DOI:
10.1039/C6RA18589D
(Paper)
RSC Adv., 2016,
6, 86082-86088
Improving SO2 capture by basic ionic liquids in an acid gas mixture (10% vol SO2) through tethering a formyl group to the anions†
Received
22nd July 2016
, Accepted 1st September 2016
First published on
5th September 2016
Abstract
Low partial pressure SO2 gas can be efficiently absorbed by ionic liquids (ILs), especially by basic ILs, through strong chemical interaction. However, it is still a challenge to achieve a high absorption capacity under these SO2 conditions. In this work, a new class of formyl-containing aprotic task-specific ILs, including [P66614][3-CHO-Indo], [P66614][2-CHO-Pyro], [P66614][4-CHO-PhCOO], and [P66614][4-CHO-PhO], were designed and prepared through an acid–base neutralization reaction. The absorption of SO2 in simulated acid gas (SO2/N2 = 10/90 vol) by these functional ILs was investigated. Compared with the formyl-free ILs, up to 100% increased capacity could be achieved by introducing a formyl group into the basic anions. In addition, the interactions between formyl-containing task-specific ILs and SO2 were investigated through a combination of multiple techniques, including FT-IR, NMR, and quantum chemical calculations. It was found that the remarkable enhancement of SO2 absorption capacity under low partial pressure was due to the C
O⋯S interaction and C–H⋯O hydrogen bonding. Furthermore, the SO2 absorption process by the ILs could be repeated several times without loss of absorption capability.
Introduction
In the past few decades, ionic liquids (ILs) have found wide spread applications as a kind of “green media” in chemical synthesis, biomass dissolution, energy production, separation science, and gas capture. They were highly emphasized by researchers because of their unique properties, such as ultra-low vapor pressure, wide liquid temperature range, non-flammability, chemical stability, and tunable structure and properties.1–7 ILs have already been widely used in the separation of acidic gases such as SO2,8,9 CO2,10–15 H2S,16–19 BF3 (ref. 20) and NOx.21,22 In these acidic gases, SO2 is a significant source of atmospheric pollution that directly threatens the environment and human health, thus minimizing the emissions of SO2 is highly important. Herein, we developed a new strategy for the highly efficient capture of SO2 by basic ILs in an acid gas mixture (a mixture of N2 and 10% volume content of SO2) through introducing a formyl group into the basic anions.
Several conventional removal processes, such as limestone scrubbing and ammonia scrubbing, have been developed for flue gas desulfurization (FGD), and been widely used in industry, where the concentration of SO2 in flue gas was up to thousands of ppm.23,24 Nevertheless, the inherent disadvantages of these technologies should not be ignored, including the production of large quantities of wastewater and useless by-products.25–28 Compared with conventional removal process, the SO2 capture through ILs will not produce useless by-products and the captured SO2 also can be collected for directly use in Claus reaction.29 In addition, ILs can be easily recycled through heating and bubbling with N2 without vast mass loss or quality change. Thus, SO2 capture by ILs is more consist with the concept of sustainability.
High-concentration SO2 has a high solubility in some ILs through physical interaction, especially in ether-functionalized ILs.30 Thereby, effective capture of SO2 from flue gas requires strong interaction between IL and SO2 because of the relatively low SO2 partial pressure in this stream. Han et al.31 reported the first example for chemical absorption of SO2 by 1,1,3,3-tetramethylguanidine lactate ([TMG][L]), the absorption capacity of which was about 1.0 mole SO2 per mole of IL at 1 bar with 8% SO2 in a gas mixture of SO2 and N2. Since then, the SO2 capture by several other kinds of ILs had been studied. For example, Zhang et al.32 reported a kind of hydroxyl ammonium ILs which can uptake SO2 through chemical interaction. Elliott et al.33 investigated the solubility of SO2 in imidazolium-based ILs through theoretical studies using Monte Carlo simulation. Wang et al.34 designed a series of halogenated ILs to improve SO2 capture through tuning the anion of ILs. Zhang et al.35 studied SO2 absorption by several TMG-based ILs. Wu et al.36 reported a kind of porous polymer for the SO2 adsorption. Other investigations using supported ionic liquid membranes (SILMs) for SO2 gas separation had also been reported.37,38
In the past few years, Wang et al.39 reported a kind of phosphonium IL trihexyl(tetradecyl) phosphonium tetrazolate ([P66614][Tetz]) which could capture 1.50 mole SO2 per mole of IL under 10% SO2, while Han et al.40 reported a kind of IL 1-(2-diethylaminoethyl)-3-methylimidazolium tetrazolate ([Et2NEMim][Tetz]) that could achieve 1.85 mole SO2 per mole of IL under same SO2 partial pressure through a combination of amine–SO2 and tetrazolate–SO2 interaction (Scheme 1). As the anion of ILs plays a key role in SO2 capture, Wang et al.41 designed another kind of anion-functionalized IL 1-ethyl-3-methylimidazolium thiocyanate ([Emim][SCN]) that could absorb 0.98 mole SO2 per mole of IL under 10% SO2 through S(anion)–S(SO2) sulfur bridge (Scheme 2). A similar result was also reported by Deng et al.42 using other [SCN]-based ILs. Kirchner et al.43 carried out an ab initio molecular dynamics study on the SO2 solvation in [Emim][SCN]. However, these functionalized ILs were based on the single-site absorption mechanism under low concentration of SO2. That is, single-site interaction can reach the 1
:
1 (SO2: anion or cation) stoichiometry for SO2 absorption.
 |
| Scheme 1 The reaction between [Et2NEMim][Tetz] and SO2.40 | |
 |
| Scheme 2 Optimized structure showing the interaction between the anion [SCN] and SO2.41 | |
Hence, how to increase the absorption site of the anion and enhance the interaction between anion with SO2 under low concentration of SO2 are essential for a breakthrough in absorption techniques. Up to now, many works of SO2 capture by ILs were mainly focused on the improvement of capacity under high SO2 concentration. Thus, development of alternative anion-functionalized ILs that are able to achieve high capacity under low SO2 partial pressure is always highly desired.
In this work, we described a new strategy for significant improving SO2 capture by basic ILs under low partial pressure (10% vol SO2 mixed with 90% vol N2) through introducing a formyl group into the basic anions. For this purpose, several kinds of formyl-containing ILs as well as formyl-free counterparts were prepared and applied to the capture of SO2 (see Scheme 3 for structures of the formyl-containing ILs). Indeed, these formyl-containing basic ILs were found to exhibit a high capacity up to 1.92 mole SO2 per mole of IL, compared with their formyl-free counterparts. The promoting role of formyl group tethered to the basic anions of ILs for the enhancement of SO2 capture under low partial pressure was studied through a combination of spectroscopic investigations and quantum chemical calculations, and the importance of the C
O⋯S interaction and C–H⋯O hydrogen bonding between formyl-based anion and SO2 was emphasized.
 |
| Scheme 3 Structures of cation and anions of the formyl-containing ILs studied in this work. | |
Experimental
Materials
Several kinds of different formyl-containing basic compounds, such as indole-3-carboxaldehyde (3-CHO-Indo), pyrrole-2-carboxaldehyde (2-CHO-Pyro), 4-formylbenzoic acid (4-CHO-PhCOOH) and p-hydroxybenzaldehyde (4-CHO-PhOH), were purchased from Sigma-Aldrich as the proton donors. Trihexyl(tetradecyl) phosphonium bromide ([P66614][Br]) was purchased from Strem Chemicals. A series of gases with different SO2 partial pressure were prepared by mixing SO2 (99.95%) and N2 (99.9993%) which were obtained from Beijing Oxygen Plant Specialty Gases Institute Co., Ltd. An anion-exchange resin (Amersep 900 OH) was obtained from Alfa Aesar. All the chemicals were in the highest purity grade possible, and were used as received unless otherwise stated.
Preparation of ILs
In a typical synthesis of a formyl-containing IL such as [P66614][3-CHO-Indo], equimolar 3-CHO-Indo was added to an ethanol solution of phosphonium hydroxide ([P66614][OH]), which was prepared from [P66614][Br] by the anion-exchange method.15,44,45 The mixture was stirred at room temperature for 24 h. Then, ethanol and water were distilled off at 60 °C under reduced pressure. All the ILs obtained were dried with P2O5 under vacuum at 60 °C for 24 h to reduce possible traces of water.
Characterizations
1H NMR and 13C NMR spectra were recorded on a Bruker spectrometer (400 MHz) in DMSO-d6 or CDCl3 with tetramethylsilane (TMS) as the standard. FT-IR spectra were recorded on a Nicolet 4700 FT-IR spectrometer. The structures of these ILs were confirmed by NMR and FT-IR spectra measurements, and no impurities were found by NMR spectra. NMR and FT-IR spectra measurements were also performed to track SO2 binding and release. The water content in these ILs was determined with a Karl Fisher titrator (Mettler Toledo DL32, Switzerland) and found to be less than 0.1 wt%. The bromide content in these ionic liquids was determined by means of a Br− selective electrode (Shanghai Precision & Scientific Instrument Co. Ltd) coupled with a saturated calomel electrode (Shanghai Precision & Scientific Instrument Co. Ltd), which showed that bromide content was lower than 0.009 mole per kilogram.
SO2 capture and release
In the measurements of SO2 absorption under low partial pressure (0.1 bar), SO2 diluted with N2 (SO2/N2 = 10/90 vol) was bubbled through about 1.0 g IL in a glass container with an inner diameter of 10 mm, and the flow rate was about 40 ml min−1. The glass container was partly immersed in a circulation water bath controlled at desirable temperature. The amount of SO2 absorbed was determined at regular intervals by the electronic balance with an accuracy of ±0.1 mg until the weight remained constant. The amount of SO2 absorbed could be calculated by subtracting the amount of IL. The standard deviation of the absorption loadings under 10% SO2 was 0.05 mole of SO2 per mole of IL. The simulated acid gases with different SO2 partial pressure were achieved by changing the flow rate ratio of SO2 and N2. After capture, the ILs were regenerated by heating or bubbling N2 through the SO2-saturated ILs. In a typical desorption of SO2, N2 at atmospheric pressure was bubbled through about 1.0 g SO2-saturated IL in a glass container, which was partly immersed in a circulation oil bath at 80 °C, and the flow rate was about 40 ml min−1. The release of SO2 was determined at regular intervals by gravimetric method.
Results and discussion
Absorption of SO2
To study the SO2 capture by basic ILs, several kinds of different formyl-containing proton donors (3-CHO-Indo, 2-CHO-Pyro, 4-CHO-PhCOOH, and 4-CHO-PhOH) were selected. These ILs were easily prepared by the acid–base neutralization between each weak proton donor and a solution of [P66614][OH] in ethanol, which was prepared by anion-exchange method. The structures of these ILs were verified by NMR and IR spectroscopy (see ESI†).
The effect of different formyl-containing ILs on the absorption of SO2 was investigated, and the results were shown in Table 1. It can be seen that the mole ratio of SO2 to [P66614][3-CHO-Indo] was 1.92 at 20 °C and 1 bar (SO2/N2 = 10/90 vol), which is close to twice of that by [P66614][Indo]. Besides, the SO2 absorption capacities by [P66614][2-CHO-Pyro] and [P66614][Pyro] were 1.87 and 0.92 mole of SO2 per mole of IL under same conditions. Thus, the results exhibit and indicate an extremely high increase of SO2 absorption capacity by the addition of a formyl group in the anion. Additionally, the capacities under the same conditions by [P66614][4-CHO-PhO] and [P66614][4-CHO-PhCOO] were 1.61 and 1.59 mole SO2 per mole of IL, whereas that by [P66614][PhO] and [P66614][PhCOO] were only 1.25 and 1.51 mole of SO2 per mole of IL, respectively. Clearly, these formyl-containing ILs exhibits a higher SO2 absorption capacity than that by formyl-free counterparts. Additionally, introducing the formyl group into ILs can also improve the solubility of other acid gases such as CO2.46
Table 1 The effect of different formyl-containing ILs on SO2 absorptiona
IL |
Capacityb |
Data reference |
10% SO2 |
100% SO2 |
SO2 capture at 20 °C and 1 bar until reach the equilibrium.
Mole of SO2 per mole of IL.
|
[P66614][3-CHO-Indo] |
1.92 |
4.24 |
This work |
[P66614][Indo] |
1.00 |
2.92 |
47
|
[P66614][2-CHO-Pyro] |
1.87 |
4.15 |
This work |
[P66614][Pyro] |
0.92 |
2.51 |
48
|
[P66614][4-CHO-PhO] |
1.61 |
3.73 |
This work |
[P66614][PhO] |
1.25 |
3.02 |
34
|
[P66614][4-CHO-PhCOO] |
1.59 |
3.59 |
This work |
[P66614][PhCOO] |
1.51 |
3.74 |
34
|
It is also clear from Table 1 that the performance of formyl group was greatly affected by the basicity of the ILs. There is no doubt that the formyl group substituent can decrease the basicity of such weak acids as indole, pyrrole, benzoic acid, and phenol. However, the charge on these anions could flow to the formyl groups and make them as an efficient interaction site. For such ILs with strongly basic anions as [P66614][Indo] (pKa = 16.2,49 in water) and [P66614][Pyro] (pKa = 16.5,50 in water), when formyl group was added into the anions, the increments under 10% SO2 were 0.92 and 0.95 mole of SO2 per mole of IL for [P66614][4-CHO-Indo] and [P66614][4-CHO-Pyro], respectively. While, when formyl group was added into such ILs with weakly basic anions as [P66614][PhO] (pKa = 9.99,51 in water) and [P66614][PhCOO] (pKa = 4.2,51 in water), the increments under 10% SO2 were only 0.36 and 0.08 mole of SO2 per mole of IL for [P66614][4-CHO-PhO] and [P66614][4-CHO-PhCOO], respectively.
The increased absorption capacities could be a result of the interactions between formyl group on the anion and SO2. For [Indo] and [Pyro], their strong basicity results in the strong C
O⋯S interaction and C–H⋯O hydrogen bonding between –CHO and SO2, thus significantly increased SO2 absorption capacities by [P66614][3-CHO-Indo] and [P66614][2-CHO-Pyro]. However, for [PhO] and [PhCOO] with low basicity, the effect of formyl group on the absorption capacity by [P66614][4-CHO-PhCOO] is low, because the C
O⋯S interaction and C–H⋯O hydrogen bonding were weak. The results indicated that the effect of formyl group on the anion was strongly affected by the basicity of the anion. Thus, the SO2 absorption can be facilely improved by formyl-containing ILs through tuning the basicity of the ILs.
Effect of temperature and pressure on SO2 absorption
The effect of temperature and partial pressure on the SO2 absorption by [P66614][3-CHO-Indo] was investigated as an example. Fig. 1a shows the effect of SO2 partial pressure on its absorption by [P66614][3-CHO-Indo] at 20 °C. It was shown that as the pressure decreased from 1.0 to 0.1 bar, the molar ratio of SO2 to [P66614][3-CHO-Indo] dropped from 4.24 to 1.92. Fig. 1b shows the temperature dependence of the SO2 absorption at 1.0 bar. It can be seen that when the temperature was increased from 20 °C to 80 °C, SO2 absorption capacity by [P66614][3-CHO-Indo] decreased from 4.24 to 1.70 mole of SO2 per mole of IL. It is indicated that the captured SO2 by [P66614][3-CHO-Indo] could be facilely stripped by heating or bubbling N2 through the IL.
 |
| Fig. 1 Effect of SO2 partial pressure (at 20 °C, a) and temperature (at 1.0 bar, b) on SO2 absorption performance by [P66614][3-CHO-Indo]. | |
In addition, multiple SO2 absorption cycles were investigated by using [P66614][3-CHO-Indo] as an example (Fig. 2). It is evident that the SO2 absorption process could be recycled for more than six times without a loss of absorption capability, indicating that SO2 absorption process by [P66614][3-CHO-Indo] was highly reversible.
 |
| Fig. 2 Six consecutive cycles of SO2 absorption and release by [P66614][3-CHO-Indo]. SO2 absorption was carried out at 20 °C and 1 bar under 100% SO2 (40 ml min−1), and desorption was performed at 80 °C and 1 bar under 100% N2 (40 ml min−1). | |
Quantum chemical calculations
To investigate the role of the two interaction sites on the anion [3-CHO-Indo] in SO2 absorption, we calculated the charge distribution of N and O atoms on the anion from Natural bond orbital (NBO)52 analysis using the Gaussian 09 program53 at B3LYP/6-31++G(p,d) level.54–56 It was shown in Fig. 3a that the Mulliken atomic charge of N and O atom in the anion [3-CHO-Indo] were −0.552 and −0.676, respectively, while that of N atom in the formyl-free anion [Indo] was −0.594.47 Compared with in [Indo], the NBO atomic charge of N atom in [3-CHO-Indo] decreased because the formyl group is a kind of electron withdrawing group. On the other hand, formyl group shared the negative charge of N atom in the anion, which enhanced interactions between formyl group and SO2, leading to the increase of absorption capacity.
 |
| Fig. 3 Optimized structures of [3-CHO-Indo] (a) and [3-CHO-Indo]–SO2 complexes (b-d) at B3LYP/6-31++G(d,p) level. (b), O atom in [3-CHO-Indo] with the closest SO2 molecule, [3-CHO-Indo]·SO2, ΔH = −97.1 kJ mol−1; (c–d), multiple-site chemical interactions between [3-CHO-Indo] and SO2 molecule: (c), N atom in [3-CHO-Indo] with the closest SO2 molecule, [3-CHO-Indo]–SO2, ΔH1 = −115.3 kJ mol−1; (d), [3-CHO-Indo]–2SO2, ΔH2 = −76.3 kJ mol−1. Note that van der Waals radii (in Å) are 1.70 (C), 1.20 (H), 1.52 (O), 1.55 (N), and 1.80 (S), respectively.57 O, red; S, yellow; N, blue; C, grey. | |
To further investigate the interactions between the formyl-containing anion [3-CHO-Indo] and SO2, we calculated the geometry and energy optimizations for the free anion, the free SO2 molecule, and each anion–SO2 complex by dispersion corrected density functional theory (DFT-D3(BJ)) at the B3LYP/6-31++G(p,d) level. The optimized structures reflecting the interactions between the anions and SO2 were shown in Fig. 3 and S1.† It can be seen from Fig. 3b and c that the intermolecular distance between O or N atoms in the anion [3-CHO-Indo] and S atom in SO2 was predicted to be 2.194 Å and 2.037 Å, which corresponds to a reduction of approximately 33.9% and 39.2% of the sum of the van der Waals radii of the two interacting atoms, respectively, indicating that the interaction of CHO–SO2 was weaker than N–SO2. The calculated SO2 absorption enthalpies for each site of [3-CHO-Indo] were −115.3 kJ mol−1 for N–SO2 and −97.1 kJ mol−1 for O–SO2, respectively. These data suggest that the interactions of each site with SO2 were both in the chemical regime due to the absorption enthalpies were larger than −50 kJ mol−1. It should be noted from Fig. 3b that hydrogen bond was formed simultaneously between H of –CHO and O of SO2, and the bond distance was predicted to be 2.252 Å.
In addition, the calculated absorption enthalpies for [3-CHO-Indo]–SO2 and [3-CHO-Indo]–2SO2 complexes were −115.3 and −76.3 kJ mol−1, respectively (Fig. 3c and d). Therefore, 1 mole of [3-CHO-Indo]-based IL could absorb about 2 mole of SO2 under low partial pressure, and formyl-containing ILs can be regarded as the most potential absorbent for SO2 absorption under 10% SO2.
Spectroscopic investigations
The interactions of formyl-containing ILs with SO2 were further investigated by FT-IR and NMR spectroscopy in order to provide new proof for the results obtained from experimental and theoretical calculations. Here, [P66614][3-CHO-Indo] was selected as a typical case for such a study.
FT-IR spectrum of [P66614][3-CHO-Indo] before and after reaction with SO2 was shown in Fig. 4. It is clear that new peaks at 1325 cm−1 and 1143 cm−1 appeared after the absorption of SO2 by [P66614][3-CHO-Indo], which could be assigned to asymmetric stretching and symmetric vibration of S
O bonds. On the other hand, the new peak at 947 cm−1 was attributed to S–O stretching,31,39 indicating SO2 chemisorption. Another new peak at 1045 cm−1 could be assigned to the vibration of O–S stretching mode of –OSO2−,58 which indicates the strong C
O⋯S interactions between the formyl group on the anion and SO2. Furthermore, a new peak at 2460 cm−1 should be assigned to the C–H⋯O
S hydrogen bonding interaction between the –CHO and SO2.47,58 Additionally, a stretching vibration peak of C–H in –CHO shifted from 2700 cm−1 to 2738 cm−1, also suggesting the C–H⋯O hydrogen bonding.59,60
 |
| Fig. 4 The FT-IR spectrum of the fresh IL [P66614][3-CHO-Indo] (grey area) and [P66614][3-CHO-Indo] after the absorption of SO2 (red line). | |
1H NMR spectra of [P66614][3-CHO-Indo] before and after reaction with SO2 was shown in Fig. 5. In comparison with the 1H NMR spectra of the fresh [P66614][3-CHO-Indo], the typical peak of formyl group in [3-CHO-Indo] moved downfield from 9.61 ppm to 9.93 ppm after reaction with SO2. Similarly, the typical peak of formyl group in the 13C NMR spectrum moved downfield from 179.9 to 185.4, also indicating the formation of C
O⋯S interaction and C–H⋯O hydrogen bonding between formyl group and SO2. Additionally, the peaks of two C atoms adjacent to the N atom on the anion [3-CHO-Indo] in the 13C NMR spectra shifted from 150.4 and 149.5 ppm to 139.0 and 137.6 ppm, respectively, which could be attributed to the strong N⋯S interaction.
 |
| Fig. 5
1H NMR (a) and 13C NMR (b) spectra of the fresh IL [P66614][3-CHO-Indo] (blue) and [P66614][3-CHO-Indo] after the absorption of SO2 (red). | |
Conclusions
In summary, a new strategy for improving SO2 uptake capacity under low partial pressure was developed through introducing a formyl group into the basic anions of ILs. Compared with the formyl-free counterparts, up to 100% increased capacity could be achieved by formyl-containing ILs. The effect of formyl group on the anion was strongly affected by the basicity of the anion. Quantum chemical calculations and spectroscopic investigations indicate that the absorption capacity increased when formyl group was introduced on the anion because of the additional C
O⋯S interaction and C–H⋯O hydrogen bonding between –CHO and SO2. Thus, this method provides a potential alternative for acid gas capture and separation such as SO2, CO2, and H2S.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (No. 21403059, 21133009), the Natural Science Foundation of Henan Province (No. 142300413213), the S&T Research Foundation of Education Department of Henan Province (No. 14A150031), and the Doctoral Scientific Research Foundation of Henan Normal University (No. qd13007).
Notes and references
- J. Dupont, R. F. de Souza and P. A. Z. Suarez, Chem. Rev., 2002, 102, 3667–3691 CrossRef CAS PubMed.
- W. Yao, H. Wang, G. Cui, Z. Li, A. Zhu, S. Zhang and J. Wang, Angew. Chem., Int. Ed., 2016, 55, 7934–7938 (
Angew. Chem.
, 2016
, 128
, 8066–8070
) CrossRef CAS PubMed.
- J. F. Huang, H. M. Luo, C. D. Liang, I. W. Sun, G. A. Baker and S. Dai, J. Am. Chem. Soc., 2005, 127, 12784–12785 CrossRef CAS PubMed.
- J. E. Bara, T. K. Carlisle, C. J. Gabriel, D. Camper, A. Finotello, D. L. Gin and R. D. Noble, Ind. Eng. Chem. Res., 2009, 48, 2739–2751 CrossRef CAS.
- M. D. Soutullo, C. I. Odom, B. F. Wicker, C. N. Henderson, A. C. Stenson and J. H. Davis, Chem. Mater., 2007, 19, 3581–3583 CrossRef CAS.
- Z. Lei, C. Dai and B. Chen, Chem. Rev., 2014, 114, 1289–1326 CrossRef CAS PubMed.
- D. Xiong, G. Cui, J. Wang, H. Wang, Z. Li, K. Yao and S. Zhang, Angew. Chem., Int. Ed., 2015, 54, 7265–7269 (
Angew. Chem.
, 2015
, 127
, 7373–7377
) CrossRef CAS PubMed.
- S. H. Ren, Y. C. Hou, W. Z. Wu, Q. Y. Liu, Y. F. Xiao and X. T. Chen, J. Phys. Chem. B, 2010, 114, 2175–2179 CrossRef CAS PubMed.
- G. Cui, W. Lin, F. Ding, X. Luo, X. He, H. Li and C. Wang, Green Chem., 2014, 16, 1211–1216 RSC.
- E. D. Bates, R. D. Mayton, I. Ntai and J. H. Davis, J. Am. Chem. Soc., 2002, 124, 926–927 CrossRef CAS PubMed.
- G. Cui, J. Wang and S. Zhang, Chem. Soc. Rev., 2016, 45, 4307–4339 RSC.
- Y. Zhang, S. Zhang, X. Lu, Q. Zhou, W. Fan and X. Zhang, Chem.–Eur. J., 2009, 15, 3003–3011 CrossRef CAS PubMed.
- C. Wang, H. Luo, D.-E. Jiang, H. Li and S. Dai, Angew. Chem., Int. Ed., 2010, 49, 5978–5981 (
Angew. Chem.
, 2010
, 122
, 6114–6117
) CrossRef CAS PubMed.
- J. F. Brennecke and B. E. Gurkan, J. Phys. Chem. Lett., 2010, 1, 3459–3464 CrossRef CAS.
- X. Luo, Y. Guo, F. Ding, H. Zhao, G. Cui, H. Li and C. Wang, Angew. Chem., Int. Ed., 2014, 53, 7053–7057 (
Angew. Chem.
, 2014
, 126
, 7173–7177
) CrossRef CAS PubMed.
- K. Huang, D. Cai, Y. Chen, Y. Wu, X. Hu and Z. Zhang, AIChE Journal, 2013, 59, 2227–2235 CrossRef CAS.
- A. J. Hossein, R. R. Mahboubeh, C. Ghotbi, J. H. Masih and A. A. Naser, J. Chem. Eng. Data, 2009, 54, 1844–1849 CrossRef.
- F. Y. Jou and A. Mather, Int. J. Thermophys., 2007, 28, 490–495 CrossRef CAS.
- M. B. Shiflett, A. M. S. Niehaus and A. Yokozeki, J. Chem. Eng. Data, 2010, 55, 4785–4793 CrossRef CAS.
- D. J. Tempel, P. B. Henderson, J. R. Brzozowski, R. M. Pearlstein and H. S. Cheng, J. Am. Chem. Soc., 2008, 130, 400–401 CrossRef CAS PubMed.
- A. L. Revelli, F. Mutelet and J. N. Jaubert, J. Phys. Chem. B, 2010, 114, 8199–8206 CrossRef CAS PubMed.
- M. B. Shiflett, A. M. S. Niehaus and A. Yokozeki, J. Phys. Chem. B, 2011, 115, 3478–3487 CrossRef CAS PubMed.
- Y. Hu, S. Naito, N. Kobayashi and M. Hasatani, Fuel, 2000, 79, 1925–1932 CrossRef CAS.
-
S. C. Chamberlain, Ph.D., thesis, Brigham Young University, 2012. http://scholarsarchive.byu.edu/etd/3319.
- H. Kikkawa, T. Nakamoto, M. Morishita and K. Yamada, Ind. Eng. Chem. Res., 2002, 41, 3028–3036 CrossRef CAS.
- X. Ma, T. Kaneko, T. Tashimo, T. Yoshida and K. Kato, Chem. Eng. Sci., 2000, 55, 4643–4652 CrossRef CAS.
- J. L. Anderson, J. K. Dixon, E. J. Maginn and J. F. Brennecke, J. Phys. Chem. B, 2006, 110, 15059–15062 CrossRef CAS PubMed.
- Z. Yang, L. He, Y. Zhao and B. Yu, Environ. Sci. Technol., 2013, 47, 1598–1605 CAS.
- K. Huang, X. Feng, X. M. Zhang, Y. T. Wu and X. B. Hu, Green Chem., 2016, 18, 1859–1863 RSC.
- S. Y. Hong, J. Im, J. Palgunadi, S. D. Lee, J. S. Lee, H. S. Kim, M. Cheong and K. D. Jung, Energy Environ. Sci., 2011, 4, 1802–1806 CAS.
- W. Wu, B. Han, H. Gao, Z. Liu, T. Jiang and J. Huang, Angew. Chem., Int. Ed., 2004, 43, 2415–2417 (
Angew. Chem.
, 2004
, 116
, 1421–1423
) CrossRef CAS PubMed.
- X. L. Yuan, S. J. Zhang and X. M. Lu, J. Chem. Eng. Data, 2007, 52, 596–599 CrossRef CAS.
- A. F. Ghobadi, V. Taghikhani and J. R. Elliott, J. Phys. Chem. B, 2011, 115, 13599–13607 CrossRef CAS PubMed.
- G. Cui, J. Zheng, X. Luo, W. Lin, F. Ding, H. Li and C. Wang, Angew. Chem., Int. Ed., 2013, 52, 10620–10624 (
Angew. Chem.
, 2013
, 125
, 10814–10818
) CrossRef CAS PubMed.
- Y. Shang, H. Li, S. Zhang, H. Xu, Z. Wang, L. Zhang and J. Zhang, Chem. Eng. J., 2011, 175, 324–329 CAS.
- L. Wu, D. An, J. Dong, Z. Zhang, B. G. Li and S. Zhu, Macromol. Rapid Commun., 2006, 27, 1949–1954 CrossRef CAS.
- Y.-Y. Jiang, Z. Zhou, Z. Jiao, L. Li, Y.-T. Wu and Z.-B. Zhang, J. Phys. Chem. B, 2007, 111, 5058–5061 CrossRef CAS PubMed.
- P. Luis, L. Neves, C. Afonso, I. Coelhoso, J. Crespo, A. Garea and A. Irabien, Desalination, 2009, 245, 485–493 CrossRef CAS.
- C. Wang, G. Cui, X. Luo, Y. Xu, H. Li and S. Dai, J. Am. Chem. Soc., 2011, 133, 11916–11919 CrossRef CAS PubMed.
- D. Yang, M. Hou, H. Ning, J. Ma, X. Kang, J. Zhang and B. Han, ChemSusChem, 2013, 6, 1191–1195 CrossRef CAS PubMed.
- C. Wang, J. Zheng, G. Cui, X. Luo, Y. Guo and H. Li, Chem. Commun., 2013, 49, 1166–1168 RSC.
- B. Yang, Q. Zhang, Y. Fei, F. Zhou, P. Wang and Y. Deng, Green Chem., 2015, 17, 3798–3805 RSC.
- D. S. Firaha, M. Kavalchuk and B. Kirchner, J. Solution Chem., 2015, 44, 838–849 CrossRef CAS PubMed.
- G. Cui, C. Wang, J. Zheng, Y. Guo, X. Luo and H. Li, Chem. Commun., 2012, 48, 2633–2635 RSC.
- K. Fukumoto, M. Yoshizawa and H. Ohno, J. Am. Chem. Soc., 2005, 127, 2398–2399 CrossRef CAS PubMed.
- F. Ding, X. He, X. Luo, W. Lin, K. Chen, H. Li and C. Wang, Chem. Commun., 2014, 50, 15041–15044 RSC.
- G. Cui, F. Zhang, X. Zhou, Y. Huang, X. Xuan and J. Wang, ACS Sustainable Chem. Eng., 2015, 3, 2264–2270 CrossRef CAS.
- G. Cui, F. Zhang, X. Zhou, H. Li, J. Wang and C. Wang, Chem.–Eur. J., 2015, 21, 5632–5639 CrossRef CAS PubMed.
- D. Virieux, A. F. Guillouzic and H. J. Cristau, Tetrahedron, 2006, 62, 3710–3720 CrossRef CAS.
- S. Inque and Y. Imanaka, J. Organomet. Chem., 1972, 35, 1–7 CrossRef.
-
W. M. Haynes, CRC Handbook of Chemistry and Physics, CRC Press/Taylor and Francis, Boca Raton, FL, 96th Edition (Internet Version), 2016 Search PubMed.
-
E. D. Glendening, A. E. Reed, J. E. Carpenter and F. Weinhold, NBO, Version 3.1 Search PubMed.
-
G. W. T. M. J. Frisch, H. B. Schlegel, G. E. Scuseria, J. R. C. M. A. Robb, G. Scalmani, V. Barone, B. Mennucci, H. N. G. A. Petersson, M. Caricato, X. Li, H. P. Hratchian, J. B. A. F. Izmaylov, G. Zheng, J. L. Sonnenberg, M. Hada, K. T. M. Ehara, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, O. K. Y. Honda, H. Nakai, T. Vreven, J. A. Montgomery Jr, F. O. J. E. Peralta, M. Bearpark, J. J. Heyd, E. Brothers, V. N. S. K. N. Kudin, R. Kobayashi, J. Normand, A. R. K. Raghavachari, J. C. Burant, S. S. Iyengar, J. Tomasi, N. R. M. Cossi, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, C. A. V. Bakken, J. Jaramillo, R. Gomperts, R. E. Stratmann, A. J. A. O. Yazyev, R. Cammi, C. Pomelli, J. W. Ochterski, K. M. R. L. Martin, V. G. Zakrzewski, G. A. Voth, J. J. D. P. Salvador, S. Dapprich, A. D. Daniels, J. B. F. O. Farkas, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed.
- A. D. Becke, Phys. Rev. A, 1988, 38, 3098–3100 CrossRef CAS.
- A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
- C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
- A. Bondi, J. Phys. Chem., 1964, 68, 441–451 CrossRef CAS.
- G. Cui, Y. Huang, R. Zhang, F. Zhang and J. Wang, RSC Adv., 2015, 5, 60975–60982 RSC.
- H. Matsuura, H. Yoshida, M. Hieda, S. y. Yamanaka, T. Harada, K. Shinya and K. Ohno, J. Am. Chem. Soc., 2003, 125, 13910–13911 CrossRef CAS PubMed.
- M. A. Blatchford, P. Raveendran and S. L. Wallen, J. Phys. Chem. A, 2003, 107, 10311–10323 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: NMR and IR data of the formyl-containing anion-functionalized ILs. See DOI: 10.1039/c6ra18589d |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.