Next Article in Journal
The Association of Grain Yield and Agronomical Traits with Genes of Plant Height, Photoperiod Sensitivity and Plastid Glutamine Synthetase in Winter Bread Wheat (Triticum aestivum L.) Collection
Next Article in Special Issue
Adjustment of the Structure of the Simplest Amino Acid Present in Nature—Glycine, toward More Environmentally Friendly Ionic Forms of Phenoxypropionate-Based Herbicides
Previous Article in Journal
SP1 and NFY Regulate the Expression of PNPT1, a Gene Encoding a Mitochondrial Protein Involved in Cancer
Previous Article in Special Issue
Carbon-Based Electrocatalyst Design with Phytic Acid—A Versatile Biomass-Derived Modifier of Functional Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tuning Functionalized Ionic Liquids for CO2 Capture

College of Chemical Engineering, Zhejiang University of Technology, Huzhou 313299, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(19), 11401; https://doi.org/10.3390/ijms231911401
Submission received: 27 August 2022 / Revised: 8 September 2022 / Accepted: 16 September 2022 / Published: 27 September 2022
(This article belongs to the Special Issue Advanced Research in Green Chemistry)

Abstract

:
The increasing concentration of CO2 in the atmosphere is related to global climate change. Carbon capture, utilization, and storage (CCUS) is an important technology to reduce CO2 emissions and to deal with global climate change. The development of new materials and technologies for efficient CO2 capture has received increasing attention among global researchers. Ionic liquids (ILs), especially functionalized ILs, with such unique properties as almost no vapor pressure, thermal- and chemical-stability, non-flammability, and tunable properties, have been used in CCUS with great interest. This paper focuses on the development of functionalized ILs for CO2 capture in the past decade (2012~2022). Functionalized ILs, or task-specific ILs, are ILs with active sites on cations or/and anions. The main contents include three parts: cation-functionalized ILs, anion-functionalized ILs, and cation-anion dual-functionalized ILs for CO2 capture. In addition, classification, structures, and synthesis of functionalized ILs are also summarized. Finally, future directions, concerns, and prospects for functionalized ILs in CCUS are discussed. This review is beneficial for researchers to obtain an overall understanding of CO2-philic ILs. This work will open a door to develop novel IL-based solvents and materials for the capture and separation of other gases, such as SO2, H2S, NOx, NH3, and so on.

1. Introduction

According to the report “State of the Global Climate 2021” recently published by the World Meteorological Organization (WMO), CO2 mole fraction reached new high (413.2 ± 0.2 ppm in 2020, while pre-industrial mole fraction of 278 ppm [1]. The increasing concentration of CO2 in the atmosphere during these centuries, especially since the 20th century, leads to the greenhouse effect and global climate change. A large amount of all human-produced CO2 emissions come from the burning of fossil fuels, such as coal, natural gas, and oil, including gasoline.
In recent decades, carbon capture, utilization, and storage (CCUS) has become one of the important technologies to reduce CO2 emissions [2]. For carbon capture, the common CCUS technologies are based on chemical sorption, physical sorption, membrane separation, calcium looping, etc. For example, aqueous monoethanolamine (30 wt%) process is the current CO2 capture technology in industry via carbamate mechanism. Although the chemical reaction methods are more efficiency, the regeneration energy consumption of these methods is high [3]. For carbon utilization, the most effective strategy is CO2 convention, including thermocatalysis, photocatalysis, or electrocatalysis, of CO2 cycloaddition reaction, CO2 reduction reaction (CO2RR), etc. [4,5]. For example, the final products of the CO2RR are widely distributed from C1 (carbon monoxide, formic acid, methane) to C2+ (ethylene, ethanol, acetone, etc.) [6]. However, the CO2 conversion via CO2RR approach is still steps away from widespread commercialization. For carbon storage, the widely used way to store captured CO2 is in deep geological formations, such as oil fields, gas fields, coal seams, and saline aquifers [7]. However, the process increases the amount of energy required by power plants. Therefore, alternative CCUS technologies with high efficiency and low energy-consumption are highly desired.
Ionic liquids (ILs) are composed of organic cations and organic or inorganic anions with melting points below 100 °C [8,9,10,11]. Their excellent properties, including extremely low vapor pressure, high thermal and chemical stability, wide liquid temperature range, high electrical conductivity and wide electrochemical window, and good solubility for both polar and non-polar compounds, make it possible for ILs to be designed according to needs. Thus, ILs are widely used as green solvents and catalysts in such fields as energy and environment [12,13,14], chemistry and chemical synthesis [15,16], sorption and separation [17], and pharmaceutics and medicine [18,19]. Compared with conventional methods, ILs, especially functionalized ILs, have been used in CCUS with great interest all over the world due to the advantages of fast absorption, high capacity, low energy-consumption, good stability, and good recyclability.
Several interesting reviews for CO2 capture by ILs have been published during the last few years. For example, Zhang et al. [20] reviewed the IL-based CO2 capture systems from structure and interaction to process. Zhang and Ji et al. [21] reported the reviewing and evaluating of ionic liquids/deep eutectic solvents for CO2 capture. However, it is crucial to review this developing field from a viewpoint of functionalization of ILs with active sites, which is beneficial for researchers to obtain an overall understanding of CO2-philic ILs and grasp the development direction.
In this critical review, we mainly focus on the development of functionalized ILs for CO2 capture in the past 10 years (2012~2022). The main contents include three parts, cation-functionalized ILs, anion-functionalized ILs, and cation-anion dual-functionalized ILs for CO2 capture (Figure 1). Besides, classification, structures, and synthesis of functionalized ILs are also summarized. Finally, future directions, concerns, and prospects for functionalized ILs in CCUS are discussed.

2. Classification, Structures, and Synthesis of Functionalized ILs

2.1. Classification and Structures of Functionalized ILs

Functionalized ILs (task-specific ILs, or functional ILs) can be simply classified into three categories according to the locations of active sites, including cation-functionalized ILs, anion-functionalized ILs, and cation-anion dual-functionalized ILs. The cation-functionalized ILs and anion-functionalized ILs can be divided into two categories according to the number of functional groups and the mechanism of CO2 capture, including single-site functionalized ILs and multiple-site functionalized ILs. It is clear that cation-anion dual-functionalized ILs are multiple-site functionalized ILs. The main reaction groups with active sites are listed in each category, such as amino, carboxylate, alkoxide, phenolate, and azolate. The structures of cations and anions for synthesis of functionalized ILs are collected in Figure 2.

2.2. Synthesis of Functionalized ILs

Generally, the synthesis of functionalized ILs includes several separated unit operations, such as quaternization, anion-exchange, acid-base neutralization, coordination, etc. According to the structure of functionalized ILs, the methods or pathway for the synthesis can be typically classified into two categories: direct methods and indirect methods. Direct methods include one of above-mentioned operations, while indirect methods include two or more of above-mentioned operations. The typical strategies for synthesis of functionalized ILs for CO2 capture can be found in Figure 3.

3. Functionalized ILs for CO2 Capture

3.1. Cation-Functionalized ILs for CO2 Capture

3.1.1. Single-Site Mechanisms

It is known that the most studied cation-functionalized ILs for CO2 capture should be amino-functionalized ILs, which were first reported by Davis et al. [22] in 2002, two decades ago. They showed that 0.5 mole of CO2 per mole of IL was captured by 1-propylamide-3-butyl imidazolium tetrafluoroborate ([apbim][BF4]) via a carbamate mechanism (2 amino: 1 CO2). Compared with conventional alkanolamine aqueous solution (30 wt% monoethanolamine) for CO2 capture, amino grafted on cations of ILs showed high thermostability [23,24], while amino grafted on ILs showed high capture capacity compared with conventional ILs [25]. Subsequently, a number of amino-grafted cation-functionalized ILs were reported for efficient CO2 capture [26,27,28,29,30]. The mechanisms of amino–CO2 reaction in ILs are similar to those in aqueous alkanolamine solutions. Compared with primary and secondary amines, tertiary amine is considered unreactive with CO2 under anhydrous conditions (Figure 4). However, He et al. [31] reported that tertiary amino-containing Li-chelated cation-functionalized ILs, [PEG150MeBu2NLi][Tf2N] and [PEG150MeTMGLi][Tf2N], could achieve high CO2 capacities, 0.66 and 0.89 mol CO2 per mol IL, respectively, via coordination with lithium ion.

3.1.2. Multiple-Site Mechanisms

Multiple functional sites on the cations are multiple amino groups. For example, Zhang et al. [32] reported CO2 capture by a dual amino-containing cation-functionalized IL, 1, 3-di (2′-aminoethyl)-2-methylimidazolium bromide (DAIL), via a 2:1 carbamate mechanism (amino: CO2). However, the synthesis of DAIL was not easy. Therefore, other kinds of polyamine-based ILs were developed through acid-base neutralization or metal coordination. Clyburne et al. [33] and Meng et al. [34] studied CO2 capture by [DETA][NO3] and [TETA][NO3] ammonium ILs, which were prepared through acid-base neutralization of diethylenetriamine (DETA) or triethylenetetramine (TETA) with nitric acid. On the other hand, Wang and Dai et al. [35] investigated the CO2 capture by a series of chelate ILs with multiple Li-coordinated amino groups on the cations, and up to 0.88 and 0.90 mol CO2 per mol IL could be captured by [Li(HDA)][Tf2N] and [Li(DOBA)][Tf2N] at 40 °C and 1 bar, respectively, via a 2:1 mechanism. Subsequently, Wang et al. [36] reported several polyamine-based ILs ([Li(TETA)][Tf2N], [Li(DETA)][Tf2N] and [Li(TEPA)][Tf2N]) and polyalcohol-based ILs ([Li(TEG)][Tf2N] and [Li(TTEG)][Tf2N]). The former could chemically absorb CO2, while the later could only physically absorb CO2. Their results showed that CO2 capacity of polyamine-based ILs increased when [Li(TTEG)][Tf2N] or [Li(TEG)][Tf2N] was added, and CO2 capacity of [Li(TEPA)][Tf2N]/[Li(TEG)][Tf2N] (weight ratio is 1:2) decreased from 2.05 to 0.83 mol per CO2 mol IL at 80 °C when CO2 concentration was reduced from 100 vol.% to 380 ppm. Recently, Yang, Xing, and Dai et al. [37] reported the tuning of stability constants of metal-amine complexes for efficient CO2 desorption.
The comparison of the absorption capacities, including molar capacities and corresponding mass capacities, of typical cation-functionalized ILs for CO2 capture are listed in Table 1.

3.2. Anion-Functionalized ILs for CO2 Capture

Compared with only amino-grafted cations for efficient CO2 capture by ILs, there are numerous kinds of functional groups grafted on anions for efficient CO2 capture. According to their reaction mechanism with CO2, the anion-functionalized ILs can be classified into two categories, including single-site mechanisms and multiple-site mechanisms.

3.2.1. Single-Site Mechanisms

Functionalized ILs with single-site on anions include amino anions, carboxylate anions, alkoxide anion, phenolate anions, and azolate anions. The typical mechanisms for the reaction of non-amino anion-CO2 can be found in Figure 5a.
(1)
Amino anion functionalized ILs
Typical amino anions are amino acid anions that are prepared via dehydrogenation (or acid-base neutralization). For example, several amino acid ILs (AAILs) ([P4444][Gly], [P4444][Ala], [P4444][β-Ala], [P4444][Ser], and [P4444][Lys]) with high viscosity were first reported by Zhang et al. [38] for CO2 capture through supporting on SiO2, and the absorption following a 2:1 carbamate pathway. However, [P66614][Met] and [P66614][Pro] with large phosphonium cations were reported by Brennecke et al. [39,40] for equimolar absorption of CO2 via a 1:1 mechanism. In order to understand the CO2 absorption mechanisms with AAILs, Xing et al. [41] showed that the actual mechanism went beyond the apparent stoichiometry. Take [Gly] and [Met] anions as the examples, although the apparent chemical stoichiometry approached 1:1 and the absorption was previously considered to simply follow the 1:1 mechanism, their results indicated that more than 20 % of the CO2 still was absorbed in the 1:2 reaction mechanism. Recently, Mehrdad et al. [42] investigated three AAILs ([BMIm][Gly], [BMIm][Ala], and [BMIm][Val]) for CO2 capture via physical and chemical sorption mechanism. However, the hydrogen bond in these AAILs resulted in high viscosity, and the viscosity increased dramatically after the absorption of CO2. Therefore, other AAILs supported on porous materials [43,44,45,46] or mixed with liquids [33,34] were reported;
(2)
Carboxylate anion-functionalized ILs (O-site)
From the 1:1 mechanism of AA ILs with CO2, the carboxylate in the AA anions provides a negative charge but seemed to not interact with CO2. Through tuning the structure of carboxylate ILs, Ils can also chemically react with CO2. 1-Butyl-3-methylimidazolium acetate ([Bmim][Ac]), reported by Maginn et al. [47], was the first carboxylate IL example for efficient CO2 capture. The reported mechanism of N-heterocyclic carbene−CO2 was verified by NMR. However, Steckel et al. [48], Shi et al. [49], and Ruiz-López et al. [50] studied the mechanism via ab initio calculations. The results indicated that for glycinate anion, interactions with the amino and carboxylic moieties involved comparable energetics. For example, Tao et al. [51] studied a series of phosphonium carboxylate ILs for CO2 capture, and butyrate IL could absorb 0.4 mol CO2 per mol IL. Yunus et al. [52] investigated ammonium carboxylate ILs with different organic acid anions for CO2 capture at high pressures and obtained the high capacities of ILs with heptanoate anions. Cheng et al. [53] correlated the data of CO2 solubility in carboxylate-based N-ethylmorpholinium ILs ([NEMH][Ac], [NEMH][Propionate], and [TEAH][Propionate]) with Pitzer’s model and the Soave–Redlich–Kwong cubic equation of state. Similarly, Umecky et al. [54] showed that acetylacetonate ILs could also chemically absorb CO2;
(3)
Alkoxide anion-functionalized ILs (O-site)
The alkoxide is an anion that forms when we remove the hydrogen atom from the –OH group of an alcohol. It was known that switchable solvents, a liquid mixture of an alcohol (e.g., pKa of ethanol in DMSO is 29.8) and a strong organic base (e.g., 1,8-diazabicyclo-[5.4.0]undec-7-ene, DBU), could chemically bind CO2 to form an alkylcarbonate salt through proton transfer from alcohol to superbase [55,56]. Thus, alcohols with the appropriate acidity can be used to synthesize alkoxide ILs through dehydrogenation. Dai et al. [57] reported a series of superbase-derived protic ILs with trifluoroethanolate (TFE, pKa = 23.5), 1-phenyl-2,2,2-trifluoroethanolate (TFPA, pKa = 23), and 2,2,3,3,4,4-hexafluoro-1,5-pentanediolate (HFPD, pKa = 23.2) anions for equimolar CO2 capture. Subsequently, Liu et al. [58] used [DBUH][TFE] (1.01 mol CO2 per mol IL) to catalyze CO2 conversion into quinazoline-2,4(1H,3H)-diones;
(4)
Phenolate anion-functionalized ILs (O-site)
With appropriate acidity (pKa = 10 in water), the phenol could be used to form the functional anion, phenolate (or phenoxide), via removal of the hydrogen atom from the –OH group of a phenol to prepare ILs for efficient CO2 capture. For example, Wang et al. [59] studied a series of phenolate anion-functionalized ILs for CO2 chemisorption. Through tuning the structure of phenolate anion with different substituents, CO2 absorption performance could be further regulated. For example, the absorption capacities of [P66614][4-Me-PhO], [P66614][4-H-PhO], [P66614][4-Cl-PhO], [P66614][4-CF3-PhO], [P66614][4-NO2-PhO], and [P66614][2,4,6-Cl-PhO] were found to be 0.91, 0.85, 0.82, 0.61, 0.30, and 0.07 mol CO2 per mol IL, respectively. The same authors also found carbonyl-substituted phenolate ILs, [P66614][4-Kt-PhO], [P66614][4-EF-PhO], and [P66614][4-CHO-PhO] could achieve 1.04, 1.03, and 1.01 mol CO2 per mol IL at 20 °C and 1 bar, respectively [60]. Additionally, they also synthesized a series of conjugated phenolate ILs and investigated their CO2 absorption performance [61]. The results showed that the molar ratios of CO2 to [P66614][PPhO] and [P66614][PCCPhO] were 0.93 and 0.96, respectively. Wu and Hu et al. [62,63] investigated that fluorinated phenolate ILs with different positions resulted in low viscosity and tunable capacity ([4-F-PhO] > [3-F-PhO] > [2-F-PhO]). Recently, Yang, Xing, and Dai et al. [64] constructed several phenolate chelate ILs for CO2 chemisorption via coordination of phenolate alkali metal salts with crown ethers. With 15-crown-5-coordinated Na+ as the cation, the CO2 uptake capacity of the phenolate anion decreased in the following order: [PhO] (0.75 mol mol−1) > [n-C3H7PhO] (0.66 mol mol−1) > [n-C8H17PhO] (0.50 mol mol−1);
(5)
Azolate anion-functionalized ILs (N-site)
Azoles, such as imidazole (Im), pyrazole (Pyrz), 1,3,4-trizole (Triz), tetrazole (Tetz), etc., are a kind of five-membered heterocycles. These azoles with high basicity were used by Dai and Wang et al. [57,65] for preparing anion-functionalized ILs via dehydrogenation for CO2 capture. There have been numerous investigations on CO2 capture by azolate ILs, including protic ILs and aprotic ILs, during this decade. Take imidazolate ILs as an example, the reported CO2 absorption capacities of [MTBDH][Im] [57], [(P2-Et)H][Im] [57], [P66614][Im] [65,66], [TMGH][Im] [67,68], [DBUH][Im] [69,70,71], and [DBNH][Im] [70] were in the rage of 0.8~1.0 at room temperature and atmospheric pressure. Different from aprotic cations or protic cations with a strong base, Oncsik and MacFarlane et al. [72] and Yang et al. [73] reported N,N-dimethylethylenediaminium azolate ILs for CO2 capture via forming carbamate species. The former authors believed that CO2 reacted with azolate anions, while the latter authors found that CO2 reacted with the cations. Thus, it is well understood that (1) the anion but not the cation has the key function in CO2 capture and (2) the mechanism follows a 1:1 stoichiometry. On the other hand, the absorption molar ratios of CO2 to IL were affected by the basicity of ILs. Wang et al. [65] showed that the absorption capacity decreased from 1.02 for [P66614][Pyrz] to 0.08 for [P66614][Tetz], when the pKa value of the azoles decreased from 19.8 for Pyrz to 8.2 for Tetz, and [P66614][Triz] was an ideal IL with desirable absorption enthalpy (−56 kJ mol−1) and high absorption capacity (0.95 mol CO2 per mol IL). Subsequently, kinds of [Triz]-based ILs were reported for CO2 capture [74,75].
The substituents on the heterocycles will affect the performance of CO2 capture by azolate ILs. Wu et al. [76] studied the reactivity of azolate anions with CO2 from the density functional theory (DFT) perspective. It was studied that the absorption capacity by imidazolate [66,69], pyrazolate ILs [77,78], and indazolate ILs [79] was affected by the substituents on the heterocycles. For example, Wang et al. [60] showed that the molar ratio of CO2 to [P66614][4-CHO-Im] could reach 1.24 at 20 °C and 1 bar via the interactions of imidazolate-CO2 and hydrogen bonding. Later, they also concluded that [P66614][4-Br-Im] was an ideal substituent imidazolate IL with desirable absorption enthalpy (−61.4 kJ mol−1), basicity (pKa is 12.2 in H2O), and CO2 capacity (0.87 mol CO2 per mol IL) [66]. It is also clearly that the substituents resulted in different basicity of azolate ILs. Recently, a light-responsive 1,3,4-trizolate IL was reported by Wang et al. [80], and the decreased capacity of CO2 was found when the IL converted from the trans to cis state. They confirmed that the entropy change was the key influencing factor. Additionally, different from the amino-functionalized ILs with the viscosity increasing during the CO2 absorption, the viscosity of the viscosity of [P66614][Im] was found to decrease from 810.4 cP to 648.7 cP after absorption of CO2. Jiang et al. [81] revealed the microscopic origin for the decrease in viscosity after CO2 absorption by [P66614][Im] via molecular dynamics (MD) simulation. Rogers et al. [76] reviewed the ILs with azolate anions due to their desired properties, including a diffuse ionic charge, tailorable asymmetry, and synthetic flexibility. Azolate anion-functionalized ILs are also reported with the name “aprotic heterocyclic anion” ([AHA]) based ILs. Maginn et al. [82] reported [P66614][2-CN-Pyr]) and [P66614][2-CF3-Pyra] could obtain ~0.9 mol CO2 per mol IL via a 1:1 mechanism. Brennecke et al. [83,84] investigated the influence of substituent groups on the reaction enthalpy of CO2–[AHA], and the estimated values range between −37 and −54 kJ mol–1, lower than that of CO2–MEA (−85 kJ mol–1). The structure and mechanism of azolate–CO2 was systematically studied via DFT calculations, [82] ab initio MD simulation [85,86,87], Monte Carlo simulation [88], first principles simulations [89,90], and other computer calculations [91,92].
Another reaction pathway was reported. It should be noted that when the anion has a certain basicity for CO2 capture, the basicity of the anion can cause it to pull out an active hydrogen atom on the imidazolium or phosphonium cation to form a carbene or zwitterionic compound (ylide), which could subsequently interact with CO2 to form a carbene–CO2 or ylide–CO2 complex, respectively. Brennecke et al. [93] selected four azolate ILs with different basicity and low basicity for [Tetz] and high basicity for [3-Triz], [4-Triz], and [2-CN-Pyr]. They quantified the amounts of cation–CO2 and anion–CO2 complexes. For [Emim][2-CN-Pyr], 59% of the C2 acidic protons are removed, leaving the carbene to react with CO2 and form the cation–CO2 complex. Chen et al. [94] showed that the more basic [AHA] anion would form carbene–CO2 via DFT. As the formed carbene–CO2 resulted in the reduced efficiency of anions, Wang et al. [95] investigated that substituted imidazolium reduced the amount of carbene–CO2 and increased the amount of azolate–CO2. For the ylide–CO2 pathway, Brennecke et al. [78,96] investigated the cation–anion and [AHA] –CO2 interactions and the quantification of ylide–CO2 in phosphonium ILs. In addition to the azolate anions, the phenolate anions and carboxylate anions can also result in carbene–CO2 in imidazolium ILs [97,98,99] or ylide–CO2 in phosphonium ILs [100], respectively.
The comparison of the absorption capacities, including molar capacities and corresponding mass capacities, of typical anion-functionalized ILs for CO2 capture via single-site mechanisms are listed in Table 2.

3.2.2. Multiple-Site Mechanism

It is known that single-site in ILs result in up to a 1:1 stoichiometry absorption capacity. However, multiple-site in ILs may not result in doubled capacity. For multiple sites sharing one negative charge, the efficiency of a site may be decreased. Besides, even if the two sites are independent or each has a negative charge, the absorption capacity may not double, due to the complex interactions in ILs. The typical multiple-site mechanisms can be found in Figure 5b.
(1)
Multiple same groups in anion-functionalized ILs
As amino group is a functional group for CO2 capture, AAILs based on amino acid anions with multiple amino were developed, including [Lys], [His], [Asn], and [Gln]. For [Lys], the molar ratios of CO2 to [P66614][Lys] [101], [N66614][Lys] [102], [C2OHmim][Lys] [103], and [N1,1,6,2O4][Lys] [104] were 1.37, 2.1, 1.68, and 1.62, respectively, via the reaction mechanism of 1:1. Different from two amino groups in one anion, CO2 capacities of several dicationic ILs with two amino acid anions [105] or azolate anions [106,107] were reported nearly twice that of the monocationic analogues. Additionally, CO2 absorption capacity of a superbase-derived diolate IL [MTBDH]+2[HFPD]2− reported by Dai et al. [57] was more than 2.04 mol CO2 per mol IL because of two alkoxide groups. Wang et al. [108] investigated the CO2 capture by a pillar[5]arene-based −10 valent carboxylate anion-functionalized phosphonium IL, [P66614]10[DCP5]. Their results showed that capacities of 5.52 mol CO2 per mol IL and 0.55 mol CO2 per mol carboxylate could be obtained through multiple-site interactions;
(2)
Pyridinolate anion-functionalized ILs
Although the N atom in neutral pyridine has poor ability for CO2 capture [109], Wang et al. [110] reported CO2 capture by a series of hydroxypyridine-based anion-functionalized ILs, including [P66614][2-Op], [P66614][4-Op], [P66614][3-Op], etc. The CO2 capacities of these ILs were more than 1 (up to 1.65) mol CO2 mol−1 IL due to the cooperative N–CO2 and O–CO2 interactions. Hao and Guan et al. [111] investigated the anion–CO2 interaction in hydroxypyridinate ILs with [P4444] or [N4444] cations via quantum chemistry calculations. A viscosity as low as 193 cP with an absorption capacity as high as 1.20 mol CO2 per mol IL were obtained by [P4444][2-Op]. Lin and Luo et al. [112] investigated a series of hydroxypyridine ILs with different kinds of cations, and the enhanced CO2 capacity up to 1.83 mol CO2 mol1 IL could be obtained at 20 °C and 1 bar via reducing cation-anion interactions. Xu et al. [113] reported that the CO2 capture capacity of ILs with [DBUH] and [TMGH] cations and hydroxypyridine anions followed the order of [2-Op] > [4-Op] > [3-Op]. The molar ratio of CO2 to [DBUH][2-Op] at 40 °C was up to ~0.90 mol CO2 per mol IL, similar to azolate ILs. In order to enhance the adsorption kinetics, CO2 capture by [2-Op]-based ILs were performed on porous supports with high capacities [114,115];
(3)
Imide anion-functionalized ILs
Imide and amide anions reported to have nucleophilic reactivities [116]. In order to improve CO2 capacity under low concentration CO2 (10 vol%), Cui and Wang et al. [117] synthesized a series of imide anion-functionalized ILs, [P4442][Suc] and [P4442][DAA]. Through pre-organization strategy, the prepared [P4442][Suc] showed a high efficiency on CO2 capture (1.65 mol CO2 mol−1 IL for 10 vol% and 1.87 mol CO2 mol−1 IL for 100 vol%) via cooperative 3 site (O-N-O)-2–CO2 interaction. Further studies showed that the electro-withdrawing phenyl group on the anion, [Ph-Suc], reduced the CO2 absorption capacity, while the electro-donating cyclohexyl group on the anion, [Cy-Suc], increased the CO2 absorption capacity (1.76 mol CO2 mol−1 IL for 10 vol% and 2.21 mol CO2 mol−1 IL for 100 vol%) via enhanced cooperation and physical interaction [118]. Additionally, the obtained imide-based ILs are stable in water, and the CO2 absorption could be improved under low water content [119]. The results of thermodynamic studies showed that the absorption was an enthalpy-driven process[120]. Wang et al. [121] reported an aminomethyl-functionalized tetrazolate IL, [P66614][MA-Tetz], with a CO2 capacity of 1.13 mol CO2 per mol IL, due to the interaction of one amino group (H-N-H) with two molecules of CO2;
(4)
Other multiple-site anion-functionalized ILs
When multiple sites are independent in an anion, they give an opportunity for improving CO2 capture. As amino and carboxylate could both interact with CO2 efficiently, Tao et al. [122] synthesized the [P4442]2[IDA] with a −2 valent amino acid anion, and the improved absorption capacity was 1.69 mol CO2 per mol IL through amino-CO2 and carboxylate-CO2 interactions. Subsequently, Pan and Zou et al. [123] reported a series of AA ILs based on −2 valent amino acid anions. Compared with hydroxyl-containing −1 valent counterparts, alkoxide anion-containing −2 valent AA ILs, [P4442]2[D-Ser] and [P4442]2[L-Ser], showed high CO2 capacity due to the interactions of amino-CO2 and alkoxide-CO2. Luo and Lin et al. [124] reported kinds of −2 valent AA ILs, [P66614]2[AA-R], where R was sulfonate (Su), carboxylate (Ac), imidazolate (Im), or indolate (Ind). The CO2 capacities of [P66614]2[AA-Su], [P66614]2[AA-Ac], [P66614]2[AA-Im], and [P66614]2[AA-Ind] were 0.49, 1.97, 1.55, and 1.45 mol CO2 per mol IL, respectively. The results presented that CO2 capacity increased first and then decreased later with the continuous increase in the activity of the anion site.
On the other hand, when multiple sites are dependent in an anion, their interactions may lead to mutual restraint for CO2 capture. For example, Liu et al. [125] synthesized a -3 valent carboxylate- hydroxypyridinate- containing anion IL, [P4444]3[2,4-OPym-5-Ac] and found that a CO2 capacity of 1.46 mol CO2 per mol IL could be obtained, lower than the theoretical value. However, Luo and Lin et al. [124] reported a −2 valent IL, [P66614]2[Am-iPA], with amino functionalized dicarboxylate anion. Their results showed that CO2 capture capacity of this IL was 2.38 mol CO2 per mol IL at 30 °C. Besides, multiple sites dependently shared one negative charge, resulting in decreased efficiency of CO2 capture. Wang and MacFarlane et al. [126] studied CO2 capture by an amino-containing hydroxypyridinate anion-functionalized ILs with the capacity of 0.87~0.99 mol CO2 per mol IL. The NMR results indicated the primary reaction of amino-CO2 and the lesser reaction of phenolate-CO2. Tao et al. [127] reported CO2 capture by amino-functionalized triazolate anion ILs, [Bmim][ATZ] and [Emim][ATZ], with a capacity as low as 0.14 and 0.13 mol CO2 per mol IL, respectively, via physical interaction.
The comparison of the absorption capacities, including molar capacities and corresponding mass capacities, of typical anion-functionalized ILs for CO2 capture via multiple-site mechanisms are listed in Table 3.

3.3. Cation-Anion Dual-Functionalized ILs for CO2 Capture

It is clear that functional groups on cations are mainly amino groups. Thus, with the combination of functional cations and functional anions, kinds of dual-functionalized ILs with multiple sites were developed for CO2 capture.

3.3.1. Amino-Based Cation and Amino-Based Anion

The early dual-functionalized ILs were ILs with amino-functionalized cations with amino acid anions. For example, [aP4443][AA] (Gly, and [Ala]) [128], [aemmim][Tau] [129], [apaeP444][AA] ([Lys], [Gly], [Ser], [Ala], [Asp], and [His]) [130], and [AEMP][AA] ([Gly], [Ala], [Pro], and [Leu]) [131], and [APmim][AA] ([Gly] and [Lys]) [132,133] were found to capture CO2 via 2:1 mechanism (amino:CO2). Jing et al. [134] used quantum chemical simulation for screening of multi-amino-functionalized ILs for CO2 capture. Their experimental results confirmed the predictions and the absorption capacities of [TETAH][Lys] (5 amino groups) and [DETAH][Lys] (4 amino groups) were 2.59 and 2.13 mol CO2 per mol IL, respectively, via 2:1 zwitterionic mechanism.

3.3.2. Amino-Based Cation and Phenolate Anion

Based on the high reactivity of phenolate anions for CO2 capture, Ye and Li et al. [135] reported that the CO2 absorption capacities of supported dual functionalized phosphonium ILs [aP4443][2-Op] and [aP4443][2-Np] were 1.57 and 1.88 mol CO2 per mol IL, respectively, via 2:1 mechanism of amino-CO2 and 1:1 mechanism of phenolate-CO2 mechanism. Recently, Wang et al. [136] reported two dual-functionalized protic ILs, dimethylethylenediamine 4-fluorophenolate ([DMEDAH][4-F-PhO]) and dimethylethylenediamine acetate ([DMEDAH][OAc]) to investigate the different chemisorption mechanisms via DFT study. Their results showed that, for [DMEDAH][4-F-PhO], phenolate-CO2 was favorable in kinetics and amino-CO2 was thermodynamically beneficial; for [DMEDAH][OAc], amino-CO2 was favorable with proton-transfer to weak acid anion.

3.3.3. Amino-Based Cation and Azolate Anion

Based on the high reactivity of azolate anions for CO2 capture, Ye and Li et al. [135] reported that the CO2 absorption capacities of supported dual functionalized phosphonium IL [aP4443][Triz] was 1.32 mol CO2 per mol IL via 1:1 mechanism of azolate-CO2 mechanism. Considering the metal coordination of amino groups as well as the CO2-philic azolate anions, Xu et al. reported a series of polyamine-based dual-functionalized azolate ILs with different structures of polyamines, metal ions, and azolate anions. For example, CO2 capacities of [Na(MDEA)2][Pyrz] [137], [K(DGA)2][Im] [138], and [K(AMP)2][Im] [139], were 0.75 (80 °C), 1.37 (60 °C), and 1.19 (60 °C) mol CO2 per mol IL, respectively, via reactions of amino-CO2 and azolate-CO2.
The comparison of the absorption capacities, including molar capacities and corresponding mass capacities, of typical cation-anion dual-functionalized ILs for CO2 capture are listed in Table 4.

4. Conclusions and Outlook

It is known that functionalized ILs started in 2002, and it has been just two decades. Due to the designable and tunable structures of ILs, functionalized ILs have developed rapidly in the past ten years (2012–2022). CO2-philic active sites can be tethered to the cations and anions, forming cation-functionalized ILs, anion-functionalized ILs, and cation-anion dual-functionalized ILs. Compared with conventional ILs for physisorption of CO2, functionalized ILs or task-specific ILs could chemically absorb CO2 through single-site mechanisms or multiple-site mechanisms. Based on the research results, we can safely conclude that efficient absorption of CO2 with a high capacity, low energy consumption, and high reversibility could be reached through tuning the structures of functionalized ILs and regulating the interactions between active sites and CO2. Nonetheless, for large-scale industrial application of IL-based CCUS technology, we also need to consider the following issues:
(1)
Reaction mechanism of functionalized IL-CO2 needs to be investigated further;
(2)
A large amount of CO2 absorption experiments was tested at room temperature and atmospheric pressure, but the temperature of flue gas is high (50~80 °C) and the concentration of CO2 is low (10~15 vol%), there is still a big gap between laboratory research and industrial application;
(3)
The selective capture of CO2 and the deactivation of functionalized ILs under other gases conditions (H2O, SO2, NOx, etc.) should be studied;
(4)
Compared with conventional absorbents such as alkanolamine aqueous solutions, pure functionalized ILs have higher viscosity and cost;
(5)
It is important to investigate capture efficiency in mass absorption capacity or gravimetric capacity in order to better comparison and realize the competitive ILs. Thus, functionalized ILs with high mass absorption capacity should be developed.
(6)
The regeneration of the ILs is also important and related to energetic consume and the absorption cost. Thus, the absorption enthalpies should be investigated.
Here are some suggestions or strategies to address the aforementioned issues:
(1)
A combination of NMR and IR analysis and chemical calculations can be used to investigate the absorption mechanisms of active sites on the ILs with CO2;
(2)
The performance of CO2 capture is affected by absorption temperature and CO2 partial pressure. Due to the tunable structure and property of ILs, design functionalized ILs with high active sites is an efficient way to help ILs applicate in industry;
(3)
H2O, SO2, NOx, etc. will lead to a decrease in the activity of ILs, especially ILs with strong basicity. Thus, these impurities should first be removed. For example, ILs with weak basicity for SO2 or NOx removal and ILs with strong basicity for CO2 removal;
(4)
Functionalized ILs with a low viscosity could be synthesized through tuning the structures of cation and anion. Besides, the viscosity of amine-containing functionalized ILs or protic ILs were reported to be increased during the absorption of CO2, while for amine-free functionalized ILs and aprotic ILs no obvious change during CO2 capture was reported due to the absence of strong hydrogen bonded networks in these ILs (Table 5);
(5)
Aqueous monoethanolamine (30 wt%) process is the current CO2 capture technology in industry with a mass capacity of ~7 wt%. It can be found in Table 1, Table 2, Table 3 and Table 4 that functionalized ILs with a high molecular weight resulted in a high molar capacity but a low mass capacity. Functionalized ILs with a high molar capacity open the door to developing functionalized ILs with a high mass capacity via combining functional sites and a small molecular weight;
(6)
High regeneration or reversibility of the ILs for CO2 capture needs weak interactions or low absorption enthalpies, which always results in low efficiency. Thus, functionalized ILs is always accompanied by high energy consumption. However, the results from CO2 capture by preorganized imide-based ILs indicate that multiple weak interactions also lead to strong adsorption and high capacity, even under low concentrations of CO2.
Therefore, continuously developing novel functional IL-based CO2-philic solvents or sorbents and systematically studying the reaction mechanism of CO2 with active sites under different conditions are the main concern of IL-based CCUS technologies in order to realize large-scale, rapid, economical, efficient, and reversible absorption of CO2 in the flue gas.

Author Contributions

Conceptualization and project administration, G.C.; writing—original draft preparation and visualization, R.Z.; writing—review and editing, G.C.; funding acquisition, G.C. and H.L.; validation, Q.K. and Z.Z.; formal analysis, B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (No. 22078294), the Zhejiang Provincial Natural Science Foundation of China (No. LZ21E80001 and LGF20E080018), and the Zhejiang Tiandi Environmental Protection Technology Co., Ltd. “Development of Ionic Liquid Absorbents for Carbon Dioxide Capture with Green and Low Energy Consumption” Technology Project (No. TD-KJ-22-007-W001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Meteorological Organization. State of the Global Climate 2021; WMO-No. 1290; WMO: Geneva, Switzerland, 2022; Available online: https://library.wmo.int/index.php?lvl=notice_display&id=22080 (accessed on 1 August 2022).
  2. Chen, S.; Liu, J.; Zhang, Q.; Teng, F.; McLellan, B.C. A critical review on deployment planning and risk analysis of carbon capture, utilization, and storage (CCUS) toward carbon neutrality. Renew. Sustain. Energy Rev. 2022, 167, 112537. [Google Scholar] [CrossRef]
  3. Chen, H.; Tsai, T.-C.; Tan, C.-S. CO2 capture using amino acid sodium salt mixed with alkanolamines. Int. J. Greenh. Gas Control 2018, 79, 127–133. [Google Scholar] [CrossRef]
  4. Bhalothia, D.; Hsiung, W.-H.; Yang, S.-S.; Yan, C.; Chen, P.-C.; Lin, T.-H.; Wu, S.-C.; Chen, P.-C.; Wang, K.-W.; Lin, M.-W.; et al. Submillisecond Laser Annealing Induced Surface and Subsurface Restructuring of Cu–Ni–Pd Trimetallic Nanocatalyst Promotes Thermal CO2 Reduction. ACS Appl. Energy Mater. 2021, 4, 14043–14058. [Google Scholar] [CrossRef]
  5. Zhao, Y.; Han, B.; Liu, Z. Ionic-Liquid-Catalyzed Approaches under Metal-Free Conditions. Accounts Chem. Res. 2021, 54, 3172–3190. [Google Scholar] [CrossRef]
  6. Yan, C.; Wang, C.-H.; Lin, M.; Bhalothia, D.; Yang, S.-S.; Fan, G.-J.; Wang, J.-L.; Chan, T.-S.; Wang, Y.-l.; Tu, X.; et al. Local synergetic collaboration between Pd and local tetrahedral symmetric Ni oxide enables ultra-high-performance CO2 thermal methanation. J. Mater. Chem. A 2020, 8, 12744–12756. [Google Scholar] [CrossRef]
  7. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef]
  8. Quintana, A.A.; Sztapka, A.M.; Ebinuma, V.D.C.S.; Agatemor, C. Enabling Sustainable Chemistry with Ionic Liquids and Deep Eutectic Solvents: A Fad or the Future? Angew. Chem. Int. Ed. 2022, 61, e202205609. [Google Scholar] [CrossRef]
  9. Fajardo, O.Y.; Bresme, F.; Kornyshev, A.A.; Urbakh, M. Electrotunable Friction with Ionic Liquid Lubricants: How Important Is the Molecular Structure of the Ions? J. Phys. Chem. Lett. 2015, 6, 3998–4004. [Google Scholar] [CrossRef]
  10. Wang, Y.; He, H.; Wang, C.; Lu, Y.; Dong, K.; Huo, F.; Zhang, S. Insights into Ionic Liquids: From Z-Bonds to Quasi-Liquids. JACS Au 2022, 2, 543–561. [Google Scholar] [CrossRef]
  11. Beil, S.; Markiewicz, M.; Pereira, C.S.; Stepnowski, P.; Thöming, J.; Stolte, S. Toward the Proactive Design of Sustainable Chemicals: Ionic Liquids as a Prime Example. Chem. Rev. 2021, 121, 13132–13173. [Google Scholar] [CrossRef]
  12. Roy, S. Ahmaruzzaman Ionic liquid based composites: A versatile materials for remediation of aqueous environmental contaminants. J. Environ. Manag. 2022, 315, 115089. [Google Scholar] [CrossRef]
  13. Tang, X.; Lv, S.; Jiang, K.; Zhou, G.; Liu, X. Recent development of ionic liquid-based electrolytes in lithium-ion batteries. J. Power Sources 2022, 542, 231792. [Google Scholar] [CrossRef]
  14. Sun, L.; Zhuo, K.; Chen, Y.; Du, Q.; Zhang, S.; Wang, J. Ionic Liquid-Based Redox Active Electrolytes for Supercapacitors. Adv. Funct. Mater. 2022, 32, 2203611. [Google Scholar] [CrossRef]
  15. Li, Z.; Han, Q.; Wang, K.; Song, S.; Xue, Y.; Ji, X.; Zhai, J.; Huang, Y.; Zhang, S. Ionic liquids as a tunable solvent and modifier for biocatalysis. Catal. Rev. 2022, 117, 1–47. [Google Scholar] [CrossRef]
  16. Tan, X.; Sun, X.; Han, B. Ionic liquid-based electrolytes for CO2 electroreduction and CO2 electroorganic transformation. Natl. Sci. Rev. 2021, 9, nwab022. [Google Scholar] [CrossRef]
  17. Hosseini, A.; Khoshsima, A.; Sabzi, M.; Rostam, A. Toward Application of Ionic Liquids to Desulfurization of Fuels: A Review. Energy Fuels 2022, 36, 4119–4152. [Google Scholar] [CrossRef]
  18. Zhuang, W.; Hachem, K.; Bokov, D.; Ansari, M.J.; Nakhjiri, A.T. Ionic liquids in pharmaceutical industry: A systematic review on applications and future perspectives. J. Mol. Liq. 2021, 349, 118145. [Google Scholar] [CrossRef]
  19. Nikfarjam, N.; Ghomi, M.; Agarwal, T.; Hassanpour, M.; Sharifi, E.; Khorsandi, D.; Ali Khan, M.; Rossi, F.; Rossetti, A.; Nazarzadeh Zare, E.; et al. Antimicrobial Ionic Liquid-Based Materials for Biomedical Applications. Adv. Funct. Mater. 2021, 31, 2104148. [Google Scholar] [CrossRef]
  20. Zeng, S.; Zhang, X.; Bai, L.; Zhang, X.; Wang, H.; Wang, J.; Bao, D.; Li, M.; Liu, X.; Zhang, S. Ionic-Liquid-Based CO2 Capture Systems: Structure, Interaction and Process. Chem. Rev. 2017, 117, 9625–9673. [Google Scholar] [CrossRef] [PubMed]
  21. Liu, Y.; Dai, Z.; Zhang, Z.; Zeng, S.; Li, F.; Zhang, X.; Nie, Y.; Zhang, L.; Zhang, S.; Ji, X. Ionic liquids/deep eutectic solvents for CO2 capture: Reviewing and evaluating. Green Energy Environ. 2020, 6, 314–328. [Google Scholar] [CrossRef]
  22. Bates, E.D.; Mayton, R.D.; Ntai, I.; Davis, J.H. CO2 capture by a task-specific ionic liquid. J. Am. Chem. Soc. 2002, 124, 926–927. [Google Scholar] [CrossRef]
  23. Heldebrant, D.J.; Yonker, C.R.; Jessop, P.G.; Phan, L. Organic liquid CO2 capture agents with high gravimetric CO2 capacity. Energy Environ. Sci. 2008, 1, 487–493. [Google Scholar] [CrossRef]
  24. Rochelle, G. Conventional amine scrubbing for CO2 capture. Science 2009, 325, 1652–1654. [Google Scholar] [CrossRef]
  25. Blanchard, L.A.; Hancu, D.; Beckman, E.J.; Brennecke, J.F. Green processing using ionic liquids and CO2. Nature 1999, 399, 28–29. [Google Scholar] [CrossRef]
  26. Sistla, Y.S.; Khanna, A. Carbon dioxide absorption studies using amine-functionalized ionic liquids. J. Ind. Eng. Chem. 2014, 20, 2497–2509. [Google Scholar] [CrossRef]
  27. Sharma, P.; Park, S.D.; Park, K.T.; Nam, S.C.; Jeong, S.K.; Yoon, Y.I.; Baek, I.H. Solubility of carbon dioxide in amine-functionalized ionic liquids: Role of the anions. Chem. Eng. J. 2012, 193–194, 267–275. [Google Scholar] [CrossRef]
  28. Gutowski, K.E.; Maginn, E.J. Amine-Functionalized Task-Specific Ionic Liquids: A Mechanistic Explanation for the Dramatic Increase in Viscosity upon Complexation with CO2 from Molecular Simulation. J. Am. Chem. Soc. 2008, 130, 14690–14704. [Google Scholar] [CrossRef]
  29. Cao, B.; Du, J.; Liu, S.; Zhu, X.; Sun, X.; Sun, H.; Fu, H. Carbon dioxide capture by amino-functionalized ionic liquids: DFT based theoretical analysis substantiated by FT-IR investigation. RSC Adv. 2016, 6, 10462–10470. [Google Scholar] [CrossRef]
  30. Sistla, Y.S.; Khanna, A. CO2 absorption studies in amino acid-anion based ionic liquids. Chem. Eng. J. 2015, 273, 268–276. [Google Scholar] [CrossRef]
  31. Yang, Z.-Z.; He, L.-N. Efficient CO2 capture by tertiary amine-functionalized ionic liquids through Li+-stabilized zwitterionic adduct formation. Beilstein J. Org. Chem. 2014, 10, 1959–1966. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, J.; Jia, C.; Dong, H.; Wang, J.; Zhang, X.; Zhang, S. A Novel Dual Amino-Functionalized Cation-Tethered Ionic Liquid for CO2 Capture. Ind. Eng. Chem. Res. 2013, 52, 5835–5841. [Google Scholar] [CrossRef]
  33. Doyle, K.A.; Murphy, L.J.; Paula, Z.A.; Land, M.A.; Robertson, K.N.; Clyburne, J.A.C. Characterization of a New Ionic Liquid and Its Use for CO2 Capture from Ambient Air: Studies on Solutions of Diethylenetriamine (DETA) and [DETAH]NO3 in Polyethylene Glycol. Ind. Eng. Chem. Res. 2015, 54, 8829–8841. [Google Scholar] [CrossRef]
  34. Hu, P.; Zhang, R.; Liu, Z.; Liu, H.; Xu, C.; Meng, X.; Liang, M.; Liang, S. Absorption Performance and Mechanism of CO2 in Aqueous Solutions of Amine-Based Ionic Liquids. Energy Fuels 2015, 29, 6019–6024. [Google Scholar] [CrossRef]
  35. Wang, C.; Guo, Y.; Zhu, X.; Cui, G.; Li, H.; Dai, S. Highly efficient CO2 capture by tunable alkanolamine-based ionic liquids with multidentate cation coordination. Chem. Commun. 2012, 48, 6526–6528. [Google Scholar] [CrossRef]
  36. Shi, G.; Zhao, H.; Chen, K.; Lin, W.; Li, H.; Wang, C. Efficient capture of CO2 from flue gas at high temperature by tunable polyamine-based hybrid ionic liquids. AIChE J. 2019, 66, e16779. [Google Scholar] [CrossRef]
  37. Suo, X.; Yang, Z.; Fu, Y.; Do-Thanh, C.-L.; Maltsev, D.; Luo, H.; Mahurin, S.M.; Jiang, D.-E.; Xing, H.; Dai, S. New-Generation Carbon-Capture Ionic Liquids Regulated by Metal-Ion Coordination. ChemSusChem 2022, 15, e202102136. [Google Scholar] [CrossRef]
  38. Zhang, J.; Zhang, S.; Dong, K.; Zhang, Y.; Shen, Y.; Lv, X. Supported Absorption of CO2 by Tetrabutylphosphonium Amino Acid Ionic Liquids. Chem. A Eur. J. 2006, 12, 4021–4026. [Google Scholar] [CrossRef] [PubMed]
  39. Gurkan, B.E.; de la Fuente, J.C.; Mindrup, E.M.; Ficke, L.E.; Goodrich, B.F.; Price, E.A.; Schneider, W.F.; Brennecke, J.F. Equimolar CO2 Absorption by Anion-Functionalized Ionic Liquids. J. Am. Chem. Soc. 2010, 132, 2116–2117. [Google Scholar] [CrossRef] [PubMed]
  40. Brennecke, J.F.; Gurkan, B.E. Ionic Liquids for CO2 Capture and Emission Reduction. J. Phys. Chem. Lett. 2010, 1, 3459–3464. [Google Scholar] [CrossRef]
  41. Yang, Q.; Wang, Z.; Bao, Z.; Zhang, Z.; Yang, Y.; Ren, Q.; Xing, H.; Dai, S. New Insights into CO2 Absorption Mechanisms with Amino-Acid Ionic Liquids. ChemSusChem 2016, 9, 806–812. [Google Scholar] [CrossRef]
  42. Noorani, N.; Mehrdad, A. CO2 solubility in some amino acid-based ionic liquids: Measurement, correlation and DFT studies. Fluid Phase Equilibria 2020, 517, 112591. [Google Scholar] [CrossRef]
  43. Ren, H.; Shen, H.; Liu, Y. Adsorption of CO2 with tetraethylammonium glycine ionic liquid modified alumina in the Rotating Adsorption Bed. J. CO2 Util. 2022, 58, 101925. [Google Scholar] [CrossRef]
  44. Zhang, W.; Gao, E.; Li, Y.; Bernards, M.T.; Li, Y.; Cao, G.; He, Y.; Shi, Y. Synergistic Enhancement of CO2 Adsorption Capacity and Kinetics in Triethylenetetrammonium Nitrate Protic Ionic Liquid Functionalized SBA. Energy Fuels 2019, 33, 8967–8975. [Google Scholar] [CrossRef]
  45. Wang, X.; Akhmedov, N.G.; Duan, Y.; Luebke, D.; Li, B. Immobilization of amino acid ionic liquids into nanoporous microspheres as robust sorbents for CO2 capture. J. Mater. Chem. A 2013, 1, 2978–2982. [Google Scholar] [CrossRef]
  46. Zheng, S.; Zeng, S.; Li, Y.; Bai, L.; Bai, Y.; Zhang, X.; Liang, X.; Zhang, S. State of the art of ionic liquid-modified adsorbents for CO2 capture and separation. AIChE J. 2022, 68, e17500. [Google Scholar] [CrossRef]
  47. Maginn, E.J. Design and Evaluation of Ionic Liquids as Novel CO2 Absorbents; Quaterly Technical Reports to DOE (Award Number: DE-FG26-04NT42122). 2005. Available online: https://www.semanticscholar.org/paper/Design-and-Evaluation-of-Ionic-Liquids-as-Novel-CO2-Maginn/ae256f654ad690eee1e3c0231154748db89adbbb#citing-papers (accessed on 1 August 2022).
  48. Steckel, J.A. Ab Initio Calculations of the Interaction between CO2 and the Acetate Ion. J. Phys. Chem. A 2012, 116, 11643–11650. [Google Scholar] [CrossRef]
  49. Shi, W.; Thompson, R.L.; Albenze, E.; Steckel, J.A.; Nulwala, H.B.; Luebke, D.R. Contribution of the Acetate Anion to CO2 Solubility in Ionic Liquids: Theoretical Method Development and Experimental Study. J. Phys. Chem. B 2014, 118, 7383–7394. [Google Scholar] [CrossRef]
  50. Harb, W.; Ingrosso, F.; Ruiz-López, M.F. Molecular insights into the carbon dioxide–carboxylate anion interactions and implications for carbon capture. Theor. Chim. Acta 2019, 138, 85. [Google Scholar] [CrossRef]
  51. Chen, F.-F.; Dong, Y.; Sang, X.-Y.; Zhou, Y.; Tao, D.-J. Physicochemical Properties and CO2 Solubility of Tetrabutylphosphonium Carboxylate Ionic Liquids. Acta Phys.-Chim. Sin. 2016, 32, 605–610. [Google Scholar] [CrossRef]
  52. Zailani, N.H.Z.O.; Yunus, N.M.; Ab Rahim, A.H.; Bustam, M.A. Experimental Investigation on Thermophysical Properties of Ammonium-Based Protic Ionic Liquids and Their Potential Ability towards CO2 Capture. Molecules 2022, 27, 851. [Google Scholar] [CrossRef]
  53. Wang, N.; Cheng, H.; Wang, Y.; Yang, Y.; Teng, Y.; Li, C.; Zheng, S. Measuring and modeling the solubility of carbon dioxide in protic ionic liquids. J. Chem. Thermodyn. 2022, 173, 106838. [Google Scholar] [CrossRef]
  54. Umecky, T.; Abe, M.; Takamuku, T.; Makino, T.; Kanakubo, M. CO2 absorption features of 1-ethyl-3-methylimidazolium ionic liquids with 2,4-pentanedionate and its fluorine derivatives. J. CO2 Util. 2019, 31, 75–84. [Google Scholar] [CrossRef]
  55. Chen, Y.; Hu, X.; Chen, W.; Liu, C.; Qiao, K.; Zhu, M.; Lou, Y.; Mu, T. High volatility of superbase-derived eutectic solvents used for CO2 capture. Phys. Chem. Chem. Phys. 2020, 23, 2193–2210. [Google Scholar] [CrossRef]
  56. Huang, Z.; Jiang, B.; Yang, H.; Wang, B.; Zhang, N.; Dou, H.; Wei, G.; Sun, Y.; Zhang, L. Investigation of glycerol-derived binary and ternary systems in CO2 capture process. Fuel 2017, 210, 836–843. [Google Scholar] [CrossRef]
  57. Wang, C.; Luo, H.; Jiang, D.-E.; Li, H.; Dai, S. Carbon Dioxide Capture by Superbase-Derived Protic Ionic Liquids. Angew. Chem. Int. Ed. 2010, 49, 5978–5981. [Google Scholar] [CrossRef]
  58. Zhao, Y.; Yu, B.; Yang, Z.; Zhang, H.; Hao, L.; Gao, X.; Liu, Z. A Protic Ionic Liquid Catalyzes CO2 Conversion at Atmospheric Pressure and Room Temperature: Synthesis of Quinazoline-2,4(1H,3H)-diones. Angew. Chem. Int. Ed. 2014, 53, 5922–5925. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, C.; Luo, H.; Li, H.; Zhu, X.; Yu, B.; Dai, S. Tuning the Physicochemical Properties of Diverse Phenolic Ionic Liquids for Equimolar CO2 Capture by the Substituent on the Anion. Chem. A Eur. J. 2012, 18, 2153–2160. [Google Scholar] [CrossRef] [PubMed]
  60. Ding, F.; He, X.; Luo, X.; Lin, W.; Chen, K.; Li, H.; Wang, C. Highly efficient CO2 capture by carbonyl-containing ionic liquids through Lewis acid-base and cooperative C-HO hydrogen bonding interaction strengthened by the anion. Chem. Commun. 2014, 50, 15041–15044. [Google Scholar] [CrossRef]
  61. Pan, M.; Cao, N.; Lin, W.; Luo, X.; Chen, K.; Che, S.; Li, H.; Wang, C. Reversible CO2Capture by Conjugated Ionic Liquids through Dynamic Covalent Carbon-Oxygen Bonds. ChemSusChem 2016, 9, 2351–2357. [Google Scholar] [CrossRef]
  62. Zhang, X.; Huang, K.; Xia, S.; Chen, Y.-L.; Wu, Y.-T.; Hu, X.-B. Low-viscous fluorine-substituted phenolic ionic liquids with high performance for capture of CO2. Chem. Eng. J. 2015, 274, 30–38. [Google Scholar] [CrossRef]
  63. Zhao, T.; Zhang, X.; Tu, Z.; Wu, Y.; Hu, X. Low-viscous diamino protic ionic liquids with fluorine-substituted phenolic anions for improving CO2 reversible capture. J. Mol. Liq. 2018, 268, 617–624. [Google Scholar] [CrossRef]
  64. Suo, X.; Yang, Z.; Fu, Y.; Do-Thanh, C.; Chen, H.; Luo, H.; Jiang, D.; Mahurin, S.M.; Xing, H.; Dai, S. CO2 Chemisorption Behavior of Coordination-Derived Phenolate Sorbents. ChemSusChem 2021, 14, 2784. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, C.; Luo, X.; Luo, H.; Jiang, D.-E.; Li, H.; Dai, S. Tuning the Basicity of Ionic Liquids for Equimolar CO2 Capture. Angew. Chem. Int. Ed. 2011, 50, 4918–4922. [Google Scholar] [CrossRef] [PubMed]
  66. Cui, G.; Zhao, N.; Wang, J.; Wang, C. Computer-Assisted Design of Imidazolate-Based Ionic Liquids for Improving Sulfur Dioxide Capture, Carbon Dioxide Capture, and Sulfur Dioxide/Carbon Dioxide Selectivity. Chem. Asian J. 2017, 12, 2863–2872. [Google Scholar] [CrossRef]
  67. Lei, X.; Xu, Y.; Zhu, L.; Wang, X. Highly efficient and reversible CO2 capture through 1,1,3,3-tetramethylguanidinium imidazole ionic liquid. RSC Adv. 2014, 4, 7052–7057. [Google Scholar] [CrossRef]
  68. Li, F.; Bai, Y.; Zeng, S.; Liang, X.; Wang, H.; Huo, F.; Zhang, X. Protic ionic liquids with low viscosity for efficient and reversible capture of carbon dioxide. Int. J. Greenh. Gas Control 2019, 90, 102801. [Google Scholar] [CrossRef]
  69. Zhu, X.; Song, M.; Xu, Y. DBU-Based Protic Ionic Liquids for CO2 Capture. ACS Sustain. Chem. Eng. 2017, 5, 8192–8198. [Google Scholar] [CrossRef]
  70. Xu, Y. CO2 absorption behavior of azole-based protic ionic liquids: Influence of the alkalinity and physicochemical properties. J. CO2 Util. 2017, 19, 1–8. [Google Scholar] [CrossRef]
  71. Gao, F.; Wang, Z.; Ji, P.; Cheng, J.-P. CO2 Absorption by DBU-Based Protic Ionic Liquids: Basicity of Anion Dictates the Absorption Capacity and Mechanism. Front. Chem. 2019, 6, 658. [Google Scholar] [CrossRef]
  72. Oncsik, T.; Vijayaraghavan, R.; MacFarlane, D.R. High CO2 absorption by diamino protic ionic liquids using azolide anions. Chem. Commun. 2018, 54, 2106–2109. [Google Scholar] [CrossRef]
  73. Wang, X.; Wu, C.; Yang, D. CO2 Absorption Mechanism by Diamino Protic Ionic Liquids (DPILs) Containing Azolide Anions. Processes 2021, 9, 1023. [Google Scholar] [CrossRef]
  74. Zhang, X.; Xiong, W.; Peng, L.; Wu, Y.; Hu, X. Highly selective absorption separation of H2S and CO2 from CH4 by novel azole-based protic ionic liquids. AIChE J. 2020, 66, e16936. [Google Scholar] [CrossRef]
  75. Zhu, X.; Song, M.; Ling, B.; Wang, S.; Luo, X. The Highly Efficient Absorption of CO2 by a Novel DBU Based Ionic Liquid. J. Solut. Chem. 2020, 49, 257–271. [Google Scholar] [CrossRef]
  76. Tang, H.; Wu, C. Reactivity of Azole Anions with CO2 from the DFT Perspective. ChemSusChem 2013, 6, 1050–1056. [Google Scholar] [CrossRef]
  77. Fu, H.; Wang, X.; Sang, H.; Fan, R.; Han, Y.; Zhang, J.; Liu, Z. The study of bicyclic amidine-based ionic liquids as promising carbon dioxide capture agents. J. Mol. Liq. 2020, 304, 112805. [Google Scholar] [CrossRef]
  78. Oh, S.; Morales-Collazo, O.; Keller, A.N.; Brennecke, J.F. Cation–Anion and Anion–CO2 Interactions in Triethyl(octyl)phosphonium Ionic Liquids with Aprotic Heterocyclic Anions (AHAs). J. Phys. Chem. B 2020, 124, 8877–8887. [Google Scholar] [CrossRef]
  79. Keller, A.N.; Bentley, C.L.; Morales-Collazo, O.; Brennecke, J.F. Design and Characterization of Aprotic N-Heterocyclic Anion Ionic Liquids for Carbon Capture. J. Chem. Eng. Data 2022, 67, 375–384. [Google Scholar] [CrossRef]
  80. Lin, W.; Pan, M.; Xiao, Q.; Li, H.; Wang, C. Tuning the Capture of CO2 through Entropic Effect Induced by Reversible Trans–Cis Isomerization of Light-Responsive Ionic Liquids. J. Phys. Chem. Lett. 2019, 10, 3346–3351. [Google Scholar] [CrossRef]
  81. Li, A.; Tian, Z.; Yan, T.; Jiang, D.-E.; Dai, S. Anion-Functionalized Task-Specific Ionic Liquids: Molecular Origin of Change in Viscosity upon CO2 Capture. J. Phys. Chem. B 2014, 118, 14880–14887. [Google Scholar] [CrossRef] [PubMed]
  82. Gurkan, B.; Goodrich, B.F.; Mindrup, E.M.; Ficke, L.E.; Massel, M.; Seo, S.; Senftle, T.P.; Wu, H.; Glaser, M.F.; Shah, J.K.; et al. Molecular Design of High Capacity, Low Viscosity, Chemically Tunable Ionic Liquids for CO2 Capture. J. Phys. Chem. Lett. 2010, 1, 3494–3499. [Google Scholar] [CrossRef]
  83. Seo, S.; Quiroz-Guzman, M.; DeSilva, M.A.; Lee, T.B.; Huang, Y.; Goodrich, B.F.; Schneider, W.F.; Brennecke, J.F. Chemically Tunable Ionic Liquids with Aprotic Heterocyclic Anion (AHA) for CO2 Capture. J. Phys. Chem. B 2014, 118, 5740–5751. [Google Scholar] [CrossRef] [PubMed]
  84. Seo, S.; Simoni, L.D.; Ma, M.; DeSilva, M.A.; Huang, Y.; Stadtherr, M.A.; Brennecke, J.F. Phase-Change Ionic Liquids for Postcombustion CO2 Capture. Energy Fuels 2014, 28, 5968–5977. [Google Scholar] [CrossRef]
  85. Sheridan, Q.R.; Mullen, R.G.; Lee, T.B.; Maginn, E.J.; Schneider, W.F. Hybrid Computational Strategy for Predicting CO2 Solubilities in Reactive Ionic Liquids. J. Phys. Chem. C 2018, 122, 14213–14221. [Google Scholar] [CrossRef]
  86. Sheridan, Q.R.; Oh, S.; Morales-Collazo, O.; Castner, J.E.W.; Brennecke, J.F.; Maginn, E.J. Liquid Structure of CO2–Reactive Aprotic Heterocyclic Anion Ionic Liquids from X-ray Scattering and Molecular Dynamics. J. Phys. Chem. B 2016, 120, 11951–11960. [Google Scholar] [CrossRef]
  87. Wu, H.; Shah, J.K.; Tenney, C.M.; Rosch, T.W.; Maginn, E.J. Structure and Dynamics of Neat and CO2-Reacted Ionic Liquid Tetrabutylphosphonium 2-Cyanopyrrolide. Ind. Eng. Chem. Res. 2011, 50, 8983–8993. [Google Scholar] [CrossRef]
  88. Mullen, R.G.; Corcelli, S.A.; Maginn, E.J. Reaction Ensemble Monte Carlo Simulations of CO2 Absorption in the Reactive Ionic Liquid Triethyl(octyl)phosphonium 2-Cyanopyrrolide. J. Phys. Chem. Lett. 2018, 9, 5213–5218. [Google Scholar] [CrossRef]
  89. Wu, C.; Senftle, T.P.; Schneider, W.F. First-principles-guided design of ionic liquids for CO2 capture. Phys. Chem. Chem. Phys. 2012, 14, 13163–13170. [Google Scholar] [CrossRef]
  90. Goel, H.; Windom, Z.W.; Jackson, A.A.; Rai, N. CO2 sorption in triethyl(butyl)phosphonium 2-cyanopyrrolide ionic liquid via first principles simulations. J. Mol. Liq. 2019, 292, 111323. [Google Scholar] [CrossRef]
  91. Firaha, D.S.; Hollóczki, O.; Kirchner, B. Computer-Aided Design of Ionic Liquids as CO2 Absorbents. Angew. Chem. Int. Ed. 2015, 54, 7805–7809. [Google Scholar] [CrossRef] [PubMed]
  92. Sheridan, Q.R.; Schneider, W.F.; Maginn, E.J. Role of Molecular Modeling in the Development of CO2–Reactive Ionic Liquids. Chem. Rev. 2018, 118, 5242–5260. [Google Scholar] [CrossRef]
  93. Oh, S.; Morales-Collazo, O.; Brennecke, J.F. Cation–Anion Interactions in 1-Ethyl-3-methylimidazolium-Based Ionic Liquids with Aprotic Heterocyclic Anions (AHAs). J. Phys. Chem. B 2019, 123, 8274–8284. [Google Scholar] [CrossRef]
  94. Hu, J.; Chen, L.; Shi, M.; Zhang, C. A quantum chemistry study for 1-ethyl-3-Methylimidazolium ion liquids with aprotic heterocyclic anions applied to carbon dioxide absorption. Fluid Phase Equilibria 2018, 459, 208–218. [Google Scholar] [CrossRef]
  95. Mei, K.; He, X.; Chen, K.; Zhou, X.; Li, H.; Wang, C. Highly Efficient CO2 Capture by Imidazolium Ionic Liquids through a Reduction in the Formation of the Carbene–CO2 Complex. Ind. Eng. Chem. Res. 2017, 56, 8066–8072. [Google Scholar] [CrossRef]
  96. Gohndrone, T.R.; Song, T.; DeSilva, M.A.; Brennecke, J.F. Quantification of Ylide Formation in Phosphonium-Based Ionic Liquids Reacted with CO2. J. Phys. Chem. B 2021, 125, 6649–6657. [Google Scholar] [CrossRef]
  97. Gurau, G.; Rodríguez, H.; Kelley, S.P.; Janiczek, P.; Kalb, R.S.; Rogers, R.D. Demonstration of Chemisorption of Carbon Dioxide in 1,3-Dialkylimidazolium Acetate Ionic Liquids. Angew. Chem. Int. Ed. 2011, 50, 12024–12026. [Google Scholar] [CrossRef]
  98. Zhang, X.; Xiong, W.; Tu, Z.; Peng, L.; Wu, Y.; Hu, X. Supported Ionic Liquid Membranes with Dual-Site Interaction Mechanism for Efficient Separation of CO2. ACS Sustain. Chem. Eng. 2019, 7, 10792–10799. [Google Scholar] [CrossRef]
  99. Hu, X.; Wang, J.; Mei, M.; Song, Z.; Cheng, H.; Chen, L.; Qi, Z. Transformation of CO2 incorporated in adducts of N-heterocyclic carbene into dialkyl carbonates under ambient conditions: An experimental and mechanistic study. Chem. Eng. J. 2021, 413, 127469. [Google Scholar] [CrossRef]
  100. Lee, T.B.; Oh, S.; Gohndrone, T.R.; Morales-Collazo, O.; Seo, S.; Brennecke, J.F.; Schneider, W.F. CO2 Chemistry of Phenolate-Based Ionic Liquids. J. Phys. Chem. B 2016, 120, 1509–1517. [Google Scholar] [CrossRef]
  101. Goodrich, B.F.; de la Fuente, J.C.; Gurkan, B.E.; Lopez, Z.K.; Price, E.A.; Huang, Y.; Brennecke, J.F. Effect of Water and Temperature on Absorption of CO2 by Amine-Functionalized Anion-Tethered Ionic Liquids. J. Phys. Chem. B 2011, 115, 9140–9150. [Google Scholar] [CrossRef]
  102. Saravanamurugan, S.; Kunov-Kruse, A.J.; Fehrmann, R.; Riisager, A. Amine-functionalized amino acid-based ionic liquids as efficient and high-capacity absorbents for CO2. ChemSusChem 2014, 7, 897–902. [Google Scholar] [CrossRef]
  103. Li, S.; Zhao, C.; Sun, C.; Shi, Y.; Li, W. Reaction Mechanism and Kinetics Study of CO2 Absorption into [C2OHmim][Lys]. Energy Fuels 2016, 30, 8535–8544. [Google Scholar] [CrossRef]
  104. Bhattacharyya, S.; Shah, F.U. Ether Functionalized Choline Tethered Amino Acid Ionic Liquids for Enhanced CO2 Capture. ACS Sustain. Chem. Eng. 2016, 4, 5441–5449. [Google Scholar] [CrossRef]
  105. Ma, J.-W.; Zhou, Z.; Zhang, F.; Fang, C.-G.; Wu, Y.-T.; Zhang, Z.-B.; Li, A.-M. Ditetraalkylammonium Amino Acid Ionic Liquids as CO2 Absorbents of High Capacity. Environ. Sci. Technol. 2011, 45, 10627–10633. [Google Scholar] [CrossRef]
  106. Zhang, Y.; Wu, Z.; Chen, S.; Yu, P.; Luo, Y. CO2 Capture by Imidazolate-Based Ionic Liquids: Effect of Functionalized Cation and Dication. Ind. Eng. Chem. Res. 2013, 52, 6069–6075. [Google Scholar] [CrossRef]
  107. Zhang, Y.; Li, T.; Wu, Z.; Yu, P.; Luo, Y. Synthesis and thermophysical properties of imidazolate-based ionic liquids: Influences of different cations and anions. J. Chem. Thermodyn. 2014, 74, 209–215. [Google Scholar] [CrossRef]
  108. Lin, W.; Cai, Z.; Lv, X.; Xiao, Q.; Chen, K.; Li, H.; Wang, C. Significantly Enhanced Carbon Dioxide Capture by Anion-Functionalized Liquid Pillar[5]arene through Multiple-Site Interactions. Ind. Eng. Chem. Res. 2019, 58, 16894–16900. [Google Scholar] [CrossRef]
  109. Singh, P.; Niederer, J.P.; Versteeg, G.F. Structure and activity relationships for amine-based CO2 absorbents-II. Chem. Eng. Res. Des. 2009, 87, 135–144. [Google Scholar] [CrossRef]
  110. Luo, X.; Guo, Y.; Ding, F.; Zhao, H.; Cui, G.; Li, H.; Wang, C. Significant Improvements in CO2 Capture by Pyridine-Containing Anion-Functionalized Ionic Liquids through Multiple-Site Cooperative Interactions. Angew. Chem. Int. Ed. 2014, 53, 7053–7057. [Google Scholar] [CrossRef]
  111. An, X.; Du, X.; Duan, D.; Shi, L.; Hao, X.; Lu, H.; Guan, G.; Peng, C. An absorption mechanism and polarity-induced viscosity model for CO2 capture using hydroxypyridine-based ionic liquids. Phys. Chem. Chem. Phys. 2016, 19, 1134–1142. [Google Scholar] [CrossRef] [PubMed]
  112. Luo, X.-Y.; Chen, X.-Y.; Qiu, R.-X.; Pei, B.-Y.; Wei, Y.; Hu, M.; Lin, J.-Q.; Zhang, J.-Y.; Luo, G.-G. Enhanced CO2 capture by reducing cation–anion interactions in hydroxyl-pyridine anion-based ionic liquids. Dalton Trans. 2019, 48, 2300–2307. [Google Scholar] [CrossRef] [PubMed]
  113. Chen, T.; Wu, X.; Xu, Y. Effects of the structure on physicochemical properties and CO2 absorption of hydroxypyridine anion-based protic ionic liquids. J. Mol. Liq. 2022, 362, 119743. [Google Scholar] [CrossRef]
  114. Cheng, J.; Li, Y.; Hu, L.; Zhou, J.; Cen, K. CO2 Adsorption Performance of Ionic Liquid [P66614][2-Op] Loaded onto Molecular Sieve MCM-41 Compared to Pure Ionic Liquid in Biohythane/Pure CO2 Atmospheres. Energy Fuels 2016, 30, 3251–3256. [Google Scholar] [CrossRef]
  115. Xue, C.; Feng, L.; Zhu, H.; Huang, R.; Wang, E.; Du, X.; Liu, G.; Hao, X.; Li, K. Pyridine-containing ionic liquids lowly loaded in large mesoporous silica and their rapid CO2 gas adsorption at low partial pressure. J. CO2 Util. 2019, 34, 282–292. [Google Scholar] [CrossRef]
  116. Breugst, M.; Tokuyasu, T.; Mayr, H. Nucleophilic Reactivities of Imide and Amide Anions. J. Org. Chem. 2010, 75, 5250–5258. [Google Scholar] [CrossRef] [PubMed]
  117. Huang, Y.; Cui, G.; Zhao, Y.; Wang, H.; Li, Z.; Dai, S.; Wang, J. Preorganization and Cooperation for Highly Efficient and Reversible Capture of Low-Concentration CO2 by Ionic Liquids. Angew. Chem. Int. Ed. 2017, 56, 13293–13297. [Google Scholar] [CrossRef]
  118. Huang, Y.; Cui, G.; Wang, H.; Li, Z.; Wang, J. Tuning ionic liquids with imide-based anions for highly efficient CO2 capture through enhanced cooperations. J. CO2 Util. 2018, 28, 299–305. [Google Scholar] [CrossRef]
  119. Huang, Y.; Cui, G.; Zhao, Y.; Wang, H.; Li, Z.; Dai, S.; Wang, J. Reply to the Correspondence on “Preorganization and Cooperation for Highly Efficient and Reversible Capture of Low-Concentration CO2 by Ionic Liquids”. Angew. Chem. 2018, 58, 386–389. [Google Scholar] [CrossRef]
  120. Huang, Y.; Cui, G.; Wang, H.; Li, Z.; Wang, J. Absorption and thermodynamic properties of CO2 by amido-containing anion-functionalized ionic liquids. RSC Adv. 2019, 9, 1882–1888. [Google Scholar] [CrossRef]
  121. Luo, X.Y.; Lv, X.Y.; Shi, G.L.; Meng, Q.; Li, H.R.; Wang, C.M. Designing amino-based ionic liquids for improved carbon capture: One amine binds two CO2. AIChE J. 2018, 65, 230–238. [Google Scholar] [CrossRef]
  122. Chen, F.-F.; Huang, K.; Zhou, Y.; Tian, Z.-Q.; Zhu, X.; Tao, D.-J.; Jiang, D.-E.; Dai, S. Multi-Molar Absorption of CO2 by the Activation of Carboxylate Groups in Amino Acid Ionic Liquids. Angew. Chem. Int. Ed. 2016, 55, 7166–7170. [Google Scholar] [CrossRef]
  123. Pan, M.; Zhao, Y.; Zeng, X.; Zou, J. Efficient Absorption of CO2 by Introduction of Intramolecular Hydrogen Bonding in Chiral Amino Acid Ionic Liquids. Energy Fuels 2018, 32, 6130–6135. [Google Scholar] [CrossRef]
  124. Chen, X.; Luo, X.; Li, J.; Qiu, R.; Lin, J. Cooperative CO2 absorption by amino acid-based ionic liquids with balanced dual sites. RSC Adv. 2020, 10, 7751–7757. [Google Scholar] [CrossRef] [PubMed]
  125. Wu, Y.; Zhao, Y.; Li, R.; Yu, B.; Chen, Y.; Liu, X.; Wu, C.; Luo, X.; Liu, Z. Tetrabutylphosphonium-Based Ionic Liquid Catalyzed CO2 Transformation at Ambient Conditions: A Case of Synthesis of α-Alkylidene Cyclic Carbonates. ACS Catal. 2017, 7, 6251–6255. [Google Scholar] [CrossRef]
  126. Pan, M.; Vijayaraghavan, R.; Zhou, F.; Kar, M.; Li, H.; Wang, C.; MacFarlane, D.R. Enhanced CO2 uptake by intramolecular proton transfer reactions in amino-functionalized pyridine-based ILs. Chem. Commun. 2017, 53, 5950–5953. [Google Scholar] [CrossRef]
  127. Zhang, Z.; Zhang, L.; He, L.; Yuan, W.-L.; Xu, D.; Tao, G.-H. Is it Always Chemical When Amino Groups Come Across CO2? Anion–Anion-Interaction-Induced Inhibition of Chemical Adsorption. J. Phys. Chem. B 2019, 123, 6536–6542. [Google Scholar] [CrossRef]
  128. Zhang, Y.; Zhang, S.; Lu, X.; Zhou, Q.; Fan, W.; Zhang, X. Dual Amino-Functionalised Phosphonium Ionic Liquids for CO2Capture. Chem. A Eur. J. 2009, 15, 3003–3011. [Google Scholar] [CrossRef]
  129. Xue, Z.; Zhang, Z.; Han, J.; Chen, Y.; Mu, T. Carbon dioxide capture by a dual amino ionic liquid with amino-functionalized imidazolium cation and taurine anion. Int. J. Greenh. Gas Control 2011, 5, 628–633. [Google Scholar] [CrossRef]
  130. Ren, J.; Wu, L.; Li, B.-G. Preparation and CO2 Sorption/Desorption of N-(3-Aminopropyl)Aminoethyl Tributylphosphonium Amino Acid Salt Ionic Liquids Supported into Porous Silica Particles. Ind. Eng. Chem. Res. 2012, 51, 7901–7909. [Google Scholar] [CrossRef]
  131. Peng, H.; Zhou, Y.; Liu, J.; Zhang, H.; Xia, C.; Zhou, X. Synthesis of novel amino-functionalized ionic liquids and their application in carbon dioxide capture. RSC Adv. 2013, 3, 6859–6864. [Google Scholar] [CrossRef]
  132. Zhou, Z.; Zhou, X.; Jing, G.; Lv, B. Evaluation of the Multi-amine Functionalized Ionic Liquid for Efficient Postcombustion CO2 Capture. Energy Fuels 2016, 30, 7489–7495. [Google Scholar] [CrossRef]
  133. Lv, B.; Jing, G.; Qian, Y.; Zhou, Z. An efficient absorbent of amine-based amino acid-functionalized ionic liquids for CO2 capture: High capacity and regeneration ability. Chem. Eng. J. 2016, 289, 212–218. [Google Scholar] [CrossRef]
  134. Jing, G.; Qian, Y.; Zhou, X.; Lv, B.; Zhou, Z. Designing and Screening of Multi-Amino-Functionalized Ionic Liquid Solution for CO2 Capture by Quantum Chemical Simulation. ACS Sustain. Chem. Eng. 2017, 6, 1182–1191. [Google Scholar] [CrossRef]
  135. Ye, C.-P.; Wang, R.-N.; Gao, X.; Li, W.-Y. CO2 Capture Performance of Supported Phosphonium Dual Amine-Functionalized Ionic Liquids@MCM-41. Energy Fuels 2020, 34, 14379–14387. [Google Scholar] [CrossRef]
  136. Ma, J.; Wang, Y.; Yang, X.; Zhu, M.; Wang, B. DFT Study on the Chemical Absorption Mechanism of CO2 in Diamino Protic Ionic Liquids. J. Phys. Chem. B 2021, 125, 1416–1428. [Google Scholar] [CrossRef] [PubMed]
  137. Qian, W.; Xu, Y.; Xie, B.; Ge, Y.; Shu, H. Alkanolamine-based dual functional ionic liquids with multidentate cation coordination and pyrazolide anion for highly efficient CO2 capture at relatively high temperature. Int. J. Greenh. Gas Control 2017, 56, 194–201. [Google Scholar] [CrossRef]
  138. Shu, H.; Xu, Y. Tuning the strength of cation coordination interactions of dual functional ionic liquids for improving CO2 capture performance. Int. J. Greenh. Gas Control 2020, 94, 102934. [Google Scholar] [CrossRef]
  139. Zema, Z.A.; Chen, T.; Shu, H.; Xu, Y. Tuning the CO2 absorption and physicochemical properties of K+ chelated dual functional ionic liquids by changing the structure of primary alkanolamine ligands. J. Mol. Liq. 2021, 344, 117983. [Google Scholar] [CrossRef]
Figure 1. A summary of different kinds of functionalized ILs for CO2 capture.
Figure 1. A summary of different kinds of functionalized ILs for CO2 capture.
Ijms 23 11401 g001
Figure 2. Structures of typical cations and anions used for designing functionalized ILs.
Figure 2. Structures of typical cations and anions used for designing functionalized ILs.
Ijms 23 11401 g002
Figure 3. Typical (a) direct methods and (b) indirect methods for synthesis of functionalized ILs for CO2 capture.
Figure 3. Typical (a) direct methods and (b) indirect methods for synthesis of functionalized ILs for CO2 capture.
Ijms 23 11401 g003
Figure 4. General mechanisms of amino-CO2 reactions for (a) primary or secondary amine, and (b) tertiary amine.
Figure 4. General mechanisms of amino-CO2 reactions for (a) primary or secondary amine, and (b) tertiary amine.
Ijms 23 11401 g004
Figure 5. Typical (a) single-site and (b) multiple-site mechanisms of non-amino anion-CO2 reactions.
Figure 5. Typical (a) single-site and (b) multiple-site mechanisms of non-amino anion-CO2 reactions.
Ijms 23 11401 g005
Table 1. Typical cation-functionalized ILs for CO2 capture.
Table 1. Typical cation-functionalized ILs for CO2 capture.
ILT (°C)P (bar)Mw (g mol–1) an CO2/n ILn CO2/kg IL bg CO2/g IL bRef.
[apbim][BF4]221269.1~0.5~1.86~0.08[22]
[aemim][BF4]301213.00.411.920.08[27]
[Bmim][Met]252287.40.421.460.06[30]
[Bmim][Pro]252253.30.321.260.06[30]
[PEG150MeBu2NLi][Tf2N]251562.50.661.170.05[31]
[PEG150MeTMGLi][Tf2N]251548.50.891.620.07[31]
DAIL c301249.21.054.21 (0.42)0.19 (0.02)[32]
[TETA][NO3] d151209.31.497.12 (2.85)0.31 (0.13)[34]
[Li(HDA)][Tf2N]401391.20.882.250.10[35]
[Li(DOBA)][Tf2N]401435.30.902.070.09[35]
[Li(TEPA)][Tf2N]800.1476.40.721.510.07[36]
[Li(TEPA)][Tf2N] e800.1476.41.954.09 (1.36)0.18 (0.06)[36]
a Molecular weight of pure IL. b Values shown in brackets are based on the total weight of IL + support or IL + solvent. c Mixed with H2O (mass ratio of IL: H2O is 10:90). d Mixed with H2O (mass ratio of IL: H2O is 40:60). e Mixed with [Li(TEG)][Tf2N] (mass ratio is 1:2).
Table 2. Typical anion-functionalized ILs for CO2 capture via single-site mechanisms.
Table 2. Typical anion-functionalized ILs for CO2 capture via single-site mechanisms.
ILT (°C)P (bar)Mw (g mol–1) an CO2/n ILn CO2/kg IL bg CO2/g IL bRef.
[P4444][Gly] c1333.5~0.6~1.80 (0.74)~0.08 (0.03)[38]
[P4444][Ala] c1347.5~0.67~1.93 (0.81)~0.08 (0.04)[38]
[P4444][β-Ala] c1347.5~0.6~1.73 (0.72)~0.08 (0.03)[38]
[P66614][Met]221632.1~0.9~1.42~0.06[39]
[P66614][Pro]221598.0~0.9~1.51~0.07[39]
[P4444][Butyrate]401346.50.41.150.05[51]
[MTBDH][TFE]231253.31.134.460.20[57]
[MTBDH][TFPA]231329.40.932.820.12[57]
[MTBDH]2[HFPD]231518.52.043.930.17[57]
[DBUH][TFE]r.t.1252.31.014.000.18[58]
[P66614][4-Me-PhO]301591.00.911.540.07[59]
[P66614][4-H-PhO]301577.00.851.470.06[59]
[P66614][4-Cl-PhO]301611.40.821.340.06[59]
[P66614][4-CF3-PhO]301645.00.610.950.04[59]
[P66614][4-NO2-PhO]301622.00.300.480.02[59]
[P66614][2,4,6-Cl-PhO]301680.30.070.100.0044[59]
[P66614][4-Kt-PhO]301619.01.041.680.07[60]
[P66614][4-EF-PhO]301649.01.031.590.07[60]
[P66614][4-CHO-PhO]301605.01.011.670.07[60]
[P66614][PPhO]201653.10.931.420.06[61]
[P66614][PCCPhO]201679.10.961.410.06[61]
[P4444][2-F-PhO]401370.50.671.810.08[62]
[P4444][3-F-PhO]401370.50.742.000.09[62]
[P4444][4-F-PhO]401370.50.842.270.10[62]
[Na(15-crown-5)][PhO]251336.40.752.230.10[64]
[Na(15-crown-5)][n-C3H7PhO]251378.40.661.740.08[64]
[Na(15-crown-5)][n-C8H17PhO]251448.60.501.110.05[64]
[MTBDH][Im]231221.31.034.650.20[57]
[(P2-Et)H][Im]231407.50.962.360.10[57]
[P66614][Im]231550.91.001.820.08[65]
[TMGH][Im]301183.31.005.460.24[67]
[DBUH][Im]251220.3~0.88~3.99~0.18[69]
[DBNH][Im]251192.30.84.160.18[70]
[DMAPAH][Im]221170.30.814.760.21[72]
[DMEDAH][Im]221156.20.774.930.22[73]
[P66614][Pyrz]231550.91.021.850.08[65]
[P66614][Tetz]231552.90.080.140.01[65]
[P66614][Triz]231551.90.951.720.08[65]
[P66614][4-CHO-Im]201578.91.242.140.09[60]
[P66614][4-Br-Im]201629.80.871.380.06[66]
[P66614][2-CN-Pyr]221575.00.91.570.07[82]
[P66614][2-CF3-Pyra]221618.90.91.450.06[82]
a Molecular weight of pure IL. b Values shown in brackets are based on the total weight of IL + support or IL + solvent. c Immobilization of IL on porous silica gel (SiO2) support (molar ratio of IL: SiO2 is 1:8).
Table 3. Typical anion-functionalized ILs for CO2 capture via multiple-site mechanisms.
Table 3. Typical anion-functionalized ILs for CO2 capture via multiple-site mechanisms.
ILT (°C)P (bar)Mw (g mol–1)n CO2/n ILn CO2/kg ILg CO2/g ILRef.
[P66614][Lys]221629.01.372.180.10[101]
[N66614][Lys]221612.12.13.430.15[102]
[C2OHmim][Lys]301272.31.686.170.27[103]
[N1,1,6,2O4][Lys]201375.61.624.310.19[104]
[N66614][His]221612.01.93.100.14[102]
[N66614][Asn]221598.02.03.340.15[102]
[N66614][Gln]221612.11.93.100.14[102]
[MTBDH]2[HFPD]231518.52.043.930.17[57]
[P66614]10[DCP5]5016019.55.520.920.04[108]
[P66614][2-Op]201578.01.582.730.12[110]
[P66614][4-Op]201578.01.492.580.11[110]
[P66614][3-Op]201578.01.382.390.11[110]
[P4442][2-Op]301325.51.404.30 0.19[112]
[N4442][2-Op]301308.51.244.02 0.18[112]
[Bmim][2-Op]301233.31.024.37 0.19[112]
[P4442OH][2-Op]301341.50.942.75 0.12[112]
[Ph-C8eim][2-Op]301379.51.694.45 0.20[112]
[Ph-C8eim][2-Op]201379.51.834.82 0.21[112]
[DBUH][2-Op]401247.3~0.86~3.48 ~0.15[113]
[TMGH][2-Op]401210.3~0.82~3.90 ~0.17[113]
[P4442][Suc]201329.51.875.68 0.25[117]
[P4442][Suc]200.1329.51.655.01 0.22[117]
[P4442][DAA]201331.51.253.77 0.17[117]
[P4442][DAA]200.1331.51.123.38 0.15[117]
[P4442][Cy-Suc]201383.62.215.76 0.25[118]
[P4442][Ph-Suc]201377.51.02.65 0.12[118]
[P66614][MA-Tetz]301581.91.131.94 0.09[121]
[P4442]2[IDA]401593.81.692.85 0.13[122]
[P4442]2[D-Ser]251565.81.061.87 0.08[123]
[P4442]2[L-Ser]251565.81.101.94 0.09[123]
[P66614]2[AA-Su]3011148.91.481.29 0.06[124]
[P66614]2[AA-Ac]3011112.81.971.77 0.08[124]
[P66614]2[AA-Im]3011120.91.551.38 0.06[124]
[P66614]2[AA-Ind]3011169.91.451.24 0.05[124]
[P4444]3[2,4-OPym-5-Ac]r.t.1931.41.461.57 0.07[125]
[P66614]2[Am-iPA]3011146.82.382.08 0.09[124]
Table 4. Typical cation-anion dual-functionalized ILs for CO2 capture.
Table 4. Typical cation-anion dual-functionalized ILs for CO2 capture.
ILT (°C)P (bar)Mw (g mol–1) an CO2/n ILn CO2/kg IL bg CO2/g IL bRef.
[aP4443][Gly] c1334.5~0.94~2.81 (1.15)~0.12 (0.05)[128]
[aP4443][Ala] c1348.5~0.92~2.64 (1.11)~0.12 (0.05)[128]
[aemmim][Tau]301264.4~0.9~3.40~0.15 [129]
[apaeP444][Lys] d251448.71.733.86 (1.93)0.17 (0.08)[130]
[apaeP444][Gly] d251377.61.293.42 (1.71)0.15 (0.08)[130]
[apaeP444][Ser] d251407.61.192.92 (1.46)0.13 (0.06)[130]
[apaeP444][Ala] d251391.61.142.91 (1.46)0.13 (0.06)[130]
[apaeP444][Asp] d251435.61.072.46 (1.23)0.11 (0.05)[130]
[apaeP444][His] d251457.61.012.21 (1.11)0.10 (0.05)[130]
[AEMP][Gly] e1218.31.506.87 (1.37)0.30 (0.06)[131]
[AEMP][Ala] e1232.31.576.76 (1.35)0.30 (0.06)[131]
[AEMP][Pro] e1258.41.545.96 (1.19)0.26 (0.05)[131]
[AEMP][Leu] e1274.41.475.36 (1.07)0.24 (0.05)[131]
[APmim][Gly] f301214.31.234.310.19[133]
[APmim][Lys] f301285.41.806.710.27[132]
[TETAH][Lys] f401292.42.598.860.39[134]
[DETAH][Lys] f401249.42.138.550.38[134]
[aP4443][2-Op]301354.51.574.440.20[135]
[aP4443][2-Np]301353.51.885.320.23[135]
[aP4443][Triz]301328.51.324.020.18[135]
[Na(MDEA)2][Pyrz]801328.40.752.280.10[137]
[K(DGA)2][Im]601316.41.374.330.19[138]
[K(AMP)2][Im]601284.41.194.180.18[139]
a Molecular weight of pure IL. b Values shown in brackets are based on the total weight of IL + support or IL + solvent. c Immobilization of the IL on porous SiO2 support (molar ratio of IL: SiO2 is 1:8), and absorption by SiO2 is subtracted. d Immobilization of the IL on porous SiO2 support (mass ratio of IL: SiO2 is 1:1). e Immobilization of the IL on porous SiO2 support (mass ratio of IL: SiO2 is 1:4). f Mixed with H2O (IL concentration: 0.5 mol L−1).
Table 5. The viscosities of typical functionalized ILs before and after CO2 capture.
Table 5. The viscosities of typical functionalized ILs before and after CO2 capture.
ILT (°C)P (bar)Viscosity of IL (cP)Viscosity of IL + CO2 (cP)Viscosity Increase (fold)Ref.
[APbim][BF4]221--Dramatic increase[22]
[P66614][Pyr]231245.4555.12.26[65]
[P66614][Oxa]231555.51145.82.06[65]
[P66614][PhO]231390.3645.41.65[65]
[P66614][Im]231810.4648.70.84[65]
[P66614][2-CN-Pyr]2513603701.03[82]
[P66614][3-CF3-Pyr]2512705001.85[82]
[P66614][Pro]201100017001.7[101]
[P66614][Met]25135033,00094[101]
[P66614][Lys]2011000280,000280[101]
[P66614][Tau]25167044,00066[101]
[P66614][2-Op]20157322734[110]
[P4442][Suc]200.19986290.63[117]
[P4442][DAA]200.16051470.24[117]
[P4442]2[IDA]40166.2961.614.5[122]
[aP4443][Gly]251713.9-Dramatic increase[128]
[Na(MDEA)2][Pyrz]5011310713.90.54[137]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, R.; Ke, Q.; Zhang, Z.; Zhou, B.; Cui, G.; Lu, H. Tuning Functionalized Ionic Liquids for CO2 Capture. Int. J. Mol. Sci. 2022, 23, 11401. https://doi.org/10.3390/ijms231911401

AMA Style

Zhang R, Ke Q, Zhang Z, Zhou B, Cui G, Lu H. Tuning Functionalized Ionic Liquids for CO2 Capture. International Journal of Molecular Sciences. 2022; 23(19):11401. https://doi.org/10.3390/ijms231911401

Chicago/Turabian Style

Zhang, Ruina, Quanli Ke, Zekai Zhang, Bing Zhou, Guokai Cui, and Hanfeng Lu. 2022. "Tuning Functionalized Ionic Liquids for CO2 Capture" International Journal of Molecular Sciences 23, no. 19: 11401. https://doi.org/10.3390/ijms231911401

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop