327 calculated harmonic and also anharmonic frequencies of
1328 [C1C1im]+ and [C2C1im]+ using the second-order perturbative
1329 method proposed by Barone56,57 in order to avoid such scaling
1330 factors of harmonic frequencies.
1331 Long tables of vibrational frequencies calculated for 1-alkyl-3-
1332 methylimidazolium cations are available in these papers. The
1333 composition of normal modes is naturally complicated because
1334 of the electronic structure of imidazolium cations. Furthermore,
1335 visual inspection of the eigenvectors on computer screen Figure 19. IR (red, transmittance scale at right) and Raman (black)
1336 implies that description of the same normal coordinate may spectra in the CH stretching range of [C4C1im][PF6] (full lines) and
1337 vary among different authors. Grondin et al.58 provided the [C4C1im][CF3SO3] (dashed lines) at room temperature.
1338 distribution of potential energy allowing for more rigorous
1339 assignment of normal modes in terms of internal coordinates. bands of CH stretching modes. The comparison made in 1364
1340 In the following, we will distinguish three spectral ranges in Figure 19 illustrates the well-known finding that this spectral 1365
1341 vibrational spectra of 1-alkyl-3-methylimidazolium cations. The range is a signature of ionic interactions:54,119,175,176 the 1366
1342 high-frequency range, 2800−3200 cm−1, exhibits a complex imidazolium ring CH stretching modes shift to lower 1367
1343 pattern of overlapped bands proper to several C−H stretching wavenumbers for ionic liquids containing stronger interacting 1368
1344 modes. The 800−1600 cm−1 range includes characteristic anions. Furthermore, the band at ∼3115 cm−1 in the IR 1369
1345 bands of imidazolium ring vibrations. The 400−800 cm−1 spectrum of [C4C1im][CF3SO3] is much broader than 1370
1346 range, which also contains ring vibrations, is interesting because [C4C1im][PF6], suggesting stronger hydrogen bonding in the 1371
1347 it provides insight on conformations of alkyl chains. (The low- former. Out-of-plane ring CH vibrations within 700−900 cm−1 1372
1348 frequency range probing intermolecular dynamics is the issue of are also sensitive to strength of anion−cation interactions, 1373
1349 section 5.) shifting to lower wavenumber with increasing anion basicity.58 1374
f18 1350 Figure 18 shows IR and Raman spectra of [C4C1im]Br and Using these γ(CH) vibrations as probes of ionic interactions is 1375
1351 [C6C1im]Br in the spectral range covering imidazolium ring better accomplished with IR than Raman spectroscopy because 1376
1352 vibrations. (There are also out-of-plane ring vibrations within of higher IR activity. Figure 19 also illustrates stronger anion 1377
1353 600−650 cm−1.) It is clear from this figure that vibrational dependence for the ring CH than alkyl CH stretching modes, as 1378
1354 spectra in this range are essentially the same irrespective of the expected from the preferred location of anions around the 1379
1355 length of the alkyl chain. Table 4 illustrates different ways that imidazolium polar head. Furthermore, the doubled-peaked 1380
1356 authors describe the nature of vibrations of the bands marked feature observed within the 3100−3200 cm−1 range is 1381
1357 1−5 in the Raman spectra of Figure 18. commonly assigned to stretching of C(2)−H (the low frequency 1382
1358 The stretching of C−H bonds of 1-alkyl-3-methylimidazo- component) and C(4),(5)−H (the high frequency component). 1383
1359 lium cations can be separated into two groups of bands: 2800− This interpretation is supported by quantum chemistry 1384
1360 3000 cm−1, belonging to CH modes of alkyl chains attached to calculations of ionic pair54,119,53,175,177,48 putting the anion 1385
f19 1361 the imidazolium ring, and 3000−3200 cm−1, belonging to CH either in front of the C(2) or the C(4),(5) atoms of the 1386

Q DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1387 imidazolium ring. These findings are considered as evidence in indicating the [CnC1im]+ conformation are not so evident in 1450
1388 favor of anion−cation hydrogen bonding, the C(2)−H···A− the IR spectrum.99 Figure 20 illustrates the frequency range 1451 f20
1389 arrangement being preferred over C(4),(5)−H···A−.
1390 The interpretation of this spectral range in terms of different
1391 strengths between C(2)−H···A− and C(4),(5)−H···A− hydrogen
1392 bonds has been disputed by Lassègues et al.59 These authors
1393 claim that the double-peaked feature above 3100 cm−1 is
1394 instead a Fermi resonance,42 resulting from mixing of
1395 vibrational states of fundamental CH ring vibrations with
1396 overtone and combination bands of ring modes whose
1397 fundamental transitions are observed in the 1550−1585 cm−1
1398 range (see Figure 18). Lassègues et al.59 grounded their
1399 interpretation by isotopic substitution of each hydrogen atom
1400 bonded to the ring in order to clear out the spectral range of
1401 3100−3200 cm−1, leaving the corresponding C−D stretching
1402 within the 2300−2400 cm−1 range. In accordance with their
1403 proposition, the band at ∼3170 cm−1 (see Figure 19) includes
1404 all C(2),(4),(5)−H vibrations, and the band at ∼3115 cm−1 is
1405 overtone and combination bands of ring modes enhanced by
1406 Fermi resonance. The alternative assignment put forward by
1407 Lassègues et al.59 roused a lively debate in the literature178,179
1408 because it implies there is no need of anion−cation hydrogen
1409 bonding to explain the vibrational spectra nor C(2)−H being
1410 stronger interacting site than C(4),(5)−H. Wulf et al.178 Figure 20. Raman spectra of [C4C1im]Br (blue) and [C6C1im]Br
1411 strengthened the usual point of view with further quantum (black) in the liquid phase at room temperature. For comparison
1412 chemistry calculations showing that C(2)−H modes are always purpose, Raman spectrum of [C2C1im]Cl (red) at T = 360 K is shown.
1413 obtained at lower frequency than C(4),(5)−H modes in many Bands marked * (620 and 735 cm−1 in [C4C1im]Br) characterize the
1414 different ion pairs. However, the cyanate-anions chosen by AA conformer; bands marked # (600 and 697 cm−1 in [C4C1im]Br)
1415 Wulf et al.178 are strongly interacting species, whereas characterize the GA conformer. Optimized structures of AA and GA
1416 Lassègues et al.179 provided a reminder that their interpretation conformers are shown.
1417 concerned imidazolium ionic liquids containing less coordinat-
1418 ing anions (e.g., [NTf2]−, [BF4]−, and [PF6]−). Nevertheless, exhibiting bands that characterize anti−anti (AA, 620 and 735 1452
1419 there is consensus that strongly coordinated anions shift CH cm−1) and gauche−anti (GA, 600 and 697 cm−1) conforma- 1453
1420 stretching modes to lower wavenumber,178,179,58 even though tions of [C4C1im]+. Quantum chemistry calculations show that 1454
1421 definitive assignment of this spectral range is an open issue in these bands belong to ring deformations coupled to CH2 1455
1422 vibrational spectroscopy of imidazolium ionic liquids. Irre- rocking motions3,185 so that the actual vibrational frequency 1456
1423 spective of the assignment of vibrational spectra, yet a more is sensitive to the butyl chain conformation. As emphasized by 1457
1424 delicate issue is the nature of hydrogen bond between Berg,3,47 Figure 20 also shows a mixture of gauche and anti 1458
1425 imidazolium cations and anions.180 conformers in the longer alkyl chain [C6C1im]+ cation. Kiefer 1459
1426 Vibrational spectroscopy has been a powerful tool to reveal and Pye155 considered other bands as signatures of three 1460
1427 molecular conformations in ionic liquids. Studies in this conformers of [C6C1im]+ in [C6C1im][HSO4] (e.g., bending 1461
1428 direction followed after the discovery of crystal polymorphism modes in 340−510 cm−1 and C−C stretching modes in 900− 1462
1429 in [C4C1im]Cl by Hayashi et al.181 and Holbrey et al.182 These 1100 cm−1 ranges). 1463
1430 authors concluded from X-ray diffraction that [C4C1im]Cl Nine conformers of [C4C1im]+ have been the relative 1464
1431 forms two crystalline phases differing in the conformation of energies calculated by quantum chemistry methods.187,188 In 1465
1432 the butyl chain. Hayashi et al.181 found that these polymorphs a re-examination of the Raman spectrum of [C4C1im][BF4], 1466
1433 exhibit different patterns in the 500−800 cm−1 range of the Holomb et al.99 were able to discriminate four conformers 1467
1434 Raman spectrum, where a group of bands (625, 730, an 790 (GG, GA, TA, and AA) coexisting in the liquid phase by 1468
1435 cm−1) and (500, 600, an 700 cm−1) is characteristic of each considering additional bands covering wider spectral range 1469
1436 crystal. All of these bands appear in the Raman spectrum of (200−1200 cm−1). Further studies by Umebayashi et al.,188 1470
1437 liquid phase of [C4C1im]Cl, so that one concludes there is a Katayanagi et al.,189 and Singh et al.190 focused on the anion 1471
1438 mixture of conformers in the ionic liquid.181,183,184 Hamaguchi dependence of relative intensities of the 600 and 620 cm−1 1472
1439 and Ozawa185 reviewed their early Raman spectroscopy studies bands. The ratio of intensities I600/I620 increases in the 1473
1440 on conformational changes of [CnC1im]+ in ionic liquids with sequence [C4C1im]I, [C4C1im]Br, and [C4C1im]Cl, so that 1474
1441 Cl−, Br−, I−, [BF4]−, and [PF6]−. In a previous review, Berg3 one concludes that gauche conformation is stabilized for 1475
1442 offered a detailed analysis on how alkyl chain conformation of stronger interacting halide anions. 1476
1443 [CnC1im]+ adds fingerprints in Raman spectra of ionic liquids. Umebayashi et al.49 and Lassègues et al.125 showed that 1477
1444 We also recommend a very detailed work recently published by [C2C1im]+ conformers with the planar or nonplanar ethyl chain 1478
1445 Endo et al.186 concerning DFT calculations of conformational can be distinguished by characteristic bands in the spectral 1479
1446 flexibility and vibrational frequencies of typical ionic liquid range of 200−500 cm−1 of the Raman spectrum. Figure 21 1480 f21
1447 forming cations (imidazolium, pyridinium, pyrrolidinium, and shows this spectral range of the Raman spectrum of [C2C1im] 1481
1448 piperidinium). It should be noted that this issue has been Cl with the bands proposed by these authors as signatures of 1482
1449 addressed mainly by Raman spectroscopy, since features the [C2C1im]+ conformation. Endo and Nishikawa191 found 1483

R DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

the [NTf2]− band marked with asterisk in Figure 22 as the 1508


intensity pattern, it is remarkable how strong the [C4C1C1im]+ 1509
Raman band is at 1515 cm−1. Figure 23 shows atomic 1510 f23

Figure 23. Atomic displacement of the two normal modes of


Figure 21. Raman spectrum of [C2C1im]Cl at T = 360 K. Bands [C4C1C1im]+ calculated at the DFT/B3LYP level of theory for the
marked * belong to the planar conformer, whereas bands marked # isolated cation. The harmonic vibrational frequencies calculated at
belong to the nonplanar conformer. 1533 and 1560 cm−1 correspond to experimental bands at 1515 and
1540 cm−1, respectively, of [C4C1C1im][NTf2] (bands marked by
arrows in Figure 22).
1484 two isomers (symmetric and asymmetric) of 1-isopropyl-3-
1485 methylimidazolium cation in the ionic liquids [i-C3C1im]Br and
1486 [i-C3C1im]I. Raman bands considered as signatures of the displacements calculated in this work at the DFT/B3LYP 1511
1487 symmetric conformation of [i-C3C1im]+ are found at ∼315 and level of theory for two normal modes given new spectral 1512
1488 688 cm−1 and for the asymmetric conformation at ∼340 and features in IR (1540 cm−1) and Raman (1515 cm−1) spectra of 1513
1489 699 cm−1.191 The spectral pattern of Raman spectra of ionic [C4C1C1im][NTf2]. The small mass of hydrogen implies large 1514
1490 liquids may become very different after phase transition when displacements of hydrogen atoms in the eigenvectors displayed 1515
1491 some bands characterizing mixture of conformers in the liquid in Figure 23. However, in terms of potential energy distribution 1516
1492 phase are absent in the crystalline phase. Section 6 discusses the calculated with the VEDA program,50 the [C4C1C1im]+ IR 1517
1493 usage of vibrational spectroscopy for studying phase transition band at 1540 cm−1 involves mainly stretching of N(1)−C(2) and 1518
1494 of ionic liquids. N(3)−C(2) bonds (29%) and bending of the methyl group and 1519
1495 Methylation at the C(2) position of imidazolium cations, the C(5)−N(3)−C(2) angle (23%). The Raman band at 1515 1520
1496 leading to [CnC1C1im]+, blocks the most important site for C− cm−1 involves stretching of C(2)−CH3, C(4)−C(5), and N(1)− 1521
1497 H···A− hydrogen bonding. Despite eliminating this site for C(2) bonds (54%) and bending of the N(1)−C(2)−N(3) angle 1522
1498 hydrogen bonding, for a given anion the ionic liquid based on (13%). Hunt et al.197 showed that the electronic structure of 1523
1499 [CnC1C1im]+ is more viscous than the [CnC1im]+ counterpart. imidazolium cations is better represented by the double bond 1524
1500 The reason for this effect is not fully understood, and different C(4)C(5) and delocalization in the N(1)−C(2)−N(3) bonds of 1525
1501 explanations have been proposed.192−195 Concerning vibra- the ring.197 Therefore, the 1515 cm−1 Raman band of 1526
1502 tional spectroscopy, the most significant difference between [C4C1C1im]+ is very intense because of vibrations of N(1)− 1527
1503 [CnC1im]+ and [CnC1C1im]+ is seen in the spectral range of C(2)−N(3) and C(2)−CH3 moieties with large polarizability 1528
1504 ring vibrations. Noack et al.196 discussed IR and Raman spectra fluctuation. Endo et al.198 discussed the effect of methylation of 1529
1505 of ionic liquids based on [CnC1im]+ and [CnC1C1im]+, n = 2 the C(2) position on the spectral ranges of C−H stretching 1530
f22 1506 and 4, with the [NTf2]− anion. Figure 22 compares vibrational modes and 550−800 cm−1. Raman frequencies of C(4),(5)−H 1531
1507 spectra of [C4C1im][NTf2] and [C4C1C1im][NTf2]. Taking stretching modes change when [C4C1im]+ is replaced by 1532
[C4C1C1im]+, shifting to lower wavenumber when the 1533
counterion is Cl−, Br−, or I−, but shifting to a higher 1534
wavenumber when the counterion is [BF4]− or [PF6]−. The 1535
relative intensities of bands that characterize AA and GA 1536
conformers (see previous Figure 20) change in going from 1537
[C4C1im]+ to [C4C1C1im]+. These findings were considered as 1538
the consequence of the local arrangement of anions around 1539
[C4C1C1im]+ being different from [C4C1im]+.198 1540
Fewer works have been dedicated to detailed assignment of 1541
vibrational frequencies of protic imidazolium ionic liquids (i.e., 1542
monoalkylimidazolium with a free N−H bond). Moschovi et 1543
al.118 discussed IR and Raman spectra as a function of 1544
temperature of [C1im][NTf2], which melts at 325 K, where 1545
[C1im]+ is the 1-H-3-methylimidazolium cation. Moschovi et 1546
al.118 addressed the issue of vibrational signatures of [NTf2]− 1547

Figure 22. IR and Raman spectra of [C4C1C1im][NTf2] (green and


conformers and the effect of the hydrogen bond on the N−H 1548

blue lines) and [C4C1im][NTf2] (red and black lines) at room and C−H vibrations of the imidazolium ring. This work 1549

temperature. The band marked with asterisk is assigned to an anion provides full tables of observed IR and Raman frequencies and 1550
normal mode. Bands marked with arrows (Raman, 1515 cm−1; IR, vibrational assignments for [C1im][NTf2]. An expected differ- 1551
1540 cm−1) belonging to [C4C1C1im]+ have the atomic displacements ence in comparison with 1,3-dialkylimidazolium cations is the 1552
of the corresponding normal modes shown in Figure 23. occurrence of the N−H stretching mode of [C1im]+ at higher 1553

S DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1554 frequency than the C−H stretching modes. The ν(N−H)


1555 mode is observed at 3287 cm−1 in the Raman spectrum of solid
1556 [C1im][NTf2], becoming a very weak band at 3281 cm−1 upon
1557 melting. The ν(N−H) mode gives intense IR bands both in
1558 solid and liquid phases of [C1im][NTf2], 3271 and 3273 cm−1,
1559 respectively, although broader in the liquid.118 Moschovi et
1560 al.199 went further in reporting IR and Raman spectra for a
1561 series of protic ionic liquids [Cnim][NTf2], n = 0−12. Insights
1562 on conformation of the alkyl chain of [Cnim]+ cations have
1563 been obtained from the Raman bands in the 820−900 cm−1
1564 range, corresponding to deformation motions of the CH3 group
1565 at the end of the chain.199 It has been found that vibrational
1566 frequencies of N−H and C−H stretching modes in protic
1567 [C nim][NTf 2 ], 199 and also C−H modes in nonprotic
1568 [CnC1im][NTf2],200 decrease as n increases, but these findings
1569 have not been considered as signatures of stronger anion−
1570 cation hydrogen bonding. These authors considered instead the
1571 role played by an intramolecular effect, that is, the positive
1572 charge on the imidazolium ring is diminished by the electron Figure 25. Characteristics IR (red, transmittance scale at right) and
1573 donor inductive effect of longer alkyl chain, so that the strength Raman (black) bands of (A) 1-allyl-3-methylimidazolium dicyanamide,
1574 of anion−cation interaction does not increase even though (B) 1-benzyl-3-methylimidazolium dicyanamide, (C) 1-(3-cyanoprop-
1575 polar/nonpolar segregation becomes better defined in ionic yl)-3-methylimidazolium bis(trifluoromethanesulfonyl)imide, and (D)
1576 liquids with the imidazolium cation of the long chain. In a 1-(2-hydroxyethyl)-3-methylimidazolium tetrafluoroborate at room
temperature.
1577 subsequent work, Moschovi et al.201 provided very detailed
1578 analyses and comparisons of vibrational spectra for nonprotic,
1579 [CnC1im][NTf2], and protic, [Cnim][NTf2], ionic liquids. of the 1-allyl-3-methylimidazolium cation, but no spectroscopic 1604
1580 Inserting a functional group in the side chain of 1-alkyl-3- signatures have been proposed for these conformers. 1605
1581 methylimidazolium cations might result in ionic liquids with Shirota et al.204 discussed the vibrational spectra of some 1606
1582 improved properties for some specific tasks (e.g., gas ionic liquids containing benzyl-substituted cations, and Xue et 1607
1583 absorption). Despite vibrational spectroscopy being a powerful al.205 compared the spectra of 1-benzyl-3-methylimidazolium 1608
1584 tool to characterize molecular interactions in these solutions bis(trifluoromethanesulfonyl)imide and a solution of benzene 1609
1585 (see section 7), detailed spectroscopic analyses and assignments in [C1C1im][NTf2]. However, these works focused on the low- 1610
1586 of vibrations for these imidazolium derivatives are more sparse frequency range obtained by femtosecond Raman-induced Kerr 1611
1587 than the usual 1-alkyl-3-methylimidazolium cations. Once the effect spectroscopy. Shirota et al.204 provided the observed 1612
1588 vibrational frequencies of [CnC1im]+ and anions are well- vibrational frequencies for bands below 700 cm−1, and they 1613
1589 established, the most important band of the added functional distinguished the vibrations belonging to anions or cations. We 1614
f24 1590 group is easily identified. Figure 24 shows molecular structures show in Figure 25B part of the vibrational spectra, including the 1615
1591 of some derivatives whose characteristic vibrations we show in most characteristic new feature of 1-benzyl-3-methylimidazo- 1616
f25 1592 Figure 25. lium dicyanamide. The sharp band observed at 1003 cm−1 in 1617
the Raman spectrum of the ionic liquid is the counterpart to the 1618
totally symmetric breathing mode of benzene. 1619
There is still no detailed analysis of vibrational spectra of an 1620
ionic liquid based on CN-functionalized imidazolium cation. 1621
Figure 25C shows the characteristic band of the CN stretching 1622
vibration in 1-(3-cyanopropyl)-3-methylimidazolium bis- 1623
(trifluoromethanesulfonyl)imide, [NC-C3C1im][NTf2], at 1624
room temperature. The vibrational frequency of the ν(CN) 1625
mode at 2251 cm−1 in [NC-C3C1im]+ lies at a significantly 1626
Figure 24. Molecular structures of some functionalized imidazolium
cations. higher wavenumber than cyanate-anions (see Figure 10 and 1627
Table 1). The ν(CN) frequency in [NC-C3C1im]+ is indeed 1628
close to the value in the molecular liquid acetonitrile (2253 1629
cm−1). 1630
1593 Xuan et al.202 studied 1-allyl-3-methylimidazolium dicyana- Knorr et al.206,207 discussed temperature effect on the IR 1631
1594 mide and 1-allyl-3-methylimidazolium chloride with assign- band corresponding to O−H stretching mode of 1-(2- 1632
1595 ments based on potential energy distributions of normal modes hydroxyethyl)-3-methylimidazolium tetrafluoroborate, [HO- 1633
1596 calculated for ion pairs at the DFT/B3LYP level of theory. C2C1im][BF4]. The ν(OH) spectral range is shown in Figure 1634
1597 Figure 25A shows IR and Raman spectra of 1-allyl-3- 25D for [HO-C2C1im][BF4] at room temperature, where the 1635
1598 methylimidazolium dicyanamide in the spectral range where strong feature at ∼3552 cm−1 in the IR spectrum belongs to the 1636
1599 the CC stretching mode is observed at 1647 cm−1 in the O−H stretching motion in O−H···F hydrogen bonded to 1637
1600 Raman spectrum. Xu et al.203 assigned a few IR bands of 1-allyl- [BF4]−. Knorr et al.206,207 found that the low-frequency tail in 1638
1601 3-methylimidazolium bicarbonate in agreement with the more the IR band grows in intensity with decreasing temperature, 1639
1602 complete assignment of ref 202. The quantum chemistry becoming a well-resolved band at ∼3402 cm−1 at 233 K. With 1640
1603 calculations of Xuan et al.202 indicated three stable conformers the support of quantum chemistry calculations of clusters of 1641

T DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1642 ions, this low-frequency component has been assigned to the N-alkylpyridinium cations can be separated in stretching 1685
1643 O−H vibration engaged in O−H···O hydrogen bond with the vibrations of C−H bonds of the alkyl chain (2800−3000 1686
1644 neighbor cation.206,207 Katsyuba et al.208 observed a weak band cm−1) and the ring (above 3000 cm−1). Frequency shift of C− 1687
1645 at ∼3425 cm−1 in the IR spectrum of [HO-C2C1im][PF6], and H stretching of N-alkylpyridinium to a lower wavenumber can 1688
1646 they also assigned it to ν(OH) in hydrogen bonding with −OH also be related to stronger anion−cation interaction.216 1689
1647 groups of neighbor cations. Hydrogen bonding between like- It has not been addressed in these works on N- 1690
1648 charge species has been also suggested for the choline cation alkylpyridinium cations whether vibrational spectroscopy 1691
1649 from the analysis of the O−H stretching region of IR spectra of could be used to identify different conformation of the alkyl 1692
1650 [Cho][NTf2].209 Knorr et al.206,207,209 pointed out that these chain. Thus, we calculated harmonic vibrational frequencies at 1693
1651 are the first spectroscopic evidence of what Weinhold and the DFT/B3LYP level of theory for gauche and anti conformers 1694
1652 Klein210 called an antielectrostatic hydrogen bond (i.e., a of an isolated [Py1,4]+ cation (see molecular structures in Figure 1695 f27
1653 hydrogen bond between ions of the same charge). It is worth 27). The calculations suggest there is indeed a doublet of 1696 f27
1654 remembering, however, that signature of anion−anion hydro-
1655 gen bond between [HSO4]−, previously found in IR and Raman
1656 spectra of crystals of the halide salts157−160 has also been found
1657 in Raman spectra of [C2C1im][HSO4] and [C4C1im][HSO4]
1658 as discussed in the previous section.156
1659 4.2.2. Pyridinium. Gale et al. discussed Raman211 and IR212
1660 spectra of N-butylpyridinium in the early period of room
1661 temperature molten salts based on mixtures of an organic
1662 cation chloride and AlCl3. Tait and Osteryoung213 found that
1663 the main difference in IR spectra between basic and acidic
1664 AlCl3/N-butylpyridinium chloride mixtures (acidic melts have
1665 AlCl3:organic salt mole ratio greater than 1) occurs in the C−H
1666 stretching region, in particular enhancement in the intensity of
1667 bands of aliphatic relative to aromatic C−H stretching modes,
1668 suggesting loss in ring aromaticity with decreasing acidity.
1669 Tables of experimental frequencies of cation vibrations are
1670 available;213 however, the most important issue of a vibrational
1671 spectroscopy study of those systems was identifying complex
1672 species AlCl4− and Al2Cl7− rather than a detailed analysis of
1673 cation normal modes.211,212
f26 1674 Figure 26 shows the fingerprint regions of IR and Raman
1675 spectra of 1-butyl-4-methylpyridinium bromide, [Py1,4][Br], Figure 27. Raman spectra of liquid [Py1,4][BF4] (black line) and solid
[Py1,4]Br (blue line) at room temperature in the region including
bands assigned to different [Py 1,4 ] + conformers. Vibrational
frequencies and relative intensities calculated for [Py1,4]+ are indicated
by green (886 cm−1, gauche) and red (917 cm−1, anti) lines. Atomic
displacement vectors calculated for these normal modes are shown for
each conformer.

Raman bands (887 and 909 cm−1 in liquid [Py1,4][BF4]; 892 1697
and 913 cm−1 in solid [Py1,4]Br) that can be used to 1698
discriminate each conformer. These bands correspond to 1699
Raman active modes with frequencies calculated at 886 and 1700
917 cm−1, respectively, for gauche and anti [Py1,4]+ conformers. 1701
Figure 27 highlights this spectral range of the Raman spectra of 1702
[Py1,4][BF4] and [Py1,4]Br together with peak position and 1703

Figure 26. IR (red) and Raman (black) spectra of 1-butyl-4- relative Raman intensities obtained from the DFT calculation. 1704

methylpyridinium tetrafluoroborate, [Py1,4][BF4], at room temper- This figure indicates there is a mixture of gauche and anti 1705
ature. Bands marked with asterisk belong to anion vibrations. For [Py1,4]+ conformers in both solid [Py1,4]Br and liquid 1706
comparison purposes, IR and Raman spectra of solid [Py1,4]Br are [Py1,4][BF4]. Recently, Endo et al.186 provided a detailed 1707
shown by green and blue lines, respectively. theoretical analysis by DFT calculations on the conformational 1708
flexibility and the corresponding vibrational frequencies of the 1709
1676 and 1-butyl-4-methylpyridinium tetrafluoroborate, [Py1,4][BF4], most common organic cations forming ionic liquids. In the case 1710
1677 where the few [BF4]− bands are marked with asterisks. The of pyridinium derivatives, the calculations suggest that the 1711
1678 vibrational modes of N-ethylpyridinium have been discussed by 480−510 and 740−820 cm−1 regions of the Raman spectrum 1712
1679 Zhao et al.214 for [BF4]− and [CF3COO]− salts, and by Zheng could also be used to identify different conformers. 1713
1680 et al.215 for the Br− salt. Sashina et al.216 discussed IR and 4.2.3. Pyrrolidinium. The relationship between vibrational 1714
1681 Raman spectra of halides salts of N-alkylpyridinium for the frequency and molecular conformation has been studied in 1715
1682 series from two to ten carbon atoms in the alkyl chain. These more detail for 1,1-dialkylpyrrolidinium than pyridinium 1716
1683 authors paid attention to the high-frequency range of 2800− cations. Castriota et al.114 reported Raman spectra in the 1717
1684 3200 cm−1 of C−H stretching modes. This spectral range for 250−1700 cm−1 range of ionic liquids based on 1-methyl-1- 1718

U DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1719 propylpyrrolidinium, [Pyr1,3]+, with the anions [NTf2]− and I−. conformers in the liquid phase. Although there is overlap of 1751
1720 However, assignment of [Pyr1,3]+ vibrations was not given bands in this region, quantum chemistry calculations indicate 1752
1721 because the issue was coordination of Li+ by [NTf2]− in a 2:1 that the 883 cm−1 Raman band is a signature of the ax 1753
1722 mixture of [Pyr1,3][NTf2] with Li[NTf2].114 The analogous conformer, whereas the 890−930 cm−1 range includes bands of 1754
1723 effect of a strong polarizing cation on the [NTf2]− vibrations both eq and ax conformers. Furthermore, the intensities of 1755
1724 has been discussed by Rocher et al.217 in [Pyr1,4][NTf2]/AlCl3 components at 892, 905, and 930 cm−1 decreases with 1756
1725 mixtures as a function of Al3+ concentration. The clear increasing temperature, while the intensity of the 883 cm−1 1757
1726 signature provided by Raman spectroscopy about coordination band remains the same, also indicating a mixture of eq and ax 1758
1727 of [NTf2]− to Li+ will be discussed in section 7. The vibrational [Pyr1,4]+ conformers in [Pyr1,4][NTf2].218 Umebayashi et al.127 1759
1728 motions of [Pyr1,3]+ and [Pyr1,4]+ have been discussed in also concluded for mixtures of eq and ax conformers in 1760
1729 several works.122,3,127,130,217,218 Besides gauche or anti con- [Pyr1,3][NTf2] and [Pyr1,4][NTf2] on the basis of the DFT/ 1761
1730 formation of the butyl chain, additional issues concerning B3LYP calculations and the temperature dependence of Raman 1762
1731 [Pyr1,4]+ include equatorial or axial position (eq or ax) of the spectra in the range shown in Figure 28. Fujii et al.130 paid 1763
1732 butyl chain relative to ring carbon atoms, and envelope or twist attention to the 250−400 cm−1 range of the Raman spectrum 1764
1733 conformation of the nonplanar ring. By comparing exper- of [Pyr1,3][N(SO2F)2], in which cation bands at 350 and 380 1765
1734 imental Raman spectrum and DFT/B3LYP calculation within cm−1 again indicate a mixture of [Pyr1,3]+ conformers. There is 1766
1735 the whole 200−1600 cm−1 range, Fujimori et al.218 concluded an interesting contrast between these studies on pyrrolidi- 1767
1736 that in [Pyr1,4][NTf2] the butyl chain and the ring are at all anti nium218,127 and the above discussion of imidazolium ionic 1768
1737 and envelope conformations, respectively, but there is mixture liquids: there is mixture of anti and gauche conformation of the 1769
1738 of eq and ax [Pyr1,4]+ conformers. butyl chain in [C4C1im]+ ionic liquids, but anti conformation of 1770
1739 It is difficult to characterize the conformers when many [Pyr1,4]+ predominates with negligible population of the gauche 1771
1740 experimental and calculated frequencies have to be compared, conformer. However, when [Pyr1,4]+ is functionalized upon 1772
1741 instead of a small frequency range including a few character- insertion of an ether group function in the long alkyl chain then 1773
1742 istics bands. Comparison with related systems with a single IR spectroscopy with support of DFT calculations indicate that 1774
1743 conformer in crystalline phase (e.g., [Pyr 1,4 ]Br) helps the side chain acquires gauche geometry.219 A distinctive aspect 1775
1744 identifying appropriate bands as the fingerprint of a given of works by Vitucci et al.121 and Mao et al.117 is that these 1776
1745 conformer.218 Bands observed at 486, 586, and 655 cm−1 in the authors considered the IR spectrum, rather than the Raman 1777
1746 Raman spectrum of solid [Pyr1,4]Br closely match the spectrum, of [Pyr1,4][NTf2]. In order to assign the IR bands by 1778
1747 frequencies calculated for [Pyr1,4]+ in ax−anti-envelope DFT calculations, Vitucci et al.121 considered the isolated ion, 1779
f28 1748 conformation. Figure 28 shows IR and Raman spectra of whereas Mao et al.117 considered an ionic pair. Tables 1780
1749 [Pyr1,4][NTf2] in the range of 860−950 cm−1 used by Fujimori containing assignment of [Pyr1,4]+ frequencies are avail- 1781
1750 et al.218 in order to reveal coexisting eq and ax [Pyr1,4]+ able.117,121 In particular, Figure 6 of ref 121 provides 1782
visualization of atomic displacements for the normal modes 1783
corresponding to the bands shown in Figure 28. 1784
4.2.4. Piperidinium. Derivatives of piperidinium encom- 1785
pass an important class of ionic liquid forming cations, but 1786
assignment of vibrational spectra has been much less discussed 1787
for piperidinium than imidazolium or pyrrolidinium cations. 1788
Shukla et al.120 compared calculated and experimental IR and 1789
Raman spectra of N-butyl-N-methylpiperidinium bis- 1790
(trifluoromethanesulfonyl)imide, [Pip1,4][NTf2] and [Pip1,4]Br, 1791
the latter being a solid with a relatively high melting point (241 1792
°C). Figure 29 shows IR and Raman spectra of [Pip1,4][NTf2] 1793 f29
and [Pip1,4]Br. As the [NTf2]− bands dominate the spectra of 1794
[Pip1,4][NTf2], we provide spectra for solid [Pip1,4]Br in order 1795

Figure 28. IR (red, transmittance scale at right) and Raman (black)


spectra of [Pyr1,4][NTf2] at room temperature in the region used to
characterize different conformers of [Pyr1,4]+. The asterisk marks the
883 cm−1 Raman band characteristic of the ax conformer. Optimized
structures are shown for an isolated [Pyr1,4]+ cation with the butyl-
chain in the equatorial (eq) or axial (ax) position in relation to the
ring. In both the structures, the butyl chain is in anti and the ring in Figure 29. IR and Raman spectra of liquid [Pip1,4][NTf2] (red and
envelope conformations. black) and solid [Pip1,4]Br (green and blue) at room temperature.

V DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 30. IR (red) and Raman (black) spectra of [C4C1C1C1N][NTf2] at room temperature in different regions covering cation bands.

1796 to highlight the [Pip1,4]+ bands. Conformations of piperidi- 1350−1550 cm−1 corresponds to C−H bending modes, and the 1847
1797 nium, like pyrrolidinium cations, include ring flexibility (chair, range of 800−1050 cm−1 corresponds to normal modes with 1848
1798 twist, or boat), eq or ax position of the butyl chain, and displacements of many atoms of the cation. The complex 1849
1799 rotational isomers of the butyl chain. The DFT/B3LYP patterns of atomic displacements calculated for the normal 1850
1800 calculations of Shukla et al.120 indicated that the more stable modes of [C1C1C1C3N]+ and [C1C1C1C6N]+ can be found in 1851
1801 conformation is chair for the piperidinium ring and gauche for ref 121. In a subsequent work, Vitucci et al.122 discussed 1852
1802 the butyl chain. The characteristic Raman bands of cisoid and spectral changes after crystallization of [C1C1C1C3N][NTf2] 1853
1803 transoid conformations of [NTf2]− were discussed by Shukla et and [C1C1C1C6N][NTf2] at ∼250 K. 1854
1804 al.,120 but they did not identify vibrations that could be used as The possibility of revealing alkyl chain conformation of 1855
1805 signatures of cation conformation. Shimizu et al.132 identified tetraalkylammonium cations by combined usage of IR spec- 1856
1806 characteristic bands of [Pip1,4]+ conformers in the range of troscopy and quantum chemistry calculations is a more 1857
1807 700−800 cm−1 of the Raman spectrum of [Pip1,4][N(SO2F)2]. challenging task in comparison with the above discussion on 1858
1808 With the support of DFT/B3LYP calculations for the chair 1-alkyl-3-methylimidazolium cations. Palumbo et al.223 used the 1859
1809 form of [Pip1,4]+, Shimizu et al.132 proposed that eq and ax DFT/B3LYP level of theory to calculate vibrational frequencies 1860
1810 conformations of [Pip1,4]+ can be identified by the Raman for six different conformers of [C1C1C1C3N]+. An appropriate 1861
1811 bands at 704 and 710 cm−1, respectively, in the crystalline spectral window for comparison between calculation and 1862
1812 phases of [Pip1,4][N(SO2F)2]. In the liquid phase, the Raman experiment is the 900−1070 cm−1 range of the IR spectrum 1863
1813 spectrum indicates a mixture of eq and ax conformers of of [C1C1C1C3N][NTf2]. The calculations indicated that the IR 1864
1814 [Pip1,4]+. feature at 1000 cm−1 (a vibration mixing C−C, C−N 1865
1815 4.2.5. Ammonium. A large variety of ionic liquids can be stretching, and CH2 wagging) is a characteristic band of the 1866
1816 prepared from mono-, bi-, tri-, and tetraalkylammonium lowest energy [C1C1C1C3N]+ conformer.223 Occurrence of 1867
1817 cations, [C i NH 3 ] + , [C i C j NH 2 ] + , [C i C j C k NH] + , and several bands in this spectral range indicates a mixture of 1868
1818 [CiCjCkClN]+, respectively, with lengths of alkyl chains conformers in the liquid phase, whereas only the lowest-energy 1869
1819 indicated by the number of carbon atoms i, j, k, and l. conformer is retained in the crystalline phase accompanied by 1870
1820 Vibrational frequencies of the prototype [NH4]+ of T d concomitant change in the spectral pattern.223 The application 1871
1821 symmetry in halide salts were already reviewed by Herzberg,42 of vibrational spectroscopy to investigate phase transitions of 1872
1822 3033 (A1), 1685 (E), 3134 (F2), and 1397 (F2) cm−1, the actual ionic liquids will be reviewed in section 6. 1873
1823 values being dependent on the counterion and temper- Domańska and Bogel-L̷ ukasik,224 in a study focusing on 1874
1824 ature.143,220,221 On the other hand, in the context of ionic thermodynamic properties of ammonium salts, provided the IR 1875
1825 liquid forming cations, carbon chains of different lengths result frequencies of solid phases of tetraalkylammonium bromide 1876
1826 in asymmetric species, and the alkyl substituents give salts in which one of the alkyl chains has been functionalized 1877
1827 conformational flexibility. Those cations with one or more with a terminal −OH group. Aparicio et al.225 used the high- 1878
1828 hydrogen atoms directly bounded to the nitrogen are called frequency range of the IR spectrum, 3100−3600 cm−1, in order 1879
1829 protic ionic liquids with properties dependent on relatively to monitor water uptake from the atmosphere by ionic liquids 1880
1830 strong anion−cation hydrogen bonds,222 whereas those based based on 2-hydroxyethyltrimethylammonium and tris(2- 1881
1831 on tetraalkylammonium cations are nonprotic ionic liquids. Let hydroxyethyl)methylammonium cations, but detailed assign- 1882
1832 us focus first on tetraalkylammonium ionic liquids, for which ment of cation vibrations was not an issue. A discussion on the 1883
1833 detailed assignment of cation vibrations became available nature of vibrations was given by Arkas et al.226 for N,N-di(2- 1884
1834 recently.121,122,223 hydroxyethyl)-N-methyl-N-alkylammonium bromides (alkyl 1885
1835 Vitucci et al.121 discussed IR spectra of [C1C1C1C3N][NTf2] chain from dodecyl to octadecyl), however, within the context 1886
1836 and [C1C1C1C6N][NTf2] with the help of quantum chemistry of liquid-crystal phase transition. 1887
1837 calculation at the DFT/B3LYP level of theory. The IR spectra Concerning −OH-functionalized ammonium cations, more 1888
1838 exhibit of course the high-frequency bands of C−H stretching detailed spectroscopy studies have been reported for choline 1889
1839 modes at 2800−3200 cm−1, but the fingerprint region is mainly (i.e., hydroxyethyl-trimethylammonium, [Cho]+). In the case of 1890
1840 dominated by the [NTf2]− bands. In accordance with Figure nonprotic tetraalkylammonium ionic liquids, an indication of 1891
1841 11, spectral windows appropriate to characterize the cations weak interaction comes from finding that simple juxtaposition 1892
1842 without [NTf2]− bands include the ranges of 800−1050 cm−1 of vibrational frequencies calculated for isolated anion and 1893
1843 and 1350−1550 cm−1. The work of Vitucci et al.121 concerned cation give reasonable agreement to experimental spectra.121 1894
f30 1844 the IR spectrum, so that we show in Figure 30 for completeness On the other hand, this is not the expectation for systems with 1895
1845 these spectral ranges for both IR and Raman spectra for a the possibility of strong hydrogen bonds such as choline cation 1896
1846 closely related system, [C4C1C1C1N][NTf2]. The range of with carboxylate170 or amino acid anions.227 Vibrational 1897

W DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1898 frequencies of the choline cation have been discussed by Tanzi


1899 et al.170 in ionic liquids with formate, propanoate, or butanoate.
1900 By comparing quantum chemistry calculations for the isolated
1901 choline and the choline−anion pair, these authors found the
1902 expected hydrogen bond induced frequency shift: ν(OH) shifts
1903 to a lower wavenumber, whereas HOC bending and CO
1904 stretching and torsion modes shift to a higher wavenumber.170
1905 The more powerful tool to calculate vibrational spectra of these
1906 strongly hydrogen-bonded ionic liquids is ab initio MD
1907 simulation, since the method takes into account many different
1908 ionic arrangements resulting from the liquid dynamics.170,227
1909 Tanzi et al.170 showed that Raman spectra of choline
1910 carboxilates exhibit a large variability of O−H stretching
1911 frequencies within 3200−3400 cm−1, whereas the less
1912 perturbed C−H stretching modes of choline cover narrower Figure 31. IR (red, transmittance scale at right) and Raman (black)
1913 range centered at ∼2985 cm−1. This spectral separation spectra of [Cho][NTf2] at room temperature. Some bands assigned to
1914 between O−H and C−H stretching modes is clear in IR choline are indicated by arrows.
1915 spectra reported by Harmon et al.228 for the simple salts
1916 choline bromide and choline iodide. The O−H stretching band [C1NH3][NO3] exhibits relatively sharp bands of C−H 1961
1917 in the IR spectrum of [Cho][NTf2] exhibits a tail with stretching modes in the range of 2800−3050 cm−1 and N−H 1962
1918 components at 3431 and 3474 cm−1 which have been assigned stretching modes seen as broad bands at 3100−3300 cm−1 1963
1919 as the signature of hydrogen bonding between neighboring because of a distribution of cation−anion hydrogen bonds of 1964
1920 choline cations.209 On the other hand, the IR spectrum different strengths.230 The intense [NO3]− Raman bands 1965
1921 reported by Campetella et al.227 for choline with the amino acid dominate the spectrum in the 400−1700 cm−1 range, where 1966
1922 anion alanine, where there is also overlap with the alanine N−H few bands (C−N stretching and CH2 bending) have been 1967
1923 stretching modes, shows a broad band covering the whole assigned to [C1NH3]+.230 1968
1924 2500−3500 cm−1 range. The environmental-induced modifica- This approach of comparing experimental spectra with DFT 1969
1925 tions of IR spectra have also been discussed by Perkins et al.229 calculations of clusters of ions was continued by Bodo et al.231 1970
1926 for the eutectic mixture of choline chloride and urea (1:2 molar for the sequence [C2NH3][NO3], [C3NH3][NO3], and 1971
1927 ratio). The ν(OH) IR band in pure choline chloride is a [C4NH3][NO3]. Figure 32 shows IR and Raman spectra of 1972 f32
1928 relatively sharp band at 3256 cm−1, which overlaps the N−H
1929 stretching modes of urea upon formation of the mixture.
1930 It is difficult to correlate vibrational frequencies in the
1931 spectral range of 700−1700 cm−1 for simple choline halides228
1932 with the frequencies observed in ionic liquids with strongly
1933 interacting carboxilate and amino acid anions. For instance, the
1934 IR band at 1088 cm−1 for choline carboxylate,170 which agrees
1935 with the value for choline halides,228 was assigned to the
1936 choline CO stretching mode in ref 170, but the corresponding
1937 band at 1091 cm−1 in choline alanine was assigned to CH rock
1938 and twisting.227 Thus, it would be interesting to mark the
1939 choline vibrational frequencies in IR and Raman spectra of an
1940 ionic liquid with a less interacting anion. Once the [NTf2]−
1941 frequencies were already discussed in the previous section, then
f31 1942 we show IR and Raman spectra of [Cho][NTf2] in Figure 31
1943 with the choline bands indicated by arrows. Harmon et al.228
1944 assigned the choline normal modes within 700−1200 cm−1,
1945 assuming a C3v point group for the (CH3)3N−CH2− moiety.
1946 Some of the IR frequencies reported for choline halides228 have
1947 a correspondent feature in the spectra shown in Figure 31 for
1948 [Cho][NTf2].
Figure 32. IR (red, transmittance scale at right) and Raman (black)
1949 The first example of the series of protic alkylammonium ionic spectra of [C3NH3][NO3] at room temperature. Bands indicated by
1950 liquids is monomethylammonium, [C1NH3]+. Raman spectra blue arrows at 828 and 868 cm−1 characterize gauche and anti
1951 and X-ray powder diffraction as a function of temperature have conformers. Optimized structures of propylammonium in gauche and
1952 been reported by Bodo et al.230 for [C1NH3][NO3], which anti conformations are shown.
1953 melts at 381 K. Calculation of vibrational spectra have been
1954 done by first using classical MD simulation of clusters of ion
1955 pairs, ([C1NH3][NO3])n, n up to 8, in order to generate [C3NH3][NO3]. Several groups of Raman bands were assigned 1973
1956 reasonable configurations, which were then optimized by to [CnNH3]+ vibrations:231 Cn−N bending (416−450 cm−1), 1974
1957 quantum chemistry calculation at the DFT/B3LYP level of Cn−N symmetric stretching (870−890 cm−1), CH2 and CH3 1975
1958 theory. This approach aims for a better representation of the wagging (∼990 cm−1), asymmetric C−C and C−N stretching 1976
1959 hydrogen bond network and the condensed phase spectrum. (∼950 cm−1), CHn rocking (1175−1195 cm−1), CHn scissoring 1977
1960 The high-frequency range of the Raman spectrum of liquid (∼1300 cm−1), and CH2 and CH3 bending (1450−1460 cm−1). 1978

X DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

1979 A specific comment concerns the band located at 870−890


1980 cm−1, which Henderson et al.232 assigned instead to C−C
1981 stretching and found at 875 cm−1 in the Raman spectrum of
1982 [C2NH3][NO3]. This assignment is in line with a detailed
1983 analysis performed by Hagemann and Bill233 of the Raman
1984 spectra of the simpler salts [C2NH3]Cl and [C2NH3]Br. These
1985 authors provided a full list of Raman frequencies and
1986 assignments for ethylammonium halides, including the
1987 complicated pattern of bands intensified by Fermi resonance
1988 in the high-frequency range of C−H and N−H stretching
1989 modes. It is worth mentioning that the ν(CC) mode at 1047
1990 cm−1 in [C3NH3]Cl is an useful probe of local structure along
1991 the phase transition,232,234 but it is hidden by the intense anion
1992 band νs(NO3) in [CnNH3][NO3]. Nevertheless, there is nice
Figure 33. IR (red, transmittance scale at right) and Raman (black)
1993 agreement between experiment and theory,231 in particular the spectra of [C2C2C2S][NTf2] at room temperature. Some bands
1994 finding that cation frequencies within 400−1700 cm−1 depend assigned to [C2C2C2S]+ are indicated by arrows.
1995 on the alkyl chain length of [CnNH3]+.
1996 Faria et al.235 calculated vibrational frequencies for different
1997 conformers of [C3NH3]+ at the DFT/B3LYP level of theory. One of the most fundamental applications of molecular 2039

1998 Temperature and pressure induced crystallization of [C3NH3]- spectroscopy42 is the calculation of thermodynamic properties 2040

1999 [NO3] helped Faria et al.235 to propose that Raman bands at of the ideal polyatomic gas model, for which exact expressions 2041

2000 828 and 868 cm−1 characterize, respectively, gauche and anti exist for the partition function of translational, rotational, and 2042

2001 conformers of [C3NH3]+ (i.e., there is a mixture of conformers vibrational degrees of freedom. Focusing here on the 2043

2002 in the liquid phase of [C3NH3][NO3]) (see Figure 32). The vibrational contribution, the knowledge of vibrational frequen- 2044

2003 high-frequency spectral range, which is separated into bands cies allows for the calculation of thermodynamic functions 2045

belonging to C−H and N−H stretching modes, is barely according to fundamental expressions of statistical mechanics 2046
2004
(e.g., for the heat capacity of harmonic oscillators):74,243 2047
2005 affected by the alkyl chain length because hydrogen bonds
2006 remain essentially the same between the polar head of C V,vib 3N − 6 ⎛ hvj ⎞2 e hvj / kT
2007 [CnNH3]+ and the nitrate anion.231 Bodo et al.236 have also = ∑ ⎜ ⎟ hv / kT
2008 compared the experimental Raman spectrum of liquid Nk j=1 ⎝ kT ⎠ (e j − 1)2 (5) 2048
2009 [C2NH3][NO3] with the power spectrum obtained from the
where k is the Boltzmann constant, h is the Planck constant, T 2049
2010 Fourier transform of the time correlation function of atomic
is temperature, N is the number of atoms, and the summation 2050
2011 velocities calculated by ab initio MD simulation. IR frequencies extends to the 3N − 6 normal modes (3N − 5 for linear 2051
2012 have been reported by Luo et al.237 for derivatives containing molecules) each one with vibrational frequency νj. Paulechka et 2052
2013 longer alkyl chains (trioctylammonium) and the more complex al.244 calculated heat capacity, energy, and entropy for the ideal 2053
2014 triphenylammonium (and also the trialkylphosphonium gas phase of [C4C1im][PF6]. Blokhin et al.245 calculated these 2054
2015 counterpart) triflate ionic liquids. thermodynamic properties for [C4C1im][NTf2], and Paulechka 2055
2016 4.2.6. Other Cations. In Berg’s review on Raman et al.244 extended the calculations for the series [CnC1im]- 2056
2017 spectroscopy of ionic liquids,3 he provided preliminary results [NTf2], n = 2, 4, 6, and 8. Following a combined usage of 2057
2018 for a system based on the tetramethylguanidinium cation, experimental IR frequencies and calculated by quantum 2058
2019 [TMGH]+. More detailed analysis of vibrations of [TMGH]+ chemistry methods, these works give full lists of vibrational 2059
2020 have been performed by Berg et al.238 and Horikawa et al.239 for frequencies νj that have actually been considered in the 2060
2021 [TMGH][NTf2], and by Berg et al. for [TMGH]Cl240 and thermodynamic calculations. Some low frequencies are not 2061
2022 [TMGH]Br.241 These works provide lists of vibrational included in the summation because they correspond to internal 2062
2023 frequencies experimentally observed and calculated by quantum rotations, which are commonly represented by an empirical 2063
2024 chemistry methods. cosine potential function for hindered rotations.74,243 Rota- 2064
2025 Carper et al.242 reported IR and Raman spectra of tional contributions to thermodynamic properties take into 2065
2026 trimethylsulfonium dicyanamide, [C1C1C1S][N(CN)2]. The account the moments of inertia given in these works.244−246 2066
2027 authors provided vibrational assignment based on quantum The quantum chemistry calculations giving the full set of 2067
2028 chemistry calculations for an ionic pair and a dimer vibrational frequencies needed in evaluating thermodynamic 2068
f33 2029 ([C1C1C1S][N(CN)2])2.242 We show in Figure 33 the properties have been done for a cation−anion pair because it is 2069
2030 fingerprint range of IR and Raman spectra of other considered that the gas phase of ionic liquids is made of ion 2070
2031 alkylsulfonium ionic liquid, [C2C2C2S][NTf2], with some pairs rather than isolated ions.247,248 Dong et al.249 were able to 2071
2032 cation bands indicated by arrows. obtain in situ IR spectrum of [C2C1im][NTf2] in the gas phase 2072

4.3. Applications
after vacuum distillation of the liquid at 150 °C. Relative 2073
intensities of some bands change in the gas phase in 2074
2033 The discussion of vibrational frequencies and normal modes comparison with the liquid phase spectrum, a finding 2075
2034 assignments for important anions and cations in the previous reproduced by calculations of IR spectra of clusters of different 2076
2035 sections already stated some of the applications of vibrational sizes.249 Obi et al.250 and Cooper et al.251 also concluded for 2077
2036 spectroscopy in studies of ionic liquids. This section aims an vaporization of [C2C1im][NTf2] as ion pairs, but these workers 2078
2037 overall view of these and further applications of vibrational obtained the IR spectrum after trapping the vapor in helium 2079
2038 spectroscopy of pure ionic liquids. nanodroplets at very low temperature250 or supersonic jet- 2080

Y DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

2081 cooled in helium carrier gas.251 The role played by methylation Vibrational assignments discussed in previous sections 2143
2082 of the imidazolium ring on the strength of interactions of the indicated there are some frequencies that characterize 2144
2083 anion to the imidazolium cation was addressed by Fournier et molecular conformations, whose difference of conformer 2145
2084 al.252 by comparing IR spectra of clusters of [C4C1im][BF4] energies according to quantum chemistry calculations might 2146
2085 and [C4C1C1im][BF4]. Both trapping techniques of helium be relatively small in comparison with thermal energy available 2147
2086 nanodroplets and matrix isolation were used by Hanke et al.253 at room temperature. Vibrational spectroscopy has been 2148
2087 to reveal individual ion contributions to the IR spectra of extensively used to estimate the relative population of 2149
2088 [C1C1C1C4N][NTf2]. Booth et al.254 identified in the gas phase conformers and to obtain the thermodynamic functions 2150
2089 several conformers of [C2C1im][NTf2] ion pairs with different characterizing conformational changes in ionic liquids. The 2151
2090 strengths of interaction with the anion according to the procedure considers equilibrium between conformers, conf1 ⇆ 2152
2091 magnitude of vibrational frequency shift of the C(2)−H conf2, with equilibrium constant K = [conf2]/[conf1], and takes 2153
2092 stretching mode. Berg et al.255 obtained in situ gas phase Raman intensities, Iconf1 and Iconf2 (i.e., areas of appropriate 2154
2093 Raman spectrum of a protic ionic liquid, 1-methylimidazolium Raman bands), as proportional to the concentration of 2155
2094 acetate. Comparison between spectra of the gas phase ionic conformers, Iconf1 = Jconf1[conf1] and Iconf 2 = Jconf2[conf2], 2156
2095 liquid and 1-methylimizadole and acetic acid indicated that where Jconf1 and Jconf2 are the Raman scattering cross sections 2157
2096 proton transfer takes place upon vaporization, so that the gas corresponding to the vibration of each conformer. From the 2158
2097 phase of this protic ionic liquid is made of the starting neutral experimental intensity ratio Iconf2/Iconf1, one obtains the relative 2159
2098 molecules.255 However, this is not a general rule for protic ionic population of conformers as long as one takes into account the 2160
2099 liquids as demonstrated by Horikawa et al.,239 who obtained IR ratio of Raman scattering cross sections Jconf2/Jconf2, which can 2161
2100 spectra of several protic ionic liquids after vaporization and be estimated by quantum chemistry methods. By recording 2162
2101 trapping in a cryogenic neon matrix. Comparison between Raman spectra as a function of temperature, one obtains the 2163
2102 spectra of the evaporated ionic liquid and the parent acids and temperature dependence of ln(Iconf2/Iconf1) and the enthalpy of 2164
2103 bases leads to the conclusion that vaporization as an ionic pair conformational change by traditional application of the van’t 2165
is preferred as larger is the difference in pKa of the parent acids d ln K ΔH o
2104 Hoff equation, 1 = − R , with the assumption that Raman
2105 and bases.239 Returning to the issue of calculating thermody- d (T ) 2166
2106 namic properties of the ideal gas phase of nonprotic scattering cross sections Jconf1 and Jconf2 do not depend on 2167
2107 imidazolium ionic liquids,245,246 the agreement between temperature. The ratio Jconf2/Jconf1 must be known in order to 2168
2108 calculation under the assumption of ionic pairs and obtain the entropy of conformational change because the 2169
2109 experimental data gives further support for the physical picture intercept of an Arrhenius plot of Raman intensities depends on 2170
2110 that nonprotic ionic liquids in the gas phase are not made of both ΔSo and Jconf2/Jconf1: 2171
2111 isolated anions and cations.
Vibrational spectroscopy is a complementary tool for thermal Iconf2 ΔH o ΔS o J
2112
ln =− + + ln conf2
2113 analysis and mass spectrometry256 to reveal whether thermal Iconf1 RT R Jconf1 (6) 2172
2114 decomposition of ionic liquids takes place and the nature of
2115 decomposition products. Berg et al.241 proposed that the Umebayashi et al.49 considered Raman bands of [C2C1im]- 2173
2116 Raman spectrum of 1,1,3,3-tetramethylguanidinium chloride, [BF4] and [C2C1im][CF3SO3] in the spectral range shown in 2174
2117 [TMGH]Cl, in the vapor phase at 225 °C was consistent with Figure 21 and obtained ΔHo ≈ 2 kJ mol−1 for the nonplanar ⇆ 2175
2118 the presence of [TMGH]+Cl− ion pairs rather than the neutral planar equilibrium of [C2C1im]+ (i.e., the nonplanar is slightly 2176
2119 molecules 1,1,3,3-tetramethylguanidine and HCl. A Raman more stable than the planar conformer). This figure of ΔHo 2177
2120 band at 2229 cm−1 was assigned to the N−H stretching mode obtained from Raman spectroscopy was practically the same as 2178
2121 experiencing significant lower frequency shift because of the calculated by quantum chemistry methods for the two 2179
2122 N−H···Cl− hydrogen bond. In a subsequent work, however, [C2C1im]+ conformers.49 In the case of [C4C1im]+, Holomb 2180
2123 Berg et al.241 corrected this interpretation since this band was et al.99 considered the intensities of Raman bands at 808, 825, 2181
2124 observed in exactly the same position in the Raman spectrum of 883, and 905 cm−1 as characteristics of four conformers 2182
2125 vapor phase of [TMGH]Br. Thus, it was proposed instead that gauche−gauche, gauche−anti, anti−gauche, and anti-anti, 2183
2126 this band belongs to CN stretching mode of dimethylcyana- respectively, and obtained the relative populations 2184
2127 mide, (CH3)2N−CN, resulting from thermal decomposition of 0.07:0.48:0.17:0.28 in the ionic liquid [C4 C1im][BF4 ]. 2185
2128 [TMGH]Cl and [TMGH]Br.241 In the temperature-jump Thermodynamic functions of conformational changes of 2186
2129 experiments of Chambreau et al.,257 the IR spectrum is [C4C1im]+ were obtained by Umebayashi et al.188 from the 2187
2130 recorded for the vapor phase generated after a discharge on a temperature-dependence of Raman intensities of bands in the 2188
2131 filament within a small sample of the ionic liquid. For instance, spectral range shown in Figure 20. These authors calculated 2189
2132 characteristic IR bands of CH3NCS indicated formation of this ΔHo, ΔSo, and ΔGo by considering an equilibrium between 2190
2133 species upon heating of [C4C1im][SCN], and Chambreau et only two [C4C1im]+ conformers, anti ⇆ gauche, for a sequence 2191
2134 al.257 proposed reaction mechanisms of thermal decomposition of ionic liquids containing chloride, bromide, and iodide. 2192
2135 for several ionic liquids based on 1-alkyl-3-methylimidazolium Umebayashi et al.188 found that both the conformers coexist in 2193
2136 cations and cyanate-anions. Liaw et al.258 obtained IR spectra of comparable quantities (i.e., ΔGo close to zero), even though 2194
2137 [C2C1im][C2SO4], [C6C1im]Cl, and [C4C1im][NTf2] before ΔHo drives the preference for the gauche [C4C1im]+ conformer 2195
2138 and after the flash point and after ignition. Changes on the as the halide anion gets smaller. This spectroscopic approach 2196
2139 spectral pattern after ignition were considered as evidence of for obtaining thermodynamic properties of conformational 2197
2140 thermal decomposition,258 so that these authors claimed that changes has been used for other cations (e.g., 1-isopropyl-3- 2198
2141 flammability of these ionic liquids is related to thermal methylimidazolium,191 N-alkyl-N-methylpyrrolidinium,126,127 2199
2142 decomposition rather than vaporization of the liquid. and [C1C2C2C3N]+).223 2200

Z DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

2201 Figure 12 showed Raman bands that are good markers of over cisoid ratio Itrans/Icis increases with the chain length is 2264
2202 different conformers of [NTf2]−, so that Raman spectroscopy different for protic and nonprotic ionic liquids: Itrans/Icis reaches 2265
2203 has been used to quantify thermodynamic properties related to a plateau when n = 4 in [Cnim][NTf2], whereas it steadily 2266
2204 conformational changes of this anion. Assuming the [NTf2]− increases up to n = 12 in [CnC1im][NTf2].201 2267
2205 equilibrium as transoid ⇆ cisoid, Fujii et al.124 used the 398 and In the case of [N(SO2F)2]−, Fujii et al.129 obtained 2268
2206 407 cm−1 Raman bands as markers of transoid and cisoid essentially the same enthalpy of conformational change as 2269
2207 conformers, respectively. These authors obtained ΔHo = 3.5 kJ [NTf2]− from the analysis of the temperature dependence of 2270
2208 mol−1 from the Raman spectra of [C2C1im][NTf2] in close the Raman bands of [C2C1im][N(SO2F)2] within 280−380 2271
2209 agreement with ab initio calculations of the isolated anion, 2.2− cm−1 (see this spectral range in Figure 14). If the [N(SO2F)2]− 2272
2210 3.1 kJ mol−1, depending on the level of theory. Lassègues et anion is in the ionic liquid based on the N-methyl-N-propyl- 2273
2211 al.125 considered the spectral range of 260−360 cm−1 shown in pyrrolidinium cation, the ΔHo of the [N(SO2F)2]− conforma- 2274
2212 Figure 12 and found that the relative population of transoid tional change is slightly higher, 6.8 kJ mol−1.130 2275
2213 [NTf2]− conformer is 75% in [C2C1im][NTf2] at room Investigating molecular conformations in condensed phase is 2276
2214 temperature but a slightly higher value for the enthalpy, ΔHo a common issue for molecular dynamics (MD) simulations of 2277
2215 = 4.5 kJ mol−1. Essentially the same ΔHo for conformational ionic liquids. Classical MD simulation relies on an assumed 2278
2216 change of [NTf2]− was obtained in N-alkyl-N-methylpyrrolidi- potential energy function including intermolecular interactions, 2279
2217 nium ionic liquids.126,127 A distinctive feature of a Raman which are normally described by Lennard-Jones potential and 2280
2218 spectroscopic study performed by Martinelli et al.116 for some Coulombic interactions between partial atomic charges and an 2281
2219 nonprotic and protic ionic liquids was that ΔHo of [NTf2]− intramolecular force field accounting for molecular vibra- 2282
2220 conformational change was also obtained from the intense tions.260 The validation of a proposed model is accomplished 2283
2221 Raman band at 740 cm−1. This band exhibits a slight by comparing thermodynamic, structural, and dynamical 2284
2222 asymmetric shape toward the low-wavenumber side, so that properties calculated by MD simulations and experimental 2285
2223 two components at 738 and 741 cm−1 resulting from the curve data.261 Nevertheless, vibrational spectra are the source of data 2286
2224 fit have been assigned to cisoid and transoid [NTf2] − for reasonable parameters of force constants needed for the 2287
2225 conformers, respectively. The analysis of the temperature intramolecular terms of the potential function. The well-known 2288
2226 dependence of the band shape gave ΔHo in reasonable force field of Canongia Lopes and Pádua (CL&P)75 has been 2289
2227 agreement with the value obtained from the analysis of the tested taking into account information on conformers 2290
2228 260−360 cm−1 spectral range.116 Plots ln(Itrans/Icis) versus T−1 distribution available from vibrational spectroscopy. Focusing 2291
2229 for bands that characterize transoid and cisoid conformers also here on the intramolecular part of the CL&P model, Vintra, this 2292
2230 offer spectroscopic signatures of the glass transition of glass- includes bond stretching, r, angle bending, θ, and torsion of 2293
2231 forming ionic liquids. Palumbo et al.259 considered IR bands in dihedral angles, ψ: 2294
2232 the range of 500−650 cm−1 to characterize [NTf2]− conformers
kr k
2233 in an ammonium-based ionic liquid: the linear Arrhenius plot Vintra = ∑ (r − req)2 + ∑ θ (θ − θeq)2 +
2234 above the glass transition temperature (Tg ∼ 210 K) changes bonds
2 angles
2
2235 slope at Tg and becomes constant indicating that the relative 4
concentration of [NTf2]− conformers does not change any k ψ ,m
2236
∑ ∑ [1 + (− 1)m cos(mψ )]
2237 longer in the glassy phase. dihedrals m = 1 2 (7) 2295
2238 Moschovi et al.118 paid attention to appropriate corrections
2239 on the raw spectral data prior the analysis of equilibrium by where re and θeq are equilibrium bond length and angle. Force 2296
2240 Raman spectroscopy. As already warned by Papatheodorou et constants of stretching and bending, kr and kθ, used in MD 2297
2241 al.2 in their review on molten salts, these corrections include simulations of ionic liquids are usually taken from previous 2298
2242 accounting for polarization by using the isotropic component of force fields (e.g., AMBER and OPLS-AA). It is worth noting 2299
2243 the Raman spectrum, the excitation wavelength dependence, that due to the harmonic oscillator model for stretching and 2300
2244 and the Boltzmann population factor. In the liquid phase of bending modes, the condensed phase effect on vibrational 2301
2245 protic [C1im][NTf2], which melts at Tm ∼ 325 K, Moschovi et frequency shift is not the issue being addressed by this kind of 2302
2246 al.118 found significantly higher enthalpy for conformational model. In fact, proper coupling between intra- and 2303
2247 change of [NTf2]− when these corrections are taken into intermolecular degrees of freedom is very dependent on 2304
2248 account, 8.5 kJmol−1, in comparison with the analysis anharmonicity of the probe oscillator.80,79 On the other hand, 2305
2249 performed using raw spectra (6.0 kJ mol−1). Taking a sequence special attention was given in the CL&P model for kψ,m 2306
2250 of protic ionic liquids [Cnim][NTf2] with increasing length of parameters. The dihedral angle terms should give potential 2307
2251 the alkyl chain (n up to 12), Moschovi et al.199 claimed that the energy profiles of torsion consistent with quantum chemistry 2308
2252 population of [NTf2]− conformers and the corresponding ΔHo calculations and the distribution of conformers resulting from 2309
2253 are signatures of the relative importance of anion−cation analyses of Raman spectra. For instance, in the case of 1-alkyl-3- 2310
2254 interaction by Coulombic forces in comparison with other methylimidazolium cations,262 the relative population of gauche 2311
2255 contributions to intermolecular forces (hydrogen bond, van der and anti conformers calculated by MD simulations was 2312
2256 Waals, and dispersion due to π−π interaction). These authors compared with the analysis of the 600−700 cm−1 spectral 2313
2257 proposed that as the alkyl chain of [Cnim]+ increases, leading to range, which includes the Raman bands that characterize cation 2314
2258 segregation of nonpolar and polar domains, then [NTf2]− conformers. 2315
2259 experiences a local environment in which the transoid Understanding macroscopic properties of ionic liquids on the 2316
2260 conformation is favored. The population of cisoid conformer basis of intermolecular interactions is a major aim that 2317
2261 increases with temperature proper to the positive ΔHo value, vibrational spectroscopy shares with other experimental 2318
2262 which however becomes less positive the longer the [Cnim]+ techniques and theoretical methods. IR spectroscopy provides 2319
2263 alkyl chain.199 On the other hand, the way that the transoid one of the most characteristic experimental evidence of the X− 2320

AA DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

2321 H···Y hydrogen bond: the vibrational frequency of the X−H nonpolar and polar domains of ionic liquids. In fact, the 2384
2322 stretching mode shifts to lower wavenumber (red shift) in combined usage of vibrational spectroscopy and quantum 2385
2323 comparison with the isolated X−H oscillator.263−267 Other chemistry calculations covers a large literature of ionic 2386
2324 spectroscopic features usually accompanying the ν(X−H) red liquids,99,117,120,146,148,52,53,202,242,268 part of it already consid- 2387
2325 shift on hydrogen bond formation include the enhancement of ered in the previous sections. On the other hand, in an attempt 2388
2326 IR intensity and broadening of this band and the increase of to correlate with melting temperature, Katsyuba et al.48 put 2389
2327 vibrational frequency (blue-shift) of the bending mode δ(X− forward the proposition that it is the anharmonicity of the 2390
2328 H). Applying vibrational spectroscopy for studying pure ionic intermolecular anion−cation vibration, rather than the 2391
2329 liquids aims insights on the strength and atomic sites involved interaction energy of an ionic pair, that determines the melting 2392
2330 in hydrogen bonds and, within a more general viewpoint, on temperature of ionic liquids. Their conclusion was based on 2393
2331 the overall nature of anion−cation interactions. Katsyuba et quantum chemistry calculation of the potential energy curve as 2394
2332 al.48 provided a detailed comparison between experimental a function of anion−cation distance. Direct evidence of the 2395
2333 frequencies and ab initio calculations for different ion pair intermolecular anion−cation vibration is the realm of far-IR and 2396
2334 configurations of 1-alkyl-3-methylimidazolium cations and low-frequency Raman spectroscopies to be discussed in the 2397
2335 [BF4]− or [PF6]−. Frequencies related to C−H stretching and next section. 2398
2336 out-of-plane vibrations of the imidazolium ring and ν(B−F) Some works went further on quantum chemistry calculations 2399
2337 orν(P−F) are sensitive to ion pair formation. Katsyuba et al.48 of vibrational spectra by considering clusters of ions instead of a 2400
2338 advocate that IR bands belonging to ν(C(4)−H) and ν(C(5)− single ionic pair. Dong et al.269 calculated the IR spectrum of 2401
2339 H) modes (∼3160 cm−1) are separated from the red-shifted [C2C1im][BF4] at the DFT/B3LYP level of theory for clusters 2402
2340 ν(C(2)−H) mode (∼3120 cm−1) because the C(2)−H···F of ions ([C2C1im][BF4])n with increasing size, n = 2, 3, 4, and 2403
2341 interaction is stronger than C(4),(5)−H···F interactions. 5. The overall appearance of theoretical IR spectrum (peak 2404
2342 This assignment of C−H stretching modes,54,119,53,175,177,48 positions, band shapes, and relative intensities, including the 2405
2343 and the alternative interpretation that the doublet of bands in FIR range) gets closer to the experimental spectrum of liquid 2406
2344 the high-frequency range of 1-alkyl-3-methylimidazolium [C2C1im][BF4] with increasing cluster size.269 Thus, there is an 2407
2345 vibrations is due to Fermi resonance,52,168,169 have already incipient hydrogen bond network making the five ion pairs 2408
2346 been addressed in section 4.2.1. In a subsequent work, cluster a microscopic model of bulk [C2C1im][BF4]. The need 2409
2347 Katsyuba et al.208 preferred to call these bands asν(Caromatic− for considering ion clusters in order to the full appearance of 2410
2348 H) being aware of the role played by Fermi resonance in the vibrational spectra being recovered by ab initio calculations is 2411
2349 spectral range of C−H stretching modes. These authors
particularly demanding for protic ionic liquids. Bodo et al.230 2412
2350 investigated imidazolium ionic liquids with a terminal −OH
calculated the Raman spectrum of methylammonium nitrate by 2413
2351 group in the side chain [HO-C2C1im]+ (see Figure 25) and
considering clusters whose structures were first generated by 2414
2352 different fluorinated anions. Vibrational frequencies of both
classical MD simulations and then optimized at the DFT/ 2415
2353 ν(Caromatic−H) and ν(O−H) decrease, and the bands become
B3LYP level of theory. It is worth remembering that such 2416
2354 broader, as the anion basicity increases.208 It is worth noting
calculations of clusters still concern harmonic vibrational 2417
2355 that the intensity of the lower frequency ν(Caromatic−H)
2356 component (3100−3130 cm−1) increases in relation to the frequencies, so that scaling factors are needed for better 2418

2357 higher frequency component of the doublet (3150−3170 agreement between calculation and experiment. Nevertheless, 2419

2358 cm−1) with increasing strength of anion−cation hydrogen the Raman spectra calculated for clusters ([C1NH3][NO3])n, n 2420

2359 bond.208 Taking for granted the assignment of these bands = 2, 4, 6, and 8, exhibit distribution of frequencies that become 2421

2360 asν(C(2)−H) and ν(C(4),(5)−H), respectively, these findings broader as the cluster size increases.230 In a subsequent work, 2422

2361 have been considered signatures of stronger H-bonding Bodo et al.231 extended the calculations of Raman spectra for 2423

2362 through the more acid C(2)−H site of dialkyl-substituted the series ethyl-, propyl-, and butylammonium nitrate. Thus, 2424

2363 imidazolium cations. Some empirical relationships between the ab initio calculation of reasonable structures of clusters of 2425

2364 vibrational frequency shift, Δν = ν(XHfree) − ν(XHbounded), and ions captures some key features of ionic interactions that are 2426

2365 the enthalpy of hydrogen bond formation, − ΔHHB, are well- important for the calculation of vibrational frequencies, in 2427

2366 known.265,266 Katsyuba et al.208 estimated −ΔHHB for hydrogen particular the strength distribution of anion−cation hydrogen 2428
2367 bonding between [HO-C2C1im]+ and fluorinated anions from bonds. 2429

2368 both ν(Caromatic−H) and ν(O−H) modes, covering a rather The nitrate anion can be used as an example of taking an 2430
2369 large range depending on the anion basicity, 1.6−10.0 kJ mol−1. anion vibration, νs(NO3), rather than C−H, N−H, or O−H 2431
2370 In contrast, Moschovi et al.199 kept the same [NTf2]− anion, stretching modes, as a probe of the difference in width in the 2432
2371 while investigating how the hydrogen bond strength depends distribution of hydrogen bond strengths between protic and 2433
2372 on the alkyl chain length of protic imidazolium cations, nonprotic ionic liquids. Figure 34 shows the νs(NO3) Raman 2434 f34
2373 [Cnim]+. They found that vibrational frequencies of both band in [C4C1im][NO3] and [C3NH3][NO3]. Stronger anion− 2435
2374 ν(Caromatic−H) and ν(N−H) modes of [Cnim]+ are red-shifted cation interaction in [C3NH3][NO3] in comparison with 2436
2375 as n increases.199 Garaga et al.200 obtained the same results for [C4C1im][NO3] implies +3.5 cm−1 shift of νs(NO3) and a 2437
2376 ν(Caromatic−H) modes in nonprotic [CnC1im][NTf2] ionic remarkable broadening and asymmetry of the band because of 2438
2377 liquids. However, these findings were assigned to the the disturbed hydrogen bond network in the protic ionic liquid. 2439
2378 intramolecular inductive effect of longer alkyl chain rather Full width at half height (fwhh) of the νs(NO3) Raman band in 2440
2379 than an effect on the strength of the anion−cation hydrogen [C4C1im][NO3] and [C3NH3][NO3] are 5.0 and 11.0 cm−1, 2441
2380 bond. respectively. For comparison purposes, fwhh is ∼15.0 cm−1 in 2442
2381 The works of Katsyuba et al.48,208 and Moschovi et al.199,201 molten NaNO3 at 600 K and ∼6.0 cm−1 in [NO3]− aqueous 2443
2382 mentioned above illustrate the application of IR and Raman solution at room temperature.270 DFT calculation performed 2444
2383 spectroscopies to infer about structural features both in by Faria et al.235 showed that νs(NO3) vibrational frequencies 2445

AB DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 5 in ref 236). In other words, the high-frequency range 2489


of the Raman spectrum of [C2NH3][NO3] essentially reflects 2490
the density of vibrational states: ν(N−H) contributes with a 2491
broad background band between 3000−3250 cm−1 proper to 2492
hydrogen bonding of various strengths, while ν(C−H) give 2493
relatively sharp peaks on this background.236 2494

5. VIBRATIONAL SPECTROSCOPY OF PURE IONIC


LIQUIDS IN THE LOW-FREQUENCY RANGE 2495

In the previous section, insights on anion−cation interactions 2496


relied on the analysis of vibrational frequency shifts and 2497
changes in intensities and band shapes in mid-infrared and 2498
Raman spectra. Attempt to obtain data more directly related to 2499
intermolecular vibrations is the realm of far-infrared (FIR) 2500
spectroscopy, 10 < ω < 400 cm−1.271 It is natural that a smaller 2501
Figure 34. Raman spectra in the range of the νs(NO3) mode of
[C3NH3][NO3] at room temperature (red) and the molten phase of number of studies on ionic liquids has been reported using FIR 2502

[C4C1im][NO3] at 313 K (black). Intensities of Raman spectra have spectroscopy because of more demanding components (e.g., 2503
been normalized. beamsplitter and detector for this spectral range).272 Never- 2504
theless, interesting results have been obtained by FIR 2505
spectroscopy, in particular, evidence in the spectral range 2506
2446 of anions in a cluster ([C3NH3][NO3])4 indeed cover the below 150 cm−1 of anion−cation hydrogen bonds. In the case 2507
2447 spectral range of bandwidth of the experimental spectrum. of Raman spectroscopy, the range ω < 150 cm−1 is usually 2508
2448 Ionic liquids whose vibrational spectra have been interpreted called the low-frequency Raman spectrum. Raman spectrom- 2509
2449 with the help of ab initio MD simulations include those based eters with double or triple monochromator, or else single stage 2510
2450 on 1-alkyl-3-methylimidazolium with cyanate anions or spectrometer with special bandpass filters, are needed to 2511
2451 acetate,33 alkylammonium with nitrate236 or bromide,89 and achieve the frequency range close to the exciting laser line. 2512
2452 choline with carboxylates170 or amino acid anions.227 Ab initio However, the intermolecular vibrational dynamics is hidden 2513
2453 MD simulations of ionic liquids go further than calculations of under the very intense quasi-elastic scattering which dominates 2514
2454 minimum energy structures by taking into account the liquid the low-frequency Raman spectrum of a liquid. On the other 2515
2455 phase dynamics and anharmonicity of vibrations. Wendler et hand, the quasi-elastic intensity decreases as the liquid is 2516
2456 al.82 found that the calculation of liquid phase IR spectra of cooled, so that intermolecular vibrations become apparent in 2517
2457 imidazolium ionic liquids exhibit a spectral pattern significantly Raman spectra of supercooled or glassy phases of ionic liquids. 2518
2458 different from the single ion pair calculation. The calculation of The revision of vibrational spectroscopy studies on crystal- 2519
2459 ion pair strongly favors the anion pointing toward the C(2)−H lization and glass transition of ionic liquids is the issue of the 2520
2460 bond of the imidazolium ring, whereas ab initio MD simulation next section. 2521
2461 explores a distribution of anions around cations. Extreme Fumino et al.273 reported the first FIR spectroscopic study 2522
2462 examples of this situation is [C2C1im][CH3COO], for which showing evidence of +C−H···A− hydrogen bonding for a series 2523
2463 quantum chemistry calculation of a single pair leads to the of ionic liquids with the [C2C1im]+ cation and different anions. 2524
2464 formation of hydrogen-bonded carbene−acetic acid with an IR An overview of similarities and differences in FIR spectra of 2525
2465 band at 2167 cm−1, which is however not seen in the aprotic and protic ionic liquids is provided by Figure 35 taken 2526 f35
2466 experimental spectrum nor in the spectrum calculated by ab from a subsequent work by Fumino et al.274 (We recommend 2527
2467 initio MD simulation of the liquid phase.33 Wendler et al.82 the review published by Fumino and Ludwig275 focusing on 2528
2468 found that including eight ionic pairs in the calculations was their FIR spectroscopy studies of ionic liquids.) Assignments of 2529
2469 enough to capture the most important dynamics on the point the featureless bands in FIR spectra demand quantum 2530
2470 of view of vibrational spectroscopy (i.e., the short-time chemistry calculations for clusters of ions in order to 2531
2471 dynamics of ions rattling inside the cages made by the disentangle the intermolecular anion−cation vibrations from 2532
2472 neighbors). intramolecular modes eventually covering the same frequency 2533
2473 Bodo et al.236 calculated the vibrational frequencies of range. 2534
2474 [C2NH3][NO3] by ab initio MD simulation of clusters of 6 or Low-frequency vibrations expected for anion−cation hydro- 2535
2475 24 ionic pairs. The experimental Raman spectrum of [C2NH3]- gen bond are the stretching of one ion against the other, ν(C− 2536
2476 [NO3] was compared to the power spectrum obtained by H···A), and the bending, δ(C−H···A). The ν(C−H···A) in 2537
2477 Fourier transforming the autocorrelation function of atomic imidazolium ionic liquids has been assigned to a FIR band 2538
2478 velocities so that Raman intensities were not reproduced in the around 100 cm−1, depending on the anion. Stronger anion− 2539
2479 fingerprint region. In this region there are different kinds of cation hydrogen bonding, according to the magnitude of red- 2540
2480 vibrations with very different Raman activities, so that they shift of high-frequency C−H stretching modes (see previous 2541
2481 should be properly weighted by polarizability fluctuations. In section), has the counterpart in the FIR spectrum as the 2542
2482 the case of choline alanine, Campetella et al.227 found frequency of the intermolecular vibration shifts to higher 2543
2483 reasonable agreement between experiment and calculation in wavenumber.274 Curve fit by Voigt functions showed that the 2544
2484 the fingerprint region of the IR spectrum since they calculated low-frequency tail of this broad FIR band overlaps the band 2545
2485 the Fourier transform of the autocorrelation function of assigned to δ(C−H···A) at 50−60 cm−1.275 On the other hand, 2546
2486 molecular dipole moments. On the other hand, the power Buffeteau et al.15 refrained from assigning the FIR band of 2547
2487 spectrum calculated for [C2NH3][NO3] reproduces the spectral imidazolium ionic liquids to a specific anion−cation vibration. 2548
2488 pattern in the region of C−H and N−H stretching modes (see Instead, these authors called the spectral feature in the FIR 2549

AC DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

reduced mass among different ion pairs. Yamada et al.278 found 2583
that the maximum of the FIR band for [CnC1im]X, n = 3, 4, 2584
and 6, does not depend on the alkyl chain length and follows 2585
the trend 72.5, 92, and 122 cm−1 for the halide anions I−, Br−, 2586
and Cl−, respectively. Yamada et al.278 argued that these 2587
frequencies are on the same proportion of the inverse of the 2588
square root of reduced mass for the vibration of anion and 2589
imidazolium ring. However, Fumino et al.274,279−281 compared 2590
the position of the FIR band for a large set of ionic liquids and 2591
showed that only the mass effect does not account for the 2592
vibrational frequency shift. These authors performed quantum 2593
chemistry calculations for clusters of ionic pairs and correlated 2594
FIR frequencies with binding energies, thus supporting the 2595
interpretation of frequency shifts as the result of different 2596
anion−cation force constants.273,274,281−283,280,277,284 Further- 2597
more, Fumino et al.282 showed that the vibrational frequency of 2598
the main spectral feature in the FIR spectrum correlates with 2599
the enthalpy of vaporization of the ionic liquid. 2600
A step forward along the combined usage of FIR spectros- 2601
copy and DFT calculations was the attempt at separating the 2602
relative contributions of hydrogen bonding and dispersion 2603
forces out of the total energy, which is predominantly 2604
Figure 35. FIR spectra of some aprotic ionic liquids based on Coulombic energy.284−286 Fumino et al.284 measured FIR 2605
imidazolium cations with [BF4]− or [NO3]− and the protic ionic liquid spectra as a function of temperature (303−353 K) for protic 2606
propylammonium nitrate. The arrow on the top of the low-frequency ammonium ionic liquids with appropriately chosen alkyl chain 2607
feature indicates the band assigned to the stretching mode of anion−
cation hydrogen-bonded. Bands in the gray area belong to intra-
lengths and anions in order to play with the relative strengths of 2608

molecular normal modes. Reproduced with permission from ref 274. hydrogen bonding and dispersion forces. Bands assigned to 2609
Copyright 2009 PCCP Owner Societies. δ(C−H···A) and ν(C−H···A) were seen in the FIR spectrum of 2610
[C6C6C6NH][CF3SO3] as two resolved peaks around 70 and 2611
130 cm−1, respectively, as temperature decreases. When 2612
2550 spectra as the density of states by analogy to low-frequency temperature increases, the FIR spectrum is dominated by a 2613
2551 spectra of glass forming liquids (see eq 10 below). Schwenzer et single broad band with maximum at 100 cm−1, which has been 2614
2552 al.144 assigned anion−cation hydrogen bond vibrations in FIR assigned to the anion interacting with the cation alkyl chains 2615
2553 spectra of [C4C1im][CF3SO3] and [C4C1im][HSO4] to bands (i.e., an ion pair determined by dispersion forces). The 2616
2554 observed at 80 and 103 cm−1, respectively. A distinctive feature temperature dependence of relative intensities of FIR bands 2617
2555 in the FIR spectrum of [C4C1im][HSO4] are additional bands assigned to local arrangements dominated by hydrogen bonds 2618
2556 that Schwenzer et al.,144 with the support of MD simulations and dispersion forces allowed for a van’t Hoff analysis of the 2619
2557 and NMR spectroscopy, assigned to hydrogen-bonded equilibrium between these two kinds of ion pairs.284 The van’t 2620
2558 [HSO4]− anions. The occurrence of anion−anion hydrogen Hoff plot indicated that the hydrogen-bonded ion pair is 2621
2559 bonding in [HSO4]− ionic liquids is in line with the proposition favored by ∼34 kJ mol−1 over the dispersion-interaction 2622
2560 of Ribeiro156 based on the analysis of high-frequency Raman dominated structure. This experimental finding pointed out the 2623
2561 bands of [HSO4]−. need for proper consideration of dispersion corrected methods 2624
2562 Fumino et al.276,277 argued that strong hydrogen bonds in the in DFT calculations of ion pair energies.284−286 2625
2563 protic ionic liquids alkylammonium nitrate imply higher Fumino et al.193 and Buffeteau et al.287 addressed the effect of 2626
2564 frequencies for anion−cation intermolecular vibrations. The methylation of the C(2)−H position of the imidazolium ring by 2627
2565 similarity between FIR spectra of alkylammonium nitrate and comparing the FIR spectra of [C 2 C 1 im][NTf 2 ] and 2628
2566 water led Fumino et al.276 to the proposition of a three- [C2C1C1im][NTf2]. FIR bands of [C4C1im][NTf2] and 2629
2567 dimensional hydrogen bond network in these protic ionic [C4C1C1im][NTf2] were found respectively at 83.5 and 79.0 2630
2568 liquids. In the case of imidazolium ionic liquids, Fumino et cm−1 by Fumino et al.193 and 82.5 and 73 cm−1 by Buffeteau et 2631
2569 al.193 proposed that [C2C1im][NTf2] is less viscous than al.287 Analogous effect of methylation of the imidazolium ring is 2632
2570 [C2C1C1im][NTf2], even though the former has the C(2)−H seen in the FIR spectra of [C4C1im][BF4] and [C4C1C1im]- 2633
2571 site available for stronger hydrogen bond because hydrogen [BF4] shown in Figure 35. In addition to the red shift of 2634
2572 bonds act like defects perturbing the charge symmetry of the vibrational frequency, Fumino et al.193 found that the intensity 2635
2573 Coulomb network. However, it is yet an unsettled issue why of the FIR band decreases because of weakening of the anion− 2636
2574 methylation of the most acidic hydrogen of the imidazolium cation hydrogen bonding as the stronger binding site C(2)−H is 2637
2575 ring results in more viscous ionic liquid. It has also been switched off in [C2C1C1im][NTf2]. In contrast, Buffeteau et 2638
2576 proposed that methylation of the C(2) site of the imidazolium al.287 have not found significant difference in the FIR intensity 2639
2577 ring restricts ionic mobility because of loss of variation of between [C4C1im][NTf2] and [C4C1C1im][NTf2] after proper 2640
2578 anion−cation arrangements,192 higher potential energy bar- normalization by the intensities of high-frequency intra- 2641
2579 rier,194 or reduced free volume.195 molecular bands. Buffeteau et al.287 also compared FIR spectra 2642
2580 In order to use the vibrational frequency shift seen in Figure of [C4C1im][NTf2] and [C4C1C1C1N][NTf2] and they found 2643
2581 35 as a signature of the strength of +C−H···A− hydrogen bond, essentially the same intensity and vibrational frequency for both 2644
2582 one must rule out that it is not a trivial effect of changing the the systems. Therefore, Buffeteau et al.287 reached the opposite 2645

AD DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

2646 conclusion, claiming that the characteristic broad band in FIR


2647 spectra should not be assigned exclusively to stretching mode of
2648 the anion−cation hydrogen bond. These authors preferred
2649 instead assigning the density of states revealed by the FIR
2650 spectrum as a kind of envelope over the solidlike lattice
2651 vibrations. This physical picture is alike the one resulting from
2652 low-frequency Raman spectroscopy studies of phase transitions
2653 of ionic liquids to be discussed in the next section.
2654 Wulf et al.288 compared FIR to low-frequency Raman
2655 spectra, and also Terahertz (THz) spectra, for the same set
2656 of ionic liquids ([C 2C 1im][SCN], [C 2C 1 im][N(CN) 2 ],
2657 [C2C1im][C2SO4], and [C2C1im][NTf2]) previously studied
2658 by Fumino et al.273 It has been found that the spectral pattern
2659 and optical constants obtained from FIR spectra of imidazolium
2660 ionic liquids agree with data obtained by THz spectrosco-
2661 py.278,283,287,288 However, the experimental setup of Wulf et
2662 al.288 did not allow obtaining reliable Raman spectra below 100
2663 cm−1 so that the comparison between FIR and low-frequency
2664 Raman was restricted to intramolecular bending modes of
2665 cations or anions observed in the range of 150−300 cm−1.
2666 Low-frequency Raman spectroscopy investigations of inter-
2667 molecular dynamics encompass a large literature with a special
2668 focus on glass-forming liquids289−293 and among them several Figure 36. Low-frequency Raman spectra of [C4C1im][NTf2]. The
2669 (high temperature) molten salts (e.g., alkali halides,294−296 top panel shows the raw spectrum, I(ω), and the inset shows the
2670 ZnCl2,297,298 BiCl3,298,299 mixtures ZnCl2−AlCl3,300 etc.).1,2 An susceptibility representation, χ″(ω) (eq 8), at room temperature. The
2671 early study of the low-frequency range was reported by Ribeiro bottom panel shows I(ω) as a function of temperature. Note different
2672 et al.301 for the low-temperature molten salt tetra(n-butyl)- intensity scales between top and bottom panels.
2673 ammonium croconate, [C4C4C4C4N]2[C5O5]·4H2O, which is a
2674 glass-forming liquid with Tg = 293 K. The first low-frequency experimentally accessible by comparing the Raman spectrum 2702
2675 Raman spectroscopy studies concerning nowadays typical ionic with g(ω) obtained by neutron scattering spectroscopy.298,305 2703
2676 liquids were reported by Ribeiro302 for [C4C1im][PF6] as a Another common representation of the low-frequency 2704
2677 function of temperature and by Iwata et al.8 for [C4C1im][BF4], spectrum is the reduced Raman intensity, Ired(ω) = χ″(ω)/ω. 2705
2678 [C4C1im][PF6], [C6C1im][PF6], and [C8C1im][PF6] at room Since the factor [n(ω)+1]−1 depends linearly with ω at low- 2706
2679 temperature. frequencies, Ired(ω) is similar to the experimental I(ω), whereas 2707
2680 In these works, the low-frequency Raman spectrum is the low-frequency bands are slightly shifted to higher 2708
2681 conveniently reported after discounting a thermal population frequencies in the χ″(ω) representation.306 In order to 2709
2682 factor from the raw experimental spectrum, I(ω), in the so- highlight that the low-frequency bands are not artifact of the 2710
2683 called susceptibility representation, χ″(ω): χ″(ω) representation, the bottom panel of Figure 36 shows the 2711

I(ω) raw Raman spectra, I(ω), of [C4C1im][NTf2] as a function of 2712


χ ″(ω) = temperature. Low-frequency Raman band assigned to inter- 2713
2684 n(ω) + 1 (8) molecular vibration is clearly seen in the raw spectrum of 2714

2685 where [C4C1im][NTf2] as the glass transition temperature, Tg = 181 2715


K, is achieved from above. The quasi-elastic component is due 2716
1 to anharmonicity and fast relaxation processes, so that its 2717
n(ω) =
2686 eℏω / kT − 1 (9) intensity decreases significantly with decreasing temper- 2718
ature.290−292,298,300,307−310 In fact, the plot of the quasi-elastic 2719
2687 The motivation behind the reduction of the low-frequency intensity versus temperature exhibits break of slope at the glass- 2720
2688 Raman spectrum is the relation between I(ω) and the density transition temperature for several ionic liquids.302,311,312 2721
2689 of vibrational states g(ω):303,304 The optical Kerr effect (OKE) spectrum is equivalent to the 2722
n(ω) + 1 depolarized Raman spectrum, the results of these techniques 2723
I(ω) = C(ω)g (ω)
(10) differing by the population factor.313 It has been found that the 2724
2690 ω
depolarization ratio is essentially constant at 0.75 in the low- 2725
2691 where C(ω) is the light-vibration coupling. The spectral pattern frequency range of the Raman spectrum of [C4C1im][PF6].302 2726
2692 of g(ω) corresponds to the power spectrum alluded for in In other words, essentially the same band shape is obtained in 2727
2693 section 3, and C(ω) accounts for how the low-frequency the low-frequency range for polarized, depolarized, or without 2728
2694 vibrations modulate the material polarizability. polarization selection of the scattered radiation. It is worth 2729
f36 2695 Figure 36 shows the raw low-frequency Raman spectrum of noting that the frequency-dependent depolarization ratio has 2730
2696 [C4C1im][NTf2] at room temperature, and the resulting χ″(ω) been found in Raman spectra of high temperature molten salts, 2731
2697 (inset in the top panel). The relatively sharp bands above 100 especially when complex molecular-like structures occur in the 2732
2698 cm−1 belongs to intramolecular modes of the [NTf2]− anion melt.298,299,314 In the case of ionic liquids, the intermolecular 2733
2699 (see Figure 11 and Table 2). If the frequency-dependence of vibrational dynamics has been much more investigated by 2734
2700 C(ω) were linear then χ″(ω) would give directly the density of OKE315−320 than low-frequency Raman spectroscopy. Both 2735
2701 vibrational states. The actual frequency dependence of C(ω) is OKE and Raman spectroscopies probe polarizability fluctua- 2736

AE DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 37. Low-frequency Raman spectra in the susceptibility representation of ionic liquids at room temperature. Left panel: [C4C1im][NTf2]
(black) and [C4C1C1C1N][NTf2] (red); right panel: [C4C1im][BF4] (black) and [C3NH3][NO3] (red). Intensities of χ″(ω) spectra have been
normalized.

2737 tions resulting from the molecular dynamics of the liquid, so The anion−cation hydrogen bond vibrations are not 2784
2738 that the spectral pattern of χ″(ω) is the same as the OKE expected to give as intense signals in Raman as in FIR spectra. 2785
2739 spectra of ionic liquids. The OKE spectrum encompasses a The distinctive features of low-frequency Raman spectra are 2786
2740 temperature-dependent relaxation contribution within the GHz made clearer in the right panel of Figure 37 showing χ″(ω) 2787
2741 range (the structural or α-relaxation) not accessible in the spectra for [C4C1im][BF4] and [C3NH3][NO3], which are two 2788
2742 Raman spectrum and the THz range of intermolecular systems corresponding to the extreme cases of FIR spectra 2789
2743 vibrations common to Raman spectroscopy (1 THz = 33.3 shown in Figure 35. The significant frequency shift in the FIR 2790
2744 cm−1).321 OKE spectroscopy is a time-resolved spectroscopy spectrum when moving from the nonprotic [C4C1im][BF4] to 2791
2745 technique, which is beyond the scope of this review, and the the protic [C3NH3][NO3] (see Figure 35) is not manifest in 2792
2746 reader is recommended previous review focusing on OKE the low-frequency Raman spectra. The χ″(ω) spectra of 2793
2747 spectroscopy of ionic liquids.322 [C4C1im][NTf2] and [C4C1im][BF4] (black lines in the right 2794
2748 Low-frequency χ″(ω) and OKE spectra, like FIR spectra of and left panels of Figure 37) are similar, except for the [NTf2]− 2795
2749 ionic liquids, demand curve fit in order to grasp any trend of normal mode at 120 cm−1, which is evidently absent in the 2796
2750 the components when changing ionic species or thermody- spectrum of the latter. On the other hand, the spectral patterns 2797
2751 namic state. Different numbers and functional forms of bands of χ″(ω) for the two ammonium ionic liquids shown by red 2798
2752 (e.g., damped harmonic oscillator) antisymmetrized Gaussian lines in the right and left panels of Figure 37 differ in relative 2799
2753 and log-normal functions have been used in a curve fitting the intensities of the components encompassing the band shape. 2800
2754 low-frequency range of OKE and Raman spectra.315−320,322 The intense component with maximum at ∼70 cm−1 in the 2801
2755 Iwata et al.8 considered two components in the fit of low- χ″(ω) spectrum of [C3NH3][NO3] has also been found by 2802
2756 frequency Raman spectra of imidazolium ionic liquids, but most Sonnleitner et al. 323 in OKE spectra of ethyl- and 2803
2757 probably the broad spectral pattern encompasses three propylammonium nitrate. This spectral feature is close to the 2804
2758 components around 20, 60, and 90 cm−1 assigned to FIR band at ∼60 cm−1, which Fumino et al.276 assigned to 2805
2759 intermolecular vibrations.312,322 (An example of this curve fit bending of the anion−cation hydrogen bond. Krüger et al.324 2806
2760 is shown below in Figure 43 for the χ″(ω) spectrum of followed the Fumino et al.276 assignment for the corresponding 2807
2761 [C2C1im][NTf2].) In line with assignments in aromatic feature observed at ∼40 cm−1 in the THz spectrum of 2808
2762 molecular liquids,321 the 90 cm−1 component has been assigned ethylammonium nitrate. However, Sonnleitner et al.323 assigned 2809
2763 to librational motion (i.e., hindered rotation) of the the band at ∼70 cm−1 to [NO3]− libration (i.e., restricted out- 2810
f37 2764 imidazolium ring. The left panel of Figure 37 illustrates this of-plane rotation about the C2 axes of [NO3]−) because OKE 2811
2765 point by comparing χ″(ω) of [C 4 C 1 im][NTf 2 ] and spectroscopy is primarily sensitive to rotational motions and to 2812
2766 [C4C1C1C1N][NTf2] at room temperature. When the aromatic the large polarizability anisotropy of the [NO3]− anion. It is 2813
2767 cation is replaced by a nonaromatic one, the χ″(ω) spectrum is worth mentioning that the additional band assigned to libration 2814
2768 less broad because of the missing 90 cm−1 component, and the of aromatic ring is also found at 60 cm−1 in the OKE spectrum 2815
2769 intramolecular [NTf2]− mode at 120 cm−1 becomes a well- when a phenyl group is attached to the longer chain of the 2816
2770 defined band in the [C4C1C1C1N][NTf2] spectrum. It is also imidazolium cation.322 The long tail extending up to 250 cm−1 2817
2771 clear from the asymmetric band shape of the χ″(ω) spectrum of in the χ″(ω) spectrum of [C3NH3][NO3] is absent in the 2818
2772 [C4C1C1C1N][NTf2] that there is at least two components at [C4C1C1C1N][NTf2] spectrum (compare red lines in right and 2819
2773 ∼20 and ∼60 cm−1, and these seem to be common spectral left panels of Figure 37). On the basis of curve fitting, 2820
2774 features whatever the combination of anion and cation. In FIR Sonnleitner et al.323 proposed that the high-frequency tail in 2821
2775 spectra, Buffeteau et al.287 found the same intensity of the OKE spectra of [C2NH3][NO3] and [C3NH3][NO3] includes 2822
2776 intermolecular band at ∼84 cm−1 for [C4C1im][NTf2] and unresolved bands due to hydrogen bond stretching and cation 2823
2777 [C4C1C1C1N][NTf2]. In contrast, when Raman intensities are librational motions. It is worth mentioning, however, early 2824
2778 normalized by the high-frequency bands of [NTf2]−, rather Raman spectroscopic studies on molten alkali nitrates in which 2825
2779 than normalization by the low-frequency band as in Figure 37, broad bands assigned to [NO3]− libration have been found at 2826
2780 low-frequency Raman spectra are more intense for imidazolium relatively high wavenumber (maxima at ∼90, ∼115, and ∼140 2827
2781 than ammonium ionic liquids,312 as one expects from larger cm−1 in molten KNO3, NaNO3, and LiNO3, respectively).325 2828
2782 fluctuation of polarizability in a system containing an aromatic Thus, the high-frequency tail in the χ″(ω) spectrum of 2829
2783 ring. [C3NH3][NO3] might also manifest an inhomogeneous 2830

AF DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

2831 distribution of librations in this protic ionic liquid, being the configurations and also by Fourier transforming the time 2894
2832 low-frequency counterpart of the previous discussion (see correlation function of atomic velocities. The density of 2895
2833 Figure 34) concerning the Raman band shape of νs(NO3). vibrational states resulting from MD simulations exhibits the 2896
2834 Despite similarities in positions of the components after trend of higher frequency shift with increasing strength of 2897
2835 curve fit of these featureless bands, low-frequency Raman and anion−cation interactions in imidazolium88 or the number of 2898
2836 IR active intermolecular modes are not necessarily the same. hydrogen-bonding sites in ammonium89 ionic liquids. In the 2899
2837 Within the context of large literature about intermolecular case of protic trialkylammonium triflate ionic liquids, N−H···O 2900
2838 vibrational dynamics of glass-forming liquids, the band at ∼20 stretching modes calculated around 160 cm−1 agree with FIR 2901
2839 cm−1 seen in low-frequency Raman spectra of the glassy phase spectra.343 However, the peaks in the theoretical density of 2902
2840 of ionic liquids (see Figure 36) is the so-called boson vibrational states are found at significantly lower frequencies 2903
2841 peak.291−293,298−300,306,308−310 It has been pointed out that than in the experimental FIR spectra.88,343 The density of states 2904
2842 the boson peak vibrations remain in the liquid state even calculated by MD simulation is more like the low-frequency 2905
2843 though the peak is hidden under the intense quasi-elastic Raman than the FIR spectrum. In other words, even though 2906
2844 scattering of the low-frequency Raman spectrum.326−329 In an polarizability fluctuations are not being taken into account in 2907
2845 OKE spectroscopy study of 1-ethyl-3-methylimidazolium the density of states, it resembles the susceptibility 2908
2846 tosylate, Li et al.330 assigned to the boson peak dynamics the representation of low-frequency Raman spectra of ionic liquids. 2909
2847 oscillatory component with period of ∼2.0 ps (i.e., ∼17 cm−1) Nevertheless, the calculations88,89,343 point out that the 2910
2848 seen in the time domain data a few degrees above the glass vibrational frequency of intermolecular dynamics is dominated 2911
2849 transition temperature (Tg = 201 K). The partial nature of by short-range interactions, and the peak position is indeed 2912
2850 longitudinal and transverse acoustic modes of the intermo- modulated by the strength of anion−cation hydrogen bond.275 2913
2851 lecular dynamics in the spatial range of wavevectors 0.1 < k < However, the actual mode composition involves a large number 2914
2852 1.0 Å−1 and frequency range 0 < ω < 100 cm−1 in viscous glass- of atoms rather than being localized only in the C−H···A 2915
2853 forming liquids has been established in many studies by moiety.88 Furthermore, the density of states calculated by MD 2916
2854 inelastic X-ray scattering spectroscopy and MD simula- simulation of the liquid phase of [C4C1im][PF6] resembles an 2917
2855 tions.331−334 Although the exact origin of the boson peak is envelope over the corresponding density of states of the 2918
2856 yet a lively debated issue, several studies have been pointing to crystalline phase,88 a finding in line with the previous 2919
2857 the nature of transverse acoustic vibrations of high wave-vectors proposition of Buffeteau et al.287 in their FIR study of ionic 2920
2858 (i.e., soundlike modes of short wavelength).335−338 It is worth liquids. At this point we reach the application of vibrational 2921
2859 stressing that projecting into soundlike modes recovers only spectroscopy in studying phase transitions of ionic liquids, an 2922
2860 part of the intermolecular vibrations of liquids because the issue to be discussed in the next section. 2923
2861 other part is random phase motion66,67,339,340 and eventually The MD simulations of Balasubramanian et al.88,89,343 aimed 2924
2862 molecular-like normal modes in high temperature molten salts the analysis of atomic displacements of low-frequency 2925
2863 in which complexes survive for a relatively long time.1,2 vibrations by the calculation of time correlation function of 2926
2864 The assignment of FIR spectra by Fumino et al.275 focused velocities, rather than time correlation functions of dipole 2927
2865 on local motions related to stretching and bending modes of moment or polarizability fluctuation. In contrast, Hu et al.344 2928
2866 ion pairs, whereas assignment of low-frequency Raman and calculated time correlation function of polarizability by MD 2929
2867 OKE spectra emphasized librational and cage-rattling motions simulation of 1-methoxyethylpyridinium dicyanamide aiming a 2930
2868 or the many body nature of the intermolecular dynamics. direct comparison with a previous OKE spectroscopy study by 2931
2869 Shirota322 considered the average frequency (i.e., the first Shirota and Castner.345 The calculations of Hu et al.344 2932
2870 moment M1 = ∫ ωI(ω)dω/∫ I(ω) dω) as a single quantitative considered polarizability fluctuations arising from reorienta- 2933
2871 parameter for the whole low-frequency range of the OKE tional dynamics and interaction-induced mechanism according 2934
2872 spectrum. It has been found that M1 obtained from OKE to the dipole−induced dipole (DID) model in analogy with the 2935
2873 spectrum of nonaromatic ionic liquids correlates with (γ/ρ)1/2, large body of literature on MD simulations of properties of 2936
2874 where γ is the surface tension and ρ is the density.322 The spectroscopic interest in molecular liquids.86,103,346 It is worth 2937
2875 correlation between M1 and (γ/ρ)1/2 is analogous to a simple noting that depolarization ratio as low as 0.1 in Raman spectra 2938
2876 harmonic oscillator result for the dependence of vibrational of high temperature molten salts1,347 prompted Madden et 2939
2877 frequency with force constant and reduced mass (k/μ)1/2, al.100,101 to propose other interaction-induced mechanisms 2940
2878 however, involving the surface tension as a bulk property resulting in polarizability fluctuation (e.g., short-range overlap, 2941
2879 related to the strength of intermolecular forces. field, and field gradient). In the case of ionic liquids, including 2942
2880 Comparison between Figures 35 and 37 indicates that the reorientation and DID mechanism is consistent with the 2943
2881 overall spectral pattern is different for FIR and low-frequency depolarization ratio of 0.75 found in low-frequency Raman 2944
2882 Raman (and OKE) spectra for a given ionic liquid. Molecular spectra.302 Reorientational and DID mechanisms in low-viscous 2945
2883 dynamics simulations have shown that the density of vibrational molecular liquids account for long- and short-time parts, 2946
2884 states resembles low-frequency Raman and OKE spectra of respectively, of the time correlation function of polarizability 2947
2885 ionic liquids.341,342 The group of Balasubramanian attempted a fluctuation.86 In other words, DID mechanism is expected to 2948
2886 theoretical approach for understanding the low-frequency manifest higher-frequency dynamics of molecules rattling in the 2949
2887 vibrations of nonprotic imidazolium88 and protic ammo- cage of neighbors. It is known that a clear time or frequency 2950
2888 nium89,343 ionic liquids. These authors used quantum chemistry separation between reorientational and interaction-induced 2951
2889 calculations of ion pairs and computer simulations of liquid and dynamics might not be strictly valid even in molecular liquids 2952
2890 crystalline phases by classical (force field based) and ab initio (e.g., CS2).86,103 Accordingly, Hu et al.344 found that 2953
2891 MD simulations. The vibrational analysis by MD simulations interaction-induced contribution extends to a relatively long- 2954
2892 has been performed by a normal-mode analysis after time of a hundred picoseconds because the cage effect and 2955
2893 diagonalization of the Hessian matrix obtained from quenched nondiffusive dynamics take longer in ionic liquids. On the other 2956

AG DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

2957 hand, cross-correlation between different mechanisms contri- Ionic liquid mesophases (i.e., liquid crystal) are formed with 2997
2958 buting to dipole and polarizability fluctuations and strong increasing length of the alkyl chain of cation or anion.350,353,354 2998
2959 coupling between reorientational and translational motions lead Among the first vibrational spectroscopy studies concerning 2999
2960 to the warning put forward by some authors323,320 that, ionic liquid phase transitions, De Roche et al.355 used Raman 3000
2961 irrespective of number and functional forms chosen in curve spectroscopy to analyze the crystal at room temperature and 3001
2962 fitting of FIR, low-frequency Raman and OKE spectra of ionic the smectic phase at 353 K of [C16C1im][PF6]. By using as 3002
2963 liquids, each component might not necessarily be related in a vibrational probes the Raman bands at 1055, 1065, and 1119 3003
2964 one-to-one basis to specific intermolecular dynamics. cm−1, related to νs(CC) and νas(CC) modes, they concluded 3004
that the hexadecyl chain is in the anti conformation for all of 3005
6. VIBRATIONAL SPECTROSCOPY FOR STUDYING the dihedrals in the crystalline phase and the crystal melts with 3006
2965 IONIC LIQUID PHASE TRANSITIONS formation of the mesophase and occurrence of gauche 3007
2966 Ionic liquids exhibit a rich phenomenology of different phases conformers. There are interesting questions concerning ionic 3008
2967 (e.g., crystalline, glassy, and supercooled or superpressed liquid crystals,356 but we will focus in this review on phase 3009
2968 liquid) which the same system may achieve depending on the transitions of medium size alkyl-chain ionic liquids, which 3010
2969 rate that the thermodynamic state is changed. Vibrational usually have low melting points. 3011
2970 spectroscopy has been instrumental in revealing structural A recurrent issue addressed by vibrational spectroscopy 3012
2971 modifications accompanying phase transitions. Most of these studies of ionic liquid phase transitions is eventual conforma- 3013
2972 studies were performed with variable temperature at atmos- tional changes indicated by the characteristic bands of different 3014
2973 pheric pressure or variable pressure at room temperature. The conformers. The possibility for the ions acquiring different 3015
2974 temperature and pressure dependence of vibrational frequen- conformations can lead to crystal polymorphism and solid− 3016
2975 cies, band shapes, and relative intensities in Raman and IR solid transitions due to interchange between conformers. 3017
2976 spectra provide insights on ion conformation or the local Hayashi et al.181 characterized two crystal polymorphs of 3018
2977 environment experienced by the ions in different phases. [C4C1im]Cl by a combined analysis of X-ray powder patterns 3019
2978 Early vibrational spectroscopic studies took advantage of the and Raman spectra. Soon after this work, Ozawa et al.184 used 3020
2979 crystalline structure to infer about liquid phase structure before Raman spectroscopy and DFT calculations to show that 3021
2980 X-ray and neutron scattering experiments were used to study rotational isomerism of the butyl chain of [C4C1im]+ implies 3022
2981 directly ionic liquids in the normal liquid phase.348−351 In this crystal polymorphism, one phase composed by anti−anti and 3023
2982 context, Dupont352 considered the C−H stretching modes in the other one composed by gauche−anti conformers (see 3024
2983 IR spectra of [C4C1im]+ based ionic liquids with different Figure 20 for the characteristic Raman bands of [CnC1im]+ 3025
2984 anions and emphasized structural similarities between solid and conformers). In a chapter of a recently published book,4 3026
2985 liquid states. The group of Hamaguchi181,183−185 compared Hamaguchi and coauthors reviewed their Raman spectroscopic 3027
2986 Raman spectra of different crystal polymorphs and liquid phase studies on structure and phase transitions of imidazolium-based 3028
2987 of [C4C1im]Cl and [C4C1im]Br, concluding that the local ionic liquids. Endo et al.198,357−359 extended the analysis of 3029
2988 structure in the liquid is similar to the one in the solid despite phase transitions to [C4C1im]+ based ionic liquids with 3030
2989 an ionic conformation being eventually selected in the different anions using Raman spectroscopy along other 3031
f38 2990 crystalline phase. We illustrate this point in Figure 38 with experimental techniques, such as NMR and differential 3032
scanning calorimetry (DSC). They reported the formation of 3033
three crystalline phases with [C4C1im]+ having anti−anti, 3034
gauche−anti, and gauche′−anti conformations. Raman spectra 3035
and DFT calculations reported by Endo et al.191 indicated the 3036
occurrence of two conformers, asymmetric and symmetric, for 3037
1-isopropyl-3-methylimidazolium, [i-C3C1im]+, in the liquid 3038
phase. However, crystalline phases of [i-C3C1im]Br and [i- 3039
C3C1im]I contain only the asymmetric conformer.198 Endo et 3040
al.191 also found similar behavior of solid−solid transitions and 3041
cation conformational changes when methylation in the C(2) 3042
atom of the imidazolium ring lead [C 4 C 1 im] + to 3043
[C4C1C1im]+based ionic liquids. Other systems for which 3044
temperature-dependent Raman spectroscopy was used to reveal 3045
conformational changes accompanying crystal polymorphism 3046
and solid−solid transition included piperidinium132 and 3047
ammonium230,232,235,360,123,128 ionic liquids. 3048
Figure 38. Raman spectra of [C4C1im][CF3SO3] in liquid and low
temperature crystalline phases. Raman intensities have been
Turning now to conformational changes of the anion, Raman 3049

normalized by the band marked with the asterisk. The inset highlights and IR spectra have been used to characterize the [NTf2]− 3050
the range of 585−635 cm−1. conformation along phase transitions of [NTf2]− based ionic 3051
liquids. It is worth mentioning that before ionic liquids 3052
becoming a broadly investigated class of compounds, vibra- 3053
2991 Raman spectra of [C4C1im][CF3SO3] in liquid and low tional spectroscopy has been extensively used for studying 3054
2992 temperature crystalline phases. There is indeed reasonable [NTf2]− conformation in polymer electrolytes.113,361 Crystal- 3055
2993 match between crystal and liquid spectra, the gauche line phases of typical ionic liquids have been found with 3056
2994 conformation of [C4C1im]+ being selected in the crystal [NTf2]− in cisoid or transoid conformations (see Figure 12 for 3057
2995 according to the occurrence of a single Raman band at 599 the characteristic Raman bands of [NTf2]− conformers). 3058
2996 cm−1 (see inset of Figure 38). Castriota et al.114 reported one of the first vibrational studies 3059

AH DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3060 on temperature-dependent phase transition of a [NTf2]− based


3061 ionic liquid, namely [Pyr1,4][NTf2], and its mixture with
3062 Li[NTf 2 ]. Raman spectroscopy provided insight about
3063 interactions between [NTf2]− and [Pyr1,4]+ and Li+ cations in
3064 crystal and liquid phases, but [NTf2]− conformation was not an
3065 issue. Herstedt et al.360 performed a study of cation and anion
3066 conformations in different solid phases of [C2C2C2C2N]-
3067 [NTf2], whose melting temperature is 377 K. The relative
3068 amount of [NTf2]− conformers in three different solid phases
3069 were evaluated on the basis of characteristic [NTf2]− Raman
3070 bands. The analysis helped understanding the disorder caused
3071 by [NTf2]− anion in the solid, thus explaining the formation of
3072 plastic crystals and, in part, accounting for the low melting
3073 point of [NTf2]− ionic liquids.362 Martinelli et al.116,363 used
3074 Raman spectroscopy to follow the evolution of [NTf2]−
3075 conformation as a function of temperature in bulk and Figure 39. Raman spectra of [C4C1C1C1N][NTf2] in liquid (298 K,
black line) and crystalline phases after slow cooling (240 K, blue line)
3076 nanoconfined ionic liquids with protic and aprotic cations. and cold crystallization (230 K, red line).
3077 Some studies also considered IR spectra as a function of
3078 temperature to follow [NTf2]− conformational changes during
3079 phase transitions.118,122,259,364 Vitucci et al.122 suggested that liquids.185,191,368−373 Hamaguchi and Ozawa185 followed the 3123
3080 the longer alkyl chain in ammonium ionic liquids stabilizes the melting of [C4C1im]Cl crystal by the time dependence of 3124
3081 lower energy cisoid [NTf2]− conformer in solid phase because Raman bands which characterize different conformations of the 3125
3082 of stronger interactions between [NTf2]− and the polar head of butyl chain. They observed an unusually long time for 3126
3083 the cation. It became apparent from vibrational spectroscopy equilibration of anti/gauche conformers in the liquid phase, 3127
3084 studies that structural rearrangements in polar and nonpolar concluding that conformational interconversion happens along 3128
3085 domains during phase transitions may be independent of each slow collective transformation of ensembles of [C4C1im]+ 3129
3086 other.235,123,128,365 The [N(SO2F)2]− is another anion for cations. In a subsequent paper, the melting process of 3130
3087 which conformational analysis in low temperature phase [C4C1im]Cl was investigated by Okajima and Hamaguchi372 3131
3088 behavior was investigated by Raman spectroscopy. Shimizu et using simultaneously the lattice vibration bands in the low- 3132
3089 al.132 found two different crystalline phases for [Pip1,4][N- frequency range and the Raman bands of butyl chain 3133
3090 (SO2F)2] with [N(SO2F)2]− either in the cisoid or transoid conformers. Figure 40 taken from their work shows Raman 3134 f40
3091 conformation (see Figure 14 for the characteristic Raman bands spectra of [C4C1im]Cl during the melting process and Raman 3135
3092 of [N(SO2F)2]− conformers). intensities as a function of time for the lattice vibration at 111 3136
3093 Kinetic effects play a central role in determining the complex cm−1 and the characteristic bands of butyl chain conformation. 3137
3094 phase behavior of ionic liquids. Lassègues et al.125 reported In this experiment, a small piece of [C4C1im]Cl single crystal 3138
3095 Raman spectra of [C2C1im][NTf2] in the crystal phase at 113 K was rapidly heated to 363 K, that is 30 K higher than the 3139
3096 obtained after slow cooling (∼1 K min−1) and glassy phase at melting point.372 The low-frequency Raman spectrum of liquid 3140
3097 113 K obtained after fast cooling (∼20 K min−1). These [C4C1im]Cl has the typical quasi-elastic scattering (see section 3141
3098 authors found that the crystal phase contained the cisoid 5), whereas the crystal phase spectrum has peaks characteristic 3142
3099 [NTf2]− conformer, whereas the glassy phase contained mainly of lattice modes. Lattice modes gradually disappear during 3143
3100 the transoid conformer. The cooling-rate dependence of melting, while the quasi-elastic scattering intensity increases. 3144
3101 [NTf2]− conformation illustrates how the thermal history However, the bottom panel of Figure 40 indicates time lag 3145
3102 determines the ionic liquid phase behavior. In fact, Raman between the disappearance of lattice modes and the 3146
3103 spectroscopy indicates that [NTf2]− achieves different con- equilibration of anti/gauche population in the liquid phase. 3147
3104 formations when partial crystallization of [C4C1C1C1N][NTf2] Thus, conformers remain in local structures, and only after 3148
3105 is obtained by slow cooling or by cold crystallization (i.e., conversion of those arrangements as a whole is the melting 3149
3106 crystallization by heating the glassy phase).123,128 The Raman process completed.185,372 Endo and Nishikawa191 have also 3150
f39 3107 spectra of [C4C1C1C1N][NTf2] shown in Figure 39 indicate observed that conformational changes are linked to the melting 3151
3108 there is mixture of [NTf2]− conformers in the normal liquid process of [i-C3C1im]I, concluding that the premelting region 3152
3109 phase, whereas the cisoid or transoid conformation is obtained observed in DSC as broad peaks corresponds to an equilibrium 3153
3110 in the crystal formed by slow cooling or cold crystallization, state distinct from liquid and crystalline states. Altogether, the 3154
3111 respectively. kinetics of ionic liquids during crystallization or melting 3155
3112 Many ionic liquids are easily supercooled, exhibiting glass processes has been assigned to complex conformational 3156
3113 transition typically around 190−210 K, so that it might be changes depending on cooperative rearrangements and a 3157
3114 difficult to obtain the low-temperature crystal.366,367 Crystalline sluggish collective dynamics. 3158
3115 phase has never been obtained for some ionic liquids along Vibrational spectroscopy has been used more recently to 3159
3116 cooling, or it has been obtained only by the cold crystallization investigate crystallization, solid−solid, and glass transition of 3160
3117 process. The ionic liquid can also become partially crystallized ionic liquids under high pressure. In analogy with temperature- 3161
3118 forming the so-called glacial state (i.e., microcrystallites dependent studies, Raman and IR spectra provide insights on 3162
3119 immersed in a matrix of supercooled liquid), for instance, conformational changes accompanying high pressure phase 3163
3120 [C4C1C1C1N][NTf2].123,128 A large number of studies using transitions. These studies concerned mainly anti/gauche 3164
3121 DSC, NMR, and Raman spectroscopy have investigated the conformations of 1-alkyl-3-methylimidazolium cati- 3165
3122 slow dynamics during the melting process of ionic ons,365,374−390 but some studies also analyzed conformations 3166

AI DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

on the compression rate.387,388 Moreover, large hysteresis of 3197


vibrational frequency and bandwidth of the high-pressure 3198
[C4C1im][CF3SO3] crystal was found when the pressure was 3199
released stepwise back to the atmospheric pressure.387 3200
Raman spectroscopy has been used to show that high 3201
pressure crystallization of the protic ionic liquid [C3NH3]- 3202
[NO3] may result in a microscopically heterogeneous 3203
sample.235,392 Figure 41A shows two spectral patterns recorded 3204 f41

Figure 40. Top panel: Raman spectra from the low-frequency region Figure 41. (A) Spectral patterns in the range of the νs(NO3) Raman
up to ∼700 cm−1 of a [C4C1im]Cl crystal during its melting process. band in two different regions of the crystalline sample of [C3NH3]-
The lattice vibration at 111 cm−1 and anti/gauche conformer bands [NO3] inside the DAC at ca. 1.5 GPa. Raman intensities have been
(625 and 603 cm−1) are indicated by dashed lines. Time starts normalized. (B) Photograph of the DAC sample chamber showing the
counting when melting begins. Bottom panel: time-dependence of [C3NH3][NO3] crystal. (C) Micro-Raman imaging using the spectral
intensity of the band at 111 cm−1 and the intensity ratio of anti/gauche region indicated by the blue square in (A). Spectra in red and green of
bands during the melting process. Arrows indicate 48, 54, and 57 s. (A) correspond to different regions of the mapping in (C). Adapted
Reproduced with permission from ref 372. Copyright 2011 The from ref 235. Copyright 2013 American Chemical Society.
Chemical Society of Japan.
by focusing the laser beam in different regions of the same 3205
sample inside the DAC after a quick increase of pressure to ca. 3206
3167 of ammonium cations 128,235, 391,392 and the [NTf 2 ] − 1.5 GPa. (Compare with the νs(NO3) Raman band of 3207
3168 anion.128,364,385,393,394 Many of these papers report the glass [C3NH3][NO3] in the normal liquid phase shown in Figure 3208
3169 transition pressure at room temperature, Pg, according to the 34.) A photograph of the sample chamber with the [C3NH3]- 3209
3170 method of measuring the pressure dependence of the [NO3] crystal is shown in Figure 41B. The distorted 3210
3171 bandwidth of ruby fluorescence spectrum.395 Stress relaxation arrangement of hydrogen-bonded ions resulting in a distribu- 3211
3172 is no longer achieved when the glass transition takes place, so tion of νs(NO3) vibrational frequencies in liquid [C3NH3]- 3212
3173 that the ruby acts like a microscopic probe of stress [NO3] was reproduced by DFT calculation of a cluster of four 3213
3174 heterogeneity in the sample. Thus, the plot of the bandwidth ionic pairs.235 Thus, distinct local structures can be arrested in 3214
3175 of ruby fluorescence spectrum as a function of pressure exhibits isles of microscopic heterogeneity under high pressure, 3215
3176 a noticeable change in slope at Pg. The Pg of 1-alkyl-3- resulting in the anomalous crystallization of [C3NH3][NO3]. 3216
3177 methylimidazolium-based ionic liquids depends on the alkyl The micro-Raman imaging in Figure 41C illustrates the spatial 3217
3178 chain length. For instance, Pg ranges from ca. 3.0 to 2.0 GPa distribution of microscopic heterogeneity of the high-pressure 3218
3179 along the series [CnC1im][BF4], as n increases from 2 to 8.386 [C3NH3][NO3] crystal. It is worth mentioning that stepwise 3219
3180 The anion also plays a role in determining the glass transition increase of pressure to 1.5 GPa also generates the microscopic 3220
3181 under high pressure, for instance, Pg is 1.6 and 2.1 GPa for heterogeneity, and the actual νs(NO3) band shape depends on 3221
3182 [C8C1im][PF6] and [C8C1im][BF4], respectively.396 Interest- the rate of increasing pressure.235,392 Concerning crystal growth 3222
3183 ingly, plots of the ruby emission bandwidth eventually suggests under high pressure, it is usually verified heterogeneous 3223
3184 that other transitions take place for pressures higher than nucleation starting up from the gasket wall in the DAC sample 3224
3185 Pg.379,386 Yoshimura et al.386 proposed the formation of other chamber. We provide as a movie showing the crystallization 3225
3186 densified structures most probably related to dynamic process of [C2C1im][NTf2] inside the DAC at ca. 1.0 GPa. 3226
3187 heterogeneity with the intrinsic hierarchy of structures relaxing Magnetic ionic liquids are composed of metal-containing 3227
3188 at different time scales. However, this is an open issue which anions with magnetic behavior, [FeCl4]− being the most 3228
3189 deserves further studies. ́
common. Garcia-Saiz et al.398 performed magnetization and 3229
3190 Some ionic liquids in glassy phase at high pressure experience Raman spectroscopy studies of [C2C1im][FeCl4] under high 3230
3191 crystallization when decompressed, this finding being the pressure. They found that the relatively low applied pressure of 3231
3192 pressure counterpart of cold crystallization when the low- 0.34 GPa is enough to modify magnetic interactions and to 3232
3193 temperature glass is heated.386,397 Kinetic of high pressure induce transition from antiferromagnetic to ferromagnetic 3233
3194 phase transition also exhibits similarities with low-temperature ordering. Furthermore, [C2C1im][FeCl4] exhibits magnetic 3234
3195 transition. It has been found that [C2C1im][CF3SO3] and hysteresis linked to liquid−solid phase transition due to the 3235
3196 [C4C1im][CF3SO3] may become a glass or a crystal depending alignment of [FeCl4]− anions. 3236

AJ DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3237 Vibrational spectroscopy under high pressure allows one to 298 K. The Raman spectrum of the glassy phase obtained by 3270
3238 obtain the volume of conformational change, ΔV, in analogy to sudden increase of pressure (2.4 GPa, 298 K) exhibits the 3271
3239 the temperature-dependent studies discussed in section 4.3. boson peak at ∼21 cm−1 and the imidazolium ring librational 3272
3240 One obtains ΔV from the pressure dependence of intensities of mode at ∼140 cm−1. The high-pressure crystal (1.3 GPa, 298 3273
3241 bands that characterize each conformer, Iconf1 and Iconf2. If one K), with consequent sharp bands in the Raman spectrum, is 3274
3242 assumes that Raman scattering cross sections for conformers 1 obtained when pressure is stepwise increased. [C4C1im]- 3275
3243 and 2 do not depend on pressure then ΔV is given by [CF3SO3] always crystallizes at low temperature under usual 3276
cooling rates,387 and the spectral pattern of the crystal at 0.1 3277
⎡ ∂ln(Iconf2/Iconf1) ⎤ MPa and 230 K does not resemble the pattern of the high- 3278
ΔV = −RT ⎢ ⎥
⎣ ∂P ⎦T (11)
pressure crystal. However, a direct qualitative interpretation of 3279
3244 different crystal structures based only on the low-frequency 3280

3245 where R, T, and P are the gas constant, temperature, and vibrational spectrum has to be done with care. Chen et al.399 3281

3246 pressure, respectively. Takekiyo et al. 378 determined obtained single-crystal X-ray diffraction data for a series of 3282

3247 ΔV(planar→nonplanar) = +1.6 cm3/mol for [C2C1im][BF4] and nitrile functionalized ionic liquids and analyzed low-frequency 3283

3248 ΔV(anti→gauche) = −0.7 cm3/mol for [C4C1im][BF4]. Capitani et Raman spectra taking into account the crystal structures. They 3284

3249 al.364,394 obtained ΔV(transoid→cisoid) for the [NTf2]− anion in found that ionic liquids having the same space group exhibit 3285

3250 [Pyr41][NTf2] and [N1116][NTf2], respectively, − 0.34 and indeed similar low-frequency Raman spectra, but they also 3286

3251 −0.41 cm3/mol. In the case of [N1116][NTf2], Capitani et al.364 warned that the analysis based only on the similarity of Raman 3287
3252 obtained ΔV(transoid→cisoid) from IR spectra, and they also spectra might lead to erroneous conclusions and it should be 3288
3253 obtained ΔV(transoid→cisoid) = +0.7 cm3/mol in a linear regime complemented by X-ray diffraction measurements.399 3289
3254 for pressures higher than Pg (i.e., above 2 GPa). These authors Few works concerning ionic liquid phase transitions have 3290
3255 also obtained ΔH for [N1116][NTf2] by temperature-dependent been published using the low-frequency range of vibrational 3291
3256 spectra, and they discussed the competition between anion− spectra. Roth et al.283 discussed the effects of hydrogen bonds 3292
3257 cation interactions, relative energy of conformers, and the anion in FIR spectra of [NTf2]− based systems with a series of 3293
3258 volume. imidazolium cations with the ring hydrogen atoms substituted 3294
3259 The low-frequency range probing lattice dynamics is of by methyl groups, including the solid phase of 1,2,3,4,5- 3295
3260 course expected to be very sensitive to ionic liquid phase pentamethylimidazolium derivative, whose melting point is 391 3296
3261 transitions. As discussed in section 5, the low-frequency Raman K. Low-frequency Raman spectroscopy has been used by 3297
3262 spectra of amorphous phases exhibit the intense quasi-elastic Okajima and Hamaguchi372 in order to follow the melting 3298
3263 scattering and the intermolecular vibrations appear as a broad process of [C4C1im]Cl (see Figure 40) and by Faria et al.235 to 3299
3264 band, the so-called boson peak, in contrast to the sharp peaks of distinguish crystalline phases of [C3NH3][NO3] at different 3300
f42 3265 crystal lattice modes. Figure 42 illustrates low-frequency Raman conditions of temperature and pressure. The low-frequency 3301
3266 spectra at different temperature and pressure conditions for range of the Raman spectrum of [C4C1C1C1N][NTf2] also 3302
3267 [C4C1im][CF3SO3] as it undergoes glass transition or indicated the glacial state123 (i.e., mixture of microcrystals and 3303
3268 crystallization. The quasi-elastic scattering dominates the supercooled liquid), according to the occurrence of sharp bands 3304
3269 Raman spectrum for the normal liquid phase at 0.1 MPa and of lattice modes on top of the broad band characteristic of 3305
amorphous phase. The ionic liquids [C4C1C1C1N][NTf2] and 3306
[C1C4C4C4N][NTf2] do not crystallize under high-pressure at 3307
room temperature, instead they undergo glass transition at 1.1 3308
and 1.3 GPa, respectively.128 Accordingly, the low-frequency 3309
Raman spectrum under high-pressure exhibits low intensity of 3310
quasi-elastic scattering and the characteristic boson peak due to 3311
intermolecular dynamics as expected for a glassy phase.128 3312
Penna et al.400 have found a correspondence between the 3313
position of intermolecular vibrational modes in the liquid state 3314
and the spectral features observed after partial crystallization of 3315
samples at low temperature or high pressure. This point is 3316
illustrated in Figure 43 with the susceptibility representation of 3317 f43
the Raman spectra of [C2C1im][NTf2] in supercooled liquid 3318
(250 K, ●) and partially crystallized (230 K, black line) 3319
phases.400 The components (blue lines) used in the curve fit 3320
(red line) of the liquid spectrum seems to correspond to 3321
broadening of sharp bands of the crystal spectrum. This finding 3322
strongly suggests that there is some kind of mesoscopic order in 3323
the supercooled ionic liquid beyond the well-known nanoscale 3324
heterogeneity of polar/nonpolar domains.400 However, it is 3325
worth noting that in many cases, e.g. [C4C1im]Cl, the broad 3326
bands in the low-frequency Raman spectrum of the glassy phase 3327
barely indicate any direct correspondence to the sharp peaks in 3328

Figure 42. Low-frequency Raman spectra of liquid, glassy, and the crystal parent spectrum. 3329
crystalline phases of [C4C1im][CF3SO3] obtained under different It is natural that there are several open questions related to 3330
conditions of temperature and pressure as indicated in the figure. phase behavior of ionic liquids proper to the complexity of 3331
Raman spectra have been shifted vertically to aid in visualization. structural and kinetic aspects of the transitions under low 3332

AK DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

spectra, may provide chemical information about interactions in 3373


the mixture.405,406 However, this approach is undermined when 3374
the reference spectrum cannot be obtained. In this context, 3375
Koch et al.407 discussed a multivariate approach for quaternary 3376
mixtures of [C2C1im][C2SO4], water, and D- and L-glucose. A 3377
word of caution is in order since not all multivariate analysis 3378
methods will provide meaningful information in the first 3379
approach, a careful choice being fundamental for the employed 3380
method and proper data pre and post processing.408 3381
Hydrogen bonding in equimolar mixtures of ionic liquids has 3382
been studied by Fumino et al.409 using FIR spectroscopy. These 3383
authors performed an investigation of imidazolium-based ionic 3384
liquids with systematic methylation in the imidazolium ring 3385
while keeping the same [NTf2]− anion409 and also the mixture 3386
Figure 43. Low-frequency Raman spectra in the susceptibility of ionic liquids with the same triethylammonium cation and 3387
representation of [C2C1im][NTf2] in supercooled liquid phase (250 methylsulfate and triflate anions.410 The underlying assumption 3388
K, ●) and after crystallization (230 K, black line). The curve fit (red in this approach is the additivity of spectra of different ionic 3389
line) and the individual components of fit (blue lines) of the
supercooled liquid spectrum are shown. Reproduced with permission liquids, an issue which has been discussed mainly using optical 3390

from ref 400. Copyright 2013 American Institute of Physics. heterodyne-detected Raman-induced Kerr effect spectrosco- 3391
py,411,319,412 and also using far409,413 and mid414 infrared 3392
spectroscopies. Cha and Kim414 considered the IR bands 3393
3333 temperature or high pressure. Studies simultaneously changing belonging to γ(CH), ν(C(2)−H), and νas(C(4,5)−H) modes to 3394
3334 temperature and pressure are on demand for covering a wider study hydrogen-bonding effects due to anion coordination in 3395
3335 region of the phase diagram of ionic liquids. These mixtures of regular or C(2)-deuterated [C4C1im]+ cation with 3396
3336 investigations would go beyond the scope of ionic liquids by different anions, Cl−, I−, [BF4]−, and [NTf2]−. If additivity were 3397
3337 shedding light on more general issues, such as the interplay valid, the spectrum of the mixture would be the sum of spectra 3398
3338 between glass transition and crystallization, nucleation and of the neat liquids weighted by the concentration. The 3399
3339 crystal growth processes, amorphous−amorphous transitions, agreement between the actual spectrum of the mixture and 3400
3340 and the melting process. Moreover, the mixture of ionic liquid the concentration-simulated spectrum for [C4C1im]I/[C4C1im] 3401
3341 with other compounds can produce hybrid materials with Cl and [C4C1im]Cl/[C4C1im]BF4 indicated that cation−anion 3402
3342 interesting properties and phase transitions. For example, Abe interactions were not significantly changed upon mixture. The 3403
3343 et al.401 used Raman spectroscopy in studying confinement of same conclusion is valid for [C4C1im][BF4]/[C8C1im][BF4] 3404
3344 water in an ionic liquid and the concentration dependence of mixture.414 Aparicio and Atilhan415 also concluded for 3405
3345 phase behavior. Furthermore, crystallization of ionic liquid from additivity of IR spectra in mixtures of 1-butyl-3-methylpyr- 3406
3346 different solvents might be an efficient method for purification idinium, [Py1,4]+, and 1-octyl-3-methylpiridinium, [Py1,8]+, with 3407
3347 and to generate new crystalline phases, as shown by Li et al.383 the [BF4]− anion. In contrast, in deuterated [C4C1im]+ 3408
3348 for an ionic liquid-methanol solution under high pressure. IR mixtures of [NTf2]− with either Cl− or I−, the hydrogen 3409
3349 and Raman spectroscopies are powerful tools for addressing bond imposed by the more coordinating halide anion caused 3410
3350 many of these issues. red shift of vibrational frequency of the ν(C(2)-D) mode and 3411
change in dipole moment derivative as inferred by the 3412
7. VIBRATIONAL SPECTROSCOPY OF IONIC LIQUID dependence of band area with concentration.414 Analogous 3413
3351 SOLUTIONS effect of mixing anions with different coordination strength has 3414
3352 An appealing feature of the ionic liquid chemistry is the been found by Fumino et al.410 in mixtures of protic 3415
3353 possibility of fine-tuning the properties by proper combination imidazolium ionic liquids. Aparicio and Atilhan415 found 3416
3354 of cations and anions or mixing different ionic liquids.402−404 nonlinear concentration dependence of vibrational frequencies 3417
3355 From the point of view of vibrational spectroscopy, mixtures of of C−H stretching modes in the range of 2800−3200 cm−1 for 3418
3356 ionic liquids with other ionic liquids, molecular solvents, ions, [Py1,4][BF4]/[Py1,4][N(CN)2] mixtures. This finding has been 3419
3357 polymers, etc. can serve as model systems for understanding the attributed to structural changes and dominance of [N(CN)2]− 3420
3358 complex balance of intermolecular interactions determining interaction, a conclusion being corroborated by molecular 3421
3359 solvent properties and the liquid structure. A large number of dynamics simulations. Miran et al.416 found anion effects in IR 3422
3360 systems has been studied using vibrational spectroscopy, but in spectra of mixtures of protic trimethylammonium ionic liquids 3423
3361 this section we limit our discussion to mixtures of ionic liquids with [NTf2]− and [HSO4]−. Therefore, one finds significant 3424
3362 with other ionic liquids, water, or other molecular solvents, spectral signatures in mixtures of ionic liquids, in particular 3425
3363 gases, and salts. The examples given in this section show that involving different anions, when there is large difference in 3426
3364 attempts to get insights on intermolecular interactions from coordinating and hydrogen bonding abilities of the ions. 3427
3365 vibrational spectroscopy of ionic liquids rely on frequency shift, Furthermore, the different coordination capacity of anions 3428
3366 band broadening, intensities variation, etc. take place in the results in different chemical environments,416 as also suggested 3429
3367 mixtures. The approach of assigning physically meaningful by the NMR measurements of the work of Cha and Kim.414 3430
3368 interpretation to spectral changes in Raman excess spectrosco- Even though mixtures of cations with different lengths of the 3431
3369 py usually demands a reference spectrum. In Raman excess alkyl chain (e.g., 1-alkyl-3-methylimidazolium cations), but the 3432
3370 spectroscopy, or other excess spectroscopies, the deviation of same anion, implies only small effects on vibrational spectra in 3433
3371 the experimental spectrum of the mixture from the ideal comparison with spectra of the parent neat liquids, the 3434
3372 counterpart, which is the weighted average of pure samples transport coefficients, excess thermodynamic properties, and 3435

AL DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3436 phase transitions may be greatly altered.417−420 In order to were exposed to water. Proper to the appearance of a Raman 3499
3437 record Raman spectra of glassy [C2C1im][NTf2] at low band at 3300 cm−1 assigned to OH stretching mode, they also 3500
3438 temperature, Penna et al.400 added a small amount (10% suggested that water molecules make a tight hydrogen bond 3501
3439 mol) of [C 6C 1im][NTf2 ] to prevent crystallization of network with the anions displacing them from their original 3502
3440 [C2C1im][NTf2]. This allowed studying the characteristic equilibrium position.440 Raman spectroscopy indeed indicates 3503
3441 boson peak in the low-frequency Raman spectra of glassy that the population of gauche conformer increases upon the 3504
3442 [C2C1im][NTf2], [C4C1im][NTf2], and [C6C1im][NTf2]. The anti conformer when a [C4C1im]+ based ionic liquid is diluted 3505
3443 boson peak frequency does not depend on the chain length, in water.190 Jeon et al.30 also observed that the population of 3506
3444 instead the frequency depends on the strength of the anion− gauche increases with respect to the anti conformer as the water 3507
3445 cation interaction when the anion is changed while keeping the content increases, but the conformers population ratio 3508
3446 same cation.400 decreases again beyond ca. 45 mol L−1. These authors 3509
3447 Mixtures of ionic liquids with molecular solvents have been attributed the increase in vibrational frequencies of νs(CH3) 3510
3448 used to probe intermolecular interactions and structure, and νas(CH3) modes of the alkyl chain to interactions, which 3511
3449 transport properties, phase transitions, and excess thermody- are mainly repulsive in nature, with the oxygen atoms of water. 3512
3450 namic properties.401,421−424 Vibrational spectroscopy may In line with this finding, Bodo et al.441 also observed a slight 3513
3451 address structural features such as nanoscale segrega- blue shift of vibrational modes related to the alkyl chain and the 3514
3452 tion,422,425,426 solvent properties and ionic liquid polar- polar head in the Raman spectra of [C4NH3][NO3] protic ionic 3515
3453 ity,427−429 and intermolecular interactions (e.g., the role played liquid in mixtures with water. 3516
3454 by hydrogen bonding).401,30,430,406,431,432 Methanol, ethanol, Danten et al.442 carried out a systematic study of [CnC1im]+ 3517
3455 ethylene glycol, dimethyl sulfoxide, and water are among the with increasing length of the alkyl chain (n = 1, 2, 4, 8), with 3518
3456 most commonly investigated molecular solvents in ionic liquids [BF4]− and [PF6]−, for water content below the limit of 3519
3457 solutions. solubility in these ionic liquids. These authors used IR and 3520
3458 As pointed out in section 2, water might be a problem while Raman spectroscopies in combination with ab initio 3521
3459 handling ionic liquids for spectroscopic studies, but the calculations for a nearly symmetrical complex made of one 3522
3460 absorption of water may provide interesting information water molecule and two anions. The experimental difference of 3523
3461 about structure and intermolecular interactions in ionic liquids. ca. 70−80 cm−1 between symmetric and antisymmetric OH 3524
3462 Andanson433 used water as a molecular probe of the effects on stretching modes was reproduced by the calculations of such 3525
3463 the liquid structure when mixing different anions in ternary complexes, but they did not find any significant effect of the 3526
3464 mixtures of [C4C1im]+ based ionic liquids with [PF6]−, Cl−, alkyl chain length on the vibrational spectra of the mixtures. 3527
3465 Br−, and [NTf2]−. These authors performed ab initio Cammarata et al.29 studied the effect of imidazolium ring 3528
3466 calculations of vibrational frequency of water’s symmetric and methylation, [C4C1im]+, [C4C1C1im]+, and 1-butyl-2,3,4,5- 3529
3467 antisymmetric stretching modes when the water molecule is tetramethylimidazolium, and the anions, [PF6]−, [SbF6]−, 3530
3468 bound to one or two anions in the presence of a cation. The [BF 4 ] − , [NTf 2 ] − , [ClO 4 ] − , [CF 3 SO 3 ] − , [NO 3 ] − , and 3531
3469 relative population of water molecules in different environ- [CF3CO2]−, on IR spectra in ATR and transmission modes 3532
3470 ments is then obtained by adjusting the calculation as a with the amount of water ranging from 2540 to 33090 ppm. 3533
3471 weighted sum of Gaussian band shapes to the experimental The enthalpy of vaporization of water molecules from the bulk 3534
3472 data. They found there was no significant segregation in the of the ionic liquid was estimated from the vibrational frequency 3535
3473 mixtures and essentially the same affinity of both the anions for shift of the antisymmetric stretching mode and resulted in the 3536
3474 water molecules in concentrations as high as 1% mol of following order for the water−anion interaction strength: 3537
3475 water.433 In contrast, Tran et al.434 showed the anion effect on [PF6]− < [SbF6]− < [BF4]− < [NTf2]− < [ClO4]− < [CF3SO3]− 3538
3476 the amount of water absorbed by the ionic liquid by studying < [NO3]− < [CF3CO2]−. 3539
3477 the near-infrared region 1400−2000 nm (ca. 5000−7142 The enhancement of the asymmetric stretching mode 3540
3478 cm−1), the absorption coefficient of water at ca. 1419 nm intensity of water molecule in ionic liquids and in some 3541
3479 being the most appropriate one to quantify water content. They electrolytic solutions has been found.442,443 Danten et al.443 3542
3480 found that [C4C1im][BF4] absorbs more water than [C4C1im]- evaluated the anion dependence of this spectral feature in 3543
3481 [PF6] or [C4C1im][NTf2] because of stronger hydrogen [C4C1im]+ ionic liquids, keeping the water content below the 3544
3482 bonding between water and [BF4]−. The view of stronger respective solubility. Using IR, Raman, and ab initio 3545
3483 [BF4]−−water hydrogen bonding, at least in the comparison calculations, they showed that the “interaction hierarchy” for 3546
3484 between [BF4]− and [PF6]−, was shared by Dominguez-Vidal et anions with water is [PF6]− < [BF4]− < [NTf2]− < [CF3SO3]−. 3547
3485 al.435 along a FIR spectroscopy study of [C4C1im]+ based ionic The trend in interaction energy has the correspondence in 3548
3486 liquids. It is worth mentioning that Fadeeva et al.28 did not find distances between water and the fluorine atoms on the 3549
3487 different molar absorptivities of water in near-infrared spectra of calculated clusters, 1.84 and 1.72 Å in [PF6]− and [BF4]−, 3550
3488 [Pyr1,4]+ ionic liquids with [CF3SO3]− or [NTf2]−. Further- respectively, or the oxygen atoms of the anions, 2.0 and 1.9 Å in 3551
3489 more, hydrogen bond strength can be assessed by the shift of [NTf2]− and [CF3SO3]−, respectively.442,443 Dahi et al.444 3552
3490 vibrational frequencies of ν(C(2)−H) and ν(C(4,5)−H) modes shared analogous conclusions concerning the anion effect on 3553
3491 of imidazolium cations in the mid-infrared region with water uptake and miscibility, resulting in a “hierarchy” similar to 3554
3492 increasing water content.176,423,406,436−439 one proposed by Danten et al.443 and Tran et al.434 while 3555
3493 There are other spectral changes when ionic liquids are keeping the same cation and water activity (ca. 0.80). Dahi et 3556
3494 mixed with water besides those modes involving the hydrogen al.444 studied the water sorption isotherms and ATR spectra of 3557
3495 atoms of the imidazolium ring. Saha and Hamaguchi440 found several ionic liquids with the protic cations [C2NH3]+, [C2im]+, 3558
3496 all anti conformation of the butyl chains in [CN-C4C1im]Cl and [C4im]+ and the nonprotic [C4C1im]+ and [C6C1im]+ with 3559
3497 and [CN-C4C1im]I crystals, but the intensities of Raman bands anions [PF6]−, [BF4]−, [CF3SO3]−, dibutylphosphate, and 3560
3498 belonging to the gauche conformer increase when the crystals bis(2-ethylhexyl)phosphate. These authors found a small effect 3561

AM DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3562 of the alkyl chain length of the cation for the same anion and dominate. FIR spectra of the protic system [C2C2C2NH]I 3625
3563 whether the cation is protic or not. They also show that reveal that the IR band at 106 cm−1, which characterizes the 3626
3564 characteristic IR bands of free water do not appear in the contact ion pair of hydrogen bonded cation−anion, remains as 3627
3565 spectrum of [C 2NH 3][CF 3SO 3], suggesting that water the protic ionic liquid is diluted in molecular solvents. The IR 3628
3566 molecules are part of a hydrogen bond network even at low band assigned to solvent-separated ion pair is found at a slightly 3629
3567 concentrations so that vibrational frequencies of symmetric and higher wavenumber, ∼150 cm−1, and the intensity ratio 3630
3568 asymmetric modes are the same as in liquid water.444 There is between these two bands as a function of temperature allows 3631
3569 strong dependence of the νas(SO3) mode of [CF3SO3]− with for a van’t Hoff analysis of the equilibrium between contact and 3632
3570 the water activity, in contrast to [PF6]−, [BF4]−, and [NTf2]−, solvent-separated ion pairs.448 Jiang et al.449 considered IR 3633
3571 whose vibrational frequencies are only weakly concentration- spectra of deuterated DMSO−[C4C1im][BF4] mixtures and 3634
3572 dependent. Strong dependence of anion vibrational frequencies found that the occurrence of polar domains is particularly 3635
3573 with the amount of water has also been reported for affected by increasing of pressure. Rodrigues and Santos450 3636
3574 [CH 3 COO] − , 34 [CF 3 COO] − , 430 [NO 3 ] − , 441 and [C- used the CO stretching mode of dimethylformamide as a probe 3637
3575 (CN)3]−.445 of the alkyl chain length effect in 1-alkyl-3-methylimidazolium 3638
3576 The ab initio calculations at the DFT/B3LYP level of theory bromide ionic liquids. These authors found frequency shift and 3639
3577 performed by Danten et al.443 for the water-[C4C1im][NTf2] change in band shape of the CO mode proportional to the 3640
3578 system considered a 1:2 complex with the [NTf2]− in the cisoid length of the alkyl chain. Wang et al.451 used IR spectroscopy 3641
3579 conformation. The anion−water conformation and stoichiom- and ab initio calculations to study hydrogen bonding of 1- 3642
3580 etry they used, however, were contrary to previous results. butylpiridinium tetrafluoroborate with D2O and deuterated 3643
3581 Yaghini et al.438 found that the anion conformation does not DMSO. The C−H stretching modes of the cation alkyl chain 3644
3582 change upon addition of water (recall that in pure [C2C1im]- exhibit a blue shift in D2O, but a red shift in DMSO, while the 3645
3583 [NTf2] and [C4C1im][NTf2] the transoid conformation is C−D stretching modes of deuterated DMSO exhibit a blue 3646
3584 preferred, see section 4.3). Moreover, Wulf et al.429 pointed out shift in the mixture. These findings were assigned to the 3647
3585 that [NTf2]− is unlikely to form 1:2 complex with water due to behavior of the C−D bonds of deuterated DMSO as electronic 3648
3586 steric reasons. When the 1:2 water−anion complexes are density acceptors and the alkyl chain atoms of the cations as 3649
3587 considered in order to represent solutions of low-water electronic density donors. 3650
3588 concentration,429,433,442,443,386 the calculations should also Other molecular solvents carrying the OH group may be 3651
3589 include two cations to take into account five body nonadditive used as a probe for anion−cation interactions, hydrogen bonds, 3652
3590 interactions.442,443 and structural features of ionic liquids. Noack et al.423 3653
3591 High concentration of water seems to disrupt the nanoscale compared the extent of conventional and unconventional 3654
3592 segregation of polar/apolar domains in ionic liquids.441 In the hydrogen bonds in vibrational frequencies of ν(C(2)−H) and 3655
3593 case of intermediate water concentration, some authors claim νas(C(4,5)−H) modes in the Raman spectra of [C2C1im]- 3656
3594 that the system is homogeneous without phase segregation, but [C2SO4] in mixtures of methanol and ethanol as cosolvents for 3657
3595 others argue in favor of formation of water clusters. If the water water. Conventional and unconventional hydrogen bonds lead 3658
3596 content in N,N-diethyl-N-methyl-N-2-methoxyethylammonium to red and blue shifts in vibrational frequencies of modes 3659
3597 tetrafluoroborate is below 80% (mol), it has been proposed that involving the hydrogen atom, respectively, this behavior being 3660
3598 water is mostly confined since Raman frequencies of symmetric related to changes in both the C−H bond length and the 3661
3599 and asymmetric modes are close to the values of free water electronic density.423,452 The authors correlated the effects on 3662
3600 molecules.401,446 The limit value of 80% is related to the ability frequency shift with the competition between dispersion and 3663
3601 of forming 1:4 [BF4]−−water complexes, and beyond that, electrostatic interactions. They also linked the balance between 3664
3602 water starts to interact with the more electronegative part of the these interactions with several macroscopic properties, such as 3665
3603 cation. The OH stretching modes appear in the spectral range excess thermodynamic properties and excess transport 3666
3604 characteristic of bulk water when the amount of water in the coefficients.423 Abe et al.422 used Raman spectroscopy, among 3667
3605 ionic liquid is above 90% (mol).401,446 These three regimes of several other techniques, to probe the relevance of length scales 3668
3606 concentration (up to 80%, 80−90%, and above 90%) has been associated with both the alkyl chain size of dialkylimidazolium 3669
3607 also identified by Yoshimura et al.424 in the plot of vibrational cations and the one imposed when adding different alcohols. 3670
3608 frequency of the νs(BF4) mode versus water content in The authors studied 1-alkyl-3-methylimidazolium-based ionic 3671
3609 [C4C1im][BF4]. On the other hand, Fumino et al.276 proposed liquids, ranging from ethyl to decyl, with primary (propanol, n- 3672
3610 that the pure protic ionic liquid [C2NH3][NO3] has a three- butanol), secondary (2-butanol), and tertiary alcohols (2- 3673
3611 dimensional hydrogen-bonded network, and Bodo et al.441 methyl-2-propanol). Besides consequences on liquid structure 3674
3612 pointed out that in [C4NH3][NO3] the H2O molecules will and phase transitions, Abe et al.422 found that the cisoid to 3675
3613 take part of the extended hydrogen bond network at any transoid ratio of [NTf2]− conformers, as revealed by the 3676
3614 concentration of water. fingerprint region of the Raman spectrum, exhibits instability 3677
3615 Vibrational spectroscopy has been used to investigate several close to a critical length of the cation alkyl chain in mixtures 3678
3616 others molecular solvent−ionic liquid mixtures. Fumino et with n-butanol, therefore, being a signature of effects of liquid− 3679
3617 al.447 combined FIR spectroscopy and ab initio calculations to liquid equilibrium because of interplay between two length 3680
3618 study the equilibrium between contact and solvent-shared ionic scales. 3681
3619 pairs of [C3C3C3NH][CF3SO3] dissolved in different molecular Singh et al.427 studied the effects of different cations and 3682
3620 solvents over a wide range of concentration. In the case of anions in IR spectra of [C4C1im][BF4], [C8C1im][BF4], and 3683
3621 solvents of a low dielectric constant (chloroform and [C8C1im][C8SO4] in mixtures with ethylene glycol. They 3684
3622 tetrahydrofuran), there is larger concentration of contact assigned the splitting into three bands of the O−H mode of 3685
3623 ionic pairs over solvent-separated ionic pairs, whereas for ethylene glycol to different chemical environments, and they 3686
3624 high dielectric constant (DMSO), solvent-separated ion pairs also found frequency shifts of ν(C−H) modes of the 3687

AN DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3688 imidazolium ring, specifically the νas(C(4,5)−H), in line with considered the (CO) mode of acetone, N,N-dimethylforma- 3751
3689 similar results obtained by Pal et al.452 The νas(S−O) and mide, and Fe(CO)5 as polarity probes. The authors did not 3752
3690 νs(C(2)−H) modes also exhibit frequency shifts in the provide parameters to quantify the polarity, but they showed 3753
3691 [C8C1im][C8SO4]−ethylene glycol mixture.427 In the case of qualitatively the correlation between the polarity and the 3754
3692 ionic liquids containing [BF4]−, it has been proposed that vibrational frequency shift of (CO). Using acetone as a probe in 3755
3693 regions rich in either ethylene glycol or ionic liquid are formed. ionic liquids with different anions, they showed that polarity of 3756
3694 Furthermore, Singh et al.427 used the blue shift of ν(OH) in [NTf2]− is smaller than [SCN]− based ionic liquids, in line with 3757
3695 order to make qualitative inference about the extent of the conclusions drawn by Wülf et al.429 using water as a probe. 3758
3696 disturbance of the hydrogen bond network of ethylene glycol. The ν(CO) frequency of Fe(CO)5 allowed the authors to 3759
3697 Comparatively, the effect is larger for [C8C1im][C8SO4], and in propose a series of decreasing polarity as the alkyl chain length 3760
3698 the case of [BF4]− based ionic liquids, the disturbance of the increases in [CnC1im][BF4], 3 ≤ n ≤ 10, and [CnC1im][PF6], 3 3761
3699 hydrogen bond network of ethylene glycol is larger as the alkyl ≤ n ≤ 8. Garcia et al.461 used frequency shifts of ν(CN) and 3762
3700 chain is longer. Pal et al.452 suggested that the interaction that νs(CD3) modes of normal and deuterated acetonitrile to 3763
3701 causes frequency shift of νas(C(4,5)−H) is of electrostatic nature, estimate acceptor (AN) and donor (DN) numbers458 of ionic 3764
3702 rather than hydrogen bonding between hydrogen atoms of the liquids on the basis of extensive data available for acetonitrile in 3765
3703 imidazolium ring and ethylene glycol. On the other hand, the common molecular solvents. It has been found consistent AN 3766
3704 fact that aggregation is not observed in the alkylsulfate-based and DN values estimated from either ν(CN) or νs(CD3) modes 3767
3705 ionic liquid might be assigned to the strong anion-ethylene of acetonitrile. The observed trend is that fluorination of anions 3768
3706 glycol hydrogen bond. This finding is in line with conclusions implies more acidic local environments (higher AN), while 3769
3707 from the IR study by Pal et al.,425 who used 1,2-propanediol as nonfluorinated anions implies more basic environments (higher 3770
3708 a probe in mixtures with [C4C1im]+ ionic liquids with [NTf2]−, DN).461 Summing up, the results of vibrational spectroscopy 3771
3709 [C1SO4]−, or [BF4]−. They pointed out from ATR measure- investigations of ionic liquids polarity indicate that the anion 3772
3710 ments that [C1SO4]− disturbs the most the hydrogen bond plays the dominant role, whereas the effect of changing the 3773
3711 network of the alcohol, followed by [NTf2]− and [BF4]−. It is length of the alkyl chain of dialkylimidazolium cations is less 3774
3712 worth noting that such order of “strength” of disturbance is important. 3775
3713 somewhat similar to the interaction strength series proposed by A well-known application of ionic liquids concerns their 3776
3714 Cammarata et al.29 and the polarity series obtained by Wulf et capacity to absorb gases. This can be a selective process aiming 3777
3715 al.429 sample purification or gas capture to remove greenhouse gases 3778
3716 Shirota et al.453 and Shimomura et al.454 used IR spectros- from the atmosphere.403,404 Vibrational spectroscopy has been 3779
3717 copy, among other techniques, to study mixtures of benzene in used as an analytical tool for quantifying the amount of 3780
3718 imidazolium ionic liquids by following the effect of the ions on absorbed gas462 or for studying the mechanism and eventual 3781
3719 the vibrational frequency of the out-of-plane C−H bending reactions between the gas molecules and ionic species.32,463 3782
3720 mode, δop(CH), of benzene. They found a significant blue shift Among the gas−ionic liquid solutions most widely investigated 3783
3721 of δop(CH), and supported by structural data and quantum by vibrational spectroscopy, SO2 and CO2 are the gases which 3784
3722 chemistry calculations, they suggested that the interaction draw more attention proper to their impact on the environ- 3785
3723 between benzene and the imidazolium ring is stronger than ment. 3786
3724 interactions between benzene rings (π···π interactions) or Zeng et al.464 carried out a systematic IR spectroscopy study 3787
3725 between benzene and hydrogen atoms (C−H···π interactions). of SO2 absorption by pyridinium-based ionic liquids with 3788
3726 Along the comparison between [C8C1im][BF4] and [C8C1im]- different lengths of the alkyl chain, while keeping the same 3789
3727 [NTf2], Shirota et al.453 assigned the larger frequency shift in [BF4]− anion, or different anions, [SCN]−, [BF4]−, and 3790
3728 the latter to weaker interaction between [NTf2]− and the [NTf2]−, while keeping the same [Py4]+ cation. In the case of 3791
3729 cation, allowing for stronger interaction between the probe [Py4][NTf2], the authors did not observe the SO2 bands after 3792
3730 molecule (benzene) and the cation. One could argue from the gas absorption because of overlapping with the [NTf2]− bands 3793
3731 results of Shimomura et al.454 that the alkyl chain length plays a at 1139 and 1352 cm−1. The symmetric and antisymmetric S− 3794
3732 less important role because the vibrational frequency shift of O stretching modes of the SO2 molecule, νs(SO) and νa(SO), 3795
3733 δop(CH) is almost the same in the [C12C1im][NTf2]−benzene are observed at 1149 and 1332 cm−1 in [Py4][BF4], and 1124 3796
3734 mixture. It is worth mentioning that benzene−ionic liquid and 1299 cm−1 in [Py4][SCN], respectively. The vibrational 3797
3735 mixtures have been studied more often by OKE spectrosco- frequencies of νs(SO) and νas(SO) in pure liquid SO2 are 1144 3798
3736 py.205,453−457 and 1336 cm−1, respectively.42 Although Zeng et al.464 did not 3799
3737 Vibrational spectra of different solutes have been used to address the issue of vibrational frequency shift between pure 3800
3738 probe the solvent ability of ionic liquids. Wülf et al.429 used liquid SO2 and SO2-ionic liquid solutions, it is worth noting 3801
3739 water as a probe of solvent polarity of imidazolium ionic liquids that νs(SO) and νa(SO) exhibit shifts to opposite directions 3802
3740 with the anions [SCN]−, [N(CN)2]−, [C2SO4]−, and [NTf2]−. when SO2 is dissolved in [Py4][BF4] and a much larger red shift 3803
3741 Their method relies on measuring the redshift of vibrational of both νs(SO) and νa(SO) in [Py4][SCN]. Variation of the 3804
3742 frequencies of stretching modes of water, whose magnitude is alkyl chain length of the pyridinium cation, while keeping the 3805
3743 then related to dielectric constant and different polarity same [BF4]− anion, has no effect on vibrational frequencies of 3806
3744 parameters (e.g., the Kamlet−Taft solvatochromic parame- SO2 modes. 3807
3745 ters).458,459 The anion increases the ionic liquid polarity in the Shang et al.465 and Huang et al.466 studied by IR 3808
3746 order: [NTf2]− < [C2SO4]− < [N(CN)2]− < [SCN]−, while the spectroscopy the SO2 absorption in ionic liquids containing 3809
3747 length of the alkyl chain of imidazolium cations has negligible tetramethylguanidinium cations and different anions. Huang et 3810
3748 effect on polarity of the investigated ionic liquids. al.446 claimed there is no chemical absorption but only physical 3811
3749 Other molecules besides water have been used to probe ionic absorption when SO2 is absorbed by [TMGH][BF4] or 3812
3750 liquid polarity using vibrational spectroscopy. Tao et al.460 [TMGH][NTf2], as the only spectral feature is occurrence of 3813

AO DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3814 two νa(SO) modes, 1376 and 1360 cm−1, observed in both hydrogen atom of the imidazolium ring, leading to the 3877
3815 ionic liquids. Shang et al.,465 dealing with the same [TMGH]+ formation of a carbene complex with CO2, 1-butyl-3- 3878
3816 cation, but the imidazolate, phenolate, and 2,2,2-trifluoroeta- methylimidazolium-2-carboxylate, was verified through IR and 3879
3817 noate anions, argued that new bands at ca. 1410 and 954 cm−1 Raman spectroscopies.32,162 Formation of such carboxylate 3880
3818 are due to S−O and S−O−H groups, respectively. The latter species leads to new bands in both the IR (ca. 792, 1323, and 3881
3819 band was also observed when SO2 was dissolved in another 1665 cm−1) and Raman (ca. 794, 1323, and 1672 cm−1) 3882
3820 imidazolate ionic liquid with the 1-(N,N-diethylaminoethyl)-3- spectra. Figure 44 taken from the work of Cabaço et al.162 3883 f44
3821 methylimidazolium cation,467 so that the authors claimed that
3822 the mechanism of SO2 uptake in these ionic liquids involves
3823 chemical absorption. Ando et al.468 and Siqueira et al.469
3824 studied both the high- and the low-frequency spectral features
3825 of Raman spectra of [C4C1im]Br after SO2 absorption. Solid
3826 [C4C1im]Br at room temperature melts upon absorption of
3827 SO2. When SO2 is absorbed by [C4C1im]Br, the relative
3828 intensities of Raman bands at ca. 600 and 620 cm−1, which
3829 characterize gauche and anti conformers (see Figure 20),
3830 become more similar to that found in [C4C1im]I. Furthermore,
3831 vibrational frequency shifts of νs(SO) and νa(SO) modes were
3832 attributed to specific charge transfer interactions between SO2
3833 and Br−. The proposed physical picture of shielding the
3834 cation−anion interactions after uptake of SO2 was supported by
3835 molecular dynamics simulations.468,469 The authors found good
3836 agreement between the density of states calculated by the
3837 Fourier transform of the autocorrelation function of velocity
Figure 44. Comparison of Raman (top) and IR (bottom) spectra of
3838 (see section 3) and the low-frequency Raman spectra at 100 K. [C4C1im][CH3COO] and its mixtures with CO2 (red) and 13CO2
3839 Kazarian et al.470 obtained ATR spectra of subcritical CO2 (blue) (mole fraction less than ca. 0.3). The arrows pinpoint the three
3840 dissolved in [C4C1im][BF4] and [C4C1im][PF6] at 40 °C and new bands assigned to 1-butyl-3-methylimidazolium-2-carboxylate.
3841 6.8 MPa. Splitting of the IR band belonging to the bending Reproduced with permission from ref 162. Copyright 2012 American
3842 mode of CO2 has been found. Such splitting of the band is due Chemical Society.
3843 to the lifting of degeneracy, with the expectation that the
3844 splitting would be larger for solvents with more pronounced indicates by arrows these IR and Raman appearing in the 3884
3845 Lewis base character. Since the splitting was more pronounced spectra of [C4C1im][CH3COO] containing CO2 or 13CO2. 3885
3846 in [C4C1im][BF4] than [C4C1im][PF6], Kazarian et al.470 Cabaço et al.162 draws attention to the fact that, at molar 3886
3847 concluded that [BF4]− is a stronger Lewis base than [PF6]−. fractions smaller than 0.3 (or CO2 pressures of the order of 6 3887
3848 Andanson et al.471 characterized CO2 solutions in [C4C1im]- MPa), the spectra do not exhibit the Fermi dyad [i.e., the 3888
3849 [PF6] under different CO2 pressure at 40 °C using ATR overtone of the CO2 bending mode in Fermi resonance with 3889
3850 spectroscopy. The authors were able to estimate the swelling of the νs(CO)].475 This indicates that the CO2 symmetry has been 3890
3851 the ionic liquid with increasing CO2 pressure, the CO2 diffusion altered in [C4C1im][CH3COO], whereas the Fermi dyad 3891
3852 within the ionic liquid, and solubility in [C4C1im][PF6]. The appears when CO2 is dissolved in ionic liquids containing 3892
3853 vibrational spectrum of [C4C1im][PF6] suggested only minor anions other than [CH3COO]−. Besides [CH3COO]− based 3893
3854 structural modifications after CO2 uptake. The most significant ionic liquids, there are reports of other anions with 3894
3855 effects of CO2 absorption on the IR spectrum were the change carboxylate474,476 or phenolate477 groups which also favor 3895
3856 in relative intensities of bands at 600 and 625 cm−1, which mechanism of chemical absorption of CO2. Furthermore, the 3896
3857 characterize the relative proportion of gauche and anti vibrational dynamics of CO2 in ionic liquids has also been 3897
3858 [C4C1im]+ conformers and frequency shift of the νs(PF) studied by time-resolved IR spectroscopy.478 3898
3859 mode of [PF6]−.471 Seki et al.472 studied supercritical CO2 Other important contexts for ionic liquid applications include 3899
3860 dissolved in [C4C1im]+ based ionic liquids with [PF6]−, [BF4]−, electrochemistry and catalysis, in which they can be used as 3900
3861 and [NTf2]−, and [Py4][BF4], using ATR spectroscopy at 50 solvents for electrodeposition (or electroplating) of materials, 3901
3862 °C and 12 MPa. In contrast to Kazarian et al.,470 Seki et al.472 electrolytes for batteries, etc. Dissolution of precursors or 3902
3863 have not found splitting of the CO2 bending mode in the CO2 catalysts (e.g., TaCl5, AlCl3, and NbCl5) or ionic species (e.g., 3903
3864 solution in [C4C1im][BF4], arguing that the doublet convolutes Li + and Na + ) may be studied using IR and Raman 3904
3865 into a single band under higher CO2 pressure. In the case of spectroscopies for a deeper understanding of solvation and 3905
3866 CO2 solution in [Py4][BF4], no frequency shift of [Py4]+ modes structural modification of the ionic liquid upon dissolution of 3906
3867 was observed, whereas in [C4C1im][BF4] there were frequency small ions. In analogy to the previously discussed AlCl3 3907
3868 shifts of the cation νs(C(4,5)−H) mode and the anion νs(BF) mixtures with imidazolium- and pyrrolidinium-based ionic 3908
3869 mode. Seki et al.472 claimed that CO2 interacts mainly with the liquids (see section 4.1), in which speciation of aluminum as 3909
3870 [BF4]− and that the new species formed CO2-[BF4]− is a [AlCl4]− or [Al2Cl7]− was inferred by IR and Raman 3910
3871 stronger base than the single [BF4]−. spectroscopies,96,217,479 spectroscopic studies have been carried 3911
3872 Vibrational spectroscopy has been extensively used to study out for other metals, for instance, tantalum and niobium, whose 3912
3873 CO2 absorption in imidazolium ionic liquids with the deposition process cannot be done in water.480,481 Babushki- 3913
3874 [CH3COO]− anion.32,162,473,474,462 The mechanism of chemical na480 studied TaCl5-[Py1,4]Cl mixtures using IR spectroscopy 3914
3875 absorption of CO2 in [C4C1im][CH3COO] based on the over a wide range of compositions at room temperature. It has 3915
3876 reaction between [CH3COO]− and the C(2)−H acidic been found that all of the vibrational modes of [Py1,4]+ are 3916

AP DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

3917 disturbed upon dissolution of the tantalum salt, the ring related to anion modes, but Forsyth et al.491 observed also the 3980
3918 breathing vibrations within 890−930 cm−1 being the most occurrence of shoulders at 534, 576, 598 651, and 1150 cm−1 in 3981
3919 affected modes exhibiting change on relative intensities and IR bands, the latter being assigned to the twisting mode of the 3982
3920 frequency shift depending on the TaCl5 concentration.480 pyrrolidinium ring. 3983
3921 Using Raman spectroscopy, Alves et al.481 studied NbCl5- Lithium cation solutions in ionic liquids have been widely 3984
3922 [C4C1im]Cl and ZnCl2-[C4C1im]Cl over a wide range of studied because of the application as electrolyte in lithium 3985
3923 compositions at room temperature. Besides speciation of batteries. Ionic liquids based on the [NTf2]− anion show 3986
3924 niobium ([NbCl6]− and Nb2Cl10) and zinc ([ZnCl4]2−, remarkable structural and dynamical changes when mixed with 3987
3925 [Zn3Cl8]2−, and [Zn4Cl10]2−), these authors reported several the lithium salt. For example, it was shown by NMR 3988
3926 changes in the Raman spectra of the mixtures in comparison measurements492 that in Li[NTf2]−[C4C1im][NTf2] mixtures 3989
3927 with the neat ionic liquid spectrum, in particular, in the C−H the Li+ self-diffusion coefficient drops by almost a third when 3990
3928 modes of the imidazolium481 as it has also been found by the concentration ranges from 2% to 22% of Li+. Nicolau et 3991
3929 Goujon et al.482 in water-ZnCl2 or water-MgCl2 mixtures with al.493 using OKE spectroscopy estimated that the viscosity of 3992
3930 [C8C1im]Cl. Andriyko et al.483 studied TiF4-[C4C1C1im][BF4] [C 4 C 1 im][NTf2 ] increases by almost ten times when 3993
3931 using IR spectroscopy over a wide range of concentrations. The approximately 40% mole fraction of Li+ is added. The overall 3994
3932 authors were able to observe spectral features, indicating the observed trend for different systems which have been studied is 3995
3933 formation of [TiF6]2− and a heteronuclear complex between that increasing Li+ content results in increase of viscosity and 3996
3934 TiF4 and the anion, [TiBF8]−. They also found decreasing density and decrease of self-diffusion coefficient and con- 3997
3935 intensity of anion bands because of a side reaction leading to ductivity.494−497 These effects on transport coefficients upon 3998
3936 BF3 evolution and formation of [TiF6]2−. addition of Li+ are accompanied by conformational changes of 3999
3937 Arellano et al.484 studied Zn[NTf2]2-[C2C1im][NTf2] the [NTf2]− anion as revealed by vibrational spectroscopy. 4000
3938 mixtures using IR spectroscopy with the support of quantum Lassègues et al.498,499 and Duluard et al.500 used Raman 4001
3939 chemical calculations. These authors claim that a new anionic spectroscopy to study mixtures of [C2C1im][NTf2] and 4002
3940 species, [Zn(NTf2)3]−, is formed upon the addition of a [C4C1im][NTf2] with Li[NTf2] over a wide range of 4003
3941 stoichiometric amount of the solute. Formation of [Zn- concentration. The authors deconvoluted the anion Raman 4004
3942 (NTf 2 ) 3 ] − results in frequency shift of [NTf 2 ] − and band with the maximum at ∼745 cm−1 into two components, 4005
3943 [C2C1im]+ modes, in particular the ν(C(2)−H) mode. Curve the lower wavenumber component being assigned to 4006
3944 fit of spectra showed significant changes, such as the number of uncoordinated (“free”) [NTf2]− and the higher wavenumber 4007
3945 components in the spectral range of 700−820 cm−1, between component to coordinated (“bounded”) [NTf2]− to Li+. These 4008
3946 spectra of mixture and the neat ionic liquid, but the authors did
authors500,498,499 evaluated the coordination number of Li+ 4009
3947 not address this issue.484 In line with the zinc coordination
equal to two and then inferring for the formation of the 4010
3948 compound, there are similar reports of stable compounds
[Li(NTf 2 ) 2 ] − species in line with the proposition of
formed between ytterbium485 and europium486 with [NTf2]−.
4011
3949
Umebayashi et al.501−503 According to Lassègues et al.498 and
Liu et al.487 studied 0.2 mol L−1 solutions of Zn[CF3SO3]2 in
4012
3950
Hardwick et al.,504 such complexes are weakly bounded, being 4013
3951 [C1im][CF3SO3], [C2C1im][CF3SO3], and [C2C1C1im]-
3952 [CF3SO3] at 393 K, using Raman spectroscopy. The authors promptly destabilized in the presence of other highly 4014

3953 considered the δs(CF3) in order to infer about the coordination coordinating solvents (e.g., diglyme, tetraglyme, and ethylene 4015

3954 of Zn2+ ions and the changes on local environment upon carbonate). As pointed out by Martins et al.,505 the formation 4016

3955 addition of the solute. The authors found that addition of Zn2+ of [Li(NTf2)2]− is also influenced by water content. The 4017

3956 causes a higher wavenumber shoulder in the δs(CF3) Raman splitting of the 745 cm−1 band depends on the Li+ molar 4018

3957 band due to those anions coordinating the Zn2+, while the fraction, and a pseudoisosbestic point is observed in the Raman 4019

3958 lower wavenumber band is assigned to uncoordinated (“free”) spectra.501 Furthermore, Umebayashi et al.506 showed that in 4020

3959 anions.487 Methylation of the imidazolium ring (i.e., from the series of alkali cations, from Li+ to Cs+, the coordination 4021

3960 [C1im]+ to [C2C1C1im]+), implies larger splitting of vibrational number grows from two to four. Monteiro et al.496 argued that 4022

3961 frequencies between free and bounded anions. The authors the formation of [Li(NTf2)2]− is the main reason for the 4023

3962 estimated the coordination number of anions around Zn2+ from decrease of Li+ ionic mobility and conductivity in Li[NTf2]− 4024
3963 the ratio between the band areas of free and bounded anions. In [C4C1C1im][NTf2] mixtures. 4025

3964 the case of ionic liquid based on [C1im]+, [C2C1im]+, and Duluard et al.492 studying [C4C1im][NTf2] mixtures with Li+ 4026
3965 [C2C1C1im]+, the most favored species are [Zn(CF3SO3)3]−, pointed out that the conformation of the butyl chain remains 4027
3966 [Zn(CF3SO3)4]2−, and [Zn(CF3SO3)5]3−, respectively.487 essentially unchanged upon dissolution of Li+. Umebayashi et 4028
3967 Oliveira et al.488 using Raman spectroscopy studied the al.503 and Lassègues et al.499 addressed the issue of [NTf2]− 4029
3968 dissolution of [NH4]+, with [CH3COO]−, Cl−, [SCN]−, and conformation around the Li+ ion, analyzing the fingerprint 4030
3969 [C2SO3]− as counterions, and Na+, with [CH3COO]− and range of the Raman spectrum of the [NTf2]− anion with the aid 4031
3970 [SCN]− as counterions, in [C2C1im][CH3COO]. The authors of quantum chemistry calculations at the DFT/B3LYP level of 4032
3971 showed that the [CH3COO]− mode at 910 cm−1 is a good theory. In both of these papers, the authors concluded that the 4033
3972 probe of the local environment as the corresponding Raman cisoid conformation is favored in the 1:1 (Li+:[NTf2]−) 4034
3973 band is sensitive to the dissolved species. Chimdi et al.489 complex, in contrast to the situation in the neat ionic liquid, 4035
3974 studied Na[N(CN) 2]−[Pyr1,1 ][N(CN) 2] mixtures using in which the transoid conformation dominates. In the case of 4036
3975 Raman spectroscopy, whereas Carstens et al.490 studied 1:2 complex, the authors498,503 suggest that the complex 4037
3976 Na[N(SO2F)2]−[Py1,4][N(SO2F)2] and Forsyth et al.491 structures involves the two anions in cisoid or transoid, or a mix 4038
3977 studied Na[NTf2]−[Pyr1,2][NTf2] mixtures using Raman and of cisoid−transoid conformations. Lassègues et al.499 claim that 4039
3978 ATR spectroscopies. Chimdi et al.489 and Carstens et al.490 the mix of [NTf2]− conformers gives better agreement to the 4040
3979 reported only frequency and intensity changes of Raman bands fingerprint region of the Raman spectra. 4041

AQ DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4042 The above-mentioned effects of Li+ dissolution on the approximation may be needed for proper assignment and for 4082
f45 4043 Raman spectrum of [C4C1im][NTf2] is illustrated in Figure 45. including important anharmonicity effects on vibrational 4083
spectra of ionic liquids (e.g., Fermi resonance). Strong ionic 4084
interactions may imply that an ab initio calculation of 4085
vibrational frequencies for an isolated ion eventually only 4086
estimates the frequencies when compared with the actual 4087
spectra of the liquid phase. Fortunately, the computational 4088
resources available today allow for quantum chemistry 4089
calculations of vibrational frequencies for clusters made of a 4090
few ion pairs. However, the approach of cluster calculation still 4091
relies on an optimized geometry of minimum energy. Thus, ab 4092
initio molecular dynamics simulations of ionic liquids open the 4093
perspective for vibrational frequency calculations with the 4094
proper account of liquid dynamics. On the experimental point 4095
of view, linear IR and Raman spectroscopies of ionic liquids, 4096
which were the focus of this review, are being recently extended 4097
for time-resolved vibrational spectroscopy. Structural fluctua- 4098

Figure 45. Comparison between Raman spectra of pure [C4C1im]- tions of the local environment experienced by a probe oscillator 4099
[NTf2] (red line), solid Li[NTf2] (black line), and the mixture cause vibrational dephasing due to energy relaxation and loss of 4100
Li[NTf2]−[C4C1im][NTf2]) at molar fraction of Li[NTf2] equal to phase (pure dephasing). The traditional approach using linear 4101
0.37 (blue line). Spectra have been normalized by the most intense IR and Raman spectroscopies by Fourier transforming the band 4102
band at each spectral window. shape in order to get time correlation functions of vibrational 4103
dephasing and reorientational dynamics74,83,85,92 has been 4104
4044 This figure shows Raman spectra of pure [C4C1im][NTf2], applied for the CN stretching mode of ionic liquids containing 4105
4045 solid Li[NTf2], and the mixture, in the range of 250−450 cm−1, cyano-anions since the corresponding Raman band is relatively 4106
4046 which characterizes the [NTf2]− conformation and the 741 free of overlapping bands.106 However, this methodology is not 4107
4047 cm−1 band, whose frequency shift and split characterizes fully appropriate when processes of multiple time ranges 4108
4048 coordinated and uncoordinated [NTf2]−. Despite identifying simultaneously contribute with homogeneous and inhomoge- 4109
4049 more than one possible arrangement for the [Li(NTf2)2]− neous mechanisms for the band shape. Thus, perspectives for 4110
4050 complex,499,503 no more than two components were assumed vibrational spectroscopy of ionic liquids include ongoing 4111
4051 under the band at ∼745 cm−1. In contrast, Pitawala et al.495,507 studies by time-resolved techniques, such as time-resolved IR 4112
4052 and Watkins et al.,508 the latter studying Mg[NTf2]2, accounted spectroscopy and coherent anti-Stokes Raman scattering 4113
4053 for three or more components while fitting this band with the (CARS),6−11 providing insights on the short-time molecular 4114
4054 support of previously reported quantum chemistry calcula- dynamics. For instance, time-resolved CARS measurements 4115
4055 tions.500,499,501−503 Pitawala et al.507 considered three bands in give dephasing times in the subpicosecond range for C−H 4116
4056 the fit of the 741 cm−1 Raman band in Li[NTf2]−[Pyr1,4]- stretching modes of 1-alkyl-3-methylimidazolium cations that 4117
4057 [NTf2] mixtures and reached the same coordination number as can be related to flexibility of molecular structure and the 4118
4058 previous works for high molar fractions of Li+. However, for strength of different sites for hydrogen bonding to the anion.10 4119
4059 low concentrations of Li+ (equal or below 0.05), the authors Time-resolved IR spectroscopy has also been used to follow the 4120
4060 obtained a coordination number as high as four. Change in time evolution of modes belonging to molecule probes 4121
4061 coordination number might have consequences in transport dissolved in ionic liquids. This allows addressing the molecular 4122
4062 coefficients, glass transition, and melting temperatures of the dynamics at the different time scales of fluctuations of the 4123
4063 mixtures.114,495−497,507 hydrogen bond structure and reorientational motions and the 4124
4064 It is worth stressing that only part of the literature relative contributions of homogeneous and inhomogeneous 4125
4065 concerning ionic liquid solutions was reviewed in this section broadening to the band shape.517,518 Thus, time-resolved 4126
4066 proper to the broadness of the theme and the large number of vibrational spectroscopy is expected to provide experimental 4127
4067 examples in which vibrational spectroscopy served only as a data related to short-time molecular dynamics allowing for 4128
4068 complementary tool to better understand some process (e.g., direct comparison to results of computer simulations of ionic 4129
4069 cellulose and carbohydrates dissolution) or for the character- liquids. Nevertheless, there is plenty of room for applications of 4130
4070 ization of new materials. It is also worth mentioning solvate linear IR and Raman spectroscopies in many areas related to 4131
4071 ionic liquids,509−516 which are systems composed from weakly the ones discussed in this review. Besides the application on 4132
4072 coordinating anions such as [NTf2]−, small cations (e.g., Li+) ionic liquid solutions discussed in this review, vibrational 4133
4073 and a chelating molecular solvent such as tetraglyme,512 not spectroscopy is a complementary technique within the broad 4134
4074 discussed in this review. Such systems are of increasing interest field of research on hybrid materials. Vibrational spectroscopy 4135
4075 as their properties might extend the range of applications of has been used in combination with other techniques to 4136
4076 ionic liquids. characterize ionic liquid interactions with nanotube and 4137
graphene,519−524 clays,525−529 polymers,530,492,531−539 carbohy- 4138
8. CONCLUDING REMARKS drates,540−544 etc. Furthermore, surface-enhanced Raman 4139
4077 In this review, we discussed the application of vibrational scattering (SERS) studies have been reported for ionic liquids 4140
4078 spectroscopy for getting information on structure and with different metals as substrates.545−549 Within the context of 4141
4079 intermolecular interactions in ionic liquids and the related studies on phase transitions, vibrational spectra of ionic liquids 4142
4080 experimental and theoretical issues. Quantum chemistry have been reported as a function of temperature, at atmospheric 4143
4081 calculations of vibrational frequencies beyond the harmonic pressure, or as a function of pressure, at room temperature. 4144

AR DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4145 Simultaneous variation of temperature and pressure aiming ionic liquids. He also spent shorter periods working in the groups of 4202
4146 more complete mapping of the phase diagrams most probably Prof. Giancarlo Ruocco in 2007 (Universitá di Roma La Sapienza) and 4203
4147 will be the subject of further vibrational spectroscopy studies of ́ A. H. Pádua in 2013 (Université Blaise Pascal, Clermont-
Prof. Agilio 4204
4148 ionic liquids. Ferrand, France). His research concerns structure and dynamics of 4205
ionic liquids with particular focus on changes observed along the glass 4206
4149 ASSOCIATED CONTENT transition and crystallization taking place under low temperature or 4207
4150 *
S Supporting Information high pressure. 4208

4151 The Supporting Information is available free of charge on the


4152 ACS Publications website at DOI: 10.1021/acs.chem- ACKNOWLEDGMENTS 4209
4153 rev.6b00461.
The authors acknowledge the Brazilian agencies CNPq and 4210
4154 IR and Raman spectra recorded in this work are available FAPESP (Grant nos. 2015/07516-8, 2015/05803-0, and 2012/ 4211
4155 as TXT files. Each file is identified by the number of the 13119-3) for fellowships and financial support. 4212
4156 figure, the ionic liquid name, and whether it corresponds
4157 to Raman or IR spectrum. The files give Raman and IR
4158 spectra covering the same spectral window as shown in REFERENCES 4213

4159 the corresponding figure of the paper. The file (1) Papatheodorou, G. N.; Yannopoulos, S. N. Light Scattering from 4214
4160 DAC_Crystallization.wmv is a movie showing the real Molten Salts: Structure and Dynamics. In Molten Salts: From 4215
4161 time crystallization process of [C2C1im][NTf2] inside Fundamentals to Applications; Gaune-Escard, M., Ed. Springer: 4216

4162 the diamond anvil cell at 0.7 GPa and room temperature Dordrecht, The Netherlands, 2002; pp 47−106. 4217
(2) Papatheodorou, G. N.; Kalampounias, A. G.; Yannopoulos, S. N. 4218
4163 (ZIP) Raman Spectroscopy of High Temperature Melts. In Molten Salts and 4219
Ionic Liquids. Never the Twain?, Gaune-Escard, M., Seddon, K. R., Eds.; 4220
4164 AUTHOR INFORMATION John Wiley & Sons: Hoboken, NJ, 2010; pp 301−340. 4221
4165 Corresponding Author (3) Berg, R. W. Raman Spectroscopy and Ab-initio Model 4222
Calculations on Ionic Liquids. Monatsh. Chem. 2007, 138, 1045−1075. 4223
4166 *E-mail: mccribei@iq.usp.br. (4) Saha, S.; Hiroi, T.; Iwata, K.; Hamaguchi, H. Raman spectroscopy 4224
4167 ORCID and the heterogeneous liquid structure In Ionic Liquids Completely 4225

4168 Vitor H. Paschoal: 0000-0002-0935-3772 UnCOILed: Critical Expert Overviews; Plechkova, N., Seddon, K. R., 4226
Eds.; Wiley: Hoboken, NJ, 2015; p 165. 4227
4169 Luiz F. O. Faria: 0000-0002-0711-4604 (5) Penalber-Johnstone, C.; Baldelli, S. Vibrational spectroscopy of 4228
4170 Mauro C. C. Ribeiro: 0000-0002-4301-5021 ionic liquids surface. In Ionic Liquids Completely UnCOILed: Critical 4229
4171 Notes Expert Overviews; Plechkova, N., Seddon, K. R., Eds.; Wiley: Hoboken, 4230
NJ, 2015; p 145. 4231
4172 The authors declare no competing financial interest. (6) Dahl, K.; Sando, G. M.; Fox, D. M.; Sutto, T. E.; Owrutsky, J. C. 4232
4173 Biographies Vibrational Spectroscopy and Dynamics of Small Anions in Ionic 4233
Liquid Solutions. J. Chem. Phys. 2005, 123, 084504. 4234
4174 Vitor Hugo Paschoal received his degree in chemistry from the State (7) Shigeto, S.; Hamaguchi, H. Evidence for Mesoscopic Local 4235
4175 University of Londrina (Brazil) in 2013 starting his Ph.D. studies Structures in Ionic Liquids: CARS Signal Spatial Distribution of 4236
4176 under the supervision of Prof. Mauro C. C. Ribeiro in early 2014. His C(n)mim PF6 (n = 4,6,8). Chem. Phys. Lett. 2006, 427, 329−332. 4237
4177 research interests are molecular dynamics simulations and spectros- (8) Iwata, K.; Okajima, H.; Saha, S.; Hamaguchi, H. Local Structure 4238
4178 copy of high-frequency collective dynamics of liquids and their glassy Formation in Alkyl-Imidazolium-Based Ionic Liquids as Revealed by 4239
4179 phases (especially ionic liquids and solutions). Linear and Nonlinear Raman Spectroscopy. Acc. Chem. Res. 2007, 40, 4240
1174−1181. 4241
4180 Luiz Felipe de Oliveira Faria received his degree in chemistry from (9) Namboodiri, M.; Kazemi, M. M.; Zeb Khan, T.; Materny, A.; 4242
4181 Federal University of Juiz de Fora in 2010 and his Ph.D. degree from Kiefer, J. Ultrafast Vibrational Dynamics and Energy Transfer in 4243
4182 University of São Paulo (Brazil) in 2015. He completed his Ph.D. Imidazolium Ionic Liquids. J. Am. Chem. Soc. 2014, 136, 6136−6141. 4244
4183 under the supervision of Prof. Dr. Mauro C. C. Ribeiro on the (10) Chatzipapadopoulos, S.; Zentel, T.; Ludwig, R.; Luetgens, M.; 4245
4184 structure and phase transitions of ionic liquids in different conditions Lochbrunner, S.; Kühn, O. Vibrational Dephasing in Ionic Liquids as a 4246
4185 of temperature and pressure. He is currently a postdoc in Prof. Mauro Signature of Hydrogen Bonding. ChemPhysChem 2015, 16, 2519− 4247
4186 C. C. Ribeiro group, and his research has been focused on the 2523. 4248
4187 structure and phase transitions of ionic liquids under high pressure (11) Kiefer, J.; Namboodiri, M.; Kazemi, M. M.; Materny, A. Time- 4249

4188 using Raman spectroscopy and X-ray scattering techniques. Resolved Femtosecond CARS of the Ionic Liquid 1-Ethyl-3- 4250
methylimidazolium Ethylsulfate. J. Raman Spectrosc. 2015, 46, 722− 4251
4189 Mauro Carlos Costa Ribeiro is an Associate Professor at the Chemistry 726. 4252
4190 Institute of the Universidade de São Paulo, IQ-USP. Mauro obtained (12) Stuchebryukov, S. D.; Rudoy, V. M. Attenuated total reflection 4253
4191 his Bachelor degree in Chemistry in 1989 from Universidade Santa spectra under conditions of weak absorption: Physical nature. Opt. 4254
4192 ́ (Santos-SP). He obtained his Master’s degree in 1992 and his
Cecilia Commun. 1997, 140, 36−40. 4255
4193 Ph.D. degree in 1995, both from Universidade de São Paulo under (13) Bertie, J. E.; Michaelian, K. H. Comparison of Infrared And 4256
4194 supervision of Prof. Paulo S. Santos. His Master’s studies concerned Raman Wave Numbers of Neat Molecular Liquids: Which is the 4257
Correct Infrared Wave Number to Use? J. Chem. Phys. 1998, 109, 4258
4195 calculations of resonant Raman spectra and his Ph.D. thesis on Raman
6764−6771. 4259
4196 spectroscopy and molecular dynamics of liquids. He started teaching at
(14) Hancer, M.; Sperline, R. P.; Miller, J. D. Anomalous Dispersion 4260
4197 IQ-USP in 1996. Mauro spent one and a half years (1996−1998) as a Effects in the IR-ATR Spectroscopy of Water. Appl. Spectrosc. 2000, 4261
4198 postdoc in the group of Prof. Paul A. Madden at Oxford University, 54, 138−143. 4262
4199 U.K., working with molecular dynamics of high-temperature molten (15) Buffeteau, T.; Grondin, J.; Lassègues, J.-C. Infrared Spectros- 4263
4200 salts. After returning to São Paulo, he started working with vibrational copy of Ionic Liquids: Quantitative Aspects and Determination of 4264
4201 spectroscopy and molecular dynamics simulation of room-temperature Optical Constants. Appl. Spectrosc. 2010, 64, 112−119. 4265

AS DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4266 (16) Mirabella, F. M. Principles, Theory and Practice of Internal Methylimidazolium Acetate and Water: Insights from IR, Raman, 4335
4267 Reflection Spectroscopy. In Handbook of Vibrational Spectroscopy, NMR Spectroscopy and Quantum Chemistry Calculations. J. Mol. Liq. 4336
4268 Chalmers, J. M., Griffiths, P. R., Eds.; John Wiley & Sons: Chichester, 2015, 210, 227−237. 4337
4269 U.K., 2002; Vol. 2, pp 1091−1102. (35) Bertie, J. E.; Zhang, S. L.; Keefe, C. D. Measurement and use of 4338
4270 (17) Fitzpatrick, J.; Reffner, J. A. Macro and Micro Internal Absolute Infrared-Absorption Intensities of Neat Liquids. Vib. 4339
4271 Reflection Accessories. In Handbook of Vibrational Spectroscopy; Spectrosc. 1995, 8, 215−229. 4340
4272 Chalmers, J. M., Griffiths, P. R., Eds.; John Wiley & Sons: Chichester, (36) Bassett, W. A. Diamond Anvil Cell, 50th Birthday. High Pressure 4341
4273 UK, 2002; pp 1103−1116. Res. 2009, 29, CP5−163. 4342
4274 (18) Bertie, J. E. Optical Constants. In Handbook of Vibrational (37) Polsky, C. H.; Valkenburg, E. V. The Diamond Anvil Cell. In 4343
4275 Spectroscopy; Chalmers, J. M., Griffiths, P. R., Eds.; John Wiley & Sons: Handbook of Vibrational Spectroscopy; Chalmers, J. M., Griffiths, P. R., 4344
4276 Chichester, UK, 2002; Vol. 1, pp 88−100. Eds.; John Wiley & Sons: Chichester, UK, 2002; Vol. 2, pp 1352− 4345
4277 (19) Bertie, J. E.; Eysel, H. H. Infrared Intensities of Liquids 0.1. 1360. 4346
4278 Determination of Infrared Optical and Dielectric-Constants by FT-IR (38) Forman, R. A.; Block, S.; Barnett, J. D.; Piermarini, G. J. 4347
4279 Using the Circle ATR Cell. Appl. Spectrosc. 1985, 39, 392−401. Pressure Measurement Made by Utilization of Ruby Sharp-Line 4348
4280 (20) Burba, C. M.; Frech, R. Existence of Optical Phonons in the Luminescence. Science 1972, 176 (4032), 284−285. 4349
4281 Room Temperature Ionic Liquid 1-ethyl-3-methylimidazolium tri- (39) Syassen, K. Ruby Under Pressure. High Pressure Res. 2008, 28, 4350
4282 fluoromethanesulfonate. J. Chem. Phys. 2011, 134, 134503. 75−126. 4351
4283 (21) Burba, C. M.; Janzen, J.; Butson, E. D.; Coltrain, G. L. Using (40) Faria, L. F. O.; Nobrega, M. M.; Temperini, M. L. A.; Ribeiro, 4352
4284 FT-IR Spectroscopy to Measure Charge Organization in Ionic Liquids. M. C. C. Ionic Liquids Based on the Bis(Trifluoromethylsulfonyl)- 4353
4285 J. Phys. Chem. B 2013, 117, 8814−8820. Imide Anion for High-Pressure Raman Spectroscopy Measurements. J. 4354
4286 (22) Burba, C. M.; Janzen, J.; Butson, E. D.; Coltrain, G. L. Raman Spectrosc. 2013, 44, 481−484. 4355
4287 Correction to ″Using FT-IR Spectroscopy to Measure Charge (41) Wilson, E. B., Jr.; Decius, J. C.; Cross, P. C. Molecular Vibrations: 4356
4288 Organization in Ionic Liquids″. J. Phys. Chem. B 2016, 120, 3591− The Theory of Infrared and Raman Vibrational Spectra; McGraw-Hill: 4357
4289 3592. New York, 1955. 4358
4290 (23) Decius, J. C. Dipolar Coupling and Molecular Vibration in (42) Herzberg, G. Molecular Spectra and Molecular Structure. II. 4359
4291 Crystals. I. General Theory. J. Chem. Phys. 1968, 49, 1387−1392. Infrared and Raman Spectra of Polyatomic Molecules; D. van Nostrand 4360
4292 (24) Frech, R.; Decius, J. C. Dipolar Coupling and Molecular Co.: Princeton, NJ, 1945; Vol. 2. 4361
4293 Vibrations in Crystals. IV. Frequency Shifts and Dipole Moment (43) Baddiel, C. B.; Janz, G. J. Molten Thiocynates: Raman Spectra 4362
4294 Derivatives. J. Chem. Phys. 1971, 54, 2374−2379. and Structure. Trans. Faraday Soc. 1964, 60 (503P), 2009−2012. 4363
4295 (25) Huddleston, J. G.; Visser, A. E.; Reichert, W. M.; Willauer, H. (44) Izgorodina, E. I.; Bernard, U. L.; MacFarlane, D. R. Ion-Pair 4364
4296 D.; Broker, G. A.; Rogers, R. D. Characterization and Comparison of Binding Energies of Ionic Liquids: Can DFT Compete with Ab Initio- 4365
4297 Hydrophilic and Hydrophobic Room Temperature Ionic Liquids Based Methods? J. Phys. Chem. A 2009, 113, 7064−7072. 4366
4298 Incorporating the Imidazolium Cation. Green Chem. 2001, 3, 156− (45) Rigby, J.; Izgorodina, E. I. New SCS- and SOS-MP2 Coefficients 4367
4299 164. Fitted to Semi-Coulombic Systems. J. Chem. Theory Comput. 2014, 10, 4368
4300 (26) Jacquemin, J.; Husson, P.; Padua, A. A. H.; Majer, V. Density 3111−3122. 4369
4301 and Viscosity of Several Pure and Water-Saturated Ionic Liquids. Green (46) Garcia, G.; Atilhan, M.; Aparicio, S. Assessment of DFT 4370
4302 Chem. 2006, 8, 172−180. Methods for Studying Acid Gas Capture by Ionic Liquids. Phys. Chem. 4371
4303 (27) Andanson, J. M.; Meng, X.; Traikia, M.; Husson, P. Chem. Phys. 2015, 17, 26875−26891. 4372
4304 Quantification of the Impact of Water as an Impurity on Standard (47) Berg, R. W.; Deetlefs, M.; Seddon, K. R.; Shim, I.; Thompson, J. 4373
4305 Physico-Chemical Properties of Ionic Liquids. J. Chem. Thermodyn. M. Raman and ab initio Studies of Simple and Binary 1-Alkyl-3- 4374
4306 2016, 94, 169−176. methylimidazolium Ionic Liquids. J. Phys. Chem. B 2005, 109, 19018− 4375
4307 (28) Fadeeva, T. A.; Husson, P.; DeVine, J. A.; Gomes, M. F. C.; 19025. 4376
4308 Greenbaum, S. G.; Castner, E. W., Jr. Interactions Between Water and (48) Katsyuba, S. A.; Zvereva, E. E.; Vidis, A.; Dyson, P. J. 4377
4309 1-Butyl-1-methylpyrrolidinium Ionic Liquids. J. Chem. Phys. 2015, 143, Application of Density Functional Theory and Vibrational Spectros- 4378
4310 064503. copy Toward the Rational Design of Ionic Liquids. J. Phys. Chem. A 4379
4311 (29) Cammarata, L.; Kazarian, S. G.; Salter, P. A.; Welton, T. 2007, 111, 352−370. 4380
4312 Molecular States of Water in Room Temperature Ionic Liquids. Phys. (49) Umebayashi, Y.; Fujimori, T.; Sukizaki, T.; Asada, M.; Fujii, K.; 4381
4313 Chem. Chem. Phys. 2001, 3, 5192−5200. Kanzaki, R.; Ishiguro, S. Evidence of Conformational Equilibrium of 1- 4382
4314 (30) Jeon, Y.; Sung, J.; Kim, D.; Seo, C.; Cheong, H.; Ouchi, Y.; Ethyl-3-methylimidazolium in its Ionic Liquid Salts: Raman Spectro- 4383
4315 Ozawa, R.; Hamaguchi, H.-o. Structural Change of 1-Butyl-3- scopic Study and Quantum Chemical Calculations. J. Phys. Chem. A 4384
4316 methylimidazolium Tetrafluoroborate + Water Mixtures Studied by 2005, 109, 8976−8982. 4385
4317 Infrared Vibrational Spectroscopy. J. Phys. Chem. B 2008, 112, 923− (50) Jamróz, M. H. Vibrational Energy Distribution Analysis 4386
4318 928. (VEDA): Scopes and Limitations. Spectrochim. Acta, Part A 2013, 4387
4319 (31) Bertie, J. E.; Lan, Z. D. Infrared Intensities of Liquids 0.20. The 114, 220−230. 4388
4320 Intensity of the OH Stretching Band of Liquid Water Revisited, and (51) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; 4389
4321 the Best Current Values of the Optical Constants of H2O(1) at 25 Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; 4390
4322 Degrees C between 15,000 and 1 cm(−1). Appl. Spectrosc. 1996, 50, Kudin, K. N.; Burant, J. C.; Millam, J. M.; et al. Gaussian 03; Gaussian, 4391
4323 1047−1057. Inc.: Wallingford, CT, 2004. 4392
4324 (32) Gurau, G.; Rodríguez, H.; Kelley, S. P.; Janiczek, P.; Kalb, R. S.; (52) Talaty, E. R.; Raja, S.; Storhaug, V. J.; Dölle, A.; Carper, W. R. 4393
4325 Rogers, R. D. Demonstration of Chemisorption of Carbon Dioxide in Raman and Infrared Spectra and ab Initio Calculations of C2−4MIM 4394
4326 1,3-Dialkylimidazolium Acetate Ionic Liquids. Angew. Chem., Int. Ed. Imidazolium Hexafluorophosphate Ionic Liquids. J. Phys. Chem. B 4395
4327 2011, 50, 12024−12026. 2004, 108, 13177−13184. 4396
4328 (33) Thomas, M.; Brehm, M.; Holloczki, O.; Kelemen, Z.; Nyulaszi, (53) Heimer, N. E.; Del Sesto, R. E.; Meng, Z. Z.; Wilkes, J. S.; 4397
4329 L.; Pasinszki, T.; Kirchner, B. Simulating the Vibrational Spectra of Carper, W. R. Vibrational Spectra of Imidazolium Tetrafluoroborate 4398
4330 Ionic Liquid Systems: 1-Ethyl-3-methylimidazolium acetate and its Ionic Liquids. J. Mol. Liq. 2006, 124, 84−95. 4399
4331 Mixtures. J. Chem. Phys. 2014, 141, 024510. (54) Katsyuba, S. A.; Dyson, P. J.; Vandyukova, E. E.; Chernova, A. 4400
4332 (34) Marekha, B. A.; Bria, M.; Moreau, M.; De Waele, I.; Miannay, V.; Vidis, A. Molecular Structure, Vibrational Spectra, and Hydrogen 4401
4333 F.-A.; Smortsova, Y.; Takamuku, T.; Kalugin, O. N.; Kiselev, M.; Bonding of the Ionic Liquid 1-Ethyl-3-methyl-1H-imidazolium 4402
4334 Idrissi, A. Intermolecular Interactions in Mixtures of 1-N-Butyl-3- Tetrafluoroborate. Helv. Chim. Acta 2004, 87, 2556−2565. 4403

AT DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4404 (55) Darling, B. T.; Dennison, D. M. The Water Vapor Molecule. (80) Buckingham, A. D. Solvent Effects in Vibrational Spectroscopy. 4473
4405 Phys. Rev. 1940, 57, 128−139. Trans. Faraday Soc. 1960, 56, 753−760. 4474
4406 (56) Barone, V. Anharmonic Vibrational Properties by a Fully (81) Schweizer, K. S.; Chandler, D. Vibrational Dephasing and 4475
4407 Automated Second-Order Perturbative Approach. J. Chem. Phys. 2005, Frequency-Shifts of Polyatomic-Molecules in Solution. J. Chem. Phys. 4476
4408 122, 014108. 1982, 76, 2296−2314. 4477
4409 (57) Barone, V.; Biczysko, M.; Bloino, J. Fully Anharmonic IR and (82) Wendler, K.; Brehm, M.; Malberg, F.; Kirchner, B.; Delle Site, L. 4478
4410 Raman Spectra of Medium-Size Molecular Systems: Accuracy and Short Time Dynamics of Ionic Liquids in AIMD-Based Power Spectra. 4479
4411 Interpretation. Phys. Chem. Chem. Phys. 2014, 16, 1759−1787. J. Chem. Theory Comput. 2012, 8, 1570−1579. 4480
4412 (58) Grondin, J.; Lassègues, J.-C.; Cavagnat, D.; Buffeteau, T.; (83) Gordon, R. G. Molecular Motion in Infrared and Raman 4481
4413 Johansson, P.; Holomb, R. Revisited Vibrational Assignments of Spectra. J. Chem. Phys. 1965, 43, 1307−1312. 4482
4414 Imidazolium-Based Ionic Liquids. J. Raman Spectrosc. 2011, 42, 733− (84) Shimizu, H. Time-Correlation Function of Molecular Random 4483
4415 743. Motion and Shape of Spectral Bands. J. Chem. Phys. 1965, 43, 2453− 4484
4416 (59) Lassègues, J.-C.; Grondin, J.; Cavagnat, D.; Johansson, P. New 2465. 4485
4417 Interpretation of the CH Stretching Vibrations in Imidazolium-Based (85) Rothschild, W. G. Dynamics of Molecular Liquids; John Wiley & 4486
4418 Ionic Liquids. J. Phys. Chem. A 2009, 113, 6419−6421. Sons: New York, 1984. 4487
4419 (60) Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids; (86) Madden, P. A. Simulation of Properties of Spectroscopic 4488
4420 Claredon: Oxford, 1987. Interest. In Molecular Dynamics Simulation of Statistical-Mechanical 4489
4421 (61) Buchner, M.; Ladanyi, B. M.; Stratt, R. M. The Short-Time (Proceedings of the International School of Physics); Hoover, W. G., 4490
4422 Dynamics of Molecular Liquids. Instantaneous-Normal-Mode Theory. Ciccoti, G., Eds. North-Holland: Amsterdam, 1986; pp 371−400. 4491
4423 J. Chem. Phys. 1992, 97, 8522−8535. (87) Brehm, M.; Kirchner, B. TRAVIS - A Free Analyzer and 4492
4424 (62) Stratt, R. M. The Instantaneous Normal-Modes of Liquids. Acc. Visualizer for Monte Carlo and Molecular Dynamics Trajectories. J. 4493
4425 Chem. Res. 1995, 28, 201−207. Chem. Inf. Model. 2011, 51, 2007−2023. 4494
4426 (63) Keyes, T. Unstable Modes in Supercooled and Normal Liquids: (88) Sarangi, S. S.; Reddy, S. K.; Balasubramanian, S. Low Frequency 4495
4427 Density-Of-States, Energy Barriers, and Self-Diffusion. J. Chem. Phys. Vibrational Modes of Room Temperature Ionic Liquids. J. Phys. Chem. 4496
4428 1994, 101, 5081−5092. B 2011, 115, 1874−1880. 4497
4429 (64) Ribeiro, M. C. C.; Madden, P. A. Instantaneous Normal Mode (89) Mondal, A.; Balasubramanian, S. Vibrational Signatures of 4498
4430 Prediction for Cation and Anion Diffusion in Ionic Melts. J. Chem. Cation-Anion Hydrogen Bonding in Ionic Liquids: A Periodic Density 4499
4431 Phys. 1997, 106 (20), 8616−8619. Functional Theory and Molecular Dynamics Study. J. Phys. Chem. B 4500
4432 (65) Ribeiro, M. C. C.; Madden, P. A. Unstable Modes in Ionic 2015, 119, 1994−2002. 4501
4433 Melts. J. Chem. Phys. 1998, 108, 3256−3263. (90) Mathias, G.; Baer, M. D. Generalized Normal Coordinates for 4502
4434 (66) Ribeiro, M. C. C.; Wilson, M.; Madden, P. A. On the the Vibrational Analysis of Molecular Dynamics Simulations. J. Chem. 4503
4435 Observation of Propagating Sound Modes at High Momentum Theory Comput. 2011, 7, 2028−2039. 4504
4436 Transfer in Viscous Liquids and Glasses. J. Chem. Phys. 1998, 108, (91) Mathias, G.; Ivanov, S. D.; Witt, A.; Baer, M. D.; Marx, D. 4505
4437 9027−9038. Infrared Spectroscopy of Fluxional Molecules from (ab Initio) 4506
4438 (67) Ribeiro, M. C. C.; Wilson, M.; Madden, P. A. The Nature of the Molecular Dynamics: Resolving Large-Amplitude Motion, Multiple 4507
4439 ″Vibrational Modes″ of the Network-Forming Liquid ZnCl2. J. Chem. Conformations, and Permutational Symmetries. J. Chem. Theory 4508
4440 Phys. 1998, 109, 9859−9869. Comput. 2012, 8, 224−234. 4509
4441 (68) Ribeiro, M. C. C.; Wilson, M.; Madden, P. A. Raman Scattering (92) Kirillov, S. A. Interactions and Picosecond Dynamics in Molten 4510
4442 in the Network Liquid ZnCl2 Relationship to the Vibrational Density Salts: A Review with Comparison to Molecular Liquids. J. Mol. Liq. 4511
4443 of States. J. Chem. Phys. 1999, 110, 4803−4811. 1998, 76, 35−95. 4512
4444 (69) Wilson, M.; Ribeiro, M. C. C. The Vibrational Motion of (93) Janz, G. J.; James, D. W. Raman Spectra and Ionic Interactions 4513
4445 ’Polymeric’ BeCl2. Mol. Phys. 1999, 96, 867−876. in Molten Nitrates. J. Chem. Phys. 1961, 35, 739−744. 4514
4446 (70) Castiglione, M. J.; Ribeiro, M. C. C.; Wilson, M.; Madden, P. A. (94) Chabanel, M.; Bencheikh, A.; Puchalska, D. Vibrational Study of 4515
4447 Al3+ Coordination in Cryolitic Melts: A Computer Simulation Study. Ionic Association in Aprotic Solvents. Part 12. Isothiocyanate 4516
4448 Z. Naturforsch., A: Phys. Sci. 1999, 54 (10-11), 605−610. M(NCS) Complexes of Non-Transition-Metal Ions. J. Chem. Soc., 4517
4449 (71) Hansen, J.-P.; McDonald, I. R. Theory of Simple Liquids, 3rd ed.; Dalton Trans. 1989, No. 11, 2193−2197. 4518
4450 Academic Press: Amsterdam, 2006. (95) Nyquist, R. A.; Putzig, C. L.; Leugers, M. A. The Handbook of 4519
4451 (72) Kubo, R. Statistical-Mechanical Theory of Irreversible Processes. Infrared and Raman Spectra of Inorganic Compounds and Organic Salt; 4520
4452 I. General Theory and Simple Applications to Magnetic and Academic Press: San Diego, 1997; Vol. 1. 4521
4453 Conduction Problems. J. Phys. Soc. Jpn. 1957, 12, 570−586. (96) Takahashi, S.; Curtiss, L. A.; Gosztola, D.; Koura, N.; Saboungi, 4522
4454 (73) Harp, G. D.; Berne, B. J. Time-Correlation Functions, Memory M. L. Molecular-Orbital Calculations and Raman Measurements for 1- 4523
4455 Functions, and Molecular Dynamics. Phys. Rev. A: At., Mol., Opt. Phys. ethyl-3-methylimidazolium Chloroaluminates. Inorg. Chem. 1995, 34, 4524
4456 1970, 2, 975−996. 2990−2993. 4525
4457 (74) McQuarrie, D. A. Statistical Mechanics; University Science (97) Janz, G. J.; Kozlowski, T. R.; Wait, S. C. Raman Spectrum and 4526
4458 Books: Sausalito, California, 2000. Vibrational Assignment for Molten Thallous Nitrate. J. Chem. Phys. 4527
4459 (75) Canongia Lopes, J. N.; Padua, A. A. H. CL&P: A Generic and 1963, 39, 1809−1812. 4528
4460 Systematic Force Field for Ionic Liquids Modeling. Theor. Chem. Acc. (98) Wait, S. C.; Ward, A. T.; Janz, G. J. AnionCation Interactions 4529
4461 2012, 131, 1129. in Molten Inorganic Nitrates: Vibrational Analysis. J. Chem. Phys. 4530
4462 (76) Anderson, P. W. A Mathematical Model for the Narrowing of 1966, 45, 133−137. 4531
4463 Spectral Lines by Exchange or Motion. J. Phys. Soc. Jpn. 1954, 9, 316− (99) Holomb, R.; Martinelli, A.; Albinsson, I.; Lassègues, J. C.; 4532
4464 339. Johansson, P.; Jacobsson, P. Ionic Liquid Structure: the Conforma- 4533
4465 (77) Wang, C. H. Spectroscopy of Condensed Media; Academic Press: tional Isomerism in 1-Butyl-3-methyl-imidazolium Tetrafluoroborate 4534
4466 Amsterdam, 1985. ([bmim][BF4]). J. Raman Spectrosc. 2008, 39, 793−805. 4535
4467 (78) Thomas, M.; Brehm, M.; Fligg, R.; Voehringer, P.; Kirchner, B. (100) Madden, P. A.; O’Sullivan, K.; Board, J. A.; Fowler, P. W. Light 4536
4468 Computing Vibrational Spectra from ab initio Molecular Dynamics. Scattering by Alkali Halide Melts: A Computer Simulation Study. J. 4537
4469 Phys. Chem. Chem. Phys. 2013, 15, 6608−6622. Chem. Phys. 1991, 94, 918−927. 4538
4470 (79) Benson, A. M.; Drickamer, H. G. Stretching Vibrations in (101) Madden, P. A.; O’Sullivan, K. F. The Raman Spectra of Alkali 4539
4471 Condensed Systems: Especially Bonds Containing Hydrogen. J. Chem. Halide Melts: A Theoretical Description. J. Chem. Phys. 1991, 95, 4540
4472 Phys. 1957, 27, 1164−1174. 1980−1990. 4541

AU DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4542 (102) Hegemann, B.; Jonas, J. Separation of Temperature and (120) Shukla, M.; Noothalapati, H.; Shigeto, S.; Saha, S. Importance 4611
4543 Density Effects on Collision-Induced Rayleigh and Raman Line- of Weak Interactions and Conformational Equilibrium in N-Butyl-N- 4612
4544 Shapes of Liquid Carbon-Disulfide. J. Chem. Phys. 1985, 82, 2845− Methylpiperidinium Bis(trifluromethanesulfonyl)imide Room Tem- 4613
4545 2855. perature Ionic Liquids: Vibrational and Theoretical Studies. Vib. 4614
4546 (103) Madden, P. A.; Tildesley, D. J. Interaction-Induced Spectrosc. 2014, 75, 107−117. 4615
4547 Contributions to Rayleigh and Allowed Raman Bands - A Simulation (121) Vitucci, F. M.; Trequattrini, F.; Palumbo, O.; Brubach, J. B.; 4616
4548 Study of CS2. Mol. Phys. 1985, 55, 969−998. Roy, P.; Paolone, A. Infrared Spectra of Bis(trifluoromethanesulfonyl)- 4617
4549 (104) Ikawa, S.; Whalley, E. Polarized and Depolarized Raman imide Based Ionic Liquids: Experiments and DFT Simulations. Vib. 4618
4550 Spectra of Liquid Carbon Disulfide at 0−10 kbar. 3. Interaction- Spectrosc. 2014, 74, 81−87. 4619
4551 Induced.nu.2 and.nu.3 Scattering and the Fluctuation of the Local (122) Vitucci, F. M.; Trequattrini, F.; Palumbo, O.; Brubach, J. B.; 4620
4552 Field. J. Phys. Chem. 1990, 94, 7834−7839. Roy, P.; Navarra, M. A.; Panero, S.; Paolone, A. Stabilization of 4621
4553 (105) Hipps, K. W.; Aplin, A. T. The Tricyanomethanide Ion: an Different Conformers of Bis(trifluoromethanesulfonyl)imide Anion in 4622
4554 Infrared, Raman, and Tunneling Spectroscopy Study Including Ammonium-Based Ionic Liquids at Low Temperatures. J. Phys. Chem. 4623
4555 Isotopic-Substitution. J. Phys. Chem. 1985, 89, 5459−5464. A 2014, 118, 8758−8764. 4624
4556 (106) Penna, T. C.; Faria, L. F. O.; Ribeiro, M. C. C. Raman Band (123) Faria, L. F. O.; Matos, J. R.; Ribeiro, M. C. C. Thermal Analysis 4625
4557 Shape Analysis of Cyanate-Anion Ionic Liquids. J. Mol. Liq. 2015, 209, and Raman Spectra of Different Phases of the Ionic Liquid 4626
4558 676−682. Butyltrimethylammonium Bis(trifluoromethylsulfonyl)imide. J. Phys. 4627
4559 (107) Sprague, J. W.; Ritchey, W. M.; Grasselli, J. G. The Synthesis Chem. B 2012, 116, 9238−9245. 4628
4560 and Infrared and Nuclear Magnetic Resonance Spectra of Ammonium (124) Fujii, K.; Fujimori, T.; Takamuku, T.; Kanzaki, R.; 4629
4561 Dicyanamide. J. Phys. Chem. 1964, 68, 431−433. Umebayashi, Y.; Ishiguro, S. I. Conformational Equilibrium of 4630
4562 (108) Reckeweg, O.; DiSalvo, F. J.; Schulz, A.; Blaschkowski, B.; Bis(Trifluoromethanesulfonyl) Imide Anion of a Room-Temperature 4631
4563 Jagiella, S.; Schleid, T. Synthesis, Crystal Structure, and Vibrational Ionic Liquid: Raman Spectroscopic Study and DFT Calculations. J. 4632
4564 Spectra of the Anhydrous Lithium Dicyanamide Li[N(CN)2]. Z. Phys. Chem. B 2006, 110, 8179−8183. 4633
4565 Anorg. Allg. Chem. 2014, 640, 851−855. (125) Lassègues, J. C.; Grondin, J.; Holomb, R.; Johansson, P. Raman 4634
4566 (109) Kiefer, J.; Noack, K.; Penna, T. C.; Ribeiro, M. C. C.; Weber, and ab initio Study of the Conformational Isomerism in the 1-Ethyl-3- 4635
4567 H.; Kirchner, B. Vibrational Signatures of Cyano Groups in methylimidazolium Bis(trifluoromethanesulfonyl)imide Ionic Liquid. J. 4636
4568 Imidazolium Ionic Liquids. Vib. Spectrosc. 2016, DOI: 10.1016/ Raman Spectrosc. 2007, 38, 551−558. 4637
4569 j.vibspec.2016.05.004. (126) Lopes, J. N. C.; Shimizu, K.; Padua, A. A. H.; Umebayashi, Y.; 4638
4570 (110) Bernhardt, E.; Henkel, G.; Willner, H. Die Tetracyanoborate Fukuda, S.; Fujii, K.; Ishiguro, S.-i. A Tale of Two Ions: The 4639
4571 M[B(CN)4], M = [Bu4N]+, Ag+, K+. Z. Anorg. Allg. Chem. 2000, 626, Conformational Landscapes of Bis(trifluoromethanesulfonyl)amide 4640
4572 560−568. and N,N-Dialkylpyrrolidinium. J. Phys. Chem. B 2008, 112, 1465− 4641
4573 (111) Rey, I.; Johansson, P.; Lindgren, J.; Lassègues, J. C.; Grondin, 1472. 4642
4574 J.; Servant, L. Spectroscopic and Theoretical Study of (CF3SO2)2N- (127) Umebayashi, Y.; Mitsugi, T.; Fujii, K.; Seki, S.; Chiba, K.; 4643
4575 (TFSI-) and (CF3SO2)2NH (HTFSI). J. Phys. Chem. A 1998, 102, Yamamoto, H.; Lopes, J. N. C.; Padua, A. A. H.; Takeuchi, M.; 4644
4576 3249−3258. Kanzaki, R.; et al. Raman Spectroscopic Study, DFT Calculations and 4645
4577 (112) Herstedt, M.; Smirnov, M.; Johansson, P.; Chami, M.; MD Simulations on the Conformational Isomerism of N-Alkyl-N- 4646
4578 Grondin, J.; Servant, L.; Lassègues, J. C. Spectroscopic Character- methylpyrrolidinium Bis-(trifluoromethanesulfonyl) Amide Ionic 4647
4579 ization of the Conformational States of the Bis- Liquids. J. Phys. Chem. B 2009, 113, 4338−4346. 4648
4580 (trifluoromethanesulfonyl)imide Anion (TFSI-). J. Raman Spectrosc. (128) Lima, T. A.; Paschoal, V. H.; Faria, L. F. O.; Ribeiro, M. C. C.; 4649
4581 2005, 36, 762−770. Ferreira, F. F.; Costa, F. N.; Giles, C. Comparing Two 4650
4582 (113) Rey, I.; Lassègues, J. C.; Grondin, J.; Servant, L. Infrared and Tetraalkylammonium Ionic Liquids. II. Phase Transitions. J. Chem. 4651
4583 Raman Study of the PEO-LiTFSI Polymer Electrolyte. Electrochim. Phys. 2016, 144, 224505. 4652
4584 Acta 1998, 43, 1505−1510. (129) Fujii, K.; Seki, S.; Fukuda, S.; Kanzaki, R.; Takamuku, T.; 4653
4585 (114) Castriota, M.; Caruso, T.; Agostino, R. G.; Cazzanelli, E.; Umebayashi, Y.; Ishiguro, S.-i. Anion Conformation of Low-Viscosity 4654
4586 Henderson, W. A.; Passerini, S. Raman Investigation of the Ionic Room-Temperature Ionic Liquid 1-Ethyl-3-methylimidazolium Bis- 4655
4587 Liquid N-methyl-N-propylpyrrolidinium bis- (fluorosulfonyl)imide. J. Phys. Chem. B 2007, 111, 12829−12833. 4656
4588 (trifluoromethanesulfonyl)imide and its Mixture with LiN(SO2CF3) (130) Fujii, K.; Seki, S.; Fukuda, S.; Takamuku, T.; Kohara, S.; 4657
4589 (2). J. Phys. Chem. A 2005, 109, 92−96. Kameda, Y.; Umebayashi, Y.; Ishiguro, S.-i. Liquid Structure and 4658
4590 (115) Kiefer, J.; Fries, J.; Leipertz, A. Experimental Vibrational Study Conformation of a Low-Viscosity Ionic Liquid, N-methyl-N-propyl- 4659
4591 of Imidazolium-Based Ionic Liquids: Raman and Infrared Spectra of 1- pyrrolidinium bis(fluorosulfonyl)imide Studied by High-Energy X-ray 4660
4592 ethyl-3-methylimidazolium bis(Trifluoromethylsulfonyl)imide and 1- Scattering. J. Mol. Liq. 2008, 143, 64−69. 4661
4593 ethyl-3-methylimidazolium ethylsulfate. Appl. Spectrosc. 2007, 61, (131) Matsumoto, K.; Oka, T.; Nohira, T.; Hagiwara, R. Poly- 4662
4594 1306−1311. morphism of Alkali Bis(fluorosulfonyl)amides (M N(SO2F)(2), M = 4663
4595 (116) Martinelli, A.; Matic, A.; Johansson, P.; Jacobsson, P.; Na, K, and Cs). Inorg. Chem. 2013, 52, 568−576. 4664
4596 Borjesson, L.; Fernicola, A.; Panero, S.; Scrosati, B.; Ohno, H. (132) Shimizu, Y.; Fujii, K.; Imanari, M.; Nishikawa, K. Phase 4665
4597 Conformational Evolution of TFSI- in Protic and Aprotic Ionic Behavior of a Piperidinium-Based Room-Temperature Ionic Liquid 4666
4598 Liquids. J. Raman Spectrosc. 2011, 42, 522−528. Exhibiting Scanning Rate Dependence. J. Phys. Chem. B 2015, 119, 4667
4599 (117) Mao, J. X.; Nulwala, H. B.; Luebke, D. R.; Damodaran, K. 12552−12560. 4668
4600 Spectroscopic and Computational Analysis of the Molecular (133) Giffin, G. A.; Laszczynski, N.; Jeong, S.; Jeremias, S.; Passerini, 4669
4601 Interactions in the Ionic Liquid Ion Pair BMP(+) TFSI(−). J. Mol. S. Conformations and Vibrational Assignments of the (Fluorosulfonyl) 4670
4602 Liq. 2012, 175, 141−147. (trifluoromethanesulfonyl)imide Anion in Ionic Liquids. J. Phys. Chem. 4671
4603 (118) Moschovi, A. M.; Ntais, S.; Dracopoulos, V.; Nikolakis, V. C 2013, 117, 24206−24212. 4672
4604 Vibrational Spectroscopic Study of the Protic Ionic Liquid 1-H-3- (134) Manning, J.; Frech, R. The Structure of Associated Ionic 4673
4605 methylimidazolium Bis(Trifluoromethanesulfonyl)imide. Vib. Spec- Species in Poly(Propylene oxide) Alkali-Metal trifluoromethanesulfo- 4674
4606 trosc. 2012, 63, 350−359. nate Complexes. Polymer 1992, 33, 3487−3494. 4675
4607 (119) Dhumal, N. R.; Noack, K.; Kiefer, J.; Kim, H. J. Molecular (135) Dissanayake, M.; Frech, R. Infrared Spectroscopic Study of the 4676
4608 Structure and Interactions in the Ionic Liquid 1-Ethyl-3-methylimida- Phases and Phase-Transitions in Poly(Ethylene Oxide) and Poly- 4677
4609 zolium Bis(Trifluoromethylsulfonyl)imide. J. Phys. Chem. A 2014, 118, (Ethylene Oxide)-Lithium Trifluoromethanesulfonate Complexes. 4678
4610 2547−2557. Macromolecules 1995, 28, 5312−5319. 4679

AV DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4680 (136) Rhodes, C. P.; Frech, R. Cation-Anion and Cation-Polymer (157) Goypiron, A.; Devillepin, J.; Novak, A. Raman and Infrared 4749
4681 Interactions in (PEO)(n)NaCF3SO3 (n = 1−80). Solid State Ionics Study of KHSO4 Crystal. J. Raman Spectrosc. 1980, 9, 297−303. 4750
4682 1999, 121, 91−99. (158) Phamthi, M.; Colomban, P.; Novak, A.; Blinc, R. Vibrational 4751
4683 (137) Alloin, F.; Hirankumar, G.; Pagnier, T. Temperature- Spectra of and Phase Transitions in Cesium Hydrogen Sulfate. J. 4752
4684 Dependent Raman Spectroscopy of Lithium Triflate-PEO Complexes: Raman Spectrosc. 1987, 18, 185−194. 4753
4685 Phase Equilibrium and Component Interactions. J. Phys. Chem. B (159) Colomban, P.; Phamthi, M.; Novak, A. Influence of Thermal 4754
4686 2009, 113, 16465−16471. and Mechanical Treatment and of Water on Structural Phase- 4755
4687 (138) Huang, W. W.; Frech, R.; Wheeler, R. A. Molecular Structures Transitions in CsHSO4. Solid State Ionics 1987, 24, 193−203. 4756
4688 and Normal Vibrations of Trifluoromethane Sulfonate (CF3SO3-) and (160) Varma, V.; Rangavittal, N.; Rao, C. N. R. A Study of Superionic 4757
4689 Its Lithium Ion Pairs and Aggregates. J. Phys. Chem. 1994, 98, 100− CsHSO4 and Cs1-XLixHSO4 by Vibrational Spectroscopy and X-Ray 4758
4690 110. Diffraction. J. Solid State Chem. 1993, 106, 164−173. 4759
4691 (139) Gejji, S. P.; Hermansson, K.; Tegenfeldt, J.; Lindgren, J. (161) Rogers, S. E.; Ubbelohde, A. R. Melting and Crystal Structure. 4760
4692 Geometry and Vibrational Frequencies of the Lithium Triflate Ion- III. Low-Melting Acid Sulfates. Trans. Faraday Soc. 1950, 46, 1051− 4761
4693 Pair: An Ab-Initio Study. J. Phys. Chem. 1993, 97, 11402−11407. 1061. 4762
4694 (140) Frech, R.; Huang, W. W. Anion-Solvent and Anion-Cation (162) Cabaço, M. I.; Besnard, M.; Danten, Y.; Coutinho, J. A. P. 4763
4695 Interactions in Lithium and Tetrabutylammonium Trifluoromethane- Carbon Dioxide in 1-Butyl-3-methylimidazolium Acetate. I. Unusual 4764
4696 sulfonate Solutions. J. Solution Chem. 1994, 23, 469−481.
Solubility Investigated by Raman Spectroscopy and DFT Calculations. 4765
4697 (141) Alia, J. M.; Edwards, H. G. M. Ion Solvation and Ion
J. Phys. Chem. A 2012, 116, 1605−1620. 4766
4698 Association in Lithium Trifluoromethanesulfonate Solutions in Three
(163) Ito, K.; Bernstein, H. J. The Vibrational Spectra of the 4767
4699 Aprotic Solvents. An FT-Raman Spectroscopic Study. Vib. Spectrosc.
Formate, Acetate, and Oxalate Ions. Can. J. Chem. 1956, 34, 170−178. 4768
4700 2000, 24, 185−200.
(164) Alcock, N. W.; Tracy, V. M.; Waddington, T. C. Acetates and 4769
4701 (142) Mikhailov, G. P. Calculation of Vibrational Spectra for
4702 Coordinated Trifluoromethanesulfonate Ion in Dipolar Aprotic Acetato-Complexes. Part 2. Spectroscopic Studies. J. Chem. Soc., Dalton 4770
4703 Solvents. J. Appl. Spectrosc. 2015, 82, 165−168. Trans. 1976, No. 21, 2243−2246. 4771
4704 (143) Varetti, E. L.; Fernandez, E. L.; Altabef, A. B. The Polarized (165) Deacon, G. B.; Phillips, R. J. Relationships Between the 4772
4705 Infrared and Raman-Spectra of Crystalline ammonium trifluorome- Carbon-Oxygen Stretching Frequencies of Carboxylato Complexes 4773
4706 thanesulfonate. Spectrochim. Acta, Part A 1991, 47, 1767−1774. and the Type of Carboxylate Coordination. Coord. Chem. Rev. 1980, 4774
4707 (144) Schwenzer, B.; Kerisit, S. N.; Vijayakumar, M. Anion Pairs in 33, 227−250. 4775
4708 Room Temperature Ionic Liquids Predicted by Molecular Dynamics (166) Nara, M.; Torii, H.; Tasumi, M. Correlation between the 4776
4709 Simulation, Verified by Spectroscopic Characterization. RSC Adv. Vibrational Frequencies of the Carboxylate Group and the Types of its 4777
4710 2014, 4, 5457−5464. Coordination to a Metal Ion: An ab initio Molecular Orbital Study. J. 4778
4711 (145) Akai, N.; Kawai, A.; Shibuya, K. Ion-Pair Structure of Phys. Chem. 1996, 100, 19812−19817. 4779
4712 Vaporized Ionic Liquid Studied by Matrix-Isolation FTIR Spectros- (167) Holloczki, O.; Firaha, D. S.; Friedrich, J.; Brehm, M.; Cybik, R.; 4780
4713 copy with DFT Calculations: A Case of 1-Ethyl-3-methylimidazolium Wild, M.; Stark, A.; Kirchner, B. Carbene Formation in Ionic Liquids: 4781
4714 Trifluoromethanesulfonate. J. Phys. Chem. A 2010, 114, 12662−12666. Spontaneous, Induced, or Prohibited? J. Phys. Chem. B 2013, 117, 4782
4715 (146) Mao, J. X.; Damodaran, K. Spectroscopic and Computational 5898−5907. 4783
4716 Analysis of the Molecular Interactions in the Ionic Liquid Emim (+) (168) Nakamoto, K. Infrared and Raman Spectra of Inorganic and 4784
4717 FAP (−). Ionics 2015, 21, 1605−1613. Coordination Compounds, Part B, 6th ed.; John Wiley & Sons: 4785
4718 (147) Voroshylova, I. V.; Teixeira, F.; Costa, R.; Pereira, C. M.; Hoboken, NJ, 2009. 4786
4719 Cordeiro, M. N. D. S. Interactions in the Ionic Liquid [EMIM][FAP]: (169) Rudolph, W. W.; Irmer, G. Raman Spectroscopic Studies and 4787
4720 a Coupled Experimental and Computational Analysis. Phys. Chem. DFT Calculations on NaCH3CO2 and NaCD3CO2 Solutions in 4788
4721 Chem. Phys. 2016, 18, 2617−2628. Water and Heavy Water. RSC Adv. 2015, 5, 21897−21908. 4789
4722 (148) Dhumal, N. R.; Kim, H. J.; Kiefer, J. Electronic Structure and (170) Tanzi, L.; Benassi, P.; Nardone, M.; Ramondo, F. Vibrations of 4790
4723 Normal Vibrations of the 1-Ethyl-3-methylimidazolium Ethyl Sulfate Bioionic Liquids by Ab Initio Molecular Dynamics and Vibrational 4791
4724 Ion Pair. J. Phys. Chem. A 2011, 115, 3551−3558. Spectroscopy. J. Phys. Chem. A 2014, 118, 12229−12240. 4792
4725 (149) Koda, S.; Nomura, H. Aqueous-Solutions of Sodium (171) Quiles, F.; Burneau, A. Infrared and Raman Spectra of 4793
4726 Methylsulfate by Raman-Scattering, NMR, Ultrasound, and Density- Alkaline-Earth and Copper(II) Acetates in Aqueous Solutions. Vib. 4794
4727 Measurements. J. Solution Chem. 1985, 14, 355−366. Spectrosc. 1998, 16, 105−117. 4795
4728 (150) Walrafen, G. E.; Young, T. F.; Irish, D. E. Raman Spectral (172) Bowron, D. T.; D’Agostino, C.; Gladden, L. F.; Hardacre, C.; 4796
4729 Studies of Molten Potassium Bisulfate and Vibrational Frequencies of Holbrey, J. D.; Lagunas, M. C.; McGregor, J.; Mantle, M. D.; Mullan, 4797
4730 S2O7 Groups. J. Chem. Phys. 1962, 37, 662−670. C. L.; Youngs, T. G. A. Structure and Dynamics of 1-Ethyl-3- 4798
4731 (151) Fehrmann, R.; Hansen, N. H.; Bjerrum, N. J. Raman-
methylimidazolium Acetate via Molecular Dynamics and Neutron 4799
4732 Spectroscopic and Spectrophotometric Study of the System K2S2O7-
Diffraction. J. Phys. Chem. B 2010, 114, 7760−7768. 4800
4733 KHSO4 in the Temperature-Range 200−450 °C. Inorg. Chem. 1983,
(173) Cabaço, M. I.; Besnard, M.; Danten, Y.; Coutinho, J. A. P. 4801
4734 22, 4009−4014.
(152) Knudsen, C. B.; Kalampounias, A. G.; Fehrmann, R.; Solubility of CO2 in 1-Butyl-3-methyl-imidazolium-trifluoro Acetate 4802
4735
4736 Boghosian, S. Thermal Dissociation of Molten KHSO4: Temperature Ionic Liquid Studied by Raman Spectroscopy and DFT Investigations. 4803
4737 Dependence of Raman Spectra and Thermodynamics. J. Phys. Chem. B J. Phys. Chem. B 2011, 115, 3538−3550. 4804
4738 2008, 112, 11996−12000. (174) Robinson, R. E.; Taylor, R. C. Raman Spectrum and 4805
4739 (153) Irish, D. E.; Chen, H. Equilibria and Proton Transfer in Vibrational Assignments for the Trifluoroacetate Ion. Spectrochim. 4806
4740 Bisulfate-Sulfate System. J. Phys. Chem. 1970, 74, 3796−3801. Acta 1962, 18, 1093−1095. 4807
4741 (154) Turner, D. J. Raman Spectral Study of Bisulphate Ion (175) Shukla, M.; Srivastava, N.; Saha, S. Theoretical and 4808
4742 Hydration. J. Chem. Soc., Faraday Trans. 2 1972, 68, 643−648. Spectroscopic Studies of 1-butyl-3-methylimidazolium iodide Room 4809
4743 (155) Kiefer, J.; Pye, C. C. Structure of the Room-Temperature Ionic Temperature Ionic Liquid: Its Differences with chloride and bromide 4810
4744 Liquid 1-Hexyl-3-methylimidazolium Hydrogen Sulfate: Conforma- Derivatives. J. Mol. Struct. 2010, 975, 349−356. 4811
4745 tional Isomerism. J. Phys. Chem. A 2010, 114, 6713−6720. (176) Jeon, Y.; Sung, J.; Seo, C.; Lim, H.; Cheong, H.; Kang, M.; 4812
4746 (156) Ribeiro, M. C. C. High Viscosity of Imidazolium Ionic Liquids Moon, B.; Ouchi, Y.; Kim, D. Structures of Ionic Liquids with 4813
4747 with the Hydrogen Sulfate Anion: A Raman Spectroscopy Study. J. Different Anions Studied by Infrared Vibration Spectroscopy. J. Phys. 4814
4748 Phys. Chem. B 2012, 116, 7281−7290. Chem. B 2008, 112, 4735−4740. 4815

AW DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4816 (177) Dhumal, N. R. Molecular Interactions in 1,3-dimethylimida- (195) Chen, Z. J.; Lee, J.-M. Free Volume Model for the Unexpected 4884
4817 zolium-bis(trifluromethanesulfonyl)imide Ionic Liquid. Chem. Phys. Effect of C2-Methylation on the Properties of Imidazolium Ionic 4885
4818 2007, 342, 245−252. Liquids. J. Phys. Chem. B 2014, 118, 2712−2718. 4886
4819 (178) Wulf, A.; Fumino, K.; Ludwig, R. Comment on ″New (196) Noack, K.; Schulz, P. S.; Paape, N.; Kiefer, J.; Wasserscheid, P.; 4887
4820 Interpretation of the CH Stretching Vibrations in Imidazolium-Based Leipertz, A. The Role of the C2 Position in Interionic Interactions of 4888
4821 Ionic Liquids″. J. Phys. Chem. A 2010, 114, 685−686. Imidazolium Based Ionic Liquids: a Vibrational and NMR 4889
4822 (179) Lassègues, J.-C.; Grondin, J.; Cavagnat, D.; Johansson, P. Reply Spectroscopic Study. Phys. Chem. Chem. Phys. 2010, 12, 14153−14161. 4890
4823 to the ″Comment on ’New Interpretation of the CH Stretching (197) Hunt, P. A.; Kirchner, B.; Welton, T. Characterising the 4891
4824 Vibrations in Imidazolium-Based Ionic Liquids’″. J. Phys. Chem. A Electronic Structure of Ionic Liquids: An Examination of the 1-butyl-3- 4892
4825 2010, 114, 687−688. methylimidazolium chloride Ion Pair. Chem. - Eur. J. 2006, 12, 6762− 4893
4826 (180) Hunt, P. A.; Ashworth, C. R.; Matthews, R. P. Hydrogen 6775. 4894
4827 Bonding in Ionic Liquids. Chem. Soc. Rev. 2015, 44, 1257−1288. (198) Endo, T.; Kato, T.; Nishikawa, K. Effects of Methylation at the 4895
4828 (181) Hayashi, S.; Ozawa, R.; Hamaguchi, H. Raman Spectra, Crystal 2 Position of the Cation Ring on Phase Behaviors and Conformational 4896
4829 Polymorphism, and Structure of a Prototype Ionic-Iiquid [bmim]Cl. Structures of Imidazolium-Based Ionic Liquids. J. Phys. Chem. B 2010, 4897
4830 Chem. Lett. 2003, 32, 498−499. 114, 9201−9208. 4898
4831 (182) Holbrey, J. D.; Reichert, W. M.; Nieuwenhuyzen, M.; Johnson, (199) Moschovi, A. M.; Dracopoulos, V.; Nikolakis, V. Inter- and 4899
4832 S.; Seddon, K. R.; Rogers, R. D. Crystal Polymorphism in 1-Butyl-3- Intramolecular Interactions in Imidazolium Protic Ionic Liquids. J. 4900
4833 Methylimidazolium Halides: Supporting Ionic Liquid Formation by Phys. Chem. B 2014, 118, 8673−8683. 4901
4834 Inhibition of Crystallization. Chem. Commun. 2003, No. 14, 1636− (200) Garaga, M. N.; Nayeri, M.; Martinelli, A. Effect of the Alkyl 4902
4835 1637. Chain Length in 1-Alkyl-3-Methylimidazolium Ionic Liquids on Inter- 4903
4836 (183) Saha, S.; Hayashi, S.; Kobayashi, A.; Hamaguchi, H. Crystal Molecular Interactions and Rotational Dynamics: A Combined 4904
4837 Structure of 1-butyl-3-methylimidazolium chloride. A Clue to the Vibrational and NMR Spectroscopic Study. J. Mol. Liq. 2015, 210, 4905
4838 Elucidation of the Ionic Liquid Structure. Chem. Lett. 2003, 32, 740− 169−177. 4906
4839 741. (201) Moschovi, A. M.; Dracopoulos, V. Structure of Protic 4907
4840 (184) Ozawa, R.; Hayashi, S.; Saha, S.; Kobayashi, A.; Hamaguchi, H. (HC(n)ImNTf(2), n = 0−12) and Aprotic (C(1)C(n)ImNTf(2), n 4908
4841 Rotational isomerism and Structure of the 1-butyl-3-methylimidazo- = 1−12) Imidazolium Ionic Liquids: A Vibrational Spectroscopic 4909
4842 lium Cation in the Ionic Liquid State. Chem. Lett. 2003, 32, 948−949. Study. J. Mol. Liq. 2015, 210, 189−199. 4910
4843 (185) Hamaguchi, H.; Ozawa, R. Structure of Ionic Liquids and Ionic (202) Xuan, X.; Guo, M.; Pei, Y.; Zheng, Y. Theoretical Study on 4911
4844 Liquid Compounds: Are Ionic Liquids Genuine Liquids in the Cation-Anion Interaction and Vibrational Spectra of 1-Allyl-3- 4912
4845 Conventional Sense? In Advances in Chemical Physics, Rice, S. A., Ed.; methylimidazolium-Based Ionic Liquids. Spectrochim. Acta, Part A 4913
4846 John Wiley & Sons, Inc.: Hoboken, 2005; Vol. 131, pp 85−104. 2011, 78, 1492−1499. 4914
4847 (186) Endo, T.; Hoshino, S.; Shimizu, Y.; Fujii, K.; Nishikawa, K. (203) Xu, H.; Liao, J.; Li, Q.; You, L.; Kang, S.; Chen, H.; He, B.; 4915
4848 Comprehensive Conformational and Rotational Analyses of the Butyl Tang, Q. Synthesis and Properties of 1-allyl-3-methyl-imidazolium 4916
4849 Group in Cyclic Cation: DFT Calculations for Imidazolium, bicarbonate Room Temperature Ionic Liquid. Mater. Res. Innovations 4917
4850 Pyridinium, Pyrrolidinium, and Piperidinium. J. Phys. Chem. B 2016, 2014, 18, 457−460. 4918
4851 120, 10336−10349. (204) Shirota, H.; Matsuzaki, H.; Ramati, S.; Wishart, J. F. Effects of 4919
4852 (187) Tsuzuki, S.; Arai, A. A.; Nishikawa, K. Conformational Analysis Aromaticity in Cations and Their Functional Groups on the Low- 4920
4853 of 1-Butyl-3-methylimidazolium by CCSD(T) level ab initio Frequency Spectra and Physical Properties of Ionic Liquids. J. Phys. 4921
4854 Calculations: Effects of Neighboring Anions. J. Phys. Chem. B 2008, Chem. B 2015, 119, 9173−9187. 4922
4855 112, 7739−7747. (205) Xue, L.; Tamas, G.; Matthews, R. P.; Stone, A. J.; Hunt, P. A.; 4923
4856 (188) Umebayashi, Y.; Hamano, H.; Tsuzuki, S.; Canongia Lopes, J. Quitevis, E. L.; Lynden-Bell, R. M. An OHD-RIKES and Simulation 4924
4857 N.; Padua, A. A. H.; Kameda, Y.; Kohara, S.; Yamaguchi, T.; Fujii, K.; Study Comparing a Benzylmethylimidazolium Ionic Liquid with an 4925
4858 Ishiguro, S.-i. Dependence of the Conformational Isomerism in 1-n- Equimolar Mixture of Dimethylimidazolium and Benzene. Phys. Chem. 4926
4859 Butyl-3-methylimidazolium Ionic Liquids on the Nature of the Halide Chem. Phys. 2015, 17, 9973−9983. 4927
4860 Anion. J. Phys. Chem. B 2010, 114, 11715−11724. (206) Knorr, A.; Ludwig, R. Cation-cation clusters in ionic liquids: 4928
4861 (189) Katayanagi, H.; Hayashi, S.; Hamaguchi, H. O.; Nishikawa, K. Cooperative hydrogen bonding overcomes like-charge repulsion. Sci. 4929
4862 Structure of an Ionic Liquid, 1-n-butyl-3-methylimidazolium iodide, Rep. 2015, 5, 17505. 4930
4863 Studied by Wide-Angle X-ray Scattering and Raman Spectroscopy. (207) Knorr, A.; Stange, P.; Fumino, K.; Weinhold, F.; Ludwig, R. 4931
4864 Chem. Phys. Lett. 2004, 392, 460−464. Spectroscopic Evidence for Clusters of Like-Charged Ions in Ionic 4932
4865 (190) Singh, D. K.; Cha, S.; Nam, D.; Cheong, H.; Joo, S.-W.; Kim, Liquids Stabilized by Cooperative Hydrogen Bonding. ChemPhysChem 4933
4866 D. Raman Spectroscopic Study on Alkyl Chain Conformation in 1- 2016, 17, 458−462. 4934
4867 Butyl-3-methylimidazolium Ionic Liquids and their Aqueous Mixtures. (208) Katsyuba, S. A.; Vener, M. V.; Zvereva, E. E.; Fei, Z.; Scopelliti, 4935
4868 ChemPhysChem 2016, 17, 3040−3046. R.; Laurenczy, G.; Yan, N.; Paunescu, E.; Dyson, P. J. How Strong Is 4936
4869 (191) Endo, T.; Nishikawa, K. Isomer Populations in Liquids for 1- Hydrogen Bonding in Ionic Liquids? Combined X-ray Crystallo- 4937
4870 Isopropyl-3-methylimidazolium Bromide and its Iodide and Their graphic, Infrared/Raman Spectroscopic, and Density Functional 4938
4871 Conformational Changes Accompanying the Crystallizing and Melting Theory Study. J. Phys. Chem. B 2013, 117, 9094−9105. 4939
4872 Processes. J. Phys. Chem. A 2008, 112, 7543−7550. (209) Knorr, A.; Fumino, K.; Bonsa, A. M.; Ludwig, R. Spectroscopic 4940
4873 (192) Hunt, P. A. Why Does a Reduction in Hydrogen Bonding Lead Evidence of ’Jumping and Pecking’ of Cholinium and H-bond 4941
4874 to an Increase in Viscosity for the 1-Butyl-2,3-dimethyl-imidazolium- Enhanced Cation-Cation Interaction in Ionic Liquids. Phys. Chem. 4942
4875 Based Ionic Liquids? J. Phys. Chem. B 2007, 111, 4844−4853. Chem. Phys. 2015, 17, 30978−30982. 4943
4876 (193) Fumino, K.; Wulf, A.; Ludwig, R. Strong, Localized, and (210) Weinhold, F.; Klein, R. A. Anti-Electrostatic Hydrogen Bonds. 4944
4877 Directional Hydrogen Bonds Fluidize Ionic Liquids. Angew. Chem., Int. Angew. Chem., Int. Ed. 2014, 53, 11214−11217. 4945
4878 Ed. 2008, 47, 8731−8734. (211) Gale, R. J.; Gilbert, B.; Osteryoung, R. A. Raman-Spectra of 4946
4879 (194) Izgorodina, E. I.; Maganti, R.; Armel, V.; Dean, P. M.; Pringle, Molten Aluminum Chloride: 1-Butylpyridinium Chloride Systems at 4947
4880 J. M.; Seddon, K. R.; MacFarlane, D. R. Understanding the Effect of Ambient Temperatures. Inorg. Chem. 1978, 17, 2728−2729. 4948
4881 the C2 Proton in Promoting Low Viscosities and High Conductivities (212) Gale, R. J.; Osteryoung, R. A. Infrared Spectral Investigations 4949
4882 in Imidazolium-Based Ionic Liquids: Part I. Weakly Coordinating of Room-Temperature Aluminum Chloride-1-Butylpyridinium Chlor- 4950
4883 Anions. J. Phys. Chem. B 2011, 115, 14688−14697. ide Melts. Inorg. Chem. 1980, 19, 2240−2242. 4951

AX DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

4952 (213) Tait, S.; Osteryoung, R. A. Infrared Study of Ambient- into the Thermal Phase Behavior of Protic Ionic Liquids. Phys. Chem. 5021
4953 Temperature Chloroaluminates as a Function of Melt Acidity. Inorg. Chem. Phys. 2012, 14, 16041−16046. 5022
4954 Chem. 1984, 23, 4352−4360. (233) Hagemann, H.; Bill, H. Raman Spectroscopic Study of 5023
4955 (214) Zhao, H.; Malhotra, S. V.; Luo, R. G. Preparation and EtNH3X (X = Cl,Br) and Several Deuterated Analogs. J. Chem. Phys. 5024
4956 characterization of three room-temperature ionic liquids. Phys. Chem. 1984, 80, 111−118. 5025
4957 Liq. 2003, 41, 487−492. (234) Prasad, P. S. R.; Bist, H. D. Raman Investigation of Order- 5026
4958 (215) Zheng, Y.; Zhuo, Z.; He, Y.; Mo, Q. Structure and Spectral Disorder Transition in Normal-Propylammonium Chloride. Solid State 5027
4959 Studies of N-Ethyl Pyridinium Bromide Ionic Liquids: DFT and ab Commun. 1990, 74, 885−888. 5028
4960 initio Study. Asian J. Chem. 2013, 25, 3233−3236. (235) Faria, L. F. O.; Penna, T. C.; Ribeiro, M. C. C. Raman 5029
4961 (216) Sashina, E. S.; Kashirskii, D. A.; Artamonova, T. V.; Myznikov, Spectroscopic Study of Temperature and Pressure Effects on the Ionic 5030
4962 L. V. 1-Alkyl-3-Methylpyridinium Halide Structural Features Studied Liquid Propylammonium Nitrate. J. Phys. Chem. B 2013, 117, 10905− 5031
4963 by Vibrational Spectroscopic Methods. Fibre Chem. 2015, 47, 171− 10912. 5032
4964 176. (236) Bodo, E.; Sferrazza, A.; Caminiti, R.; Mangialardo, S.; 5033
4965 (217) Rocher, N. M.; Izgorodina, E. I.; Ruether, T.; Forsyth, M.; Postorino, P. A prototypical Ionic Liquid Explored by ab initio 5034
4966 MacFarlane, D. R.; Rodopoulos, T.; Horne, M. D.; Bond, A. M. Molecular Dynamics and Raman Spectroscopy. J. Chem. Phys. 2013, 5035
4967 Aluminium Speciation in 1-Butyl-1-Methylpyrrolidinium Bis- 139, 144309. 5036
4968 (trifluoromethylsulfonyl)amide/AlCl3Mixtures. Chem. - Eur. J. 2009, (237) Luo, J.; Conrad, O.; Vankelecom, I. F. J. Physicochemical 5037
4969 15, 3435−3447. Properties of Phosphonium-Based and Ammonium-Based Protic Ionic 5038
4970 (218) Fujimori, T.; Fujii, K.; Kanzaki, R.; Chiba, K.; Yamamoto, H.; Liquids. J. Mater. Chem. 2012, 22, 20574−20579. 5039
4971 Umebayashi, Y.; Ishiguro, S.-i. Conformational Structure of Room (238) Berg, R. W.; Riisager, A.; Van Buu, O. N.; Fehrmann, R.; 5040
4972 Temperature Ionic Liquid N-Butyl-N-Methyl-pyrrolidinium Bis- Harris, P.; Tomaszowska, A. A.; Seddon, K. R. Crystal Structure, 5041
4973 (trifluoromethanesulfonyl)imide: Raman Spectroscopic Study and Vibrational Spectroscopy and ab Initio Density Functional Theory 5042
4974 DFT Calculations. J. Mol. Liq. 2007, 131−132, 216−224. Calculations on the Ionic Liquid forming 1,1,3,3-Tetramethylguanidi- 5043
4975 (219) Trequattrini, F.; Palumbo, O.; Gatto, S.; Appetecchi, G. B.; nium bis{(trifluoromethyl)sulfonyl}amide. J. Phys. Chem. B 2009, 113, 5044
4976 Paolone, A. A Computational and Experimental Study of the 8878−8886. 5045
4977 Conformers of Pyrrolidinium Ionic Liquid Cations Containing an (239) Horikawa, M.; Akai, N.; Kawai, A.; Shibuya, K. Vaporization of 5046
4978 Ethoxy Group in the Alkyl Side Chain. Adv. Chem. 2016, 2016, 1−9. Protic Ionic Liquids Studied by Matrix-Isolation Fourier Transform 5047
4979 (220) Menzies, A. C.; Mills, H. R. Raman Effect and Temperature I - Infrared Spectroscopy. J. Phys. Chem. A 2014, 118, 3280−3287. 5048
4980 Ammonium Chloride, Bromide, Iodide. Proc. R. Soc. London, Ser. A (240) Berg, R. W.; Riisager, A.; Fehrmann, R. Formation of an Ion- 5049
4981 1935, 148 (A864), 407−422. Pair Molecule with a Single NH+···Cl− Hydrogen Bond: Raman 5050
4982 (221) Holmes, F. T. Raman Spectrum of Crystalline Ammonium Spectra of 1,1,3,3-Tetramethylguanidinium Chloride in the Solid State, 5051
4983 Chloride. J. Chem. Phys. 1936, 4, 88−90. in Solution, and in the Vapor Phase. J. Phys. Chem. A 2008, 112, 5052
4984 (222) Greaves, T. L.; Drummond, C. J. Protic Ionic Liquids: Evolving 8585−8592. 5053
4985 Structure-Property Relationships and Expanding Applications. Chem. (241) Berg, R. W.; Riisager, A.; Van Buu, O. N.; Kristensen, S. B.; 5054
4986 Rev. 2015, 115, 11379−11448. Fehrmann, R.; Harris, P.; Brunetti, A. C. X-ray Crystal Structure, 5055
4987 (223) Palumbo, O.; Vitucci, F. M.; Trequattrini, F.; Paolone, A. A Raman Spectroscopy, and Ab Initio Density Functional Theory 5056
4988 Study of the Conformers of the N,N-Diethyl-N-Methyl-N-Propylam- Calculations on 1,1,3,3-Tetramethylguanidinium Bromide. J. Phys. 5057
4989 monium Ion by Means of Infrared Spectroscopy and DFT Chem. A 2010, 114, 13175−13181. 5058
4990 Calculations. Vib. Spectrosc. 2015, 80, 11−16. (242) Carper, W. R.; Langenwalter, K.; Nooruddin, N. S.; Kullman, 5059
4991 (224) Domańska, U.; Bogel-L̷ ukasik, R. Physicochemical Properties M. J.; Gerhard, D.; Wasserscheid, P. Aggregation Models of Potential 5060
4992 and Solubility of Alkyl-(2-hydroxyethyl)-dimethylammonium Bro- Cyclical Trimethylsulfonium Dicyanamide Ionic Liquid Clusters. J. 5061
4993 mide. J. Phys. Chem. B 2005, 109, 12124−12132. Phys. Chem. B 2009, 113, 2031−2041. 5062
4994 (225) Aparicio, S.; Atilhan, M.; Khraisheh, M.; Alcalde, R. Study on (243) Frenkel, M. L.; Kabo, G. J.; Marsh, K. N.; Roganov, G. N.; 5063
4995 Hydroxylammonium-Based Ionic Liquids. I. Characterization. J. Phys. Wilhoit, R. C. Thermodynamics of Organic Compounds in the Gas Phase; 5064
4996 Chem. B 2011, 115, 12473−12486. Thermodynamics Research Center: Texas, 1994; Vol. 1. 5065
4997 (226) Arkas, M.; Tsiourvas, D.; Paleos, C. M.; Skoulios, A. Smectic (244) Paulechka, Y. U.; Kabo, G. J.; Blokhin, A. V.; Vydrov, O. A.; 5066
4998 Mesophases from Dihydroxy Derivatives of Quaternary Alkylammo- Magee, J. W.; Frenkel, M. Thermodynamic Properties of 1-butyl-3- 5067
4999 nium Salts. Chem. - Eur. J. 1999, 5, 3202−3207. methylimidazolium Hexafluorophosphate in the Ideal Gas State. J. 5068
5000 (227) Campetella, M.; Bodo, E.; Caminiti, R.; Martino, A.; D’Apuzzo, Chem. Eng. Data 2003, 48, 457−462. 5069
5001 F.; Lupi, S.; Gontrani, L. Interaction and Dynamics of Ionic Liquids (245) Blokhin, A. V.; Paulechka, Y. U.; Strechan, A. A.; Kabo, G. J. 5070
5002 Based on Choline and Amino Acid Anions. J. Chem. Phys. 2015, 142, Physicochemical Properties, Structure, and Conformations of 1-Butyl- 5071
5003 234502. 3-methylimidazolium Bis(Trifluoromethanesulfonyl)imide C(4)mim 5072
5004 (228) Harmon, K. M.; Avci, G. F.; Thiel, A. C. Hydrogen-Bonding NTf2 Ionic Liquid. J. Phys. Chem. B 2008, 112, 4357−4364. 5073
5005 Part 21. Infrared Spectral Study of the High-Temperature Phases of (246) Paulechka, Y. U.; Kabo, G. J.; Emel’yanenko, V. N. Structure, 5074
5006 Choline Bromide and Choline Iodide. J. Mol. Struct. 1986, 145, 83−91. Conformations, Vibrations, and Ideal-Gas Properties of 1-Alkyl-3- 5075
5007 (229) Perkins, S. L.; Painter, P.; Colina, C. M. Molecular Dynamic methylimidazolium bis(trifluoromethylsulfonyl)imide Ionic Pairs and 5076
5008 Simulations and Vibrational Analysis of an Ionic Liquid Analogue. J. Constituent Ions. J. Phys. Chem. B 2008, 112, 15708−15717. 5077
5009 Phys. Chem. B 2013, 117, 10250−10260. (247) Leal, J. P.; Esperanca, J. M. S. S.; Minas da Piedade, M. E.; 5078
5010 (230) Bodo, E.; Postorino, P.; Mangialardo, S.; Piacente, G.; Canongia Lopes, J. N.; Rebelo, L. P. N.; Seddon, K. R. The Nature of 5079
5011 Ramondo, F.; Bosi, F.; Ballirano, P.; Caminiti, R. Structure of the Ionic Liquids in the Gas Phase. J. Phys. Chem. A 2007, 111, 6176− 5080
5012 Molten Salt Methyl Ammonium Nitrate Explored by Experiments and 6182. 5081
5013 Theory. J. Phys. Chem. B 2011, 115, 13149−13161. (248) Armstrong, J. P.; Hurst, C.; Jones, R. G.; Licence, P.; Lovelock, 5082
5014 (231) Bodo, E.; Mangialardo, S.; Ramondo, F.; Ceccacci, F.; K. R. J.; Satterley, C. J.; Villar-Garcia, I. J. Vapourisation of Ionic 5083
5015 Postorino, P. Unravelling the Structure of Protic Ionic Liquids with Liquids. Phys. Chem. Chem. Phys. 2007, 9, 982−990. 5084
5016 Theoretical and Experimental Methods: Ethyl-, Propyl- and (249) Dong, K.; Zhao, L.; Wang, Q.; Song, Y.; Zhang, S. Are Ionic 5085
5017 Butylammonium Nitrate Explored by Raman Spectroscopy and DFT Liquids Pairwise in Gas Phase? A Cluster Approach and in situ IR 5086
5018 Calculations. J. Phys. Chem. B 2012, 116, 13878−13888. Study. Phys. Chem. Chem. Phys. 2013, 15, 6034−6040. 5087
5019 (232) Henderson, W. A.; Fylstra, P.; De Long, H. C.; Trulove, P. C.; (250) Obi, E. I.; Leavitt, C. M.; Raston, P. L.; Moradi, C. P.; Flynn, S. 5088
5020 Parsons, S. Crystal Structure of the Ionic Liquid EtNH3NO3-Insights D.; Vaghjiani, G. L.; Boatz, J. A.; Chambreau, S. D.; Douberly, G. E. 5089

AY DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5090 Helium Nanodroplet Isolation and Infrared Spectroscopy of the (269) Dong, K.; Song, Y.; Liu, X.; Cheng, W.; Yao, X.; Zhang, S. 5157
5091 Isolated Ion-Pair 1-Ethyl-3-methylimidazolium bis- Understanding Structures and Hydrogen Bonds of Ionic Liquids at the 5158
5092 (trifluoromethylsulfonyl)imide. J. Phys. Chem. A 2013, 117, 9047− Electronic Level. J. Phys. Chem. B 2012, 116, 1007−1017. 5159
5093 9056. (270) Kato, T.; Takenaka, T. Raman-Study of Rotational Motion and 5160
5094 (251) Cooper, R.; Zolot, A. M.; Boatz, J. A.; Sporleder, D. P.; Stearns, Vibrational Dephasing Dynamics of NO3- in Molten Nitrates. Mol. 5161
5095 J. A. IR and UV Spectroscopy of Vapor-Phase Jet-Cooled Ionic Liquid Phys. 1985, 54, 1393−1414. 5162
5096 [emim]+ [Tf2N]−: Ion Pair Structure and Photodissociation Dynamics. (271) Griffiths, P. R. Far-Infrared Spectroscopy. In Handbook of 5163
5097 J. Phys. Chem. A 2013, 117, 12419−12428. Vibrational Spectroscopy; Chalmers, J. M., Griffiths, P. R., Eds.; John 5164
5098 (252) Fournier, J. A.; Wolke, C. T.; Johnson, C. J.; McCoy, A. B.; Wiley & Sons: Chichester, UK, 2002; Vol. 1, pp 229−239. 5165
5099 Johnson, M. A. Comparison of the Local Binding Motifs in the (272) Griffiths, P. R.; Homes, C. C. Instrumentation for Far-Infrared 5166

5100 Imidazolium-Based Ionic Liquids [EMIM][BF4] and [EMMIM][BF4] Spectroscopy. In Handbook of Vibrational Spectroscopy; Chalmers, J. 5167

5101 Through Cryogenic Ion Vibrational Predissociation Spectroscopy: M., Griffiths, P. R., Eds.; John Wiley & Sons: Chichester, UK, 2002; 5168

Unraveling the Roles of Anharmonicity and Intermolecular Vol. 1, pp 326−336. 5169


5102
(273) Fumino, K.; Wulf, A.; Ludwig, R. The Cation-Anion 5170
5103 Interactions. J. Chem. Phys. 2015, 142, No. 064306.
Interaction in Ionic Liquids Probed by Far-Infrared Spectroscopy. 5171
5104 (253) Hanke, K.; Kaufmann, M.; Schwaab, G.; Wolke, C. T.;
Angew. Chem., Int. Ed. 2008, 47, 3830−3834. 5172
5105 Gorlova, O.; Johnson, M. A.; Kar, B. P.; Sander, W.; Sanchez-Garcia,
(274) Fumino, K.; Wulf, A.; Ludwig, R. The Potential Role of 5173
5106 E.; Havenith, M. Understanding the Ionic Liquid [NC4111][NTf2] Hydrogen Bonding in Aprotic and Protic Ionic Liquids. Phys. Chem. 5174
5107 from Individual Building Blocks: An IR-Spectroscopic Study. Phys. Chem. Phys. 2009, 11, 8790−8794. 5175
5108 Chem. Chem. Phys. 2015, 17, 8518−8529. (275) Fumino, K.; Ludwig, R. Analyzing the Interaction Energies 5176
5109 (254) Booth, R. S.; Annesley, C. J.; Young, J. W.; Vogelhuber, K. M.; between Cation and Anion in Ionic Liquids: the Subtle Balance 5177
5110 Boatz, J. A.; Stearns, J. A. Identification of Multiple Conformers of the between Coulomb Forces and Hydrogen Bonding. J. Mol. Liq. 2014, 5178
5111 Ionic Liquid [emim][tf2n] in the Gas Phase Using IR/UV Action 192, 94−102. 5179
5112 Spectroscopy. Phys. Chem. Chem. Phys. 2016, 18, 17037−17043. (276) Fumino, K.; Wulf, A.; Ludwig, R. Hydrogen Bonding in Protic 5180
5113 (255) Berg, R. W.; Canongia Lopes, J. N.; Ferreira, R.; Rebelo, L. P. Ionic Liquids: Reminiscent of Water. Angew. Chem., Int. Ed. 2009, 48, 5181
5114 N.; Seddon, K. R.; Tomaszowska, A. A. Raman Spectroscopic Study of 3184−3186. 5182
5115 the Vapor Phase of l-Methylimidazolium Ethanoate, a Protic Ionic (277) Fumino, K.; Reichert, E.; Wittler, K.; Hempelmann, R.; 5183
5116 Liquid. J. Phys. Chem. A 2010, 114, 10834−10841. Ludwig, R. Low-Frequency Vibrational Modes of Protic Molten Salts 5184
5117 (256) Materazzi, S.; Vecchio, S. Evolved Gas Analysis by Infrared and Ionic Liquids: Detecting and Quantifying Hydrogen Bonds. 5185
5118 Spectroscopy. Appl. Spectrosc. Rev. 2010, 45, 241−273. Angew. Chem., Int. Ed. 2012, 51, 6236−6240. 5186
5119 (257) Chambreau, S. D.; Schenk, A. C.; Sheppard, A. J.; Yandek, G. (278) Yamada, T.; Tominari, Y.; Tanaka, S.; Mizuno, M.; Fukunaga, 5187
5120 R.; Vaghjiani, G. L.; Maciejewski, J.; Koh, C. J.; Golan, A.; Leone, S. R. K. Vibration Modes at Terahertz and Infrared Frequencies of Ionic 5188
5121 Thermal Decomposition Mechanisms of Alkylimidazolium Ionic Liquids Consisting of an Imidazolium Cation and a Halogen Anion. 5189
5122 Liquids with Cyano-Functionalized Anions. J. Phys. Chem. A 2014, Materials 2014, 7, 7409−7422. 5190
5123 118, 11119−11132. (279) Wulf, A.; Fumino, K.; Ludwig, R. Spectroscopic Evidence for 5191
5124 (258) Liaw, H.-J.; Chen, C.-C.; Chen, Y.-C.; Chen, J.-R.; Huang, S.- an Enhanced Anion-Cation Interaction from Hydrogen Bonding in 5192
5125 K.; Liu, S.-N. Relationship Between Flash Point of Ionic Liquids and Pure Imidazolium Ionic Liquids. Angew. Chem., Int. Ed. 2010, 49, 449− 5193
5126 Their Thermal Decomposition. Green Chem. 2012, 14, 2001−2008. 453. 5194
5127 (259) Palumbo, O.; Trequattrini, F.; Vitucci, F. M.; Navarra, M. A.; (280) Fumino, K.; Wittler, K.; Ludwig, R. The Anion Dependence of 5195
5128 Panero, S.; Paolone, A. An Infrared Spectroscopy Study of the the Interaction Strength between Ions in Imidazolium-Based Ionic 5196
5129 Conformational Evolution of the Bis(trifluoromethanesulfonyl) imide Liquids Probed by Far-Infrared Spectroscopy. J. Phys. Chem. B 2012, 5197
5130 Ion in the Liquid and in the Glass State. Adv. Condens. Matter Phys. 116, 9507−9511. 5198
5131 2015, 2015, 176067. (281) Fumino, K.; Fossog, V.; Wittler, K.; Hempelmann, R.; Ludwig, 5199

5132 (260) Dommert, F.; Wendler, K.; Berger, R.; Delle Site, L.; Holm, C. R. Dissecting Anion-Cation Interaction Energies in Protic Ionic 5200

Force Fields for Studying the Structure and Dynamics of Ionic Liquids. Angew. Chem., Int. Ed. 2013, 52, 2368−2372. 5201
5133
(282) Fumino, K.; Wulf, A.; Verevkin, S. P.; Heintz, A.; Ludwig, R. 5202
5134 Liquids: A Critical Review of Recent Developments. ChemPhysChem
Estimating Enthalpies of Vaporization of Imidazolium-Based Ionic 5203
5135 2012, 13, 1625−1637.
Liquids from Far-Infrared Measurements. ChemPhysChem 2010, 11, 5204
5136 (261) Borodin, O. Polarizable Force Field Development and
1623−1626. 5205
5137 Molecular Dynamics Simulations of Ionic Liquids. J. Phys. Chem. B
(283) Roth, C.; Peppel, T.; Fumino, K.; Köckerling, M.; Ludwig, R. 5206
5138 2009, 113, 11463−11478. The Importance of Hydrogen Bonds for the Structure of Ionic 5207
5139 (262) Canongia Lopes, J. N.; Padua, A. A. H. Using Spectroscopic Liquids: Single-Crystal X-ray Diffraction and Transmission and 5208
5140 Data on Imidazolium Cation Conformations to Test a Molecular Attenuated Total Reflection Spectroscopy in the Terahertz Region. 5209
5141 Force Field for Ionic Liquids. J. Phys. Chem. B 2006, 110, 7485−7489. Angew. Chem., Int. Ed. 2010, 49, 10221−10224. 5210
5142 (263) Allerhand, A.; Schleyer, P. v. R. Solvent Effects in Infrared (284) Fumino, K.; Fossog, V.; Stange, P.; Paschek, D.; Hempelmann, 5211
5143 Spectroscopic Studies of Hydrogen Bonding. J. Am. Chem. Soc. 1963, R.; Ludwig, R. Controlling the Subtle Energy Balance in Protic Ionic 5212
5144 85, 371−380. Liquids: Dispersion Forces Compete with Hydrogen Bonds. Angew. 5213
5145 (264) Pimentel, G. C.; McClellan, A. L. The Hydrogen Bond; W. H. Chem., Int. Ed. 2015, 54, 2792−2795. 5214
5146 Freeman & Co.: San Francisco, 1960. (285) Ludwig, R. The Effect of Dispersion Forces on the Interaction 5215
5147 (265) Jeffrey, G. A. An Introduction to Hydrogen Bonding; Oxford Energies and Far Infrared Spectra of Protic Ionic Liquids. Phys. Chem. 5216
5148 University Press: Oxford, 1997. Chem. Phys. 2015, 17, 13790−13793. 5217
5149 (266) Steiner, T. The Hydrogen Bond in the Solid State. Angew. (286) Zaitsau, D. H.; Emel’yanenko, V. N.; Stange, P.; Schick, C.; 5218
5150 Chem., Int. Ed. 2002, 41, 48−76. Verevkin, S. P.; Ludwig, R. Dispersion and Hydrogen Bonding Rule: 5219
5151 (267) Buckingham, A. D.; Del Bene, J. E.; McDowell, S. A. C. The Why the Vaporization Enthalpies of Aprotic Ionic Liquids Are 5220
5152 Hydrogen Bond. Chem. Phys. Lett. 2008, 463, 1−10. Signifcantly Lager than those of Protic Ionic Liquids. Angew. Chem., 5221
5153 (268) Pitawala, J.; Scheers, J.; Jacobsson, P.; Matic, A. Physical Int. Ed. 2016, 55, 11682−11686. 5222
5154 Properties, Ion-Ion Interactions, and Conformational States of Ionic (287) Buffeteau, T.; Grondin, J.; Danten, Y.; Lassègues, J.-C. 5223
5155 Liquids with Alkyl-Phosphonate Anions. J. Phys. Chem. B 2013, 117, Imidazolium-Based Ionic Liquids: Quantitative Aspects in the Far- 5224
5156 8172−8179. Infrared Region. J. Phys. Chem. B 2010, 114, 7587−7592. 5225

AZ DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5226 (288) Wulf, A.; Fumino, K.; Ludwig, R.; Taday, P. F. Combined (309) Kojima, S.; Novikov, V. N. Correlation of Temperature 5294
5227 THz, FIR and Raman Spectroscopy Studies of Imidazolium-Based Dependence of Quasielastic-Light-Scattering Intensity and Alpha- 5295
5228 Ionic Liquids Covering the Frequency Range 2−300 cm(−1). Relaxation Time. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 5296
5229 ChemPhysChem 2010, 11, 349−353. 222−227. 5297
5230 (289) Winterling, G. Very-Low-Frequency Raman-Scattering in (310) Yannopoulos, S. N.; Papatheodorou, G. N.; Fytas, G. Low- 5298
5231 Vitreous Silica. Physi. Rev. B 1975, 12, 2432−2440. Energy Excitations in Noncrystalline Arsenic Trioxide. J. Chem. Phys. 5299
5232 (290) Gochiyaev, V. Z.; Malinovsky, V. K.; Novikov, V. N.; Sokolov, 1997, 107, 1341−1349. 5300
5233 A. P. Structure of the Rayleigh Line Wing in Highly Viscous Liquids. (311) Ribeiro, M. C. C. Low-frequency Raman Spectra and Fragility 5301
5234 Philos. Mag. B 1991, 63, 777−787. of Imidazolium Ionic Liquids. J. Chem. Phys. 2010, 133, 024503. 5302
5235 (291) Kojima, S. Low-Frequency Raman Investigation of the Liquid- (312) Ribeiro, M. C. C. Intermolecular Vibrations and Fast 5303
5236 Glass Transition in Glycerol. Phys. Rev. B: Condens. Matter Mater. Phys. Relaxations in Supercooled Ionic Liquids. J. Chem. Phys. 2011, 134, 5304
5237 1993, 47, 2924−2927. 244507. 5305
5238 (292) Sokolov, A. P.; Kisliuk, A.; Quitmann, D.; Kudlik, A.; Rossler, (313) Kinoshita, S.; Kai, Y.; Yamaguchi, M.; Yagi, T. Direct 5306
5239 E. The Dynamics of Strong and Fragile Glass Formers: Vibrational and Comparison Between Ultrafast Optical Kerr-Effect and High- 5307
5240 Relaxation Contributions. J. Non-Cryst. Solids 1994, 172−174, 138− Resolution Light Scattering Spectroscopy. Phys. Rev. Lett. 1995, 75, 5308
5241 153. 148−151. 5309
5242 (293) Alba-Simionesco, C.; Krakoviack, V.; Krauzman, M.; Migliardo, (314) Yannopoulos, S. N. The Frequency-Dependent Depolarization 5310
5243 P.; Romain, F. Low-Wavenumber Raman Scattering of Molecular Ratio of the Low-Frequency Raman Scattering of Two Inorganic 5311
5244 Glass-Forming Liquids. J. Raman Spectrosc. 1996, 27, 715−721. Systems in Their Glassy, Supercooled, and Molten States. J. Chem. 5312
5245 (294) Papatheodorou, G. N.; Kalogrianitis, S. G.; Mihopoulos, T. G.; Phys. 2000, 113, 5868−5872. 5313
5246 Pavlatou, E. A. Isotropic and Anisotropic Raman Scattering from (315) Hyun, B. R.; Dzyuba, S. V.; Bartsch, R. A.; Quitevis, E. L. 5314
5247 Molten LiCl-CsCl Mixtures: Composition and Temperature Effects. J. Intermolecular Dynamics of Room-Temperature Ionic Liquids: 5315
5248 Chem. Phys. 1996, 105, 2660−2667. Femtosecond Optical Kerr Effect Measurements on 1-Alkyl-3- 5316
5249 (295) Dracopoulos, V.; Papatheodorou, G. N. Isotropic and methylimidazolium Bis((trifluoromethyl)sulfonyl)imides. J. Phys. 5317
5250 Anisotropic Raman Scattering from Molten Alkali-Metal Fluorides. Chem. A 2002, 106, 7579−7585. 5318
5251 Phys. Chem. Chem. Phys. 2000, 2, 2021−2025. (316) Giraud, G.; Gordon, C. M.; Dunkin, I. R.; Wynne, K. The 5319
5252 (296) Kirillov, S. A.; Pavlatou, E. A.; Papatheodorou, G. N. Effects of Anion and Cation Substitution on the Ultrafast Solvent 5320
5253 Instantaneous Collision Complexes in Molten Alkali Halides: Dynamics of Ionic Liquids: A Time-Resolved Optical Kerr-Effect 5321
5254 Picosecond Dynamics from Low-Frequency Raman Data. J. Chem. Spectroscopic Study. J. Chem. Phys. 2003, 119, 464−477. 5322
5255 Phys. 2002, 116, 9341−9351. (317) Rajesh Rajian, J.; Li, S. F.; Bartsch, R. A.; Quitevis, E. L. 5323
5256 (297) Aliotta, F.; Maisano, G.; Migliardo, P.; Vasi, C.; Wanderlingh, Temperature-Dependence of the Low-Frequency Spectrum of 1- 5324
5257 F.; Smith, G. P.; Triolo, R. Vibrational Dynamics of Glassy and Molten pentyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide Stud- 5325
5258 ZnCl2. J. Chem. Phys. 1981, 75, 613−618. ied by Optical Kerr Effect Spectroscopy. Chem. Phys. Lett. 2004, 393, 5326
5259 (298) Yannopoulos, S. N.; Papatheodorou, G. N. Critical 372−377. 5327
5260 Experimental Facts Pertaining to Models and Associated Universalities (318) Shirota, H.; Funston, A. M.; Wishart, J. F.; Castner, E. W. 5328
5261 for Low-Frequency Raman Scattering in Inorganic Glass Formers. Ultrafast Dynamics of Pyrrolidinium Cation Ionic Liquids. J. Chem. 5329
5262 Phys. Rev. B: Condens. Matter Mater. Phys. 2000, 62, 3728−3734. Phys. 2005, 122, 184512. 5330
5263 (299) Kirillov, S. A.; Yannopoulos, S. N. Charge-Current (319) Xiao, D.; Rajian, J. R.; Cady, A.; Li, S.; Bartsch, R. A.; Quitevis, 5331
5264 Contribution to Low-Frequency Raman Scattering from Glass- E. L. Nanostructural Organization and Anion Effects on the 5332
5265 Forming Ionic Liquids. Phys. Rev. B: Condens. Matter Mater. Phys. Temperature Dependence of the Optical Kerr Effect Spectra of 5333
5266 2000, 61, 11391−11399. Ionic Liquids. J. Phys. Chem. B 2007, 111, 4669−4677. 5334
5267 (300) Kalampounias, A. G.; Papatheodorou, G. N.; Yannopoulos, S. (320) Russina, O.; Triolo, A.; Gontrani, L.; Caminiti, R.; Xiao, D.; 5335
5268 N. Light Scattering from Glass-Forming Molten Salts. Z. Naturforsch., Hines, L. G., Jr.; Bartsch, R. A.; Quitevis, E. L.; Pleckhova, N.; Seddon, 5336
5269 A: Phys. Sci. 2002, 57, 65−70. K. R. Morphology and Intermolecular Dynamics of 1-Alkyl-3- 5337
5270 (301) Ribeiro, M. C. C.; de Oliveira, L. F. C.; Goncalves, N. S. Boson methylimidazolium Bis{(trifluoromethane)sulfonyl}amide Ionic 5338
5271 Peak in the Room-Temperature Molten Salt Tetra(N-Butyl)- Liquids: Structural and Dynamic Evidence of Nanoscale Segregation. 5339
5272 ammonium Croconate. Phys. Rev. B: Condens. Matter Mater. Phys. J. Phys.: Condens. Matter 2009, 21, 424121. 5340
5273 2001, 63, 104303. (321) Zhong, Q.; Fourkas, J. T. Optical Kerr Effect Spectroscopy of 5341
5274 (302) Ribeiro, M. C. C. Correlation Between Quasielastic Raman Simple Liquids. J. Phys. Chem. B 2008, 112, 15529−15539. 5342
5275 Scattering and Configurational Entropy in an Ionic Liquid. J. Phys. (322) Shirota, H. Comparison of Low-Frequency Spectra between 5343
5276 Chem. B 2007, 111, 5008−5015. Aromatic and Nonaromatic Cation Based Ionic Liquids Using 5344
5277 (303) Shuker, R.; Gammon, R. W. Raman-Scattering Selection-Rule Femtosecond Raman-Induced Kerr Effect Spectroscopy. ChemPhy- 5345
5278 Breaking and Density of States in Amorphous Materials. Phys. Rev. sChem 2012, 13, 1638−1648. 5346
5279 Lett. 1970, 25, 222−225. (323) Sonnleitner, T.; Turton, D. A.; Hefter, G.; Ortner, A.; 5347
5280 (304) Shuker, R.; Gammon, R. W. Low-Frequency Vibrational Light Waselikowski, S.; Walther, M.; Wynne, K.; Buchner, R. Ultra- 5348
5281 Scattering in Viscous Liquids. J. Chem. Phys. 1971, 55, 4784−4788. Broadband Dielectric and Optical Kerr-Effect Study of the Ionic 5349
5282 (305) Surovtsev, N. V.; Sokolov, A. P. Frequency Behavior of Raman Liquids Ethyl and Propylammonium Nitrate. J. Phys. Chem. B 2015, 5350
5283 Coupling Coefficient in Glasses. Phys. Rev. B: Condens. Matter Mater. 119, 8826−8841. 5351
5284 Phys. 2002, 66, 054205−054211. (324) Krüger, M.; Bründermann, E.; Funkner, S.; Weingärtner, H.; 5352
5285 (306) Yannopoulos, S. N.; Andrikopoulos, K. S.; Ruocco, G. On the Havenith, M. Communications: Polarity Fluctuations of the Protic 5353
5286 Analysis of the Vibrational Boson Peak and Low-Energy Excitations in Ionic Liquid Ethylammonium Nitrate in the Terahertz Regime. J. 5354
5287 Glasses. J. Non-Cryst. Solids 2006, 352, 4541−4551. Chem. Phys. 2010, 132, 101101. 5355
5288 (307) Buchenau, U.; Zorn, R. A Relation Between Fast and Slow (325) Clarke, J. H. R. Raman Spectra of Lattice Vibrations in Liquid 5356
5289 Motions in Glassy and Liquid Selenium. Europhys. Lett. 1992, 18, and Solid Monovalent Metal Nitrates. Chem. Phys. Lett. 1969, 4, 39− 5357
5290 523−528. 42. 5358
5291 (308) Krüger, M.; Soltwisch, M.; Petscherizin, I.; Quitmann, D. (326) Kirillov, S. A.; Nielsen, O. F. Boson Peak in the Low- 5359
5292 Light-Scattering from Disorder and Glass-Transition-Dynamics in Frequency Raman Spectra of an Ordinary Liquid. J. Mol. Struct. 2000, 5360
5293 GeSBr2. J. Chem. Phys. 1992, 96, 7352−7363. 526, 317−321. 5361

BA DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5362 (327) Surovtsev, N. V.; Pugachev, A. M.; Nenashev, B. G.; (348) Wilkes, J. S.; Zaworotko, M. J. Air and Water Stable 1-Ethyl-3- 5429
5363 Malinovsky, V. K. Low-frequency Raman Scattering in As2S3 Glass methylimidazolium Based Ionic Liquids. J. Chem. Soc., Chem. Commun. 5430
5364 Former around the Liquid-glass Transition. J. Phys.: Condens. Matter 1992, 13, 965−967. 5431
5365 2003, 15, 7651−7662. (349) Gordon, C. M.; Holbrey, J. D.; Kennedy, A. R.; Seddon, K. R. 5432
5366 (328) Kalampounias, A. G.; Yannopoulos, S. N.; Papatheodorou, G. Ionic Liquid Crystals: Hexafluorophosphate Salts. J. Mater. Chem. 5433
5367 N. A Low-Frequency Raman Study of Glassy, Supercooled and Molten 1998, 8, 2627−2636. 5434
5368 Silica and the Preservation of the Boson Peak in the Equilibrium (350) Holbrey, J. D.; Seddon, K. R. The Phase Behaviour of 1-Alkyl- 5435
5369 Liquid State. J. Non-Cryst. Solids 2006, 352, 4619−4624. 3-methylimidazolium Tetrafluoroborates; Ionic Liquids and Ionic 5436
5370 (329) Kalampounias, A. G. Low-energy Excitations in a Low-Viscous Liquid Crystals. J. Chem. Soc., Dalton Trans. 1999, No. 13, 2133−2140. 5437
5371 Glass-Forming Liquid. Bull. Mater. Sci. 2013, 36, 607−611. (351) MacFarlane, D. R.; Meakin, P.; Sun, J.; Amini, N.; Forsyth, M. 5438
5372 (330) Li, J.; Wang, I.; Fruchey, K.; Fayer, M. D. Dynamics in Pyrrolidinium Imides: A New Family of Molten Salts and Conductive 5439

5373 Supercooled Ionic Organic Liquids and Mode Coupling Theory Plastic Crystal Phases. J. Phys. Chem. B 1999, 103, 4164−4170. 5440

5374 Analysis. J. Phys. Chem. A 2006, 110, 10384−10391. (352) Dupont, J. On the Solid, Liquid and Solution Structural 5441

(331) Balucani, U.; Zoppi, M. Dynamics of the Liquid State; Organization of Imidazolium Ionic Liquids. J. Braz. Chem. Soc. 2004, 5442
5375
5376 Clarendon Press: Oxford, 1994. 15, 341−350. 5443

(332) Sette, F.; Krisch, M. H.; Masciovecchio, C.; Ruocco, G.; (353) Goossens, K.; Lava, K.; Nockemann, P.; Van Hecke, K.; Van 5444
5377
Meervelt, L.; Driesen, K.; Görller-Walrand, C.; Binnemans, K.; 5445
5378 Monaco, G. Dynamics of Glasses and Glass-Forming Liquids Studied
Cardinaels, T. Pyrrolidinium Ionic Liquid Crystals. Chem. - Eur. J. 5446
5379 by Inelastic X-ray Scattering. Science 1998, 280, 1550−1555.
2009, 15, 656−674. 5447
5380 (333) Ruocco, G.; Sette, F. High-Frequency Vibrational Dynamics in
(354) Binnemans, K. Ionic Liquid Crystals. Chem. Rev. 2005, 105, 5448
5381 Glasses. J. Phys.: Condens. Matter 2001, 13, 9141−9164. 4148−4204. 5449
5382 (334) Monaco, G. High-Resolution Inelastic X-Ray Scattering to (355) De Roche, J.; Gordon, C. M.; Imrie, C. T.; Ingram, M. D.; 5450
5383 Study the High-Frequency Atomic Dynamics of Disordered Systems. Kennedy, A. R.; Lo Celso, F.; Triolo, A. Application of 5451
5384 C. R. Phys. 2008, 9, 608−623. Complementary Experimental Techniques to Characterization of the 5452
5385 (335) Shintani, H.; Tanaka, H. Universal Link between the Boson Phase Behavior of [C16mim][PF6] and [C14mim][PF6]. Chem. 5453
5386 Peak and Transverse Phonons in Glass. Nat. Mater. 2008, 7, 870−877. Mater. 2003, 15, 3089−3097. 5454
5387 (336) Chumakov, A. I.; Monaco, G.; Monaco, A.; Crichton, W. A.; (356) Goossens, K.; Lava, K.; Bielawski, C. W.; Binnemans, K. Ionic 5455
5388 Bosak, A.; Rüffer, R.; Meyer, A.; Kargl, F.; Comez, L.; Fioretto, D.; Liquid Crystals: Versatile Materials. Chem. Rev. 2016, 116, 4643− 5456
5389 et al. Equivalence of the Boson Peak in Glasses to the Transverse 4807. 5457
5390 Acoustic van Hove Singularity in Crystals. Phys. Rev. Lett. 2011, 106, (357) Endo, T.; Kato, T.; Tozaki, K.-i.; Nishikawa, K. Phase 5458
5391 225501. Behaviors of Room Temperature Ionic Liquid Linked with Cation 5459
5392 (337) Kumar, P.; Wikfeldt, K. T.; Schlesinger, D.; Pettersson, L. G. Conformational Changes: 1-Butyl-3-methylimidazolium Hexafluoro- 5460
5393 M.; Stanley, H. E. The Boson Peak in Supercooled Water. Sci. Rep. phosphate. J. Phys. Chem. B 2010, 114, 407−411. 5461
5394 2013, 3, 1980. (358) Endo, T.; Murata, H.; Imanari, M.; Mizushima, N.; Seki, H.; 5462
5395 (338) Chumakov, A. I.; Monaco, G. Understanding the Atomic Nishikawa, K. NMR Study of Cation Dynamics in Three Crystalline 5463
5396 Dynamics and Thermodynamics of Glasses: Status and Outlook. J. States of 1-Butyl-3-methylimidazolium Hexafluorophosphate Exhibit- 5464
5397 Non-Cryst. Solids 2015, 407, 126−132. ing Crystal Polymorphism. J. Phys. Chem. B 2012, 116, 3780−3788. 5465
5398 (339) Buchenau, U.; Wischnewski, A.; Richter, D.; Frick, B. Is the (359) Endo, T.; Nishikawa, K. Thermal Phase Behavior of 1-Butyl-3- 5466
5399 Fast Process at the Glass Transition Mainly Due to Long Wavelength Methylimidazolium Hexafluorophosphate: Simultaneous Measure- 5467
5400 Excitations? Phys. Rev. Lett. 1996, 77, 4035−4038. ments of the Melting of Two Polymorphic Crystals by Raman 5468
5401 (340) Sokolov, A. P. Vibrations at the boson peak: Random- and Spectroscopy and Calorimetry. Chem. Phys. Lett. 2013, 584, 79−82. 5469
5402 Coherent-Phase Contributions. J. Phys.: Condens. Matter 1999, 11, (360) Herstedt, M.; Henderson, W. A.; Smirnov, M.; Ducasse, L.; 5470
5403 A213−A218. Servant, L.; Talaga, D.; Lassègues, J. C. Conformational Isomerism and 5471
5404 (341) Urahata, S. M.; Ribeiro, M. C. C. Single Particle Dynamics in Phase Transitions in Tetraethylammonium bis- 5472
5405 Ionic Liquids of 1-alkyl-3-methylimidazolium Cations. J. Chem. Phys. (trifluoromethanesulfonyl)imide Et4NTFSI. J. Mol. Struct. 2006, 783, 5473
5406 2005, 122, 024511. 145−156. 5474
5407 (342) Cavalcante, A. d. O.; Ribeiro, M. C. C.; Skaf, M. S. (361) Bakker, A.; Gejji, S.; Lindgren, J.; Hermansson, K.; Probst, M. 5475

5408 Polarizability Effects on the Structure and Dynamics of Ionic Liquids. M. Contact Ion Pair Formation and Ether Oxygen Coordination in the 5476
5409 J. Chem. Phys. 2014, 140, 114108. Polymer Electrolytes M[N(CF3SO2)2]2PEOn for M = Mg, Ca, Sr and 5477

5410 (343) Sunda, A. P.; Mondal, A.; Balasubramanian, S. Atomistic Ba. Polymer 1995, 36, 4371−4378. 5478

Simulations of Ammonium-Based Protic Ionic Liquids: Steric Effects (362) Henderson, W. A.; Herstedt, M.; Young, V. G.; Passerini, S.; 5479
5411
De Long, H. C.; Trulove, P. C. New Disordering Mode for TFSI- 5480
5412 on Structure, Low-Frequency Vibrational Modes and Electrical
anions: The Nonequilibrium, Plastic Crystalline Structure of 5481
5413 Conductivity. Phys. Chem. Chem. Phys. 2015, 17, 4625−4633.
Et4NTFSI. Inorg. Chem. 2006, 45, 1412−1414. 5482
5414 (344) Hu, Z.; Huang, X.; Annapureddy, H. V. R.; Margulis, C. J.
(363) Martinelli, A. Conformational Changes and Phase Behaviour in 5483
5415 Molecular Dynamics Study of the Temperature-Dependent Optical the Protic Ionic Liquid 1-Ethylimidazolium Bis- 5484
5416 Kerr Effect Spectra and Intermolecular Dynamics of Room Temper- (trifluoromethylsulfonyl)imide in the Bulk and Nano-Confined State. 5485
5417 ature Ionic Liquid 1-Methoxyethylpyridinium Dicyanoamide. J. Phys. Eur. J. Inorg. Chem. 2015, 2015, 1300−1308. 5486
5418 Chem. B 2008, 112, 7837−7849. (364) Capitani, F.; Gatto, S.; Postorino, P.; Palumbo, O.; 5487
5419 (345) Shirota, H.; Castner, E. W. Physical Properties and Trequattrini, F.; Deutsch, M.; Brubach, J. B.; Roy, P.; Paolone, A. 5488
5420 Intermolecular Dynamics of an Ionic Liquid Compared with Its The Complex Dance of the Two Conformers of Bis- 5489
5421 Isoelectronic Neutral Binary Solution. J. Phys. Chem. A 2005, 109, (trifluoromethanesulfonyl)imide as a Function of Pressure and 5490
5422 9388−9392. Temperature. J. Phys. Chem. B 2016, 120, 1312. 5491
5423 (346) Elola, M. D.; Ladanyi, B. M. Polarizability Response in Polar (365) Saouane, S.; Norman, S. E.; Hardacre, C.; Fabbiani, F. P. A. 5492
5424 Solvents: Molecular-Dynamics Simulations of Acetonitrile And Pinning Down the Solid-State Polymorphism of the Ionic Liquid 5493
5425 Chloroform. J. Chem. Phys. 2005, 122, 224506. [bmim][PF6]. Chem. Sci. 2013, 4, 1270−1280. 5494
5426 (347) McGreevy, R. L. Experimental Studies of the Structure and (366) Xu, W.; Cooper, E. I.; Angell, C. A. Ionic liquids: Ion 5495
5427 Dynamics of Molten Alkali and Alkaline-Earth Halides. Solid State Mobilities, Glass Temperatures, and Fragilities. J. Phys. Chem. B 2003, 5496
5428 Phys. 1987, 40, 247−325. 107, 6170−6178. 5497

BB DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5498 (367) Mudring, A.-V. Solidification of Ionic Liquids: Theory and (386) Yoshimura, Y.; Shigemi, M.; Takaku, M.; Yamamura, M.; 5567
5499 Techniques. Aust. J. Chem. 2010, 63, 544−564. Takekiyo, T.; Abe, H.; Hamaya, N.; Wakabayashi, D.; Nishida, K.; 5568
5500 (368) Nishikawa, K.; Wang, S. L.; Katayanagi, H.; Hayashi, S.; Funamori, N.; et al. Stability of the Liquid State of Imidazolium-Based 5569
5501 Hamaguchi, H. O.; Koga, Y.; Tozaki, K. I. Melting and Freezing Ionic Liquids under High Pressure at Room Temperature. J. Phys. 5570
5502 Behaviors of Prototype Ionic Liquids, 1-Butyl-3-methylimidazolium Chem. B 2015, 119, 8146−8153. 5571
5503 Bromide and Its Chloride, Studied by Using a Nano-Watt Differential (387) Faria, L. F. O.; Ribeiro, M. C. C. Phase Transitions of Triflate- 5572
5504 Scanning Calorimeter. J. Phys. Chem. B 2007, 111, 4894−4900. Based Ionic Liquids under High Pressure. J. Phys. Chem. B 2015, 119, 5573
5505 (369) Nishikawa, K.; Wang, S.; Tozaki, K.-i. Rhythmic Melting and 14315−14322. 5574
5506 Crystallizing of Ionic Liquid 1-butyl-3-methylimidazolium bromide. (388) Li, H.; Wang, Z.; Chen, L.; Wu, J.; Huang, H.; Yang, K.; Wang, 5575
5507 Chem. Phys. Lett. 2008, 458, 88−91. Y.; Su, L.; Yang, G. Kinetic Effect on Pressure-Induced Phase 5576
5508 (370) Nishikawa, K.; Tozaki, K.-i. Intermittent Crystallization of an Transitions of Room Temperature Ionic Liquid, 1-Ethyl-3-methyl- 5577
5509 Ionic Liquid: 1-isopropyl-3-methylimidazolium bromide. Chem. Phys. imidazolium Trifluoromethanesulfonate. J. Phys. Chem. B 2015, 119, 5578
5510 Lett. 2008, 463, 369−372. 14245−14251. 5579
5511 (371) Nishikawa, K.; Wang, S.; Endo, T.; Tozaki, K.-i. Melting and (389) Yoshimura, Y.; Takekiyo, T.; Abe, H.; Hamaya, N. High- 5580
5512 Crystallization Behaviors of an Ionic Liquid, 1-Isopropyl-3-methyl- Pressure Phase Behavior of the Room Temperature Ionic Liquid 1- 5581
5513 imidazolium Bromide, Studied by Using Nanowatt-Stabilized Differ- ethyl-3-methylimidazolium Nitrate. J. Mol. Liq. 2015, 206, 89−94. 5582
5514 ential Scanning Calorimetry. Bull. Chem. Soc. Jpn. 2009, 82, 806−812. (390) Zhu, X.; Yuan, C.; Li, H.; Zhu, P.; Su, L.; Yang, K.; Wu, J.; 5583
5515 (372) Okajima, H.; Hamaguchi, H. Unusually Long trans/gauche Yang, G.; Liu, J. Successive Disorder to Disorder Phase Transitions in 5584
5516 Conformational Equilibration Time during the Melting Process of Ionic Liquid [HMIM][BF4] Under High Pressure. J. Mol. Struct. 2016, 5585
5517 BmimC1, a Prototype Ionic Liquid. Chem. Lett. 2011, 40, 1308−1309. 1106, 70−75. 5586
5518 (373) Imanari, M.; Fujii, K.; Endo, T.; Seki, H.; Tozaki, K.-i.; (391) Imai, Y.; Abe, H.; Goto, T.; Takekiyo, T.; Yoshimura, Y. 5587
5519 Nishikawa, K. Ultraslow Dynamics at Crystallization of a Room- Pressure-Induced Raman Spectral Changes of N,N,diethyl-N-methyl- 5588
5520 Temperature Ionic Liquid, 1-Butyl-3-methylimidazolium Bromide. J. N-(2-methoxyethyl) Ammonium Tetrafluoroborate. High Pressure Res. 5589
5521 Phys. Chem. B 2012, 116, 3991−3997. 2009, 29, 536−541. 5590
5522 (374) Chang, H.-C.; Chang, C.-Y.; Su, J.-C.; Chu, W.-C.; Jiang, J.-C.; (392) Capitani, F.; Fasolato, C.; Mangialardo, S.; Signorelli, S.; 5591
5523 Lin, S. H. Conformations of 1-butyl-3-methylimidazolium chloride Gontrani, L.; Postorino, P. Heterogeneity of Propyl-ammonium 5592
5524 Probed by High Pressure Raman Spectroscopy. Int. J. Mol. Sci. 2006, 7, Nitrate Solid Phases Obtained Under High Pressure. J. Phys. Chem. 5593
5525 417−424. Solids 2015, 84, 13−16. 5594
5526 (375) Chang, H.-C.; Jiang, J.-C.; Su, J.-C.; Chang, C.-Y.; Lin, S. H. (393) Yoshimura, Y.; Takekiyo, T.; Imai, Y.; Abe, H. Pressure- 5595
5527 Evidence of Rotational Isomerism in 1-Butyl-3-methylimidazolium Induced Spectral Changes of Room-Temperature Ionic Liquid, N,N- 5596
5528 Halides: A Combined High-Pressure Infrared and Raman Spectro- Diethyl-N-methyl-N-(2-methoxyethyl)ammonium Bis- 5597
5529 scopic Study. J. Phys. Chem. A 2007, 111, 9201−9206. (trifluoromethylsulfonyl)imide, DEME TFSI. J. Phys. Chem. C 2012, 5598
5530 (376) Su, L.; Li, M.; Zhu, X.; Wang, Z.; Chen, Z.; Li, F.; Zhou, Q.; 116, 2097−2101. 5599
5531 Hong, S. In Situ Crystallization of Low-Melting Ionic Liquid BMIM (394) Capitani, F.; Trequattrini, F.; Palumbo, O.; Paolone, A.; 5600
5532 PF(6) under High Pressure up to 2 GPa. J. Phys. Chem. B 2010, 114, Postorino, P. Phase Transitions of PYR14-TFSI as a Function of 5601
5533 5061−5065. Pressure and Temperature: the Competition between Smaller Volume 5602
5534 (377) Russina, O.; Fazio, B.; Schmidt, C.; Triolo, A. Structural and Lower Energy Conformer. J. Phys. Chem. B 2016, 120, 2921− 5603
5535 Organization and Phase Behaviour of 1-Butyl-3-methylimidazolium 2928. 5604
5536 Hexafluorophosphate: An High Pressure Raman Spectroscopy Study. (395) Piermari, G. J.; Block, S.; Barnett, S. Hydrostatic Limits in 5605
5537 Phys. Chem. Chem. Phys. 2011, 13, 12067−12074. Liquids and Solids to 100 kbar. J. Appl. Phys. 1973, 44, 5377−5382. 5606
5538 (378) Takekiyo, T.; Imai, Y.; Hatano, N.; Abe, H.; Yoshimura, Y. (396) Ribeiro, M. C. C.; Pádua, A. A. H.; Gomes, M. F. C. Glass 5607
5539 Conformational Preferences of Two Imidazolium-Based Ionic Liquids Transition of Ionic Liquids Under High Pressure. J. Chem. Phys. 2014, 5608
5540 at High Pressures. Chem. Phys. Lett. 2011, 511, 241−246. 140, 244514. 5609
5541 (379) Su, L.; Zhu, X.; Wang, Z.; Cheng, X.; Wang, Y.; Yuan, C.; (397) Yoshimura, Y.; Abe, H.; Imai, Y.; Takekiyo, T.; Hamaya, N. 5610
5542 Chen, Z.; Ma, C.; Li, F.; Zhou, Q.; et al. In Situ Observation of Decompression-Induced Crystal Polymorphism in a Room-Temper- 5611
5543 Multiple Phase Transitions in Low-Melting Ionic Liquid BMIM BF4 ature Ionic Liquid, N,N-Diethyl-N-methyl-N-(2-methoxyethyl) Am- 5612
5544 under High Pressure up to 30 GPa. J. Phys. Chem. B 2012, 116, 2216− monium Tetrafluoroborate. J. Phys. Chem. B 2013, 117, 3264−3269. 5613
5545 2222. (398) García-Saiz, A.; de Pedro, I.; Blanco, J. A.; González, J.; 5614
5546 (380) Yoshimura, Y.; Abe, H.; Takekiyo, T.; Shigemi, M.; Hamaya, Fernández, J. R. Pressure Effects on Emim[FeCl4], a Magnetic Ionic 5615
5547 N.; Wada, R.; Kato, M. Superpressing of a Room Temperature Ionic Liquid with Three-Dimensional Magnetic Ordering. J. Phys. Chem. B 5616
5548 Liquid, 1-Ethyl-3-methylimidazolium Tetrafluoroborate. J. Phys. Chem. 2013, 117, 3198−3206. 5617
5549 B 2013, 117, 12296−12302. (399) Chen, H.-K.; Srivastava, N.; Saha, S.; Shigeto, S. Complement- 5618
5550 (381) Abe, H.; Takekiyo, T.; Hatano, N.; Shigemi, M.; Hamaya, N.; ing Crystallography with Ultralow-Frequency Raman Spectroscopy: 5619
5551 Yoshimura, Y. Pressure-Induced Frustration−Frustration Process in 1- Structural Insights into Nitrile-Functionalized Ionic Liquids. Chem- 5620
5552 Butyl-3-methylimidazolium Hexafluorophosphate, a Room-Temper- PhysChem 2016, 17, 93−97. 5621
5553 ature Ionic Liquid. J. Phys. Chem. B 2014, 118, 1138−1145. (400) Penna, T. C.; Faria, L. F. O.; Matos, J. R.; Ribeiro, M. C. C. 5622
5554 (382) Li, J.; Su, L.; Zhu, X.; Li, H.; Cheng, X.; Li, L. Decompression- Pressure and Temperature Effects on Intermolecular Vibrational 5623
5555 Induced Disorder to Order Phase Transition in Low-Melting Ionic Dynamics of Ionic Liquids. J. Chem. Phys. 2013, 138, 104503. 5624
5556 Liquid [OMIM][PF6]. Chin. Sci. Bull. 2014, 59, 2980−2986. (401) Abe, H.; Takekiyo, T.; Aono, M.; Kishimura, H.; Yoshimura, 5625
5557 (383) Li, H.; Su, L.; Zhu, X.; Cheng, X.; Yang, K.; Yang, G. In Situ Y.; Hamaya, N. Polymorphs in Room-Temperature Ionic Liquids: 5626
5558 Crystallization of Ionic Liquid [Emim][PF6] from Methanol Solution Hierarchical Structure, Confined Water and Pressure-Induced 5627
5559 under High Pressure. J. Phys. Chem. B 2014, 118, 8684−8690. Frustration. J. Mol. Liq. 2015, 210, 200−214. 5628
5560 (384) Ren, Y.; Li, H.; Zhu, X.; Chen, L.; Su, L.; Yang, K.; Yang, G.; (402) Rogers, R. D.; Seddon, K. R. Ionic liquids–solvents of the 5629
5561 Wang, H. Pressure-induced Amorphization of Ionic Liquid [HMIM]- future? Science 2003, 302 (5646), 792−793. 5630
5562 [PF6]. Chem. Phys. Lett. 2015, 629, 8−12. (403) Plechkova, N. V.; Seddon, K. R. Applications of Ionic Liquids 5631
5563 (385) Wu, J.; Zhu, X.; Li, H.; Su, L.; Yang, K.; Cheng, X.; Yang, G.; in the Chemical Industry. Chem. Soc. Rev. 2008, 37, 123−150. 5632
5564 Liu, J. Combined Raman Scattering and X-ray Diffraction Study of (404) Chatel, G.; Pereira, J. F. B.; Debbeti, V.; Wang, H.; Rogers, R. 5633
5565 Phase Transition of the Ionic Liquid [BMIM][TFSI] Under High D. Mixing Ionic Liquids − “Simple Mixtures” or “Double Salts”? Green 5634
5566 Pressure. J. Solution Chem. 2015, 44, 2106−2116. Chem. 2014, 16, 2051. 5635

BC DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5636 (405) Kiefer, J.; Molina, M. M.; Noack, K. The Peculiar Nature of (425) Pal, A.; Saini, M.; Kumar, B. Volumetric, Ultrasonic and 5704
5637 Molecular Interactions between an Imidazolium Ionic Liquid and Spectroscopic (FT-IR) Studies for the Binary Mixtures of Imidazolium 5705
5638 Acetone. ChemPhysChem 2012, 13, 1213−1220. Based ILs with 1,2-propanediol. Fluid Phase Equilib. 2016, 411, 66−73. 5706
5639 (406) Zhang, Q.-G.; Wang, N.-N.; Yu, Z.-W. The Hydrogen Bonding (426) Roth, C.; Appelhagen, A.; Jobst, N.; Ludwig, R. Micro- 5707
5640 Interactions between the Ionic Liquid 1-Ethyl-3-Methylimidazolium heterogeneities in Ionic-Liquid-Methanol Solutions Studied by FTIR 5708
5641 Ethyl Sulfate and Water. J. Phys. Chem. B 2010, 114, 4747−4754. Spectroscopy, DFT Calculations and Molecular Dynamics Simu- 5709
5642 (407) Koch, H.; Noack, K.; Will, S. Raman Excess Spectroscopy vs lations. ChemPhysChem 2012, 13, 1708−1717. 5710
5643 Principal Component Analysis: Probing the Intermolecular Inter- (427) Singh, T.; Rao, K. S.; Kumar, A. Polarity Behaviour and 5711
5644 actions Between Chiral Molecules and Imidazolium-Based Ionic Specific Interactions of Imidazolium-Based Ionic Liquids in Ethylene 5712
5645 Liquids. Phys. Chem. Chem. Phys. 2016, 18, 28370−28375. Glycol. ChemPhysChem 2011, 12, 836−845. 5713
5646 (408) Brereton, R. G. Applied Chemometrics for Scientists; John Wiley (428) Köddermann, T.; Wertz, C.; Heintz, A.; Ludwig, R. The 5714
5647 & Sons: Wiltshire, 2007. Association of Water in Ionic Liquids: A Reliable Measure of Polarity. 5715
5648 (409) Fumino, K.; Peppel, T.; Geppert-Rybczyńska, M.; Zaitsau, D. Angew. Chem., Int. Ed. 2006, 45, 3697−3702. 5716
5649 H.; Lehmann, J. K.; Verevkin, S. P.; Köckerling, M.; Ludwig, R. The (429) Wülf, A.; Köddermann, T.; Wertz, C.; Heintz, A.; Ludwig, R. 5717
5650 Influence of Hydrogen Bonding on the Physical Properties of Ionic Water Vibrational Bands as a Polarity Indicator in Ionic Liquids. Z. 5718
5651 Liquids. Phys. Chem. Chem. Phys. 2011, 13, 14064−14075. Phys. Chem. 2006, 220, 1361−1376. 5719
5652 (410) Fumino, K.; Bonsa, A. M.; Golub, B.; Paschek, D.; Ludwig, R. (430) Zhang, Q.-G.; Wang, N.-N.; Wang, S.-L.; Yu, Z.-W. Hydrogen 5720
5653 Non-Ideal Mixing Behaviour of Hydrogen Bonding in Mixtures of Bonding Behaviors of Binary Systems Containing the Ionic Liquid 1- 5721
5654 Protic Ionic Liquids. ChemPhysChem 2015, 16, 299−304. Butyl-3-methylimidazolium Trifluoroacetate and Water/Methanol. J. 5722
5655 (411) Xiao, D.; Rajian, J. R.; Li, S.; Bartsch, R. A.; Quitevis, E. L. Phys. Chem. B 2011, 115, 11127−11136. 5723
5656 Additivity in the Optical Kerr Effect Spectra of Binary Ionic Liquid (431) Joshi, R.; Pasilis, S. P. The Effect of Tributylphosphate and 5724
5657 Mixtures: Implications for Nanostructural Organization. J. Phys. Chem. Tributyl Phosphine Oxide On Hydrogen Bonding Interactions 5725
5658 B 2006, 110, 16174−16178. between Water and the 1-Ethyl-3-Methylimidazolium Cation in 1- 5726
5659 (412) Quitevis, E. L.; Bardak, F.; Xiao, D.; Hines, L. G.; Son, P.; Ethyl-3-Methylimidazolium Bis(Trifluoromethylsulfonyl)Imide. J. Mol. 5727
5660 Bartsch, R. A.; Yang, P.; Voth, G. A. OKE Spectroscopy and Molecular Liq. 2015, 209, 381−386. 5728
5661 Dynamics Simulations of Nonpolar and Polar Molecules in Ionic (432) Chang, H. C.; Jiang, J. C.; Tsai, W. C.; Chen, G. C.; Lin, S. H. 5729
5662 Liquids. In Ionic Liquids: Science and Applications; Visser, A. E., Bridges, Hydrogen Bond Stabilization in 1,3-Dimethylimidazolium Methyl 5730
5663 N. J., Rogers, R. D., Eds.; American Chemical Society: Washington, Sulfate and 1-Butyl-3-Methylimidazolium Hexafluorophosphate 5731
5664 DC, 2012; Vol. 1117, pp 271−287. Probed by High Pressure: The Role of Charge-Enhanced C−H···O 5732
5665 (413) Fumino, K.; Reimann, S.; Ludwig, R. Probing Molecular Interactions in the Room-Temperature Ionic Liquid. J. Phys. Chem. B 5733
5666 Interaction in Ionic Liquids by Low Frequency Spectroscopy: 2006, 110, 3302−3307. 5734
5667 Coulomb Energy, Hydrogen Bonding and Dispersion Forces. Phys. (433) Andanson, J.-M.; Beier, M. J.; Baiker, A. Binary Ionic Liquids 5735
5668 Chem. Chem. Phys. 2014, 16, 21903−21929. with a Common Cation: Insight into Nanoscopic Mixing by Infrared 5736
5669 (414) Cha, S.; Kim, D. Anion Exchange in Ionic Liquid Mixtures. Spectroscopy. J. Phys. Chem. Lett. 2011, 2, 2959−2964. 5737
5670 Phys. Chem. Chem. Phys. 2015, 17, 29786−29792. (434) Tran, C. D.; De Paoli Lacerda, S. H.; Oliveira, D. Absorption of 5738
5671 (415) Aparicio, S.; Atilhan, M. Mixed Ionic Liquids: The Case of Water by Room-Temperature Ionic Liquids: Effect of Anions on 5739
5672 Pyridinium-Based Fluids. J. Phys. Chem. B 2012, 116, 2526−2537. Concentration and State of Water. Appl. Spectrosc. 2003, 57, 152−157. 5740
5673 (416) Miran, M. S.; Yasuda, T.; Susan, M. A. B. H.; Dokko, K.; (435) Dominguez-Vidal, A.; Kaun, N.; Ayora-Canada, M. J.; Lendl, B. 5741
5674 Watanabe, M. Binary Protic Ionic Liquid Mixtures as a Proton Probing Intermolecular Interactions in Water/Ionic Liquid Mixtures 5742
5675 Conductor: High Fuel Cell Reaction Activity and Facile Proton by Far-Infrared Spectroscopy. J. Phys. Chem. B 2007, 111, 4446−4452. 5743
5676 Transport. J. Phys. Chem. C 2014, 118, 27631−27639. (436) Marekha, B. A.; Koverga, V. A.; Moreau, M.; Kiselev, M.; 5744
5677 (417) Navia, P.; Troncoso, J.; Romaní, L. Excess Magnitudes for Takamuku, T.; Kalugin, O. N.; Idrissi, A. Intermolecular Interactions, 5745
5678 Ionic Liquid Binary Mixtures with a Common Ion. J. Chem. Eng. Data Ion Solvation, and Association in Mixtures of 1-N-Butyl-3-Methyl- 5746
5679 2007, 52, 1369−1374. imidazolium Hexafluorophosphate and Gamma-Butyrolactone: In- 5747
5680 (418) Navia, P.; Troncoso, J.; Romaní, L. Viscosities for Ionic Liquid sights from Raman Spectroscopy. J. Raman Spectrosc. 2015, 46, 339− 5748
5681 Binary Mixtures with a Common Ion. J. Solution Chem. 2008, 37, 677− 352. 5749
5682 688. (437) Cha, S.; Ao, M.; Sung, W.; Moon, B.; Ahlström, B.; Johansson, 5750
5683 (419) Clough, M. T.; Crick, C. R.; Gräsvik, J.; Hunt, P. A.; P.; Ouchi, Y.; Kim, D. Structures of Ionic Liquid-Water Mixtures 5751
5684 Niedermeyer, H.; Welton, T.; Whitaker, O. P. A Physicochemical Investigated by IR and NMR Spectroscopy. Phys. Chem. Chem. Phys. 5752
5685 Investigation of Ionic Liquid Mixtures. Chem. Sci. 2015, 6, 1101−1114. 2014, 16, 9591−9601. 5753
5686 (420) Annat, G.; Forsyth, M.; MacFarlane, D. R. Ionic Liquid (438) Yaghini, N.; Pitawala, J.; Matic, A.; Martinelli, A. Effect of 5754
5687 MixturesVariations in Physical Properties and Their Origins in Water on the Local Structure and Phase Behavior of Imidazolium- 5755
5688 Molecular Structure. J. Phys. Chem. B 2012, 116, 8251−8258. Based Protic Ionic Liquids. J. Phys. Chem. B 2015, 119, 1611−1622. 5756
5689 (421) Katayanagi, H.; Nishikawa, K.; Shimozaki, H.; Miki, K.; Westh, (439) Singh, T.; Kumar, A. Cation−Anion−Water Interactions in 5757
5690 P.; Koga, Y. Mixing Schemes in Ionic Liquid-H2O Systems: A Aqueous Mixtures of Imidazolium Based Ionic Liquids. Vib. Spectrosc. 5758
5691 Thermodynamic Study. J. Phys. Chem. B 2004, 108, 19451−19457. 2011, 55, 119−125. 5759
5692 (422) Abe, H.; Fukushima, R.; Onji, M.; Hirayama, K.; Kishimura, (440) Saha, S.; Hamaguchi, H. Effect of Water on the Molecular 5760
5693 H.; Yoshimura, Y.; Ozawa, S. Two-Length Scale Description of Structure and Arrangement of Nitrile-Functionalized Ionic Liquids. J. 5761
5694 Hydrophobic Room-Temperature Ionic Liquid-Alcohol Systems. J. Phys. Chem. B 2006, 110, 2777−2781. 5762
5695 Mol. Liq. 2016, 215, 417−422. (441) Bodo, E.; Mangialardo, S.; Capitani, F.; Gontrani, L.; Leonelli, 5763
5696 (423) Noack, K.; Leipertz, A.; Kiefer, J. Molecular Interactions and F.; Postorino, P. Interaction of a Long Alkyl Chain Protic Ionic Liquid 5764
5697 Macroscopic Effects in Binary Mixtures of an Imidazolium Ionic and Water. J. Chem. Phys. 2014, 140, 204503. 5765
5698 Liquid with Water, Methanol, and Ethanol. J. Mol. Struct. 2012, 1018, (442) Danten, Y.; Cabaço, M. I.; Besnard, M. Interaction of Water 5766
5699 45−53. Highly Diluted in 1-Alkyl-3-methylimidazolium Ionic Liquids with the 5767
5700 (424) Yoshimura, Y.; Kimura, H.; Okamoto, C.; Miyashita, T.; Imai, PF6(−) and BF4(−) Anions. J. Phys. Chem. A 2009, 113, 2873−2889. 5768
5701 Y.; Abe, H. Glass Transition Behaviour of Ionic Liquid, 1-Butyl-3- (443) Danten, Y.; Cabaço, M. I.; Besnard, M. Interaction of Water 5769
5702 methylimidazolium Tetrafluoroborate−H2O Mixed Solutions. J. Chem. Diluted in 1-butyl-3-methylimidazolium Ionic Liquids by Vibrational 5770
5703 Thermodyn. 2011, 43, 410−412. Spectroscopy Modeling. J. Mol. Liq. 2010, 153, 57−66. 5771

BD DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5772 (444) Dahi, A.; Fatyeyeva, K.; Chappey, C.; Langevin, D.; Rogalsky, (460) Tao, G.-h.; Zou, M.; Wang, X.-h.; Chen, Z.-y.; Evans, D. G.; 5839
5773 S. P.; Tarasyuk, O. P.; Marais, S. Water Sorption Properties of Room- Kou, Y. Comparison of Polarities of Room-Temperature Ionic Liquids 5840
5774 Temperature Ionic Liquids over the Whole Range of Water Activity Using FT-IR Spectroscopic Probes. Aust. J. Chem. 2005, 58, 327. 5841
5775 and Molecular States of Water in These Media. RSC Adv. 2015, 5, (461) Garcia, H. C.; de Oliveira, L. F. C.; Nicolau, B. G.; Ribeiro, M. 5842
5776 76927−76938. C. C. Raman Spectra of Acetonitrile in Imidazolium Ionic Liquids. J. 5843
5777 (445) Romanos, G. E.; Zubeir, L. F.; Likodimos, V.; Falaras, P.; Raman Spectrosc. 2010, 41, 1720−1724. 5844
5778 Kroon, M. C.; Iliev, B.; Adamova, G.; Schubert, T. J. S. Enhanced CO2 (462) Shiflett, M. B.; Kasprzak, D. J.; Junk, C. P.; Yokozeki, A. Phase 5845
5779 Capture in Binary Mixtures of 1-Alkyl-3-methylimidazolium Tricya- Behavior of {carbon dioxide+[bmim][Ac]} Mixtures. J. Chem. 5846
5780 nomethanide Ionic Liquids with Water. J. Phys. Chem. B 2013, 117, Thermodyn. 2008, 40, 25−31. 5847
5781 12234−12251. (463) Hollóczki, O.; Kelemen, Z.; Könczöl, L.; Szieberth, D.; 5848
5782 (446) Yoshimura, Y.; Goto, T.; Abe, H.; Imai, Y. Existence of Nearly- Nyulászi, L.; Stark, A.; Kirchner, B. Significant Cation Effects in 5849
5783 Free Hydrogen Bonds in an Ionic Liquid, N,N-Diethyl-N-methyl-N- Carbon Dioxide-Ionic Liquid Systems. ChemPhysChem 2013, 14, 315− 5850
5784 (2-methoxyethyl) Ammonium Tetrafluoroborate-Water at 77 K. J. 320. 5851
5785 Phys. Chem. B 2009, 113, 8091−8095. (464) Zeng, S.; Gao, H.; Zhang, X.; Dong, H.; Zhang, X.; Zhang, S. 5852
5786 (447) Fumino, K.; Stange, P.; Fossog, V.; Hempelmann, R.; Ludwig, Efficient and Reversible Capture of SO2 by Pyridinium-based Ionic 5853
5787 R. Equilibrium of Contact and Solvent-Separated Ion Pairs in Mixtures Liquids. Chem. Eng. J. 2014, 251, 248−256. 5854
5788 of Protic Ionic Liquids and Molecular Solvents Controlled by Polarity. (465) Shang, Y.; Li, H.; Zhang, S.; Xu, H.; Wang, Z.; Zhang, L.; 5855
5789 Angew. Chem., Int. Ed. 2013, 52, 12439−12442. Zhang, J. Guanidinium-based Ionic Liquids for Sulfur Dioxide 5856
5790 (448) Fumino, K.; Fossog, V.; Stange, P.; Wittler, K.; Polet, W.; Sorption. Chem. Eng. J. 2011, 175, 324−329. 5857
5791 Hempelmann, R.; Ludwig, R. Ion Pairing in Protic Ionic Liquids (466) Huang, J.; Riisager, A.; Wasserscheid, P.; Fehrmann, R. 5858

5792 Probed by Far-Infrared Spectroscopy: Effects of Solvent Polarity and Reversible Physical Absorption of SO2 by Ionic Liquids. Chem. 5859
5793 Temperature. ChemPhysChem 2014, 15, 2604−2609. Commun. 2006, 39, 4027−4029. 5860

5794 (449) Jiang, J.-C.; Lin, K.-H.; Li, S.-C.; Shih, P.-M.; Hung, K.-C.; Lin, (467) Yang, D.; Hou, M.; Ning, H.; Ma, J.; Kang, X.; Zhang, J.; Han, 5861

5795 S. H.; Chang, H.-C. Association Structures of Ionic Liquid/DMSO B. Reversible Capture of SO2 through Functionalized Ionic Liquids. 5862

Mixtures Studied by High-Pressure Infrared Spectroscopy. J. Chem. ChemSusChem 2013, 6, 1191−1195. 5863
5796
(468) Ando, R. A.; Siqueira, L. J. A.; Bazito, F. C.; Torresi, R. M.; 5864
5797 Phys. 2011, 134, 044506.
(450) Rodrigues, F.; Santos, P. S. Effect of the Chain Length in the Santos, P. S. The Sulfur Dioxide−1-Butyl-3-Methylimidazolium 5865
5798
Structure of Imidazolic Ionic Liquids and Dimethylformamide Bromide Interaction: Drastic Changes in Structural and Physical 5866
5799
Properties. J. Phys. Chem. B 2007, 111, 8717−8719. 5867
5800 Solutions Probed by Raman Spectroscopy. Vib. Spectrosc. 2010, 54,
(469) Siqueira, L. J. A.; Ando, R. m. A.; Bazito, F. F. C.; Torresi, R. 5868
5801 123−126.
M.; Santos, P. S.; Ribeiro, M. C. C. Shielding of Ionic Interactions by 5869
5802 (451) Wang, N.-N.; Zhang, Q.-G.; Wu, F.-G.; Li, Q.-Z.; Yu, Z.-W.
Sulfur Dioxide in an Ionic Liquid. J. Phys. Chem. B 2008, 112, 6430− 5870
5803 Hydrogen Bonding Interactions Between a Representative Pyridinium-
6435. 5871
5804 Based Ionic Liquid [BuPy][BF4] and Water/Dimethyl Sulfoxide. J.
(470) Kazarian, S. G.; Briscoe, B. J.; Welton, T. Combining Ionic 5872
5805 Phys. Chem. B 2010, 114, 8689−8700.
Liquids and Supercritical Fluids: In Situ ATR-IR Study of CO2 5873
5806 (452) Pal, A.; Kumar, B.; Singh Kang, T. Effect of Structural
Dissolved in Two Ionic Liquids at High Pressures. Chem. Commun. 5874
5807 Alteration of Ionic Liquid on their Bulk and Molecular Level
2000, No. 20, 2047−2048. 5875
5808 Interactions with Ethylene Glycol. Fluid Phase Equilib. 2013, 358, (471) Andanson, J.-M.; Jutz, F.; Baiker, A. Supercritical CO2/Ionic 5876
5809 241−249. Liquid Systems: What Can We Extract from Infrared and Raman 5877
5810 (453) Shirota, H.; Kakinuma, S.; Itoyama, Y.; Umecky, T.;
Spectra? J. Phys. Chem. B 2009, 113, 10249−10254. 5878
5811 Tak am u ku , T . Effec ts of T e tr aflu or ob or ate and Bis- (472) Seki, T.; Grunwaldt, J.-D.; Baiker, A. In Situ Attenuated Total 5879
5812 (trifluoromethylsulfonyl)amide Anions on the Microscopic Structures Reflection Infrared Spectroscopy of Imidazolium-Based Room- 5880
5813 of 1-Methyl-3-octylimidazolium-Based Ionic Liquids and Benzene Temperature Ionic Liquids under “Supercritical” CO2. J. Phys. Chem. 5881
5814 Mixtures: A Multiple Approach by ATR-IR, NMR, and Femtosecond B 2009, 113, 114−122. 5882
5815 Raman-Induced Kerr Effect. J. Phys. Chem. B 2016, 120, 513−526. (473) Besnard, M.; Cabaço, M. I.; Vaca Chávez, F.; Pinaud, N.; 5883
5816 (454) Shimomura, T.; Takamuku, T.; Yamaguchi, T. Clusters of Sebastião, P. J.; Coutinho, J. A. P.; Mascetti, J.; Danten, Y. CO2 in 1- 5884
5817 Imidazolium-Based Ionic Liquid in Benzene Solutions. J. Phys. Chem. B Butyl-3-methylimidazolium Acetate. 2. NMR Investigation of Chem- 5885
5818 2011, 115, 8518−8527. ical Reactions. J. Phys. Chem. A 2012, 116, 4890−4901. 5886
5819 (455) Xue, L.; Tamas, G.; Quitevis, E. L. Comparative OHD-RIKES (474) Blath, J.; Deubler, N.; Hirth, T.; Schiestel, T. Chemisorption of 5887
5820 Study of the Low-Frequency (0−250 cm − 1) Vibrational Dynamics Carbon Dioxide in Imidazolium Based Ionic Liquids with Carboxylic 5888
5821 of Dibenzyl- and Monobenzyl-Substituted Imidazolium Ionic Liquids Anions. Chem. Eng. J. 2012, 181−182, 152−158. 5889
5822 and Benzene/Dimethylimidazolium Mixtures. ACS Sustainable Chem. (475) Makino, T. In situ Raman Study of Dissolved Carbon-Dioxide 5890
5823 Eng. 2016, 4, 514−524. Induced Changes of Imidazolium-Based Ionic Liquids. J. Phys. Conf. 5891
5824 (456) Lynden-Bell, R. M.; Xue, L.; Tamas, G.; Quitevis, E. L. Local Ser. 2010, 215, 012068. 5892
5825 Structure and Intermolecular Dynamics of an Equimolar Benzene and (476) Gurkan, B. E.; de la Fuente, J. C.; Mindrup, E. M.; Ficke, L. E.; 5893
5826 1,3-dimethylimidazolium bis[(trifluoromethane)sulfonyl]amide Mix- Goodrich, B. F.; Price, E. A.; Schneider, W. F.; Brennecke, J. F. 5894
5827 ture: Molecular Dynamics Simulations and OKE Spectroscopic Equimolar CO2 Absorption by Anion-Functionalized Ionic Liquids. J. 5895
5828 Measurements. J. Chem. Phys. 2014, 141, 044506. Am. Chem. Soc. 2010, 132, 2116−2117. 5896
5829 (457) Shirota, H. Intermolecular/Interionic Vibrations of 1-Methyl- (477) Lee, T. B.; Oh, S.; Gohndrone, T. R.; Morales-Collazo, O.; 5897
5830 3-n-octylimidazolium Tetrafluoroborate Ionic Liquid and Benzene Seo, S.; Brennecke, J. F.; Schneider, W. F. CO2 Chemistry of 5898
5831 Mixtures. J. Phys. Chem. B 2013, 117, 7985−7995. Phenolate-Based Ionic Liquids. J. Phys. Chem. B 2016, 120, 1509− 5899
5832 (458) Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 1517. 5900
5833 3rd ed.; Wiley-VCH Verlag: Weinheim, Germany, 2003. (478) Giammanco, C. H.; Kramer, P. L.; Yamada, S. A.; Nishida, J.; 5901
5834 (459) Reichardt, C. Pyridinium N-Phenoxide Betaine Dyes and their Tamimi, A.; Fayer, M. D. Coupling of Carbon Dioxide Stretch and 5902
5835 Application to the Determination of Solvent Polarities Part 29 - Bend Vibrations Reveals Thermal Population Dynamics in an Ionic 5903
5836 Polarity of Ionic Liquids Determined Empirically by Means of Liquid. J. Phys. Chem. B 2016, 120, 549−556. 5904
5837 Solvatochromic Pyridinium N-Phenolate Betaine Dyes. Green Chem. (479) Rodopoulos, T.; Smith, L.; Horne, M. D.; Rüther, T. 5905
5838 2005, 7, 339−351. Speciation of Aluminium in Mixtures of the Ionic Liquids [C3mpip]- 5906

BE DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

5907 [Ntf2] and [C4mpyr][Ntf2] with AlCl3: An Electrochemical and (497) Le, M. L. P.; Alloin, F.; Strobel, P.; Leprêtre, J.-C.; Pérez del 5976
5908 NMR Spectroscopy Study. Chem. - Eur. J. 2010, 16, 3815−3826. Valle, C.; Judeinstein, P. Structure-Properties Relationships of Lithium 5977
5909 (480) Babushkina, O. B. Phase Behaviour and FTIR Spectra of Ionic Electrolytes Based on Ionic Liquid. J. Phys. Chem. B 2010, 114, 894− 5978
5910 Liquids: The Mixtures of 1-Butyl-1-methylpyrrolidinium Chloride and 903. 5979
5911 TaCl5. Z. Naturforsch., A: Phys. Sci. 2008, 63, 66−72. (498) Lassègues, J.-C.; Grondin, J.; Talaga, D. Lithium Solvation in 5980
5912 (481) Alves, M. B.; Santos, V. O., Jr.; Soares, V. C. D.; Suarez, P. A. Bis(Trifluoromethanesulfonyl)imide-Based Ionic Liquids. Phys. Chem. 5981
5913 Z.; Rubim, J. C. Raman Spectroscopy of Ionic Liquids Derived from 1- Chem. Phys. 2006, 8, 5629−5632. 5982
5914 n-Butyl-3-methylimidazolium Chloride and Niobium Chloride or Zinc (499) Lassègues, J. C.; Grondin, J.; Aupetit, C.; Johansson, P. 5983
5915 Chloride Mixtures. J. Raman Spectrosc. 2008, 39, 1388−1395. Spectroscopic Identification of the Lithium Ion Transporting Species 5984
5916 (482) Goujon, N.; Byrne, N.; Walsh, T. R.; Forsyth, M. The in LiTFSI-Doped Ionic Liquids. J. Phys. Chem. A 2009, 113, 305−314. 5985
5917 Influence of Water and Metal Salt on the Transport and Structural (500) Duluard, S.; Grondin, J.; Bruneel, J.-L.; Pianet, I.; Grélard, A.; 5986
5918 Properties of 1-Octyl-3-methylimidazolium Chloride. Aust. J. Chem. Campet, G.; Delville, M.-H.; Lassègues, J.-C. Lithium Solvation and 5987
5919 2015, 68, 420. Diffusion in the 1-Butyl-3-methylimidazolium Bis- 5988
5920 (483) Andriyko, Y.; Andriiko, A.; Babushkina, O. B.; Nauer, G. E. (trifluoromethanesulfonyl)imide Ionic Liquid. J. Raman Spectrosc. 5989
5921 Electrochemistry of TiF4 in 1-butyl-2,3-dimethylimidazolium tetra- 2008, 39, 627−632. 5990
5922 fluoroborate. Electrochim. Acta 2010, 55, 1081−1089. (501) Umebayashi, Y.; Mitsugi, T.; Fukuda, S.; Fujimori, T.; Fujii, K.; 5991
5923 (484) Arellano, I. H.; Huang, J.; Pendleton, P. Computational Kanzaki, R.; Takeuchi, M.; Ishiguro, S.-I. Lithium Ion Solvation in 5992
5924 Insights into the Molecular Interaction and Ion-Pair Structures of a Room-Temperature Ionic Liquids Involving Bis- 5993
5925 Novel Zinc-Functionalized Ionic Liquid, [Emim][Zn(TFSI)3]. Spec- (trifluoromethanesulfonyl) Imide Anion Studied by Raman Spectros- 5994
5926 trochim. Acta, Part A 2016, 153, 6−15. copy and DFT Calculations. J. Phys. Chem. B 2007, 111, 13028− 5995
5927 (485) Mudring, A.-V.; Babai, A.; Arenz, S.; Giernoth, R. The 13032. 5996
5928 “Noncoordinating” Anion Tf2N− Coordinates to Yb2+: A Structurally (502) Umebayashi, Y.; Mori, S.; Fujii, K.; Tsuzuki, S.; Seki, S.; 5997
5929 Characterized Tf2N− Complex from the Ionic Liquid [mppyr]- Hayamizu, K.; Ishiguro, S.-i. Raman Spectroscopic Studies and Ab 5998
5930 [Tf2N]. Angew. Chem., Int. Ed. 2005, 44, 5485−5488. Initio Calculations on Conformational Isomerism of 1-Butyl-3- 5999
5931 (486) Tang, S.; Babai, A.; Mudring, A.-V. Europium-Based Ionic methylimidazolium Bis-(trifluoromethanesulfonyl)amide Solvated to 6000
5932 Liquids as Luminescent Soft Materials. Angew. Chem., Int. Ed. 2008, 47, a Lithium Ion in Ionic Liquids: Effects of the Second Solvation Sphere 6001
5933 7631−7634. of the Lithium Ion. J. Phys. Chem. B 2010, 114, 6513−6521. 6002
5934 (487) Liu, Z.; El Abedin, S. Z.; Endres, F. Electrochemical and (503) Umebayashi, Y.; Hamano, H.; Seki, S.; Minofar, B.; Fujii, K.; 6003
5935 Spectroscopic Study of Zn(II) Coordination and Zn Electrodeposition Hayamizu, K.; Tsuzuki, S.; Kameda, Y.; Kohara, S.; Watanabe, M. 6004
5936 in Three Ionic Liquids with the Trifluoromethylsulfonate Anion, Liquid Structure of and Li+ Ion Solvation in Bis- 6005
5937 Different Imidazolium Ions and Their Mixtures with Water. Phys. (trifluoromethanesulfonyl)amide Based Ionic Liquids Composed of 6006
5938 Chem. Chem. Phys. 2015, 17, 15945−15952. 1-Ethyl-3-methylimidazolium and N-Methyl-N-propylpyrrolidinium 6007
5939 (488) Oliveira, F. S.; Cabrita, E. J.; Todorovic, S.; Bernardes, C. E. S.; Cations. J. Phys. Chem. B 2011, 115, 12179−12191. 6008
5940 Canongia Lopes, J. N.; Hodgson, J. L.; MacFarlane, D. R.; Rebelo, L. (504) Hardwick, L. J.; Holzapfel, M.; Wokaun, A.; Novák, P. Raman 6009
5941 P. N.; Marrucho, I. M. Mixtures of the 1-Ethyl-3-methylimidazolium Study of Lithium Coordination in EMI-TFSI Additive Systems as 6010
5942 Acetate Ionic Liquid with Different Inorganic Salts: Insights into their Lithium-Ion Battery Ionic Liquid Electrolytes. J. Raman Spectrosc. 6011
5943 Interactions. Phys. Chem. Chem. Phys. 2016, 18, 2756−2766. 2007, 38, 110−112. 6012
5944 (489) Chimdi, T.; Gunzelmann, D.; Vongsvivut, J.; Forsyth, M. A (505) Martins, V. L.; Nicolau, B. G.; Urahata, S. M.; Ribeiro, M. C. 6013
5945 study of phase behavior and conductivity of mixtures of the organic C.; Torresi, R. M. Influence of the Water Content on the Structure 6014
5946 ionic plastic crystal N-methyl-N-methyl-pyrrolidinium dicyanamide and Physicochemical Properties of an Ionic Liquid and Its Li + 6015
5947 with sodium dicyanamide. Solid State Ionics 2015, 272, 74−83. Mixture. J. Phys. Chem. B 2013, 117, 8782−8792. 6016
5948 (490) Carstens, T.; Lahiri, A.; Borisenko, N.; Endres, F. [Py1,4]FSI- (506) Umebayashi, Y.; Yamaguchi, T.; Fukuda, S.; Mitsugi, T.; 6017
5949 NaFSI-Based Ionic Liquid Electrolyte for Sodium Batteries: Na+ Takeuchi, M.; Fujii, K.; Ishiguro, S.-i. Raman Spectroscopic Study on 6018
5950 Solvation and Interfacial Nanostructure on Au(111). J. Phys. Chem. C Alkaline Metal Ion Solvation in 1-Butyl-3-methylimidazolium Bis- 6019
5951 2016, 120, 14736−14741. (trifluoromethanesulfonyl)amide Ionic Liquid. Anal. Sci. 2008, 24, 6020
5952 (491) Forsyth, M.; Chimdi, T.; Seeber, A.; Gunzelmann, D.; Howlett, 1297−1304. 6021
5953 P. C. Structure and Dynamics in an Organic Ionic Plastic Crystal, N- (507) Pitawala, J.; Kim, J.-K.; Jacobsson, P.; Koch, V.; Croce, F.; 6022
5954 ethyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)amide, Matic, A. Phase Behaviour, Transport Properties, and Interactions in 6023
5955 Mixed with a Sodium Salt. J. Mater. Chem. A 2014, 2, 3993−4003. Li-Salt Doped Ionic Liquids. Faraday Discuss. 2012, 154, 71−80. 6024
5956 (492) Duluard, S.; Grondin, J.; Bruneel, J. L.; Campet, G.; Delville, (508) Watkins, T.; Buttry, D. A. Determination of Mg2+ Speciation in 6025
5957 M. H.; Lassègues, J. C. Lithium Solvation in a PMMA Membrane a TFSI−-Based Ionic Liquid With and Without Chelating Ethers Using 6026
5958 Plasticized by a Lithium-Conducting Ionic Liquid Based on 1-Butyl-3- Raman Spectroscopy. J. Phys. Chem. B 2015, 119, 7003−7014. 6027
5959 Methylimidazolium Bis(Trifluoromethanesulfonyl)Imide. J. Raman (509) Deepa, M.; Agnihotry, S. A.; Gupta, D.; Chandra, R. Ion- 6028
5960 Spectrosc. 2008, 39, 1189−1194. Pairing Effects and Ion−Solvent−Polymer Interactions in LiN- 6029
5961 (493) Nicolau, B. G.; Sturlaugson, A.; Fruchey, K.; Ribeiro, M. C. C.; (CF3SO2)2−PC−PMMA Electrolytes: A FTIR Study. Electrochim. 6030
5962 Fayer, M. D. Room Temperature Ionic Liquid-Lithium Salt Mixtures: Acta 2004, 49, 373−383. 6031
5963 Optical Kerr Effect Dynamical Measurements. J. Phys. Chem. B 2010, (510) Seo, D. M.; Boyle, P. D.; Sommer, R. D.; Daubert, J. S.; 6032
5964 114, 8350−8356. Borodin, O.; Henderson, W. A. Solvate Structures and Spectroscopic 6033
5965 (494) Li, Z.; Smith, G. D.; Bedrov, D. Li+ Solvation and Transport Characterization of LiTFSI Electrolytes. J. Phys. Chem. B 2014, 118, 6034
5966 Properties in Ionic Liquid/Lithium Salt Mixtures: A Molecular 13601−13608. 6035
5967 Dynamics Simulation Study. J. Phys. Chem. B 2012, 116, 12801− (511) Mandai, T.; Tsuzuki, S.; Ueno, K.; Dokko, K.; Watanabe, M. 6036
5968 12809. Pentaglyme-K Salt Binary Mixtures: Phase Behavior, Solvate 6037
5969 (495) Pitawala, J.; Martinelli, A.; Johansson, P.; Jacobsson, P.; Matic, Structures, and Physicochemical Properties. Phys. Chem. Chem. Phys. 6038
5970 A. Coordination and Interactions in a Li-Salt Doped Ionic Liquid. J. 2015, 17, 2838−2849. 6039
5971 Non-Cryst. Solids 2015, 407, 318−323. (512) Mandai, T.; Yoshida, K.; Ueno, K.; Dokko, K.; Watanabe, M. 6040
5972 (496) Monteiro, M. J.; Bazito, F. F. C.; Siqueira, L. J. a.; Ribeiro, M. Criteria for Solvate Ionic Liquids. Phys. Chem. Chem. Phys. 2014, 16, 6041
5973 C. C.; Torresi, R. M. Transport Coefficients, Raman Spectroscopy, and 8761−8772. 6042
5974 Computer Simulation of Lithium Salt Solutions in an Ionic Liquid. J. (513) Zhang, C.; Ueno, K.; Yamazaki, A.; Yoshida, K.; Moon, H.; 6043
5975 Phys. Chem. B 2008, 112, 2102−2109. Mandai, T.; Umebayashi, Y.; Dokko, K.; Watanabe, M. Chelate Effects 6044

BF DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

6045 in Glyme/Lithium Bis(trifluoromethanesulfonyl)amide Solvate Ionic (532) Chaurasia, S. K.; Singh, R. K.; Chandra, S. Structural and 6114
6046 Liquids. I. Stability of Solvate Cations and Correlation with Electrolyte Transport Studies on Polymeric Membranes of PEO Containing Ionic 6115
6047 Properties. J. Phys. Chem. B 2014, 118, 5144−5153. Liquid, EMIM-TY: Evidence of Complexation. Solid State Ionics 2011, 6116
6048 (514) Aguilera, L.; Xiong, S.; Scheers, J.; Matic, A. A Structural Study 183, 32−39. 6117
6049 of LiTFSI−tetraglyme Mixtures: From Diluted Solutions to Solvated (533) Bradley, L. C.; Gupta, M. Encapsulation of Ionic Liquids within 6118
6050 Ionic Liquids. J. Mol. Liq. 2015, 210, 238−242. Polymer Shells via Vapor Phase Deposition. Langmuir 2012, 28, 6119
6051 (515) Ueno, K.; Tatara, R.; Tsuzuki, S.; Saito, S.; Doi, H.; Yoshida, 10276−10280. 6120
6052 K.; Mandai, T.; Matsugami, M.; Umebayashi, Y.; Dokko, K.; et al. Li+ (534) Ramesh, S.; Lu, S.-C. Enhancement of Ionic Conductivity and 6121
6053 Solvation in Glyme-Li Salt Solvate Ionic Liquids. Phys. Chem. Chem. Structural Properties by 1-butyl-3-methylimidazolium trifluorometha- 6122
6054 Phys. 2015, 17, 8248−8257. nesulfonate Ionic Liquid in poly(vinylidene fluoride-hexafluoropropy- 6123
6055 (516) Angell, C. A.; Ansari, Y.; Zhao, Z. Ionic Liquids: Past, Present lene)-Based Polymer Electrolytes. J. Appl. Polym. Sci. 2012, 126, 6124
6056 and Future. Faraday Discuss. 2012, 154, 9−27. E484−E492. 6125
6057 (517) Kramer, P. L.; Giammanco, C. H.; Fayer, M. D. Dynamics of (535) He, L. H.; Sun, J.; Wang, X. X.; Wang, C. D.; Song, R.; Hao, Y. 6126
6058 Water, Methanol, and Ethanol in a Room Temperature Ionic Liquid. J. M. Facile and Effective Promotion of Crystalline Phase in Poly- 6127
6059 Chem. Phys. 2015, 142, 212408. (Vinylidene Fluoride) Via The Incorporation of Imidazolium Ionic 6128
6060 (518) Wong, D. B.; Giammanco, C. H.; Fenn, E. E.; Fayer, M. D. Liquids. Polym. Int. 2013, 62, 638−646. 6129
6061 Dynamics of Isolated Water Molecules in a Sea of Ions in a Room (536) Shalu; Chaurasia, S. K.; Singh, R. K.; Chandra, S. Electrical, 6130
6062 Temperature Ionic Liquid. J. Phys. Chem. B 2013, 117, 623−635. Mechanical, Structural, and Thermal Behaviors of Polymeric Gel 6131
6063 (519) Yu, B.; Liu, Z. L.; Zhou, F.; Liu, W. M.; Liang, Y. M. A Novel Electrolyte Membranes of Poly(vinylidene fluoride-co-hexafluoropro- 6132
6064 Lubricant Additive Based on Carbon Nanotubes for Ionic Liquids. pylene) with the Ionic Liquid 1-Butyl-3-Methylimidazolium Tetra- 6133
6065 Mater. Lett. 2008, 62, 2967−2969. fluoroborate Plus Lithium Tetrafluoroborate. J. Appl. Polym. Sci. 2015, 6134
6066 (520) Wang, J. Y.; Chu, H. B.; Li, Y. Why Single-Walled Carbon 132,10.1002/app.41456. 6135
6067 Nanotubes Can Be Dispersed in Imidazolium-Based Ionic Liquids. (537) Yuan, C. S.; Zhu, X.; Su, L.; Yang, D. Y.; Wang, Y. Q.; Yang, K.; 6136
6068 ACS Nano 2008, 2, 2540−2546. Cheng, X. R. Preparation and Characterization of a Novel Ionic 6137
6069 (521) Tamailarasan, P.; Ramaprabhu, S. Carbon Nanotubes- Conducting Foam-Type Polymeric Gel Based on Polymer PVdF-HFP 6138
6070 Graphene-Solidlike Ionic Liquid Layer-Based Hybrid Electrode and Ionic Liquid EMIM TFSI. Colloid Polym. Sci. 2015, 293, 1945− 6139
6071 Material for High Performance Supercapacitor. J. Phys. Chem. C 1952. 6140
6072 2012, 116, 14179−14187. (538) Vitucci, F. M.; Palumbo, O.; Trequattrini, F.; Brubach, J. B.; 6141
6073 (522) Kleinschmidt, A. C.; Donato, R. K.; Perchacz, M.; Benes, H.; Roy, P.; Meschini, I.; Croce, F.; Paolone, A. Interaction of 1-butyl-1- 6142
6074 Stengl, V.; Amico, S. C.; Schrekker, H. S. ″Unrolling″ Multi-Walled
methylpyrrolidinium bis(trifluoromethanesulfonyl)imide with an Elec- 6143
6075 Carbon Nanotubes with Ionic Liquids: Application as Fillers in Epoxy-
trospun PVdF Membrane: Temperature Dependence of the 6144
6076 Based Nanocomposites. RSC Adv. 2014, 4, 43436−43443.
Concentration of the Anion Conformers. J. Chem. Phys. 2015, 143, 6145
6077 (523) Attri, P.; Bhatia, R.; Arora, B.; Kumar, N.; Park, J. H.; Baik, K.
094707. 6146
6078 Y.; Lee, G. J.; Kim, I. T.; Koo, J. H.; Choi, E. H. Molecular Interactions
(539) Bahadur, I.; Momin, M. I. K.; Koorbanally, N. A.; Sattari, M.; 6147
6079 Between Carbon Nanotubes and Ammonium Ionic Liquids and their
Ebenso, E. E.; Katata-Seru, L. M.; Singh, S.; Ramjugernath, D. 6148
6080 Catalysis Properties. Mater. Res. Bull. 2014, 58, 6−9.
Interactions of polyvinylpyrrolidone with Imidazolium Based Ionic 6149
6081 (524) Jiang, F. L.; Li, C.; Fu, H. Y.; Wang, C. Y.; Guo, X. J.; Jiang, Z.;
6082 Wu, G. Z.; Chen, S. M. Temperature-Induced Molecular Rearrange- Liquids: Spectroscopic and Density Functional Theory studies. J. Mol. 6150

6083 ment of an Ionic Liquid Confined in Nanospaces: An in Situ X-ray Liq. 2016, 213, 13−16. 6151

6084 Absorption Fine Structure Study. J. Phys. Chem. C 2015, 119, 22724− (540) Kiefer, J.; Obert, K.; Fries, J.; Bösmann, A.; Wasserscheid, P.; 6152

6085 22731. Leipertz, A. Determination of Glucose and Cellobiose Dissolved in the 6153
6086 (525) Ding, Y. S.; Xiong, R. Y.; Wang, S. S.; Zhang, X. M. Aggregative Ionic Liquid 1-Ethyl-3-Methylimidazolium Acetate Using Fourier 6154
6087 Structure of the 1-Octadecyl-3-Methylimidazolium Cation in rhe Transform Infrared Spectroscopy. Appl. Spectrosc. 2009, 63, 1041− 6155
6088 Interlayer of Montmorillonite and Its Effect on Polypropylene Melt 1049. 6156
6089 Intercalation. J. Polym. Sci., Part B: Polym. Phys. 2007, 45, 1252−1259. (541) Kunov-Kruse, A. J.; Riisager, A.; Saravanamurugan, S.; Berg, R. 6157
6090 (526) Ha, J. U.; Xanthos, M. Sequential Modification of Cationic and W.; Kristensen, S. B.; Fehrmann, R. Revisiting the Bronsted Acid 6158
6091 Anionic Nanoclays with Ionic Liquids. Green Chem. Lett. Rev. 2011, 4, Catalysed Hydrolysis Kinetics of Polymeric Carbohydrates in Ionic 6159
6092 103−107. Liquids by in situ ATR-FTIR Spectroscopy. Green Chem. 2013, 15, 6160
6093 (527) Ye, W.; He, S. L.; Ding, L. P.; Yao, Y. J.; Wan, L.; Miao, S. D.; 2843−2848. 6161
6094 Xu, J. Z. Supported Ionic-Liquid ″Semi-Heterogeneous Catalyst″: An (542) Zhang, X.; Ma, J.; Ji, Z.; Yang, G. H.; Zhou, X.; Xu, F. Using 6162
6095 Interfacial Chemical Study. J. Phys. Chem. C 2013, 117, 7026−7038. Confocal Raman Microscopy to Real-Time Monitor Poplar Cell Wall 6163
6096 (528) Siqueira, L. J. A.; Constantino, V. R. L.; Camilo, F. F.; Torresi, Swelling and Dissolution During Ionic Liquid Pretreatment. Microsc. 6164
6097 R. M.; Temperini, M. L. A.; Ribeiro, M. C. C.; Izumi, C. M. S. Probing Res. Tech. 2014, 77, 609−618. 6165
6098 the Local Environment of Hybrid Materials Designed from Ionic (543) Papanyan, Z.; Roth, C.; Wittler, K.; Reimann, S.; Ludwig, R. 6166
6099 Liquids and Synthetic Clay by Raman Spectroscopy. Spectrochim. Acta, The Dissolution of Polyols in Salt Soltuions and Ionic Liquids at 6167
6100 Part A 2014, 122, 469−475. Molecular Level: Ions, Counter Ions and Hofmeister Effects. 6168
6101 (529) Lv, G. C.; Li, Z. H.; Jiang, W. T.; Chang, P. H.; Liao, L. B. ChemPhysChem 2013, 14, 3667−3671. 6169
6102 Interlayer Configuration of Ionic Liquids in a Ca-Montmorillonite as (544) Papanyan, Z.; Roth, C.; Paschek, D.; Ludwig, R. Understanding 6170
6103 Evidenced by FTIR, TG-DTG, and XRD Analyses. Mater. Chem. Phys. the Dissolution of Polyols by Ionic Liquids Using the Example of a 6171
6104 2015, 162, 417−424. Well-Defined Model Compound. ChemPhysChem 2011, 12, 2400− 6172
6105 (530) Martinelli, A.; Matic, A.; Jacobsson, P.; Börjesson, L.; Navarra, 2404. 6173
6106 M. A.; Panero, S.; Scrosati, B. A Structural Study on Ionic-Liquid- (545) Brandão, C. R. R.; Costa, L. A. F.; Breyer, H. S.; Rubim, J. C. 6174
6107 Based Polymer Electrolyte Membranes. J. Electrochem. Soc. 2007, 154, Surface-enhanced Raman Scattering (SERS) of a Copper Electrode in 6175
6108 G183−G187. 1-N-Butyl-3-Methylimidazolium Tetrafluoroborate Ionic Liquid. Elec- 6176
6109 (531) Costa, L. T.; Siqueira, L. J. A.; Nicolau, B. G.; Ribeiro, M. C. C. trochem. Commun. 2009, 11, 1846−1848. 6177
6110 Raman Spectra of Polymer Electrolytes Based on Poly(Ethylene (546) Rubim, J. C.; Trindade, F. A.; Gelesky, M. A.; Aroca, R. F.; 6178
6111 Glycol)dimethylether, Lithium Perchlorate, and the Ionic Liquid 1- Dupont, J. Surface-Enhanced Vibrational Spectroscopy of Tetrafluor- 6179
6112 Butyl-3-methylimidazolium Hexafluorophosphate. Vib. Spectrosc. 2010, oborate 1-N-Butyl-3-Methylimidazolium (BMIBF4) Ionic Liquid on 6180
6113 54, 155−158. Silver Surface. J. Phys. Chem. C 2008, 112, 19670−19675. 6181

BG DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

6182 (547) Santos, V. O., Jr.; Leite, I. R.; Brolo, A. G.; Rubim, J. C. The
6183 Electrochemical Reduction of CO2 on a Copper Electrode in 1-N-
6184 Butyl-3-Methylimidazolium Tetrafluoroborate (BMI.BF4) Monitored
6185 by Surface-Enhanced Raman Scattering (SERS). J. Raman Spectrosc.
6186 2016, 47, 674−680.
6187 (548) Santos, V. O., Jr.; Alves, M. B.; Carvalho, M. S.; Suarez, P. A.
6188 Z.; Rubim, J. C. Surface-Enhanced Raman Scattering at the Silver
6189 Electrode/Ionic Liquid (BMIPF6) Interface. J. Phys. Chem. B 2006,
6190 110, 20379−20385.
6191 (549) Harroun, S. G.; Abraham, T. J.; Prudhoe, C.; Zhang, Y.;
6192 Scammells, P. J.; Brosseau, C. L.; Pye, C. C.; Singer, R. D.
6193 Electrochemical Surface-Enhanced Raman Spectroscopy (E-SERS) of
6194 Novel Biodegradable Ionic Liquids. Phys. Chem. Chem. Phys. 2013, 15,
6195 19205−19212.

BH DOI: 10.1021/acs.chemrev.6b00461
Chem. Rev. XXXX, XXX, XXX−XXX

You might also like