Abstract
Ionic liquids (ILs) have been widely explored as alternative solvents for carbon dioxide (CO2) capture and utilization. However, most of these processes are under pressures significantly higher than atmospheric level, which not only levies additional equipment and operation costs, but also makes the large-scale CO2 capture and conversion less practical. In this study, we rationally designed glycol ether-functionalized imidazolium, phosphonium and ammonium ILs containing acetate (OAc–) or Tf2N– anions, and found these task-specific ILs could solubilize up to 0.55 mol CO2 per mole of IL (or 5.9 wt% CO2) at room temperature and atmospheric pressure. Although acetate anions enabled a better capture of CO2, Tf2N– anions are more compatible with alcohol dehydrogenase (ADH), which is a key enzyme involved in the cascade enzymatic conversion of CO2 to methanol. Our promising results indicate the possibility of CO2 capture under ambient pressure and its enzymatic conversion to valuable commodity.
Keywords: carbon dioxide capture, solubility, ionic liquid, decarbonization
1. Introduction
As carbon dioxide (CO2) concentration is drastically rising in atmosphere due to fossil fuel burning, carbon capture, sequestration and utilization (CCSU) has become a viable option for CO2 reduction and its conversion to valuable commodities such as formate/formic acid, formaldehyde, CO, bicarbonate, methanol, carboxylation products, and sugars, etc. (1–3). In particular, CO2 reduction to methanol is an attractive route for preparing C1 building blocks, and can be achieved by chemical, electrochemical, photochemical, or enzymatic methods (2, 4). First three methods suffer from issues such as low selectivity and difficulty in preparing effective catalysts; for example, electrochemical methods need excessive overpotentials and yield mixtures of products (2, 5). On the other hand, enzymatic approaches have the advantages of high specificity, high selectivity and being environmentally benign. However, enzymatic CO2 reduction in aqueous solutions is limited by poor CO2 solubility in water (2.9 g/L at 25 °C) (6). Recently, various ionic liquid (IL) systems have evolved into emerging tools for capturing high concentrations of CO2 (up to equal mole of the solvent). However, these ILs either are not compatible with enzymes (i.e. dehydrogenases), or chemically convert CO2 to an unreactive substrate for enzymatic reductions.
Post-combustion capture of CO2 can be achieved through a number of methods such as absorption, adsorption, cryogenic distillation, membrane separation, and microalgae growth (7). Conventional chemical absorption of CO2 using amine solutions (e.g. monoethanolamine) are effective but have several drawbacks including equipment corrosion, amine volatility, high construction cost, amine degradation by SO2, NO2, and O2 in the flue gas, and high energy input for absorbent regeneration (7, 8). The mixture of amines and ILs were found efficient for CO2 capture, but monoethanolamine-carbamate salt began to precipitate from ILs (9). ILs have several favorable properties for CO2 capture including low volatility, high thermal and chemical stability, various solvent polarity, and tunable structures for either physical or chemical gas absorption. CO2 capture by ILs has been extensively reviewed by several groups (10–13). A process simulation study suggests that CO2 absorption using 1-butyl-3-methylimidazolium acetate ([BMIM][OAc]) reduces the energy loss by 16%, the investment by 11% and equipment cost by 12% when compared with aqueous monoethanolamine process (14). This section intends to provide a brief overview of examples that are closely relevant to this project. In general, CO2 has a much higher solubility in ILs than other gases (e.g. CO, O2, N2, CH4, and C2H4), especially in ILs with fluorinated cations or anions; however, ILs have minimum solubility in CO2 (15–17). The use of fluorinated ILs raises some environmental concerns due to their poor biodegradability. CO2 dissolution in ILs can be through physisorption/physical interactions (e.g. electrostatic, van der Waals, and hydrogen bonding), or chemisorption/chemical reactions (18).
IL anions typically impose strong influence on CO2 solubility (19). MD simulations indicate strong CO2 interaction with anions (e.g. PF6−) but weak interaction with cations (e.g. BMIM+) (20, 21). Another MD simulations (22) suggest that CO2 molecules interact preferentially with the polar region of ILs, and coordinate with anions in a decreasing coordination order of BF4– > Br– > OAc– > prolinate (BF4– has multiple coordination sites and small size). A decreasing CO2 solubility corresponds to the order of anions: methide > Tf2N− > OTf− > PF6− ~ BF4− > dca− > NO3− (25 °C with the same BMIM+ cation; Tf2N− = bis(trifluoromethylsulfonyl)imide; dca− = dicyanamide) (23), PF6− > BF4− > NO3− > EtSO4− (24, 25), NO3− > SCN− (26), OAc− > Tf2N− > PF6− > BF4− (27), and OAc− >> MeSO4− > Cl− (28). The Jalili group (29, 30) found CO2 solubility following a similarly decreasing order of Tf2N– > OTf– > PF6– > BF4– in 1-(2-hydroxyethyl)-3-methylimidazolium based ILs at temperatures between 303.15 and 353.15 K and pressures up to 1.3 MPa. Ayad et al. (31) reported that ILs carrying the anion of tricyanomethanide [C(CN)3]− led to a higher CO2 solubility than other anions ([C(CN)3]− > BF4− > Tf2N− > PF6− > SCN−). Yokozeki et al. (32) observed similar CO2 solubilities in ILs based on CF3COO–, PF6–, and BF4– anions. Zhang et al. (33) reported ~15% higher CO2 solubility in 1-hexyl-3-methylimidazolium tris(pentafluoroethyl)trifluorophosphate than in [HMIM][Tf2N]. Fluorinated anions tend to dissolve more CO2; hydrogen-bond basicity of anions is not correlated with the solubility (e.g. CO2 solubility is similar in ILs containing BF4– and dca− anions (34)); the order of anions could depend on the temperature (23). CO2 can act as both a weak Lewis acid and a Lewis base; C-F bonds interact with Lewis acidic carbon atoms of CO2 molecules (35). However, Carvalho et al. (36) argued that fluorinated molecules including ILs are not superior to eicosane in dissolving CO2 except some fluorinated aromatic compounds. The issues with fluorinated compounds are their high cost and negative environment impact (11, 37, 38).
ILs carrying acetate anion (such as [BMIM][OAc]) could absorb less CO2 than aqueous monoethanolamine but more than physical absorption (39); different chemisorption mechanisms have been put forward including deprotonation of imidazolium C-2 position leading to the coupling of carbene with CO2 forming carboxylate (40), the formation of non-volatile (41) or electron donor-acceptor (oxygens of acetate as Lewis base interacting with CO2) complexes (42, 43), multidentate binding between CO2 and acetate (44, 45), carboxylate-induced ylide intermediate (of phosphonium) resulting in chemisorption of CO2 (46) (which is similar to the role of phenolate (47)), acetate interacting with water and CO2 to form bicarbonate (via ylide intermediate in some cases) (39, 46). Shiflett and co-workers (32, 41, 48) observed ~27 mol% (or 7.7 wt%) CO2 dissolved in [EMIM][OAc] and [BMIM][OAc] at 25 °C and 0.1 MPa. Interestingly, they noticed very low vapor pressure above the mixture containing up to 20 mol% CO2, implying the formation of non-volatile complex (but it is reversible). Similarly, Shaahmadi et al. (49) determined the CO2 solubility in [BMIM][OAc] at atmospheric conditions (25 °C and 0.99 bar) as 27.3 mol% but observed a much lower solubility (1.5 mol%) in [BMIM][BF4]. The Wallen group (45) pointed out the important role of cooperative C−H···O interaction as extra stabilizing factor in addition to Lewis acid–Lewis base interaction between CO2 and carbonyl groups with H atoms attached to carbonyl carbon or α-carbon atom. Their ab initio calculations indicate that methyl acetate has a strong interaction with CO2; they suggest that acetylation and incorporation of S=O bonds might be effective methods to design CO2-philic solvents. On the other hand, [BMIM][CF3COO] could dissolve much less CO2 (e.g. 1.8 mol% at 25 °C and 0.1 MPa) due to the physisorption process (32, 48, 50). Ab initio calculations suggest that both oxygen atoms in acetate or trifluoroacetate interact with CO2 while acetate is a stronger Lewis base than trifluoroacetate (42). However, Dupont and co-workers (51) argued that basic IL anions (e.g. acetate) do not act as nucleophiles to form covalent anion-CO2 adducts, but promote the formation of carbonate/bicarbonate due to the presence of trace water in ILs. Butyrate anion-containing ILs were found having high CO2 absorption capability than the acetate type (52).
The influence of cations on CO2 solubility is usually less drastic than anions. In general, a higher CO2 solubility can be found in an IL containing a longer alkyl chain (such as BMIM+ vs EMIM+, where EMIM+ refers to 1-ethyl-3-methylimidazolium) (25, 53–56); a longer alkyl chain grafted to pyrrolidinium exhibited minimum impact on CO2 dissolution under low pressures, but slightly increased CO2 solubility at higher pressures (> 5 MPa) (57). Short ether-chain or hydroxyl-functionalized ILs have complicated effects on CO2 dissolution (58). van Ginderen et al. (59) pointed out that ether oxygen atoms act as Lewis base to interact with the carbon atom (as Lewis acid) in CO2 forming a 1:1 van der Waals complex. ILs carrying long branched-alkyl chains or ether groups (such as Ecoeng 500) have increased free volume and thus exhibited a high affinity toward CO2 (which was comparable to [HMIM][Tf2N] at all pressures) (38). Sharma et al. (60) reported a high CO2 solubility in single ether-functionalized imidazolium ILs ([CH3OCH2-MIM]X), reaching ~0.9 mole CO2 per mole of IL at 30 °C; CO2 solubility increases with the type of anions as BF4− < dca− ~ PF6− ~ OTf− < Tf2N−. However, the Brennecke group (61) verified that such high CO2 chemisorption results are in error, and ether-functionalized ILs only physically absorb CO2. The Jalili group (29, 30) concluded that the solubility of CO2 in a number of 1-(2-hydroxyethyl)-3-methylimidazolium based ILs was greater than corresponding [EMIM]+-based ILs with the same anions. However, after comparing their solubility data with literature values, the Henni group (27, 62) implied that hydroxyl-functionalized ILs reduced CO2 solubility while ether-functionalized ammonium ILs increased the solubility. On the other hand, some ether- and hydroxyl-functionalized ILs showed no impact on CO2 solubility. Oligo(ethylene glycol)-grafted imidazolium ILs displayed similar dissolution power for CO2 as alkyl-substituted ILs (63). The Gomes group (64) suggested that in ester or ester ether-functionalized imidazolium ILs could dissolve similar amounts of CO2, methane and ethane as alkyl-substituted ILs. The same group (65) also found that CO2 solubility was not influenced by hydroxyl or ester groups in ammonium ILs, and was even reduced (up to 48 % decrease in mole fraction) in functionalized pyridinium ILs; an unfavorable entropic contribution to the Gibbs energy of solvation was likely the reason. Makino et al. (66) reported that an ester-functionalized IL actually reduced CO2 solubility compared with non-functionalized ILs while a short-ether-chain-grafted IL showed minimum impact on CO2 solubility on the mole fraction scale, but a higher molarity-scale solubility due to a smaller molar volume of functionalized IL. The same group (67) revisited the effect of ether-functionalization of both cations and anions on CO2 dissolution, and found the anion functionalization to introduce alkoxy sulfates was more effective to increase CO2 solubility on both the mole fraction and molarity scales. In the presence of superbases, hydroxy-containing surfactants and polyethylene glycols could dissolve CO2 via chemical absorption (68), and hydroxy-containing ILs could absorb equimolar CO2 reversibly (69).
Most physisorption studies in ILs applied relatively high pressures (>10 bar), but this is not practical for commercial-scale processes due to the large consumption of energy and high capital cost. The partial pressure of CO2 in flue gases is even below atmospheric pressure. At ambient pressure and temperature, the CO2 solubility in conventional ILs is only up to ~3.3 mol% or 0.3 wt% (based on low-pressure Henry’s constants of 30–50 bar for common PF6− and Tf2N− types of ILs (70)). The practical CO2 absorption conditions should be close to flue gas temperature (~45 °C) and pressure (~0.15 bar) (71).
Based on the forementioned rationales, this study focused on the design and synthesis of ether-functionalized phosphonium and ammonium-based ILs paired with acetate or Tf2N− anions for CO2 dissolution under ambient conditions. We further examined the compatibility of alcohol dehydrogenase (ADH) in 0.5 M aqueous solutions of these ILs, which provided some feasibility data about how these ILs could be further employed in the enzymatic conversion of captured CO2.
2. Materials and methods
2.1. Materials
3-Aminopropyltrimethoxysilane (1), 3-aminopropyltriethoxysilane (2), and N-(6-aminohexyl) aminopropyltrimethoxysilane (3) were acquired from Gelest Inc. (Morrisville, PA). Chemglass Cajon® Airfree® Bubbler with O.D.×H = 9×220 mm (Supplier No. AF-0085–01; VWR Catalog Number 80064–292), ammonium acetate (ultrapure), and tributylphosphine (95%, Acros Organics) were purchased from VWR (Radnor, PA). Triethylene glycol monomethyl ether (≥97.0% GC), 1-(2-bromoethoxy)-2-(2-methoxyethoxy)ethane (≥94.0%), 1,2-epoxy-2-methylpropane (≥99%, known as isobutylene oxide,1,1-dimethyloxirane, and 2,2-dimethyloxirane), and imidazole (≥99.0%) were obtained from TCI America (Portland, OR). Polyethylene glycol monomethyl ether with M.W. ~350 (PEG-350) and benzenesulfonyl chloride (≥98%) was acquired from BeanTown Chemical (Hudson, NH). Lithium bis(trifluoromethylsulfonyl)imide provided by VWR (Radnor, PA) is the product of Biosynth International Inc. (Oakbrook Terrace, IL). Amberlyst® A-26 (OH) ion exchange resin was produced by Alfa Aesar (Tewksbury, MA). Alcohol dehydrogenase (ADH) from Saccharomyces cerevisiae (catalog #A7011) as lyophilized powder (contains buffer salts) with ≥300 units/mg protein, and β-nicotinamide adenine dinucleotide (NADH), reduced disodium salt hydrate was purchased from Sigma-Aldrich (St. Louis, MO).
2.2. Synthesis and characterization of functionalized ILs
See Supporting Information for reaction schemes (Schemes S1–S8) and NMR characterizations.
2.3. CO2 solubility determination
The CO2 solubility in each solvent was determined by the bubbler setup shown in Figure 1. The empty bubbler was pre-weighed before adding 2.0–3.0 mL solvent, which was weighed again to determine the mass of each solvent. CO2 (99.9%) was bubbled into the solvent through a regulator to control the flow rate at about 0.05 SCFM (standard cubic feet per minute). The bubbler was kept under ambient conditions (22 °C and atmospheric pressure). The bubbler containing the solvent and absorbed CO2 was weighed on a balance every 15 min to observe the mass change over 1–3 h period. The absorption completed in 30–45 min in most cases. The mass of CO2 captured was determined by subtracting the mass of bubbler and solvent before absorption from the total mass (bubbler + solvent + CO2) after the absorption. The mass of CO2 in each solvent was further converted into CO2 solubility (wt%) and mole of CO2 per mol of solvent by using the mass of each solvent.
Figure 1.

Flow diagram of CO2 capture in ILs by using a bubbler.
2.4. Alcohol dehydrogenase (ADH) activity
A pre-determined amount of IL (final concentration as 0.5 M) was mixed with phosphate buffer (0.1 M, pH 6.5), 50 μL of 60 mM formaldehyde, and 50 μL of 20 mM NADH to make 980 μL solution. The absorbance of this mixture at 340 nm was determined as the blank. The reduction of formaldehyde to methanol was initiated by adding 20 μL of 500 mg/L ADH solution in phosphate buffer (0.1 M, pH 6.5) at room temperature (21 °C). The NADH absorbance in the reaction mixture at 340 nm was measured every 5 min to determine the initial reaction rate during the linear range. The molar absorptivity of NADH at 340 nm used in this study was 6,317 M–1 cm–1 at 25 °C and pH 7.8 (72). Relative ADH activity was calculated by dividing the ADH activity in 0.5 M solvent by the activity in phosphate buffer (0.1 M, pH 6.5).
3. Results and discussion
3.1. Design and synthesis of functionalized ILs
As discussed in Introduction, ether-functionalized ILs (i.e. containing alkoxy groups) tend to improve the CO2 solubility, while hydroxy groups likely reduce or have no impact on the solubility. Poly(ethylene glycol)s (PEGs) typically contain multiple alkoxy groups; in addition, they are inexpensive solvents that have negligible vapor pressure, high chemical and thermal stability, and low toxicity. CO2 solubility in PEGs increases with the molar mass of PEGs and pressure; however, longer PEG chains eventually lead to high solvent viscosity, therefore, up to PEG-400 is considered the optimum solvent (73). It was found that the mixing of PEG-600 or PEG-1000 with [BMIM][PF6] could lead to higher viscosities than individual components (so called ‘”hyperviscosity”) due to extensive hydrogen-bonding, but not in the case of PEG-200 and PEG-400 (74). Adding an alcohol such as 1-pentanol into PEG-200 could increase the CO2 solubility (75). The addition of PEG-200 into an IL (i.e. [Choline][Pro]) could considerably improve CO2 absorption and desorption rates (76). PEG-derived (77) or PEG-incorporated membranes (78) allowed a high CO2 permeability and selectivity due to the CO2-philic polyether. However, CO2 solubility decreased in aqueous diethanolamine with the increase in PEG-400 concentration (79). Therefore, this study incorporated multiple alkoxy groups in IL cations paired with Tf2N– or OAc– anions; both anions are known CO2-philic in nature as discussed in Introduction.
Through the combination of previously established methods (see Supporting Information) (80–85), we synthesized glycol-functionalized ILs (see Table 1) based on imidazolium (4), phosphonium (5–9) and alkylammonium (10 and 11) paired with Tf2N– and/or OAc– anions. As shown in Table 1, Tf2N–-based ILs could be easily dried in vacuum oven to have 0.1 wt% or lower water, while the acetate type still contained 1.6–5.0 wt% water even after one week of extensive drying possibly due to strong hydrogen-bonding interaction between water and the acetate. Ether-functionalized ILs tend to have relatively lower viscosities (dynamic viscosities of 53–131 mPa s and kinematic viscosities of 41–109 mm2/s at 30 °C) than alkyl analogues; due to higher flexibility of ether chains than rigid alkyl chains (86), the incorporation of ether chain in IL structures reduces intermolecular correlation (especially tail-tail segregation) and cation-anion specific interactions (87, 88).
Table 1.
CO2 solubility in functionalized solvents at room temperature and atmospheric pressure
| Structure | Water content (wt%)a | Dynamic viscosity, kinematic viscosity, and density (30 °C)b | CO2 solubility (wt%) | Mole of CO2 per mol of solvent | |
|---|---|---|---|---|---|
| 1 |
|
0.017 | 1.5535 mPa s 1.5496 mm2/s 1.0025 g/cm3 |
16.2 | 0.65 |
| 2 |
|
0.031 | 1.8538 mPa s 1.9697 mm2/s 0.9411 g/cm3 |
10.2 | 0.52 |
| 3 |
|
0.084 | 12.268 mPa s 12.637 mm2/s 0.9708 g/cm3 |
2.1 | 0.14 |
| 4 |
|
4.303 | 97.273 mPa s 88.385 mm2/s 1.1006 g/cm3 (11.17 wt% H2O) |
0.74 | 0.065 |
| 5 |
|
0.111 | 96.275 mPa s 78.277 mm2/s 1.2299 g/cm3 |
1.5 | 0.21 |
| 6 |
|
0.131 | 92.915 mPa s 76.811 mm2/s 1.2097 g/cm3 |
1.5 | 0.24 |
| 7 |
|
4.951 | 65.337 mPa s 66.359 mm2/s 0.9846 g/cm3 |
5.9 | 0.55 |
| 8 |
|
2.383 | 61.536 mPa s 60.838 mm2/s 1.0115 g/cm3 |
3.9 | 0.53 |
| 9 |
|
0.111 | 130.70 mPa s 109.27 mm2/s 1.1961 g/cm3 |
1.5 | 0.29 |
| 10 |
|
1.649 | 53.527 mPa s 50.194 mm2/s 1.0664g/cm3 |
5.3 | 0.37 |
| 11 |
|
0.09 | 54.900 mPa s 41.348 mm2/s 1.3278 g/cm3 |
1.3 | 0.15 |
Note:
The water content was determined by the coulometric Karl Fischer titration at 22 °C using Hydranal® Coulomat AG as the analyte.
The viscosity and density data at 30 °C were acquired by an Anton Paar SVM 3000 viscometer.
3.2. CO2 solubility in ILs
Conventional chemical absorption of CO2 using amine solutions (e.g. monoethanolamine) results in the formation of carbamate (89). Theoretically, 1 mole of primary amine could absorb 0.5 mol CO2 to form carbamate (Figure 2) (90) while an excess CO2 dissolved is due to physical absorption (91). On other other hand, Brennecke and co-workers (92) found that amino acid-anions could react with CO2 at nearly 1:1 molar ratio to form carbamic acid (Figure 2a) while physical absorption of 1 bar CO2 at 25 °C in amino acid-based ILs only accounted for 1–3% of total CO2 uptake. In addition, Poly(dimethyl siloxane) (PDMS) (93, 94) and PDMS terminated with trimethylsilyl groups (95) are known as CO2-philic solvents. Aminosilicones are commercially produced for textile and personal care application; aminosilicones and their mixtures with glycol could dissolve a significant amount of CO2 (up to 21.8 wt% at 1 bar and 40 °C, or up to ~9 wt% at 0.1 bar and 45 °C) via chemisorption (71, 96, 97). Therefore, by using the bubbler setup illustrated in Figure 1 at room temperature and atmospheric pressure, we evaluated three commercial aminosilanes (see Table 1), namely 3-aminopropyltrimethoxysilane (1), 3-aminopropyltriethoxysilane (2), and N-(6-aminohexyl)aminopropyltrimethoxysilane (3), as substitutes of conventional aqueous amines because these aminosilanes are less volatile, low in viscosities (1.6–12.3 mPa s at 30 °C), and have both amino and silane CO2-philic groups. The chemical absorption establishes the baseline for our physical absorption study using functionalized ILs. 3-Aminopropyltrimethoxysilane (1) showed a high capability of CO2 absorption with 16.2 wt% CO2 or 0.65 mol CO2 per mole of solvent, which clearly demonstrated the chemical absorption. 3-Aminopropyltriethoxysilane (2) displayed a lower CO2 solubility of 0.52 mol CO2 per mole of solvent due to the bulkiness of trialkoxysilane group. Despite having two amine groups (primary and secondary), N-(6-aminohexyl)aminopropyltrimethoxysilane (3) only absorbed 0.14 mol CO2 per mole of solvent because the mixture quickly turned into thick gel.
Figure 2.

Reactions of amine with carbon dioxide.
IL 4 consists of acetate anion and imidazolium cation functionalized with both tert-butanol and triethylene glycol ether, and exhibits weak ability of solubilizing CO2 (0.065 mol CO2 per mole of IL). The likely reason of poor CO2 solubility is the presence of hydroxy group as discussed earlier. Ether-functionalized phosphonium-based ILs (5-9) enabled much higher CO2 solubility (0.21–0.55 mol CO2 per mole of IL) at ambient conditions; interestingly, a longer glycol chain (6 vs 5; 8 vs 7) had a small impact on the CO2 dissolution. However, the change of anion from Tf2N– to OAc– drastically increased the CO2 solubility (from 0.24 in IL 6 to 0.55 in IL 7; from 0.29 in IL 9 to 0.53 in IL 8). A similar trend was observed in glycol-functionalized ammonium-based ILs (10 and 11), where the CO2 solubility increased from 0.15 in Tf2N–-based IL 11 to 0.37 in acetate-based IL 10. Therefore, even at room temperature and atmospheric pressure, glycol-functionalized ILs, especially those carrying acetate anions, could dissolve a considerable amount of carbon dioxide. A longer glycol chain on the cation imposes a minimum impact on the CO2 solubility (ILs 5, 6 and 9; ILs 7 and 8). However, different cation cores could have a drastic influence on the CO2 dissolution; for example, phosphonium IL 7 could dissolve much more CO2 than ammonium analogue 10 (0.55 vs 0.37 mol CO2 per mol IL). Most CO2 solubility data in ILs were reported at high pressures (>10 bar); there are limited literature results at atmospheric pressure that can be compared with. The CO2 solubility in [EMIM][OAc] and [BMIM][OAc] at 25 °C and 0.1 MPa were reported to be about 27 mol% despite of much lower solubilities in [BMIM][BF4] (1.5 mol%) (32, 41, 48, 49) and in [BMIM][CF3COO] (1.8 mol%) (32, 48, 50). Our acetate-based IL 10 is comparable to [BMIM][OAc] and [EMIM][OAc] for CO2 capture while ILs 7 and 8 could dissolve ~30 mol% more CO2. In addition, our glycol-functionalized ILs are less viscous (7, 8 and 10: 54–65 mPa s at 30 °C) than [EMIM][OAc] (105.3 mPa s at 30 °C) (98) and [BMIM][OAc] (274 mPa s at 30 °C) (99). The CO2 solubility in [BMIM][Tf2N] at 25 °C and 1.01 bar was reported as 3.1 mol% (19), so it is comparable with functionalized Tf2N– based ILs 5, 6, 9 and 11. Since oxygen (O2) and nitrogen (N2) are much less soluble in ILs compared with CO2 (19), this IL method is highly applicable to direct air capture of CO2.
The sorption of CO2 was examined by FT-IR spectra to determine the nature of chemisorption and/or physisorption. As shown in Figure 3, CO2 in 3-aminopropyltrimethoxysilane (1) displayed two characteristic IR bands at 1684 and 2338 cm–1. The 1684 cm–1 band corresponds to the 2° carbamate carbonyl stretching following steps in Figure 2; previous studies have suggested that secondary carbamates absorb at 1705–1722 cm–1 (100), and ammonium carbamate absorbs at 1684 cm–1 (101). The 2338 cm–1 band represents CO2 antisymmetric stretching mode, which indicates the physisorption of CO2 (102–104). CO2 in 3-aminopropyltriethoxysilane (2) showed similar IR bands (see Figure S1 in Supporting Information). In neat IL 7, there is a strong IR band at 1575 cm–1 for acetate carbonyl, which became weakened when CO2 was dissolved while the band at 1634 cm–1 was strengthened. The 1634 cm–1 band could be due to the interactions of acetate with CO2 via different mechanisms such as the formation of non-volatile (41) or electron donor-acceptor (oxygens of acetate as Lewis base interacting with CO2) complexes (42, 43), multidentate binding between CO2 and acetate (44, 45), carboxylate-induced ylide intermediate (of phosphonium) resulting in chemisorption of CO2 (46) (which is similar to the role of phenolate (47)). In IL 7, CO2 also exhibited the physisorption band at 2336 cm–1. In two other acetate-based ILs (10 and 8, see Figures S1 and S2 respectively in Supporting Information), similar IR bands were observed except IL 10 showed a minimum band at 2338 cm–1. In Tf2N–-based IL 9, only the physisorption of CO2 band at 2338 cm–1 was shown, and the band at 1684 cm–1 was absent indicating no complexing interaction between CO2 and Tf2N– anion (54). Carbonate and bicarbonate ions in aqueous solutions typically show strong infrared band in the region of 1300−1400 cm−1 (104). Figures 3, S1 and S2 suggest the absence of carbonate and bicarbonate anions in these solvents.
Figure 3.

FT-IR spectra of CO2 dissolution in 3-aminopropyltrimethoxysilane (1) and IL 7 (The spectra were obtained through using a Thermo-Fisher Scientific Nicolet™ iS™ 5 FT-IR spectrometer equipped with an iD5 ATR (attenuated total reflection) accessory with 32 scans at a resolution of 2 cm–1 followed by 5-point smoothing.).
Other than CO2, trace amounts of nitrogen oxides (NOx) and sulfur dioxide (SO2) can also be found in flue gases from burning fossil fuels. The solubility of SO2 in Tf2N– based ILs tends to be much higher than that of CO2, where the Henry law’s constants of SO2 physical absorption is more than one order of magnitude lower than that of CO2 (105). In an acetate-carrying IL, [EMIM][OAc], SO2 competes with CO2 in dissolution, leading to a large decrease of CO2 solubility (as much as 25%) in the presence of SO2 (106). Anthony and co-workers (19) found the nitrous oxide (N2O) solubility is similar to that of CO2 in [BMIM][Tf2N] due to quadrupole moments in both types of molecules. In summary, the presence of SO2 and NOx reduces the CO2 solubility in ILs, especially with water in ILs (as ILs are generally hygroscopic) (107). Therefore, SO2 and NOx should be removed through the modifications of combustion process and the post-combustion treatment.
3.3. Alcohol dehydrogenase (ADH) activity in IL solutions
Carbon dioxide can be converted to methanol by a cascade enzymatic reaction involving three dehydrogenases (see Figure 4): formate dehydrogenase (FateDH) converting CO2 to formate/formic acid, formaldehyde dehydrogenase (FaldDH) converting formate/formic acid to formaldehyde, and then alcohol dehydrogenase (ADH) converting formaldehyde to methanol (108). In each of these three reactions, reduced nicotinamide adenine dinuncleotide (NADH) acts as the terminal electron donor; with suitable electron donors, dehydrogenases are able to catalyze the reverse reactions (i.e. reduction). However, these reactions are usually carried out in aqueous buffers, where the solubility of CO2 is very small. The low substrate (CO2) availability becomes the bottleneck of the overall cascade reaction. In addition, the first reaction is usually considerably slower (e.g. 30 times (109)) than its reverse reaction (i.e. the oxidation of formate to CO2), partly due to the low substrate solubility (110), but also due to low affinity of CO2 with the enzyme (109).
Figure 4.

Enzymatic conversion of CO2 to methanol using in situ regeneration of NADH. (Abbreviations of enzymes: FateDH—formate dehydrogenase; FaldDH—formaldehyde dehydrogenase; ADH—alcohol dehydrogenase; GDH— glucose dehydrogenase).
To demonstrate the concept of carrying out these reactions in functionalized IL solutions, we examined the last step of cascade reaction (ADH-catalyzed reduction of formaldehyde to methanol) in 0.5 M IL solutions (Table 2). Despite their chemical absorption power towards CO2, aminosilanes (1–3) at 0.5 M concentration essentially completely inhibited the ADH activity possibly due to the strongly basic nature of amine groups. As shown in Table 2, Tf2N–-based ILs (5, 6, 9, and 11) are less inhibitory (59–85% relative activity) than acetate-containing ILs (4, 7, 8, and 10) (1.1–51% relative activity). Interestingly, a longer glycol chain in Tf2N–-based ILs (5, 6, and 9) led to a lower ADH activity (85% > 72% > 59%). Overall, ILs 5 and 6 showed the best performance in terms of high CO2 solubility (0.21 and 0.24 mol CO2 per mole of IL, respectively) and high ADH activity (85% and 72% relative activity, respectively). IL 10 performed the best among all acetate-based ILs for CO2 capture (5.3 wt% CO2, or 0.37 mol CO2 per mole of IL) and ADH relative activity (36%). Therefore, strong CO2-philic solvents often cause enzyme deactivating, and an optimized solvent for both CO2 capture and enzyme compatibility can be obtained by designing the IL structure. It is well known that enzymes are more compatible with hydrophobic ILs (e.g. carrying Tf2N– and PF6– anions) than hydrophilic ILs containing basic anions such as acetate; a high concentration of acetate anion could form strong hydrogen bonds with enzymes to inactivate them (111, 112). Eckstein and co-workers (113) found that alcohol dehydrogenase from Lactobacillus brevis showed a higher reduction activity in a biphasic system with [BMIM][Tf2N] than with methyl tert-butyl ether (MTBE). The Kroutil group (114) studied the reduction of ketones by alcohol dehydrogenase ADH-‘A’ from Rhodococcus ruber in aqueous ILs, and found 20% (v/v) [BMIM][Tf2N] (~0.52 M) as a biphasic system reduced the enzyme activity to <60% of that in aqueous buffer while 20% (v/v) [BMIM][OAc] (~1.1 M) or [EMIM][OAc] (~1.3 M) as a homogeneous phase only slightly reduced the activity. Weibels et al. (115) compared the activity of yeast alcohol dehydrogenase in 0.5 M IL solutions during ethanol oxidation, and suggested that the enzymatic efficiency followed the Hofmeister series for anions: Cl– > Br– > EtSO4– > OTf– > BF4– > dca– > SCN–, and for cations: Me4N+ > Cholinium+ > EMIM+ > Et4N+ > Bu4N+ > guanidinium+ > BMIM+. Bekhouche and co-workers (116) reported that formate dehydrogenase lost >50% of its activity in 10% (v/v) [BMIM][OAc] (~0.55 M). D’Oronzo et al. (117) observed that formate dehydrogenase’s oxidation activities were preserved well in up to 40% (v/v) [BMIM][CH3SO3] but dropped quickly in >5% (v/v) [BMIM][OAc] (>0.28 M). It was rationalized that IL ions could have a direct binding with the protein resulting in conformational changes, and IL anions can bind with NAD+ to disrupt the cofactor from binding with formate dehydrogenase. A more systematic evaluation of different dehydrogenases in different concentrations of various ILs is needed to fully understand how ILs influence the enzyme activity and stability.
Table 2.
Relative alcohol dehydrogenase (ADH) activity in 0.5 M solution
| Structure | Relative ADH activity in 0.5 M solution c | |
|---|---|---|
| 1 |
|
0.54% |
| 2 |
|
0.0% |
| 3 |
|
0.0% |
| 4 |
|
51% |
| 5 |
|
85% |
| 6 |
|
72% |
| 7 |
|
1.1% |
| 8 |
|
17% |
| 9 |
|
59% |
| 10 |
|
36% |
| 11 |
|
75% |
Note: The relative ADH activity was calculated from the division of ADH activity in the presence of 0.5 M silylether or IL by ADH activity in phosphate buffer (0.1 M, pH 6.5).
To further illustrate the conflicting ability of solvents in dissolving CO2 and maintaining high ADH activity, data in Tables 1 and 2 are plotted in Figure 5 to suggest that more CO2-philic solvents are generally less ADH-compatible, especially when the CO2 solubility is expressed in moles (except several solvents such as 3 and 4 in Table 1). N-(6-aminohexyl) aminopropyltrimethoxysilane (3) contains two amine groups and is probably too basic for ADH to maintain a high activity. Dual-functionalized IL (4) is more ADH-compatible (51% relative activity) than other acetate-based ILs (1.1–17%) because its hydroxy group interacts with acetate anion to minimize its enzyme-inactivating nature; for the same reason, the hydroxy group weakens the ability of acetate in dissolving CO2. The Henni group (27, 62) pointed out that hydroxy-functionalized ILs reduced CO2 solubility while ether-functionalized ammonium ILs increased the solubility.
Figure 5.

Correlation of solvent compatibility with ADH and CO2 solubility.
4. Conclusions
Glycol ether-functionalized phosphonium and ammonium ILs carrying acetate and Tf2N– anions could dissolve up to 0.55 mol CO2 per mole of IL (or 5.9 wt% CO2) at room temperature and atmospheric pressure. Although acetate anions promoted a higher CO2 solubility, Tf2N– anions allowed a higher compatibility with ADH. These preliminary results imply the feasibility of using functionalized ILs to capture CO2 and convert it to methanol via an enzymatic route. Potential challenges of using ILs for CO2 capture at an industrial scale include the high IL costs, relatively high IL viscosities, and IL biodegradability. Therefore, further studies should focus on the recycling and reusing of these ILs, and their potential environmental impact.
Supplementary Material
Funding
This material is partly based upon work supported by the National Science Foundation under Grant No. [2244638]. Research reported in this publication was partly supported by the National Institute of General Medical Sciences of the National Institutes of Health under Award Number R15GM143682. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.
Footnotes
Disclosure statement
The authors declare no competing financial interest.
REFERENCES
- [1].Appel AM; Bercaw JE; Bocarsly AB; Dobbek H; DuBois DL; Dupuis M; Ferry JG; Fujita E; Hille R; Kenis PJA; Kerfeld CA; Morris RH; Peden CHF; Portis AR; Ragsdale SW; Rauchfuss TB; Reek JNH; Seefeldt LC; Thauer RK; Waldrop GL Frontiers, opportunities, and challenges in biochemical and chemical catalysis of CO2 fixation. Chem. Rev 2013, 113, 6621–6658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [2].Shi J; Jiang Y; Jiang Z; Wang X; Wang X; Zhang S; Han P; Yang C Enzymatic conversion of carbon dioxide. Chem. Soc. Rev 2015, 44, 5981–6000. [DOI] [PubMed] [Google Scholar]
- [3].Shukla SK; Khokarale SG; Bui TQ; Mikkola J-PT Ionic Liquids: Potential materials for carbon dioxide capture and utilization. Front. Mater 2019, 6, 42. [Google Scholar]
- [4].Kaczur JJ; Yang H; Liu Z; Sajjad SD; Masel RI A review of the use of immobilized ionic liquids in the electrochemical conversion of CO2. J. Carbon Res 2020, 6, 33. [Google Scholar]
- [5].Reda T; Plugge CM; Abram NJ; Hirst J Reversible interconversion of carbon dioxide and formate by an electroactive enzyme. Proc. Natl. Acad. Sci. U. S. A 2008, 105, 10654–10658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [6].Yalkowski SH; He Y; Jain P Handbook of aqueous solubility data second edition. 2nd ed.; CRC Press: Boca Raton, 2010; 1620. [Google Scholar]
- [7].Aghaie M; Rezaei N; Zendehboudi S A systematic review on CO2 capture with ionic liquids: Current status and future prospects. Renewable Sustainable Energy Rev. 2018, 96, 502–525. [Google Scholar]
- [8].Ramdin M; de Loos TW; Vlugt TJH State-of-the-art of CO2 capture with ionic liquids. Ind. Eng. Chem. Res 2012, 51, 8149–8177. [Google Scholar]
- [9].Camper D; Bara JE; Gin DL; Noble RD Room-temperature ionic liquid−amine solutions: Tunable solvents for efficient and reversible capture of CO2. Ind. Eng. Chem. Res 2008, 47, 8496–8498. [Google Scholar]
- [10].Hu Y-F; Liu Z-C; Xu C-M; Zhang X-M The molecular characteristics dominating the solubility of gases in ionic liquids. Chem. Soc. Rev 2011, 40, 3802–3823. [DOI] [PubMed] [Google Scholar]
- [11].Karadas F; Atilhan M; Aparicio S Review on the use of ionic liquids (ILs) as alternative fluids for CO2 capture and natural gas sweetening. Energy Fuels 2010, 24, 5817–5828. [Google Scholar]
- [12].Zhang J; Sun J; Zhang X; Zhao Y; Zhang S The recent development of CO2 fixation and conversion by ionic liquid. Greenhouse Gas Sci. Technol 2011, 1, 142–159. [Google Scholar]
- [13].Zhang X; Zhang X; Dong H; Zhao Z; Zhang S; Huang Y Carbon capture with ionic liquids: overview and progress. Energy Environ. Sci 2012, 5, 6668–6681. [Google Scholar]
- [14].Shiflett MB; Drew DW; Cantini RA; A. Y Carbon dioxide capture using ionic liquid 1-butyl-3-methylimidazolium acetate. Energy Fuels 2010, 24, 5781–5789. [Google Scholar]
- [15].Sumon KZ; Henni A Ionic liquids for CO2 capture using COSMO-RS: Effect of structure, properties and molecular interactions on solubility and selectivity. Fluid Phase Equilib. 2011, 310, 39–55. [Google Scholar]
- [16].Zhao H; Xia S; Ma P Use of ionic liquids as ‘green’ solvents for extractions. J. Chem. Technol. Biotechnol 2005, 80, 1089–1096. [Google Scholar]
- [17].Li G; Zhou D; Xu Q; Qiao G; Yin J Effect of different co-solvents on solubility of ILs [Emim]Ac, [Bmim][NTf2] in scCO2. J. Mol. Liq 2019, 284, 102–109. [Google Scholar]
- [18].Cui G; Wang J; Zhang S Active chemisorption sites in functionalized ionic liquids for carbon capture. Chem. Soc. Rev 2016, 45, 4307–4339. [DOI] [PubMed] [Google Scholar]
- [19].Anthony JL; Anderson JL; Maginn EJ; Brennecke JF Anion effects on gas solubility in ionic liquids. J. Phys. Chem. B 2005, 109, 6366–6374. [DOI] [PubMed] [Google Scholar]
- [20].Cadena C; Anthony JL; Shah JK; Morrow TI; Brennecke JF; Maginn EJ Why is CO2 so soluble in imidazolium-based ionic liquids? J. Am. Chem. Soc 2004, 126, 5300–5308. [DOI] [PubMed] [Google Scholar]
- [21].Shi W; Maginn EJ Atomistic simulation of the absorption of carbon dioxide and water in the ionic liquid 1-n-hexyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([hmim][Tf2N]. J. Phys. Chem. B 2008, 112, 2045–2055. [DOI] [PubMed] [Google Scholar]
- [22].Neumann JG; Stassen H Anion effect on gas absorption in imidazolium-based ionic liquids. J. Chem. Inf. Model 2020, 60, 661–666. [DOI] [PubMed] [Google Scholar]
- [23].Aki SNVK; Mellein BR; Saurer EM; Brennecke JF High-pressure phase behavior of carbon dioxide with imidazolium-based ionic liquids. J. Phys. Chem. B 2004, 108, 20355–20365. [Google Scholar]
- [24].Blanchard LA; Gu Z; Brennecke JF High-pressure phase behavior of ionic liquid/CO2 systems. J. Phys. Chem. B 2001, 105, 2437–2444. [Google Scholar]
- [25].Lim B-H; Choe W-H; Shim J-J; Ra CS; Tuma D; Lee H; Lee CS High-pressure solubility of carbon dioxide in imidazolium-based ionic liquids with anions [PF6] and [BF4]. Korean J. Chem. Eng 2009, 26, 1130–1136. [Google Scholar]
- [26].Mokhtarani B; Khatun AN; Mafi M; Sharifi A; Mirzaei M Experimental study on the solubility of carbon dioxide in nitrate and thiocyanate-based ionic liquids. J. Chem. Eng. Data 2016, 61, 1262–1269. [Google Scholar]
- [27].Nonthanasin T; Henni A; Saiwan C Solubility of carbon dioxide in five promising ionic liquids. Chem. Eng. Trans 2013, 35, 1417–1422. [Google Scholar]
- [28].Yim J-H; Ha S-J; Lim JS Measurement and correlation of CO2 solubility in 1-butyl-3-methylimidazolium ([BMIM]) cation-based ionic liquids: [BMIM][Ac], [BMIM][Cl], [BMIM][MeSO4]. J. Supercrit. Fluids 2018, 138, 73–81. [Google Scholar]
- [29].Jalili AH; Mehdizadeh A; Shokouhi M; Sakhaeinia H; Taghikhani V Solubility of CO2 in 1-(2-hydroxyethyl)-3-methylimidazolium ionic liquids with different anions. J. Chem. Thermodyn 2010, 42, 787–791. [Google Scholar]
- [30].Shokouhi M; Adibi M; Jalili AH; Hosseini-Jenab M; Mehdizadeh A Solubility and diffusion of H2S and CO2 in the ionic liquid 1-(2-hydroxyethyl)-3-methylimidazolium tetrafluoroborate. J. Chem. Eng. Data 2010, 55, 1663–1668. [Google Scholar]
- [31].Ayad A; Negadi A; Mutelet F Carbon dioxide solubilities in tricyanomethanide-based ionic liquids: Measurements and PC-SAFT modeling. Fluid Phase Equilib. 2018, 469, 48–55. [Google Scholar]
- [32].Yokozeki A; Shiflett MB; Junk CP; Grieco LM; Foo T Physical and chemical absorptions of carbon dioxide in room-temperature ionic liquids. J. Phys. Chem. B 2008, 112, 16654–16663. [DOI] [PubMed] [Google Scholar]
- [33].Zhang X; Liu Z; Wang W Screening of ionic liquids to capture CO2 by COSMO-RS and experiments. AIChE J. 2008, 54, 2717–2728. [Google Scholar]
- [34].Galán Sánchez LM; Meindersma GW; de Haan AB Solvent properties of functionalized ionic liquids for CO2 absorption. Chem. Eng. Res. Design 2007, 85, 31–39. [Google Scholar]
- [35].Raveendran P; Wallen SL Exploring CO2-philicity: Effects of stepwise fluorination. J. Phys. Chem. B 2003, 107, 1473–1477. [Google Scholar]
- [36].Carvalho PJ; Kurnia KA; Coutinho JAP Dispelling some myths about the CO2 solubility in ionic liquids. Phys. Chem. Chem. Phys 2016, 18, 14757–14771. [DOI] [PubMed] [Google Scholar]
- [37].Sarbu T; Styranec TJ; Beckman EJ Design and synthesis of low cost, sustainable CO2-philes. Ind. Eng. Chem. Res 2000, 39, 4678–4683. [Google Scholar]
- [38].Muldoon MJ; Aki SNVK; Anderson JL; Dixon JK; Brennecke JF Improving carbon dioxide solubility in ionic liquids. J. Phys. Chem. B 2007, 111, 9001–9009. [DOI] [PubMed] [Google Scholar]
- [39].Huang J; Rüther T Why are ionic liquids attractive for CO2 absorption? An overview. Aust. J. Chem 2009, 62, 298–308. [Google Scholar]
- [40].Gurau G; Rodríguez H; Kelley SP; Janiczek P; Kalb RS; Rogers RD Demonstration of chemisorption of carbon dioxide in 1,3-dialkylimidazolium acetate ionic liquids. Angew. Chem. Int. Ed 2011, 50, 12024–12026. [DOI] [PubMed] [Google Scholar]
- [41].Shiflett MB; Kasprazak DJ; Junk CP; Yokozeki A Phase behavior of {carbon dioxide + [bmim][Ac]} mixtures. J. Chem. Thermodyn 2008, 40, 25–31. [Google Scholar]
- [42].Carvalho PJ; Álvarez VH; Schröder B; Gil AM; Marrucho IM; Aznar M; Santos LMNBF; Coutinho JAP Specific solvation interactions of CO2 on acetate and trifluoroacetate imidazolium based ionic liquids at high pressures. J. Phys. Chem. B 2009, 113, 6803–6812. [DOI] [PubMed] [Google Scholar]
- [43].Mattedi S; Carvalho PJ; Coutinho JAP; Alvarez VH; Iglesias M High pressure CO2 solubility in N-methyl-2-hydroxyethylammonium protic ionic liquids. J. Supercrit. Fluids 2011, 56, 224–230. [Google Scholar]
- [44].Beckman EJ A challenge for green chemistry: designing molecules that readily dissolve in carbon dioxide. Chem. Commun 2004, 1885–1888. [DOI] [PubMed] [Google Scholar]
- [45].Raveendran P; Wallen SL Cooperative C−H···O hydrogen bonding in CO2-Lewis base complexes: Implications for solvation in supercritical CO2. J. Am. Chem. Soc 2002, 124, 12590–12599. [DOI] [PubMed] [Google Scholar]
- [46].Yeadon DJ; Jacquemin J; Plechkova NV; Maréchal M; Seddon KR Induced protic behaviour in aprotonic ionic liquids by anion basicity for efficient carbon dioxide capture. ChemPhysChem 2020, 21, 1369–1374. [DOI] [PubMed] [Google Scholar]
- [47].Lee TB; Oh S; Gohndrone TR; Morales-Collazo O; Seo S; Brennecke JF; Schneider WF CO2 chemistry of phenolate-based ionic liquids. J. Phys. Chem. B 2016, 120, 1509–1517. [DOI] [PubMed] [Google Scholar]
- [48].Shiflett MB; Yokozeki A Phase behavior of carbon dioxide in ionic liquids: [emim][Acetate], [emim][Trifluoroacetate], and [emim][Acetate] + [emim][Trifluoroacetate] mixtures. J. Chem. Eng. Data 2009, 54, 108–114. [Google Scholar]
- [49].Shaahmadi F; Hashemi Shahraki B; Farhadi A The solubility of carbon dioxide and density for binary mixtures of 1-butyl-3-methylimidazolium acetate and 1-butyl-3-methylimidazolium tetrafluoroborate. J. Chem. Eng. Data 2019, 64, 584–593. [Google Scholar]
- [50].Isabel Cabaço M; Besnard M; Danten Y; Coutinho JAP Solubility of CO2 in 1-butyl-3-methyl-imidazolium-trifluoro acetate ionic liquid studied by Raman spectroscopy and DFT investigations. J. Phys. Chem. B 2011, 115, 3538–3550. [DOI] [PubMed] [Google Scholar]
- [51].Zanatta M; Simon NM; Dupont J The nature of carbon dioxide in bare ionic liquids. ChemSusChem 2020, 13, 3101–3109. [DOI] [PubMed] [Google Scholar]
- [52].Yunus NM; Halim NH; Wilfred CD; Murugesan T; Lim JW; Show PL Thermophysical properties and CO2 absorption of ammonium-based protic ionic liquids containing acetate and butyrate anions. Processes 2019, 7, 820. [Google Scholar]
- [53].Safarov J; Guluzade A; Hassel E Carbon dioxide solubility in 1-ethyl-3-methylimidazolium trifluoromethanesulfonate or 1-butyl-3-methylimidazolium trifluoromethanesulfonate ionic liquids. J. Chem. Eng. Data 2020, 65, 1060–1067. [Google Scholar]
- [54].Yim J-H; Park KW; Oh B-K; Lim JS CO2 solubility in 1,1,2,2-tetrafluoroethanesulfonate anion-based ionic liquids: [EMIM][TFES], [BMIM][TFES], and [BNMIM][TFES]. J. Chem. Eng. Data 2020, 65, 617–627. [Google Scholar]
- [55].Manic MS; Queimada AJ; Macedo EA; Najdanovic-Visak V High-pressure solubilities of carbon dioxide in ionic liquids based on bis(trifluoromethylsulfonyl)imide and chloride. J. Supercrit. Fluids 2012, 65, 1–10. [Google Scholar]
- [56].Kim YS; Choi WY; Jang JH; Yoo K-P; Lee CS Solubility measurement and prediction of carbon dioxide in ionic liquids. Fluid Phase Equilib. 2005, 228–229, 439–445. [Google Scholar]
- [57].Yim J-H; Song HN; Lee B-C; Lim JS High-pressure phase behavior of binary mixtures containing ionic liquid [HMP][Tf2N], [OMP][Tf2N] and carbon dioxide. Fluid Phase Equilib. 2011, 308, 147–152. [Google Scholar]
- [58].Makino T; Kanakubo M CO2 absorption property of ionic liquid and CO2 permselectivity for ionic liquid membrane. J. Jpn. Pet. Inst 2016, 59, 109–117. [Google Scholar]
- [59].van Ginderen P; Herrebout WA; van der Veken BJ van der Waals complex of dimethyl ether with carbon dioxide. J. Phys. Chem. A 2003, 107, 5391–5396. [Google Scholar]
- [60].Sharma P; Choi S-H; Park S-D; Baek I-H; Lee G-S Selective chemical separation of carbondioxide by ether functionalized imidazolium cation based ionic liquids. Chem. Eng. J 2012, 181–182, 834–841. [Google Scholar]
- [61].Seo S; Quiroz-Guzman M; Gohndrone TR; Brennecke JF Comment on “Selective chemical separation of carbon dioxide by ether functionalized imidazolium cation based ionic liquids”, by Pankaj Sharma, Soo-Hyun Choi, Sang-Do Park, Il-Hyun Baek, and Gil-Sun Lee (Chem. Eng. J., 2012, 181–182, p. 834–841). Chem. Eng. J 2014, 245, 367–369. [Google Scholar]
- [62].Nath D; Henni A Solubility of carbon dioxide (CO2) in four bis (trifluoromethyl-sulfonyl)imide ([Tf2N]) based ionic liquids. Fluid Phase Equilib. 2020, 524, 112757. [Google Scholar]
- [63].Bara JE; Gabriel CJ; Lessmann S; Carlisle TK; Finotello A; Gin DL; Noble RD Enhanced CO2 separation selectivity in oligo(ethylene glycol) functionalized room-temperature ionic liquids. Ind. Eng. Chem. Res 2007, 46, 5380–5386. [Google Scholar]
- [64].Deng Y; Morrissey S; Gathergood N; Delort A-M; Husson P; Gomes MFC The presence of functional groups key for biodegradation in ionic liquids: Effect on gas solubility. ChemSusChem 2010, 3, 377–385. [DOI] [PubMed] [Google Scholar]
- [65].Deng Y; Husson P; Delort A-M; Besse-Hoggan P; Sancelme M; Gomes MFC Influence of an oxygen functionalization on the physicochemical properties of ionic liquids: Density, viscosity, and carbon dioxide solubility as a function of temperature. J. Chem. Eng. Data 2011, 56, 4194–4202. [Google Scholar]
- [66].Makino T; Kanakubo M; Umecky T CO2 solubilities in ammonium bis(trifluoromethanesulfonyl)amide ionic liquids: Effects of ester and ether groups. J. Chem. Eng. Data 2014, 59, 1435–1440. [Google Scholar]
- [67].Kanakubo M; Makino T; Taniguchi T; Nokami T; Itoh T CO2 solubility in ether functionalized ionic liquids on mole fraction and molarity scales. ACS Sustainable Chem. Eng 2016, 4, 525–535. [Google Scholar]
- [68].Bhawna AP; Dhingra D; Pandey S Can common liquid polymers and surfactants capture CO2? J. Mol. Liq 2019, 277, 594–605. [Google Scholar]
- [69].Wang C; Mahurin SM; Luo H; Baker GA; Li H; Dai S Reversible and robust CO2 capture by equimolar task-specific ionic liquid–superbase mixtures. Green Chem. 2010, 12, 870–874. [Google Scholar]
- [70].Baltus RE; Culbertson BH; Dai S; Luo H; DePaoli DW Low-pressure solubility of carbon dioxide in room-temperature ionic liquids measured with a quartz crystal microbalance. J. Phys. Chem. B 2004, 108, 721–727. [Google Scholar]
- [71].Perry RJ; Grocela-Rocha TA; O’Brien MJ; Genovese S; Wood BR; Lewis LN; Lam H; Soloveichik G; Rubinsztajn M; Kniajanski S; Draper S; Enick RM; Johnson JK; Xie H; Tapriyal D Aminosilicone solvents for CO2 capture. ChemSusChem 2010, 3, 919–930. [DOI] [PubMed] [Google Scholar]
- [72].McComb RB; Bond LW; Burnett RW; Keech RC; Bowers GN Determination of the molar absorptivity of NADH. Clin. Chem 1976, 22, 141–150. [PubMed] [Google Scholar]
- [73].Li J; Ye Y; Chen L; Qi Z Solubilities of CO2 in poly(ethylene glycols) from (303.15 to 333.15) K. J. Chem. Eng.Data 2012, 57, 610–616. [Google Scholar]
- [74].Trivedi S; Pandey S Interactions within a [ionic liquid + poly(ethylene glycol)] mixture revealed by temperature-dependent synergistic dynamic viscosity and probe-reported microviscosity. J. Phys. Chem. B 2011, 115, 7405–7416. [DOI] [PubMed] [Google Scholar]
- [75].Hou M; Liang S; Zhang Z; Song J; Jiang T; Han B Determination and modeling of solubility of CO2 in PEG200 + 1-pentanol and PEG200 + 1-octanol mixtures. Fluid Phase Equilib. 2007, 258, 108–114. [Google Scholar]
- [76].Li X; Hou M; Zhang Z; Han B; Yang G; Wang X; Zou L Absorption of CO2 by ionic liquid/polyethylene glycol mixture and the thermodynamic parameters. Green Chem. 2008, 10, 879–884. [Google Scholar]
- [77].Cong H; Yu B Aminosilane cross-linked PEG/PEPEG/PPEPG membranes for CO2/N2 and CO2/H2 separation. Ind. Eng. Chem. Res 2010, 49, 9363–9369. [Google Scholar]
- [78].Amooghin AE; Sanaeepur H; Moghadassi A; Kargari A; Ghanbari D; Mehrabadi ZS Modification of ABS membrane by PEG for capturing carbon dioxide from CO2/N2 streams. Sep. Sci. Technol 2010, 45, 1385–1394. [Google Scholar]
- [79].Archane A; Fürst W; Provost E Influence of poly(ethylene oxide) 400 (PEG400) on the absorption of CO2 in diethanolamine (DEA)/H2O systems. J. Chem. Eng. Data 2011, 56, 1852–1856. [Google Scholar]
- [80].Zhao H; Harter GA; Martin CJ “Water-like” dual-functionalized ionic liquids for enzyme activation. ACS Omega 2019, 4, 15234–15239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [81].Franklin MS; Larm NE; Baker GA; Zhao H Functionalized ionic liquids for lignite dissolution and treatment. J. Chem. Technol. Biotechnol 2021, 96, 3273–3281. [Google Scholar]
- [82].Zhao H; Afriyie LO; Larm NE; Baker GA Glycol-functionalized ionic liquids for hight-emperature enzymatic ring-opening polymerization. RSC Adv. 2018, 8, 36025–36033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [83].Zhao H; Martin CJ; Larm NE; Baker GA; Trujillo TC Enzyme activation by water-mimicking dual-functionalized ionic liquids. Mol. Catal 2021, 515, 111882. [Google Scholar]
- [84].Wender PA; Badham NF; Conway SP; Floreancig PE; Glass TE; Houze JB; Krauss NE; Lee D; Marquess DG; McGrane PL; Meng W; Natchus MG; Shuker AJ; Sutton JC; Taylor RE The pinene path to taxanes. 6. A concise stereocontrolled synthesis of Taxol. J. Am. Chem. Soc 1997, 119, 2757–2758. [Google Scholar]
- [85].Tang S; Baker GA; Ravula S; Jones JE; Zhao H PEG-functionalized ionic liquids for cellulose dissolution and saccharification. Green Chem. 2012, 14, 2922–2932. [Google Scholar]
- [86].Siqueira LJA; Ribeiro MCC Alkoxy chain effect on the viscosity of a quaternary ammonium ionic liquid: Molecular dynamics simulations. J. Phys. Chem. B 2009, 113, 1074–1079. [DOI] [PubMed] [Google Scholar]
- [87].Smith GD; Borodin O; Li L; Kim H; Liu Q; Bara JE; Gin DL; Nobel R A comparison of ether- and alkyl-derivatized imidazolium-based room-temperature ionic liquids: a molecular dynamics simulation study. Phys. Chem. Chem. Phys 2008, 10, 6301–6312. [DOI] [PubMed] [Google Scholar]
- [88].Ganapatibhotla LVNR; Zheng J; Roy D; Krishnan S PEGylated imidazolium ionic liquid electrolytes: Thermophysical and electrochemical properties. Chem. Mater 2010, 22, 6347–6360. [Google Scholar]
- [89].Lv B; Guo B; Zhou Z; Jing G Mechanisms of CO2 capture into monoethanolamine solution with different CO2 loading during the absorption/desorption processes. Environ. Sci. Technol 2015, 49, 10728–10735. [DOI] [PubMed] [Google Scholar]
- [90].Bates ED; Mayton RD; Ntai I; Davis JH Jr. CO2 capture by a task-specific ionic liquid. J. Am. Chem. Soc 2002, 124, 926–927. [DOI] [PubMed] [Google Scholar]
- [91].Zhang J; Zhang S; Dong K; Zhang Y; Shen Y; Lv X Supported absorption of CO2 by tetrabutylphosphonium amino acid ionic liquids. Chem. Eur. J 2006, 12, 4021–4026. [DOI] [PubMed] [Google Scholar]
- [92].Gurkan BE; de la Fuente JC; Mindrup EM; Ficke LE; Goodrich BF; Price EA; Schneider WF; Brennecke JF Equimolar CO2 absorption by anion-functionalized ionic liquids. J. Am. Chem. Soc 2010, 132, 2116–2117. [DOI] [PubMed] [Google Scholar]
- [93].Shah VM; Hardy BJ; Stern SA Solubility of carbon dioxide, methane, and propane in silicone polymers. Effect of polymer backbone chains. J. Polym. Sci. Part B: Polym. Phys 1993, 31, 313–317. [Google Scholar]
- [94].Qiang W; Zhao L; Gao X; Liu T; Liu Z; Yuan W-K; Hua D Dual role of PDMS on improving supercritical CO2 foaming of polypropylene: CO2-philic additive and crystallization nucleating agent. J. Supercrit. Fluids 2020, 163, 104888. [Google Scholar]
- [95].Lee JJ; Cummings SD; Beckman EJ; Enick RM; Burgess WA; Doherty MD; O’Brien MJ; Perry RJ The solubility of low molecular weight poly(dimethyl siloxane) indense CO2 and its use as a CO2-philic segment. J. Supercrit. Fluids 2017, 119, 17–25. [Google Scholar]
- [96].Perry RJ, Aminosilicone systems for post-combustion CO2 capture, In Absorption-Based Post-combustion Capture of Carbon Dioxide, 1st ed.; Feron PHM, Woodhead Publishing: 2016; 121–144. [Google Scholar]
- [97].Westendorf T; Farnum R; Rubinsztajn G; Wood B; Perry R; McDermott J; Enick R; Tapriyal D Measurement of CO2 diffusivity in phase-changing aminosilicone CO2 capture solvent. Energy Fuels 2018, 32, 6901–6909. [Google Scholar]
- [98].Freire MG; Teles ARR; Rocha MAA; Schroder B; Neves CMSS; Carvalho PJ; Evtuguin DV; Santos LMNBF; Coutinho JAP Thermophysical characterization of ionic liquids able to dissolve biomass. J. Chem. Eng. Data 2011, 56, 4813–4822. [Google Scholar]
- [99].Hiraga Y; Kato A; Sato Y; Smith R Densities at pressures up to 200 MPa and atmospheric pressure viscosities of ionic liquids 1-ethyl-3-methylimidazolium methylphosphate, 1-ethyl-3-methylimidazolium diethylphosphate, 1-butyl-3-methylimidazolium acetate, and 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide. J. Chem. Eng. Data 2015, 60, 876–885. [Google Scholar]
- [100].Pinchas S; Ben-Ishai D The carbonyl absorption of carbamates and 2-oxazolidones in the infrared region. J. Am. Chem. Soc 1957, 79, 4099–4104. [Google Scholar]
- [101].Khanna RK; Moore MH Carbamic acid: molecular structure and IR spectra. Spectrochim. Acta, Part A 1999, 55, 961–967. [DOI] [PubMed] [Google Scholar]
- [102].Shi Z; Sanguinito S; Goodman A; Jessen K; Tsotsis TT Investigation of mass transfer and sorption in CO2/brine/rock systems via in situ FT-IR. Ind. Eng. Chem. Res 2020, 59, 20181–20189. [Google Scholar]
- [103].Goodman AL; Campus LM; Schroeder KT Direct evidence of carbon dioxide sorption on Argonne premium coals using Attenuated Total Reflectance-Fourier Transform infrared spectroscopy. Energy Fuels 2005, 19, 471–476. [Google Scholar]
- [104].Falk M; Miller AG Infrared spectrum of carbon dioxide in aqueous solution. Vib. Spectrosc 1992, 4, 105–108. [Google Scholar]
- [105].Anderson JL; Dixon JaNeille K.; Maginn EJ; Brennecke JF Measurement of SO2 solubility in ionic liquids. J. Phys. Chem. B 2006, 110, 15059–15062. [DOI] [PubMed] [Google Scholar]
- [106].Li X; Zhang L; Zheng Y; Zheng C Effect of SO2 on CO2 absorption in flue gas by ionic liquid 1-ethyl-3-methylimidazolium acetate. Ind. Eng. Chem. Res 2015, 54, 8569–8578. [Google Scholar]
- [107].Wang L; Zhang Y; Liu Y; Xie H; Xu Y; Wei J SO2 absorption in pure ionic liquids: Solubility and functionalization. J. Hazard. Mater 2020, 392, 122504. [DOI] [PubMed] [Google Scholar]
- [108].Obert R; Dave BC Enzymatic conversion of carbon dioxide to methanol: Enhanced methanol production in silica sol-gel matrices. J. Am. Chem. Soc 1999. 121, 12192–12193. [Google Scholar]
- [109].Rusching U; Müller U; Willnow P; Höpner T CO2 reduction to formate by NADH catalysed by formate dehydrogenase from Pseudomonas oxalaticus. Eur. J. Biochem 1976, 70, 325–330. [DOI] [PubMed] [Google Scholar]
- [110].Zhang Z; Muschiol J; Huang Y; Sigurdardóttir SB; von Solms N; Daugaard AE; Wei J; Luo J; Xu B; Zhang S; Pinelo M Efficient ionic liquid-based platform for multi-enzymatic conversion of carbon dioxide to methanol. Green Chem. 2018, 20, 4339–4348. [Google Scholar]
- [111].Zhao H Protein stabilization and enzyme activation in ionic liquids: specific ion effects. J. Chem. Technol. Biotechnol 2016, 91, 25–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Zhao H Methods for stabilizing and activating enzymes in ionic liquids — A review. J. Chem. Tech. Biotechnol 2010, 85, 891–907. [Google Scholar]
- [113].Eckstein M; Filho MV; Liese A; Kragl U Use of an ionic liquid in a two-phase system to improve an alcohol dehydrogenase catalysed reduction. Chem. Commun 2004, 1084–1085. [DOI] [PubMed] [Google Scholar]
- [114].de Gonzalo G; Lavandera I; Durchschein K; Wurm D; Faber K; Kroutil W Asymmetric biocatalytic reduction of ketones using hydroxy-functionalised water-miscible ionic liquids as solvents. Tetrahedron: Asymmetry 2007, 18, 2541–2546. [Google Scholar]
- [115].Weibels S; Syguda A; Herrmann C; Weingärtner H Steering the enzymatic activity of proteins by ionic liquids. A case study of the enzyme kinetics of yeast alcohol dehydrogenase. Phys. Chem. Chem. Phys 2012, 14, 4635–4639. [DOI] [PubMed] [Google Scholar]
- [116].Bekhouche M; Blum LJ; Doumèche B Contribution of dynamic and static quenchers for the study of protein conformation in ionic liquids by steady-state fluorescence spectroscopy. J. Phys. Chem. B 2012, 116, 413–423. [DOI] [PubMed] [Google Scholar]
- [117].D’Oronzo E; Secundo F; Minofar B; Kulik N; Pometun AA; Tishkov VI Activation/inactivation role of ionic liquids on formate dehydrogenase from Pseudomonas sp. 101 and its mutated thermostable form. ChemCatChem 2018, 10, 3247–3259. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
