Abstract
As part of CO2 capture strategies, ionic liquid-based CO2 absorbents have gained attention for their tunable properties to lower the energy costs for CO2 capture. In this study, a series of borate-based nonamine functionalized ionic liquids (ILs), incorporated with magnesium acetylacetonate, were developed and investigated for its CO2 capture capability at moderate temperature under ambient pressure. Nuclear magnetic resonance and Fourier transform infrared spectroscopy confirmed the successful incorporation of acetylacetonate ligands into the fluorinated-lithium borate ionic liquids. Comprehensive analyses of the physical and thermochemical properties revealed that the synthesized ILs remain stable below 200 °C, with the borate structure and acetylacetonate ligands intact. The ILs functionalized with fluorinated alcohol and magnesium acetylacetonate enhance the CO2 uptake capacity by 55% in comparison with the original lithium borate ILs, suggesting the enhanced cooperative interactions responsible for improved CO2 capture performance. The carbon capture mechanism was identified to proceed via physical absorption, as evidenced by minimal changes in the characterization results and viscosity after CO2 absorption. The enthalpy of CO2 absorption (ΔH a) for the synthesized ILs were determined experimentally by using differential scanning calorimetry to be in the range from −12.4 kJ mol–1 to −18.9 kJ mol–1, which are much lower than that of conventional amine solutions (e.g., MEA: −82 kJ mol–1) and amine-based ILs ((e.g., [Bmim][Ac]: −45.8 kJ mol–1). These findings suggest that lithium borate-acetylacetonate ILs offer a promising approach for a CO2 capture system under ambient conditions.


1. Introduction
The emissions of increasing amounts of greenhouse gases have contributed to ongoing global warming issue, with carbon dioxide (CO2) being the most critical substance. According to the annual report of the National Oceanic and Atmospheric Administration (NOAA), the global average concentration of CO2 reached a record high of 422.8 ppm in 2024. If no substantial actions are implemented to reduce emission of CO2, it is projected that the global temperature could increase significantly beyond 2 °C by the end of the century. − Carbon dioxide capture, utilization, and storage (CCUS) technologies have been developed to have the potential to reduce the CO2 emissions by 20%. , Amine-based solvents are the most commonly used in CCUS postcombustion CO2 capture method due to their high reactivity and thermal stability. , However, the high energy requirement by the solvent regeneration, equipment corrosion issues, solvent degradation, environmental hazards, and high capital costs limit the applications. − Therefore, this has prompted the development of alternative absorbents with lower energy requirements to address these limitations.
CO2 capture through the usage of ionic liquids (ILs) has gained increasing attentions, with numerous publications investigating CO2 absorption using various types of ILs. − ILs are organic salts that composed entirely of cations and anions that exist in the liquid state at or near room temperature, with melting point below 100 °C. − Their low melting points arise from the use of bulky, asymmetric cations and weakly coordinating anions, which disrupt crystal packing and allow the salt to remain in liquid. , Compared to conventional aqueous amine-based solvents such as monoethanolamine (MEA), which rely on chemisorption through strong acid–base reactions, ILs offer several distinct advantages as absorbents of CO2, including low saturated vapor pressure, a wide temperature range, excellent thermal and chemical stability, and high tunability with the ease of changing cation and anion combinations, catering to the specific requirements by the applications. − Unlike MEA or the solid adsorbents, such as amine-functionalized mesoporous silica and CaO, , which requires high regeneration energy due to strong CO2 binding, many ILs promote weaker, reversible interactions with CO2. The absorption enthalpy of CO2 with ILs can vary widely, ranging from weak to strong interactions depending on their chemical structure and absorption mechanisms, which may involve physical, chemical, or a combination of both. These advantages make ILs promising alternatives for efficient CO2 capture.
ILs can be categorized into two different types: (1) amino-functionalized and (2) nonamine-based ILs. The ILs consist of different combinations of cations and anions, with imidazolium-based being the most extensively studied as cations, followed by ammonium, phosphonium, piperidinium, and pyrrolidinium. − These cations are paired with a wide variety of anions, such as fluorinated-based (i.e., [BF4]−, [PF6]− and [TfO]−), halides (i.e., [Cl]−, [Br]− and [I]−), and amino acid-based anions (i.e., [Gly], [Lys] and [Ala]). Orhan et al. investigated the effects of various cation–anion combinations and found that the imidazolium-based IL [Bmim][Ac] exhibited the highest CO2 equilibrium loading of 0.89 mol CO2 per mol amine, followed by [Bmim][Tf2N] with 0.77 mol CO2 per mol amine at 2 bar.
To enhance the CO2 absorption capacity at mild conditions, amine functional groups have been incorporated into IL structures. ,, The amine groups can be integrated into the cation (cation-functionalized IL), the anion (anion-functionalized IL), or both (dual-functionalized IL) in the ILs. This strategic modification introduces active sites that enhance CO2 interaction, selectivity, and absorption kinetics while maintaining recyclability. − The anion of an IL has been identified as a critical factor in determining CO2 absorption performance, as CO2 solubility is strongly influenced by the nature and reactivity of the anion. − While cation-functionalized ILs can chemically bind CO2, they often exhibit a reaction stoichiometry of 1:2 for CO2 absorption which may limit process efficiency and increase solvent consumption. In contrast, anion functionalized with amine groups has demonstrated improved stoichiometry to 1:1 CO2: IL molar ratio. Gurkan et al. showed that ILs based on trihexyl(tetradecyl)phosphonium [P66614] as the cation paired with amino acid-derived anions ([Gly]−, [Lys]− and [Met]−) have achieved this stoichiometry ratio, with a CO2 uptake of 1.26, 1.60, 0.90 mol CO2 per mol IL respectively. ,, These results presented the dominant role of the anion in enhancing the CO2 capture capacity. However, the chemisorption of CO2 by amine-containing ILs typically leads to a significant increase in viscosity. For instance, the viscosity of [P66614][Gly] increased from 360 to 17000 cP upon CO2 uptake. The drastic rise in viscosity is attributed to the complexed formation of a hydrogen-bonded network between the IL and CO2. Despite the remarkable CO2 absorption performance of amine-functionalized ILs, the desorption process requires substantial energy due to strong chemical bonding, byproduct formation, and high viscosity.
To overcome these issues, recent studies have explored whether nonamine functionalized ILs can achieve efficient CO2 capture while avoiding the formation of hydrogen-bonded networks. Imidazolium-based ILs such as [Bmim][PF6], and [Cnmim][Tf2N] primarily undergo physical absorption. Blanchard et al. reported [Bmim][PF6] exhibited 0.19 mol CO2 per mol IL at 1 bar, with physisorption interaction between the IL and CO2. The exploration of high-pressure conditions to improve CO2 solubility based on Henry’s Law were further investigated by studying the phase behavior of CO2 in [Bmim][PF6] under varying pressure conditions. ,,− Their findings indicated that the IL exhibited a CO2 solubility of 0.231 mol of CO2 per mol of IL at 15.17 bar, which significantly increased to 0.729 mol of CO2 per mol of IL at 95.67 bar. Similarly, Shin et al. found that [C2mim][Tf2N] absorbed 0.24 mol CO2 per mol IL at 1 bar, compared to 0.30 mol CO2 per mol IL for [C8mim][Tf2N], indicating that longer alkyl chains enhance uptake. Both ILs exhibited sharp increases at elevated pressure, reaching 0.761 mol CO2 per mol IL and 0.845 mol CO2 per mol IL, respectively.
In addition to physical absorption, some nonamine ILs undergo chemical absorption. Dai et al. reported that superbase-derived protic IL, [MTBD][Im], exhibited rapid and reversible CO2 uptake with equimolar absorption capacity. In a subsequent study of anion reactivity, the phosphonium hydroxide-derived ILs were shown to allow fine-tuning of CO2 absorption efficiency through the selection of weak proton donors with different pK a values. For instance, the absorption capacities of [P66614][Pyrr], [P66614][Im], [P66614][Ind], [P66614][Triz], [P66614][BenTriz], and [P66614][Tetz] were 1.02, 1.00, 0.98, 0.95, 0.17, and 0.08 mol of CO2 per mole of IL, respectively. Importantly, these ILs maintained low viscosity during absorption, supporting rapid mass transfer and offering the potential for scalable CO2 capture applications.
Thermodynamic comparisons further illustrate the key differences. Amine-functinalized IL, such as [P66614][Met], has an enthalpy of −64 kJ mol–1, whereas nonamine ILs undergoing chemical absorption ([P66614][Pyrr]: −91 kJ mol–1 ;[P66614][Im]: −89.9 kJ mol–1) display even stronger interactions. In contrast, [C2mim][Tf2N], a representative of nonamine ILs that undergoes physical absorption, is around −7.81 kJ mol–1. Physical absorption is governed mainly by van der Waals forces and allows for easier desorption due to the relatively low energy needed to break these interactions. Consequently, these findings suggest that nonamine functionalized ILs has emerged as an alternative energy to address the issues of viscosity and energy requirements.
In this work, the focus is the development of a new type of nonamine functionalized ILs and testing for their potential capability for low energy CO2 capture. Harada et al. first proposed that lithium sodium borate (Li0.5Na0.5)3BO3 molten salt can achieve a high CO2 capacity of 7.3 mmol g–1 in 1 min of reaction with CO2 because of the fast reaction kinetics through enhanced ion mobility and mass transport, which allows CO2 molecules to diffuse throughout the material. Meanwhile, Guzman Gonzalez et al. synthesized a series of lithium-based ILs for the applications as single-component electrolytes in lithium batteries. This study focuses on the synthesis and characterization of lithium borate-based ILs, including two novel nonamines functionalized ILs, alongside comparative evaluation with the reference ILs. The CO2 absorption capacity was evaluated at ambient temperature to investigate the feasibility of the material. In addition, we examined key material properties, including viscosity and thermal stability, as they are critical factors influencing the CO2 absorption performance. With the developed ILs, the energy consumption could be significantly reduced, with operating absorption temperature at room temperature, unlike molten salts that require high temperatures for absorption.
2. Experimental Section
2.1. Materials
Hexane (≥96%), anhydrous magnesium sulfate (MgSO4, ≥ 98%), 2,2,2-trifluoroethanol (≥98%), and tetrahydrofuran (THF, ≥ 99.5%) were purchased from FUJIFILM Wako Pure Chemical Co., Ltd., Japan. Triethylene glycol monomethyl ether (≥98%), n-butyl lithium (ca. 15% in hexane, ca. 1.6 mol/L), and diethyl ether (≥99.5%) were purchased from Tokyo Chemical Industry (TCI) Co., Ltd., Japan. Borane tetrahydrofuran complex solution (BH3-THF, 1.0 M in THF) and magnesium acetylacetonate (Mg(acac)2, ≥ 98%) were purchased from Sigma-Aldrich. All of the above chemicals except hexane were used without further purification. Hexane was dried with anhydrous MgSO4 before synthesis. The experiments utilized deionized water. N2 (≥99.995%) and CO2 (≥99.5%) were supplied by Suzuki Shokan Japan.
2.2. Synthesis Procedure
Four different types of lithium borate-based ILs were prepared with variation with the inclusion of fluorinated alcohol (FA) or Mg(acac)2. The chemical structures of the ILs synthesized in this study are illustrated as shown in Figure S1, with the reference lithium borate IL labeled as LBIL, lithium borate IL with FA labeled as LBIL-FA, lithium borate IL with only Mg(acac)2 labeled as LBIL-Ac, and lithium borate IL with both FA and Mg(acac)2 labeled as LBIL-FA-Ac. The reference LBIL and LBIL-FA were previously reported by Guzmán-González et al.
2.2.1. Synthesis of LBILs
A previously reported methodology was used for the synthesis of the ILs. Five ml of dried hexane and triethylene glycol monomethyl ether (10 mmol, 3.6 mL) were added into a 100 mL three-neck flask. The system was purged with a N2 flow during the reaction. The solution was stirred and subsequently cooled in a liquid N2 bath, preventing solidification of the system. BH3-THF (5 mmol, 5 mL) was carefully added dropwise as the reaction is highly exothermic. The reaction mixture was then slowly warmed to room temperature and stirred for 30 min. Afterward, the reaction mixture was cooled again in a liquid N2 bath. Subsequently, 2,2,2-trifluoroethanol (5 mmol, 3.6 mL) as FA (RF–OH), was added dropwise, and the system was warmed to room temperature again for 1 h. After the reaction, it was cooled again in a liquid N2 bath and carefully added 5 mmol of n-butyl lithium (RC-Li). The reaction mixture was stirred at room temperature for another 2 h. Finally, the mixture was washed with cold diethyl ether and placed in a rotary evaporator at 40 °C for 24 h under vacuum atmosphere. It was further dried in an oil bath at 50 °C for 24 h under vacuum. The obtained final product will be present as a transparent liquid. The reference ionic liquid LBIL was synthesized using the same synthesis method without the addition step of FA. The synthetic route is shown in Figure .
1.
Synthetic route for reference LBIL and LBIL-FA.
2.2.2. Synthesis of LBILs with Mg(acac)2
The same synthesis method will be used for synthesizing LBIL-Ac and the LBIL-FA-Ac compound. The addition of Mg(acac)2 was incorporated during the IL synthesis, after the reaction with the organolithium compound. 0.15 g of Mg(acac)2 powder was stirred and dissolved in 5 mL of THF. Afterward, 1 mL of Mg(acac)2 in THF solution was slowly added into the reaction mixture and mixed for 2 h before proceeding for washing and drying steps. Yellowish liquid samples were obtained as the final products. The synthetic route for addition of Mg(acac)2 is shown in Figure .
2.

Synthetic route for references LBIL-Ac and LBIL-FA-Ac.
2.3. Material Characterization and Physicochemical Properties of Ionic Liquid
2.3.1. Characterization
The series of lithium-borate ILs were characterized using 1H and 13C Nuclear Magnetic Resonance (NMR) (Avance III 400 MHz, Bruker) and Fourier transform infrared spectroscopy (FT-IR) (Nicolet iS20, Thermo Fischer Scientific). Chloroform-d (≥99.6%, TCI) was used as the NMR solvent for all the samples tested. Infrared spectra were collected using the KBr pellet method with data collected over 64 scans at wavelengths ranging from 400 to 4000 cm–1 and a spectral resolution of 4 cm–1.
2.3.2. Viscosity
The viscosity of the samples was analyzed by using a modular compact rheometer (MCR302e, Anton Paar). The viscosity measurements were conducted at 20 °C, and the shear rates were varied from 1 to 1000 s–1 to capture the viscosity behavior.
2.3.3. Thermal Properties
The thermal onset temperature (T onset) and decomposition temperature (T decomp) were measured using a thermogravimetric analyzer (TGA-550, TA Instruments). Average weights of 5–10 mg for every sample were placed on a platinum sample pan for testing with a gas flow rate of 100 mL min–1. The sample was first pretreated to remove any residual solvents and moisture present at 70 °C for 2 h under 100% N2 gas before starting the actual analysis. For determination of T onset and T decomp, the system was ramped from 30 to 350 °C at a ramping rate of 5 °C/min under N2 atmosphere. The experimental temperatures of thermal decomposition are presented in terms of weight loss (%) and temperature (°C). T onset was obtained from the intersection of the baseline weight and the tangent of the weight dependence on the temperature curve as decomposition occurs.
Absorption enthalpy (ΔH a) were determined through differential scanning calorimetry (DSC) using a simultaneous DSC/thermogravimetric analyzer (SDT 650, TA Instruments). One to two droplets of IL sample were placed on a platinum sample pan for the testing with a gas flow rate of 100 mL min–1, with the same pretreatment procedure. CO2 was performed at 30 °C under 100% CO2 for 60 min.
2.3.4. CO2 Capture
CO2 capture performance was evaluated through thermogravimetric measurements, which were conducted on a platinum sample pan using a thermogravimetric analyzer (TGA-550, TA Instruments) with a gas flow rate of 100 mL min–1. Approximately 10 mg of sample underwent the same pretreatment procedure as the analysis of thermal properties. For CO2 uptake measurements, the sample was exposed to 100% CO2, with the absorption temperature performed at 30 °C for 60 min. Desorption of the residual was followed by an additional 60 min at 55 °C under 100 N2 flow.
3. Results and Discussion
3.1. Role of Magnesium Acetylacetonate in CO2 Absorption
Magnesium acetylacetonate was chosen to be incorporated into the lithium borate-based ILs due to its possible ability to enhance CO2 absorption through increased active sites to capture CO2 uptake and improved chemical stability. Acetylacetonate is a bidentate ligand derived from acetyl acetone and forms stable octahedral complexes through both oxygen atoms that create a six-membered chelate ring that supports the stability and reactivity of the metal complex. One of the key characteristics of acetylacetonate is keto–enol tautomerization. Tautomerization involves the transfer of a hydrogen atom from one molecular site to another, leading to changes in the molecular skeleton, electron density distribution, and chemical properties. − The enolate form of acetylacetonate has negatively charged oxygen and nucleophilic carbon, both of which can interact with the electrophilic carbon in CO2, forming a new C–C bond. , The ligand exists in equilibrium between its keto and enol forms, with the enol form being stabilized by intramolecular hydrogen bonding. The nucleophilic hydroxyl group (−OH) and the adjacent electron-rich CC bond make the enol form particularly reactive toward CO2, enabling hydrogen bonding interactions between the −OH group and CO2. , With the cooperative interaction between the magnesium center and acetylacetonate ligand, Mg2+ acts as a Lewis acid that accepts the electrons and coordinates with CO2, where CO2 donates electrons from the oxygen atoms or the π-electrons in the system. Meanwhile, the acetylacetonate ligands function as Lewis base, contributing to a CO2 “push-pull” mechanism of electron density, facilitating electron redistribution and promoting efficient physisorption.
The hypothesized reaction mechanism for the incorporation of acetylacetonate ligands involves substitution of the n-butyl group on the boron atom with the acetylacetonate ligand, forming a new boron–oxygen bond. Depending on steric and electronic factors, the ligand may coordinate to boron in a monodentate form via one oxygen atom. The main benefit of incorporation of magnesium acetylacetonate is that not only increases the number of active sites for CO2 capture but also enhances the material’s stability and tunability, as the ligand structure can be modified by varying the cation and anion. ,
3.2. Characterization
Based on the NMR results, significant changes in the chemical shifts were observed after incorporation with magnesium acetylacetonate (Figures S2–S9, Table ). A new chemical shift appeared at 2.02–2.05 ppm, and 2.66–2.73 ppm, which are attributable to the acetylacetonate ligand. The 2.02–2.05 ppm region is consistent with the methyl group attached to OCH3 of the acetylacetonate ligand. The peak signals at 2.66–2.73 ppm correspond to the protons to the carbonyl group (CO) of the conjugated acetylacetonate ligand. This downfield shift arises from the deshielding effect of the adjacent CO group, which withdraws electron density from the alpha-protons, causing their signal to appear at a higher chemical shift. Additionally, the chemical shifts of the BOCH2CF3 were observed to have an upfield shift from 3.81 to 3.73 ppm upon magnesium acetylacetonate incorporation. This change can be explained by a coordination induced redistribution of electron density, which was reported by Butera et al. on how the metal ion coordination affects the H NMR chemical shift values. The electron distribution at borate-ether oxygen atoms alters the polarization environment experienced by the CH2 protons, which resulted in slight changes in chemical shift. In the 13C NMR spectrum, the disappearance of the signals between 10 and 20 ppm after incorporation indicates that the alkyl chain of the ionic liquid is no longer present in the product.
1. Chemical Shift Assignment of the 1H NMR Spectra of LBILs.
| LBIL/LBIL-FA | δ (ppm) | LBIL-Ac/LBIL-FA-Ac | δ (ppm) |
|---|---|---|---|
| Terminal-CH3 | 0.918–0.955 | Terminal-CH3 | 0.882–0.956 |
| CH2 | 1.256–1.556 | CH2 | 1.256–1.559 |
| OCH3 (triethylene glycolmethyl ether) | 3.388 | OCH3 (acetylacetonate) | 2.020–2.050 |
| (OCH2CH2)3O | 3.550–3.725 | α-CH2 | 2.660–2.730 |
| BOCH2CF3 | 3.809–3.833 | OCH3 (triethylene glycolmethyl ether | 3.388 |
| Overlapping (OCH2CH2)3O and BOCH2CF3 | 3.560–3.720 |
The FTIR spectrum of the synthesized ionic liquids confirms their chemical structure, with key vibrational bands at 2880, 1456, 1350, 1283, 1104 cm–1 assigned to the triethylene glycol monomethyl ether chain and borate core − (Figure , Table ). A broad feature near 1653 to 1658 cm–1 was present in all the samples and is consistent with the residual water rather than the structural vibrations of the ILs reported in other literature. , Both LBIL and LBIL-FA samples exhibited strong peak signals at 1283 cm–1, which is attributed primarily to C–O stretching vibrations of the borate-ether framework. After the incorporation of Mg(acac)2, the intensity of this band significantly reduced in LBIL-Ac and LBIL-FA-Ac, consistent with coordination of Mg2+ to the oxygen atoms of the ether chains, which reduces the vibration activity of C–O groups. Furthermore, an additional peak appeared at 1616 cm–1 in LBIL-Ac and LBIL-FA-Ac samples, likely due to the acetylacetonate ligand. The stability properties of the metal complex arises from the conjugated π system of acetylacetonate ligand. The absence of distinct CO stretching frequency, typically found around 1700 cm–1 in the FTIR spectra is a key indicator of delocalized structure. , The presence of a peak at 1616 cm–1 is characteristic of a chelated acetylacetonate ligand and does not correspond to a discrete CO or CC stretching vibration. − Instead, its position and intensity are consistent with a coupled vibrational mode of the chelated acetylacetonate ligand involving the conjugated O–C–C–C–O system. The CC bond within the acetylacetonate ligand represents the delocalization of π electrons across the O–C–C–C–O chain.
3.
FT-IR spectra of LBILs under N2 at 25 °C.
2. FT-IR Peak Assignment of LBILs.
| FT-IR peaks (cm –1 ) | Assignments | FT-IR peaks (cm –1 ) | Assignments |
|---|---|---|---|
| 3396 | v(O–H) | 1283 | v(C–O) |
| 2880 | v(C–H) | 1248 | v(C–F), v(C–O) |
| 1653–1658 | δ(H–O–H), v as(CO3 2–) | 1200 | v(B–O), v(C–O) |
| 1616 | v(CC), v(CO) | 1103–1105 | v(B–O), v(C–O–C) |
| 1455–1457 | δ(CH2), δ(CH3) | 1028 | v(C–O) |
| 1350 | δ(CH3), v(B–O) |
3.3. Viscosity
Viscosity is one of the important parameters for ILs as it significantly influences the CO2 mass transfer efficiency and the energy required for transport and regeneration. Lower viscosity facilitates faster mass transfer, allowing the CO2 molecules to diffuse more rapidly within the liquid, thereby enhancing absorption rates. In contrast, high viscosity ILs hinder CO2 diffusion, creating slower absorption kinetics and a lower overall efficiency. The viscosities of the LBILs samples were measured across various shear rates, as shown in Figure . The average viscosity of LBIL and LBIL-FA was 47.37 mPa.s and 33.62 mPa.s, respectively, indicating that the inclusion of FA has reduced the viscosity. Yang et al. has reported 2,2,2-trifluoroethanol can reduce the viscosity of ILs. Furthermore, further reduction of the viscosity was observed in LBIL-FA-Ac (12.39 mPa.s) with the inclusion of both FA and Mg(acac)2. The acetylacetonate ligand from Mg(acac)2 has replaced the hydrocarbon groups present in the IL, which disrupted the hydrophobic interactions and made the structure less bullky. The effect is conceptually similar to how Pd(acac)2 reduces heavy oil viscosity by breaking down the large hydrocarbon molecules. Nonetheless, these ILs samples exhibited a low viscosity comparable to conventional solvents and other ionic liquids reported in the literature (Figure S10).
4.

Viscosity of LBILs samples across the shear rate from 1 to 1000 s–1.
3.4. Thermal Properties
Determining the decomposition temperature of ILs is essential to establishing suitable operating conditions for CO2 absorption. Typically, the thermal stability of ILs is characterized by the onset temperature, T onset, which is determined from the intersection of the baseline weight and the tangent to the weight-loss curve during decomposition. , However, ILs were found to degrade at a temperature significantly lower than the T onset, making it crucial to assess whether the material can withstand the temperature required for CO2 desorption and regeneration. , If decomposition occurs at or below the operating temperature, then the material would be unsuitable for prolonged use. The TGA curve observed two distinct stages of weight loss. The first stage shows gradual weight reduction between 105 and 220 °C, attributed to evaporation of residual solvent and moisture, together with the onset of ether chain and acetylacetonate ligand degradation. This is consistent with reports showing that the ether chains and the chelate rings of acetylacetonate degrading around 200 °C. , Farooq et al. also have synthesized a series of acetylacetonate-based magnetic ionic liquids (MILs) and tested that the thermal stabilities of these MILs were at the lower range of 130 to 215 °C. The second stage corresponds to the major decomposition step, where there is complete breakdown of the ionic liquid structure. T onset was determined to be around 240 °C, while the decomposition temperature occurred at 275 °C (Figure a).
5.

Decomposition results for LBIL-FA-Ac measured from a starting temperature of 25 to 350 °C at a heating rate of 5 °C/min under 100% N2, with determination of (a) T onset and T decomp; (b) FT-IR spectra of LBIL-FA-Ac for a thermal analysis test at a temperature from 25 to 350 °C under 100% N2 flow.
To investigate the structural changes during heating, FT-IR was performed under N2 at temperature varied from 25 to 300 °C. The IR spectra of LBIL-FA-Ac revealed a gradual decrease in peak intensity as temperature increased (Figure b). The intensity of the characteristic CC bond associated with Mg(acac)2 at 1610 cm–1 was gradually decreased as temperature increased and completely disappeared above 200 °C. The peak at 1100 cm–1, associated with ether chains and acetylacetonate, also decreases with increasing temperature, likely possible ether chain breakdown and acetylacetonate ligand decomposition. These FT-IR results align well with the staged decomposition observed in TGA, suggesting that the IL maintains structural integrity only at temperatures below 200 °C, where the borate complex and acetylacetonate remain stable.
3.5. Absorption Enthalpy
The enthalpy of the CO2 absorption provides insights into the nature of interactions between the IL and CO2. DSC revealed absorption enthalpy values in the range −12.4 kJ mol–1 to −18.9 kJ mol–1 for the LBILs synthesized in this study, with the highest value of −18.9 kJ mol–1 for LBIL-FA-Ac (Figure S11). The values are comparable to those reported for other ILs, such as [Bmim][BF4] (−15.9 kJ mol–1), as shown in Figure S12. ,
3.6. CO2 Capture Performance
3.6.1. Effects of Mg(acac)2 and Fluorinated Alcohols
The CO2 absorption–desorption profiles of the LBILs samples reveal clear differences in the working capacity (Figure ). The results demonstrated that LBIL-FA-Ac containing both FA and Mg(acac)2, exhibited the highest absorption capacity of 0.9375 mmol g–1 in 60 min of absorption time. This demonstrates that the combination of FA and Mg(acac)2 enhances the level of CO2 uptake. Comparing LBIL and LBIL-FA-Ac, the combination resulted in a 55% increase in the CO2 absorption capacity. The introduction of Mg(acac)2 significantly enhanced the CO2 uptake capacity. According to the findings discovered by Sampson et al, Mg2+ plays a critical role in stabilizing electron-rich [CO2]− intermediates and facilitating C–O bond cleavage. This behavior is attributed to the metal–ligand cooperativity, where intramolecular interactions between a metal center and a nearby ligand site work synergistically for CO2 activation. , Such systems are described as offering a “push-pull” mechanism, with Lewis acidic metal centers stabilizing the negative charge on CO2. , In addition, it was reported that substituents that decrease ligand basicity favors CO2 capture. Therefore, the Mg2+ center may act as a Lewis acid, providing additional coordination sites to interact with the electron-rich oxygen of CO2 within the IL. This coordination likely promotes physisorption within the IL matrix by stabilizing the presence of CO2 through donor–acceptor interactions. Similarly, the effect of FA on the absorption of CO2 can also be assessed by comparing LBIL with LBIL-FA and LBIL-Ac with LBIL-FA-Ac. The presence of FA moderately increased the rate of CO2 uptake. Hazra et al. reported that materials with interconnected discrete pores embedded with −CF3 functionality exhibit strong CO2-MOF interaction. The surface functionalized with −CF3 groups created hydrogen bonding and Lewis acid–base interactions with the fluorine atom of −CF3 group. Based on these reports, the moderate enhancement of uptakes is likely due to the presence of the trifluoromethyl (−CF3) group, which can help with the facilitation of CO2 dissolution and interactions. During the desorption phase, all samples exhibited a decline in CO2 uptake, with LBIL-FA-Ac also demonstrating the fastest release. The working capacity of LBIL-FA-Ac remained the highest among the samples, with minimal residual, while LBIL still retained some absorbed CO2 after 60 min of desorption. Additionally, the mass loss during the desorption for LBIL-FA-Ac was approximately only 1.5%, suggesting good thermal stability of the IL.
6.

CO2 absorption–desorption results for the LBIL samples. Absorption test was performed at 30 °C under 100% CO2 and desorption was performed at 55 °C under 100% N2.
The CO2 uptake capacity achieved for LBIL-FA-Ac (0.94 mmol g–1) is higher than that of several nonamine functionalized ILs reported in other literature (Figure S13), such as [C2min][Tf2N] (0.61 mmol g–1) and [Bmim][BF4] (0.42 mmol g–1). , Notably, this value is comparable to the CO2 uptake of these ILs measured under higher pressures, with [Bmin][Tf2N] (0.90 mmol g–1) and [Bmim][BF4] (0.80 mmol g–1), demonstrating that LBIL-FA-Ac achieves superior CO2 capture performance under mild ambient conditions.
3.6.2. Effects of Concentration of Mg(acac)2
To examine the effect of concentration of Mg(acac)2 in LBIL-FA-Ac samples on CO2 uptake capacity, the concentration was varied from 0.03 to 0.15 g of Mg(acac)2 (Figure ). LBILs samples of 0.03 and 0.05 g Mg(acac)2 were observed to reach a maximum working capacity of approximately 0.90 mmol g–1 at 60 min absorption time. However, further increasing the Mg(acac)2 concentration to 0.15 g resulted in a decline in uptake capacity. Goodwin et al. found that high concentrations of metal salts in ionic liquids can undergo ion aggregation and network formation beyond a critical gelation threshold. In such regimes, the formation of percolating ion networks and extended reduces the availability of free, mobile metal centers capable of coordination. , Since Mg2+ plays a central role in providing the coordination sites for CO2, oversaturation of Mg(acac)2 may disrupt the IL structure with fewer Lewis acid sites, making them inaccessible for CO2 interactions environments, which could explain the observed decline in CO2 uptake beyond a concentration of 0.05 g Mg(acac)2.
7.

CO2 absorption results for LBIL-FA-Ac samples with various concentrations of Mg(acac)2 per mL of THF from 0.03 to 0.15 g. Absorption test was performed at 30 °C under 100% CO2.
3.6.3. Pressure Dependence of CO2 Capture Performance
The CO2 capture performance of LBIL-FA-Ac was evaluated at pressures ranging from 0.05 to 1 bar to assess the uptake capacity at reduced pressures (Figure ). The CO2 uptake capacity exhibits a clear pressure-dependent trend. The absorption capacity increases as the partial pressure of CO2 rises, from 0.26 mmol g–1 at 0.05 bar to 0.9375 mmol g–1 at 1 bar. Notably, at 0.15 bar, which is typically the CO2 partial pressure in postcombustion flue gas, was able to achieve CO2 uptake capacity of 0.60 mmol g–1. Comparing with other imidazolium-based ILs reported in the literature, such as [Bmim][Tf2N] (0.22 mmol g–1 at 4.2 bar), [Bmim][Pf6] (0.39 mmol g–1 at 5.29 bar), [Emim][Tf2N] (0.13 mmol g–1 at 2.13 bar), the LBIL-FA-Ac IL demonstrates significantly higher CO2 uptake capacity at muchs lower pressures. , This superior low-pressure performance highlights its potential for industrial applications in which CO2 is present at partial pressures well below ambient conditions.
8.

CO2 absorption test at various pressures ranging from 0.05 to 1 bar with absorption temperature at 30 °C for 60 min.
3.6.4. Performance of CO2 Cycle Stability
The CO2 uptake capacity of the material was evaluated over five consecutive cycles to evaluate its reusability (Figure ). The results show a steady decline with each subsequent cycle, reaching 0.389 mmol g–1 by the fifth cycle. This represents almost 60% loss in capacity from the initial value. The significant decrease in the CO2 uptake capacity indicates that the material has poor reusability and long-term stability under the tested conditions. This suggests that the material might undergo physical degradation of the material or blockage of active sites during the repeated absorption–desorption cycles. The reduced reusability of LBIL-FA-Ac is likely attributed to minor evaporation losses during the CO2 cycle test. Although pure ionic liquids are generally nonvolatile, functionalized or blended IL systems can exhibit gradual mass loss due to evaporation of volatile cocomponents or loosely bound ligands. During the desorption process at 55 °C, slight evaporation or partial loss of Mg(acac)2 associated species may occur. Since Mg(acac)2 plays a crucial role in CO2 uptake capacity, this could be the observed decline in CO2 uptake upon regeneration. Similar behavior was observed and reported by Taylor et al., who discussed the effect of evaporation of the volatile components in the IL-based systems, which has affected its CO2 uptake capacity. Their study demonstrated that the IL-glycol mixtures suffer from significant volatility and encapsulation of the mixtures have mitigated evaporation losses by around 40%, and able to maintain CO2 capture capacity. Encapsulating the LBIL-FA-Ac system could be a promising strategy to improve its reusability to mitigate the evaporation of crucial parts of the ionic liquids systems. While the long-term stability and reusability of the LBIL-FA-Ac sample were not ideal, the present study provides a valuable foundation for understanding its CO2 capture performance. These results highlight opportunities for further improvement, such as optimizing regeneration conditions, tuning the ligand environment by increasing chelate strength, or using supported/hybrid systems to enhance material immobilization and stability.
9.

CO2 cycle with absorption temperature at 30 °C under 100% CO2 for 60 min and desorption was performed at 55 °C under 100% N2 for 60 min.
3.7. CO2 Absorption Mechanism
Most ILs reported are found to be predominantly absorb CO2 physically, though chemical absorption can also occur depending on the cation and anion present. To investigate the underlying mechanism of CO2 absorption in LBILs samples, the interaction between IL and CO2 was examined using FT-IR (Figure ) and 13C NMR (Figure ) spectra at ambient temperature by bubbling the LBIL samples with CO2. The FT-IR spectra of the CO2 saturated sample showed no significant new peaks compared to the lean sample, but upon CO2 absorption, a notable change was observed in the spectrum. The band at 1616 cm–1, corresponding to the conjugated CC and CO bonds of the chelated acetylacetonate ligand, disappeared, while the band intensity around 1650 cm–1 has increased. This suggests that the CO2 molecules may perturb the electron density within the acetylacetonate chelate and disrupt the vibrational coupling, causing a redistribution of the band intensity. The observed blue-shifting of the acetylacetonate-related vibration is consistent with weak CO2-ligand interactions that alter local polarization and vibrational coupling. Similar blue-shifts of carbonyl type bands upon CO2 exposure have been reported, where these are often attributed to dielectric effects that are significantly greater than specific interactions between CO2 and carbonyl groups. − The broadening and decreased intensity in aldehydic proton and methyl acetate vibrational were attributed to intermolecular C–H···O as well as dielectric and donor–acceptor interactions between CO2 and electron-rich oxygen sites. ,
10.
FT-IR spectra for LBILs samples of the CO2 lean and CO2 saturated samples.
11.
13C NMR results for LBILs samples of CO2 lean and CO2 saturated.
The band at 1620 cm–1 is typically associated with CO stretching vibrations and is often used as an indicator of bicarbonate (HCO3–) formation, which was not observed in the CO2 saturated samples. Similarly, the band at 1452 cm–1 may be attributed to asymmetric C–O stretching of carbonate species. In order to confirm any formation of carbonate species, further investigation using 13C NMR results revealed no new peaks in the 160 ppm region, which would be a significant characteristic of carbonate species. The absence of these signals strongly indicates that no chemical reaction with CO2 occurred. Furthermore, the viscosity after the CO2 absorption was also investigated. The viscosities of both the CO2 lean and CO2 saturated samples either decreased or remained similar for all ILs (Figure S14). In the case of LBIL and LBIL-FA, the viscosity was shown to decrease, as CO2 molecules dissolve physically into the ionic liquid. Experimental studies have shown that physical absorption of CO2 weakens ion–ion interactions, resulting in a decrease in viscosity. On the other hand, LBIL-FA-Ac samples exhibited an average viscosity of 13.48 mPa.s in the CO2 lean sample and 12.53 mPa.s after CO2 saturation. This is because of the effect of the acetylacetonate ligand, which has already dispersed the molecular interactions within the ILs, resulting in minimal changes after CO2 saturation. This suggests that no significant formation of carbonate species occurred, which could otherwise have led to viscosity changes.
The absorption enthalpy values of the material can also be good indicators of the determination of the absorption mechanism. For all LBILs, the enthalpy values obtained (Sec ) were relatively low (<40 kJ mol–1), indicating that CO2 interacts with the ILs predominantly through physical absorption rather than chemisorption. Together, these results suggest that the CO2 absorption in LBIL samples proceeds mainly via physical absorption.
4. Conclusions
In this study, a lithium borate-based ionic liquid incorporating Mg(acac)2 was developed and evaluated for its CO2 capture performance. The introduction of Mg(acac)2 provided additional cooperative sites, enhancing the interaction between CO2 and the ionic liquid through a cooperative push–pull mechanism. While the overall CO2 uptake of the LBILs was lower than that of conventional, the designed ionic liquid demonstrated comparable or superior performance to other reported ionic liquids under milder conditions, highlighting its potential for low energy CO2 capture. Moreover, the relatively weak interaction between CO2 and the ionic liquid facilitates the easy desorption of CO2 without requiring high thermal input. The balance between moderate uptake and energy efficiency could offer promise for practical CO2 capture applications. Future efforts may focus on optimizing the ionic liquid formulation or integrating with porous supports to further enhance the stability and performance.
Supplementary Material
Acknowledgments
The study was carried out with the technical supports on NMR analyses by Core Facility Center at Institute of Science Tokyo. We thank Mr. Moin Khwaja, a Ph.D candidate in our research group, for his support in the preparation of the cover artwork illustration.
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.5c09731.
Chemical structures of LBILs; NMR spectra of synthesized LBILs; DSC data for LBILs; comparison of viscosity and CO2 uptake capacity with other materials; viscosity measurements for LBILs upon CO2 saturation (PDF)
The authors declare no competing financial interest.
References
- Lindsey, R. Climate Change: atmospheric carbon dioxide 2025. See the following: https://www.climate.gov/news-features/understanding-climate/climate-change-atmospheric-carbon-dioxide.
- Lindsey R., Dahlman L.. Climate change: Global temperature. Climate. gov. 2025;16:1–5. [Google Scholar]
- Allen, M. ; Antwi-Agyei, P. ; Aragon-Durand, F. ; Babiker, M. ; Bertoldi, P. ; Bind, M. ; Brown, S. ; Buckeridge, M. ; Camilloni, I. ; Cartwright, A. . et al. Technical Summary: Global warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty; IPCC, 2019. [Google Scholar]
- Joshi M., Hawkins E., Sutton R., Lowe J., Frame D.. Projections of when temperature change will exceed 2°C above pre-industrial levels. Nature Climate Change. 2011;1:407–412. doi: 10.1038/nclimate1261. [DOI] [Google Scholar]
- Hasan M. F., First E. L., Boukouvala F., Floudas C. A.. A multi-scale framework for CO2 capture, utilization, and sequestration: CCUS and CCU. Comput. Chem. Eng. 2015;81:2–21. doi: 10.1016/j.compchemeng.2015.04.034. [DOI] [Google Scholar]
- Tapia J. F. D., Lee J.-Y., Ooi R. E., Foo D. C., Tan R. R.. A review of optimization and decision-making models for the planning of CO2 capture, utilization and storage (CCUS) systems. Sustainable Production and Consumption. 2018;13:1–15. doi: 10.1016/j.spc.2017.10.001. [DOI] [Google Scholar]
- Raganati F., Ammendola P.. CO2 post-combustion capture: a critical review of Current technologies and future directions. Energy Fuels. 2024;38:13858–13905. doi: 10.1021/acs.energyfuels.4c02513. [DOI] [Google Scholar]
- Liang Z., Fu K., Idem R., Tontiwachwuthikul P.. Review on current advances, future challenges and consideration issues for post-combustion CO2 capture using amine-based absorbents. Chinese journal of chemical engineering. 2016;24:278–288. doi: 10.1016/j.cjche.2015.06.013. [DOI] [Google Scholar]
- Songolzadeh M., Soleimani M., Takht Ravanchi M., Songolzadeh R.. Carbon dioxide separation from flue gases: a technological review emphasizing reduction in greenhouse gas emissions. Scientific World Journal. 2014;2014:828131. doi: 10.1155/2014/828131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borhani T. N., Abbasi M. R., Hosseinpour M., Salimi M., Afkhamipour M., Oko E., Campbell K. S., Kahllaghi N.. CO2 absorption-desorption cycles: Progress, gaps, and future. Carbon Capture Science & Technology. 2024;13:100325. doi: 10.1016/j.ccst.2024.100325. [DOI] [Google Scholar]
- Meng F., Meng Y., Ju T., Han S., Lin L., Jiang J.. Research progress of aqueousamine solution for CO2 capture: A review. Renewable and Sustainable Energy Reviews. 2022;168:112902. doi: 10.1016/j.rser.2022.112902. [DOI] [Google Scholar]
- Numpilai T., Pham L. K. H., Witoon T.. Advances in ionic liquid technologies for CO2 capture and conversion: a comprehensive review. Ind. Eng. Chem. Res. 2024;63:19865–19915. doi: 10.1021/acs.iecr.4c02072. [DOI] [Google Scholar]
- Dean P. M., Pringle J. M., MacFarlane D. R.. Structural analysis of low melting organic salts: perspectives on ionic liquids. Phys. Chem. Chem. Phys. 2010;12:9144–9153. doi: 10.1039/c003519j. [DOI] [PubMed] [Google Scholar]
- Dubey A., Arora A.. Advancements in carbon capture technologies: A review. Journal of Cleaner Production. 2022;373:133932. doi: 10.1016/j.jclepro.2022.133932. [DOI] [Google Scholar]
- Hayes R., Warr G. G., Atkin R.. Structure and nanostructure in ionic liquids. Chem. Rev. 2015;115:6357–6426. doi: 10.1021/cr500411q. [DOI] [PubMed] [Google Scholar]
- Lei Z., Chen B., Koo Y.-M., MacFarlane D. R.. Introduction: ionic liquids. 2017;117:6633. doi: 10.1021/acs.chemrev.7b00246. [DOI] [PubMed] [Google Scholar]
- Watanabe M., Thomas M. L., Zhang S., Ueno K., Yasuda T., Dokko K.. Application of ionic liquids to energy storage and conversion materials and devices. Chem. Rev. 2017;117:7190–7239. doi: 10.1021/acs.chemrev.6b00504. [DOI] [PubMed] [Google Scholar]
- McDaniel J. G., Yethiraj A.. Understanding the properties of ionic liquids: Electrostatics, structure factors, and their sum rules. J. Phys. Chem. B. 2019;123:3499–3512. doi: 10.1021/acs.jpcb.9b00963. [DOI] [PubMed] [Google Scholar]
- Halliday C., Hatton T. A.. Sorbents for the capture of CO2 and other acid gases: a review. Ind. Eng. Chem. Res. 2021;60:9313–9346. doi: 10.1021/acs.iecr.1c00597. [DOI] [Google Scholar]
- 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. doi: 10.1021/acs.chemrev.7b00072. [DOI] [PubMed] [Google Scholar]
- Faisal Elmobarak W., Almomani F., Tawalbeh M., Al-Othman A., Martis R., Rasool K.. Current status of CO2 capture with ionic liquids: Development and progress. Fuel. 2023;344:128102. doi: 10.1016/j.fuel.2023.128102. [DOI] [Google Scholar]
- Mulk W. U., Ali S. A., Shah S. N., Shah M. U. H., Zhang Q.-J., Younas M., Fatehizadeh A., Sheikh M., Rezakazemi M.. Breaking boundaries in CO2 capture: Ionic liquid-based membrane separation for post-combustion applications. Journal of CO2 Utilization. 2023;75:102555. doi: 10.1016/j.jcou.2023.102555. [DOI] [Google Scholar]
- Yuksel Orhan O.. Effects of various anions and cations in ionic liquids on CO2 capture. J. Mol. Liq. 2021;333:115981. doi: 10.1016/j.molliq.2021.115981. [DOI] [Google Scholar]
- 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. doi: 10.1002/cssc.201300691. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/ja017593d. [DOI] [PubMed] [Google Scholar]
- 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. Chemical engineering journal. 2012;193-194:267–275. doi: 10.1016/j.cej.2012.04.015. [DOI] [Google Scholar]
- Carvalho P. J., Alvarez V. H., Marrucho I. M., Aznar M., Coutinho J. A.. High pressure phase behavior of carbon dioxide in 1-butyl-3-methylimidazolium bis (trifluoromethylsulfonyl) imide and 1-butyl-3-methylimidazolium dicyanamide ionic liquids. Journal of Supercritical Fluids. 2009;50:105–111. doi: 10.1016/j.supflu.2009.05.008. [DOI] [Google Scholar]
- Wu G., Liu Y., Liu G., Pang X.. The CO2 absorption in flue gas using mixed ionic liquids. Molecules. 2020;25:1034. doi: 10.3390/molecules25051034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahmady A., Hashim M. A., Aroua M. K.. Experimental investigation on the solubility and initial rate of absorption of CO2 in aqueous mixtures of methyldiethanolamine with the ionic liquid 1-butyl-3-methylimidazolium tetrafluoroborate. Journal of Chemical & Engineering Data. 2010;55:5733–5738. doi: 10.1021/je1006949. [DOI] [Google Scholar]
- 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. doi: 10.1021/ja909305t. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/jp2015534. [DOI] [PubMed] [Google Scholar]
- Goodrich B. F., De La Fuente J. C., Gurkan B. E., Zadigian D. J., Price E. A., Huang Y., Brennecke J. F.. Experimental measurements of amine-functionalized anion-tethered ionic liquids with carbon dioxide. Industrial & engineering chemistry research. 2011;50:111–118. doi: 10.1021/ie101688a. [DOI] [Google Scholar]
- Blanchard L. A., Gu Z., Brennecke J. F.. High-pressure phase behavior of ionic liquid/CO2 systems. J. Phys. Chem. B. 2001;105:2437–2444. doi: 10.1021/jp003309d. [DOI] [Google Scholar]
- Aki S. N., Mellein B. R., Saurer E. M., Brennecke J. F.. High-pressure phase behavior of carbon dioxide with imidazolium-based ionic liquids. J. Phys. Chem. B. 2004;108:20355–20365. doi: 10.1021/jp046895+. [DOI] [Google Scholar]
- Shin E.-K., Lee B.-C., Lim J. S.. High-pressure solubilities of carbon dioxide in ionic liquids: 1-alkyl-3-methylimidazolium bis (trifluoromethylsulfonyl) imide. Journal of Supercritical Fluids. 2008;45:282–292. doi: 10.1016/j.supflu.2008.01.020. [DOI] [Google Scholar]; Blanchard L. A., Gu Z., Brennecke J. F.. High-pressure phase behavior of ionic liquid/CO2 systems. J. Phys. Chem. B. 2001;105(12):2437–2444. doi: 10.1021/jp003309d. [DOI] [Google Scholar]
- Carvalho P. J., Alvarez V. H., Marrucho I. M., Aznar M., Coutinho J. A.. High carbon dioxide solubilities in trihexyltetradecylphosphonium-based ionic liquids. Journal of Supercritical Fluids. 2010;52:258–265. doi: 10.1016/j.supflu.2010.02.002. [DOI] [Google Scholar]
- Wang C., Luo H., Jiang D., Li H., Dai S.. Carbon Dioxide Capture by Superbase-Derived Protic Ionic Liquids. Angew. Chem., Int. Ed. 2010;49(34):5978–5981. doi: 10.1002/anie.201002641. [DOI] [PubMed] [Google Scholar]
- Wang C., Luo X., Luo H., Jiang D., Li H., Dai S.. Tuning the Basicity of Ionic Liquids for Equimolar CO2 Capture. Angew. Chem., Int. Ed. 2011;50(21):4918–4922. doi: 10.1002/anie.201008151. [DOI] [PubMed] [Google Scholar]
- Freire M. G., Neves C. M., Carvalho P. J., Gardas R. L., Fernandes A. M., Marrucho I. M., Santos L. M., Coutinho J. A.. Mutual solubilities of water and hydrophobic ionic liquids. J. Phys. Chem. B. 2007;111:13082–13089. doi: 10.1021/jp076271e. [DOI] [PubMed] [Google Scholar]
- Harada T., Halliday C., Jamal A., Hatton T. A.. Molten ionic oxides for CO 2 capture at medium to high temperatures. Journal of Materials Chemistry A. 2019;7:21827–21834. doi: 10.1039/C9TA09122J. [DOI] [Google Scholar]
- Guzman-Gonzalez G., Alvarez-Tirado M., Olmedo-Martinez J. L., Picchio M. L., Casado N., Forsyth M., Mecerreyes D.. Lithium Borate Ionic Liquids as Single-Component Electrolytes for Batteries. Adv. Energy Mater. 2023;13:2202974. doi: 10.1002/aenm.202202974. [DOI] [Google Scholar]
- Sodhi R. K., Paul S.. An overview of metal acetylacetonates: Developing areas/routes to new materials and applications in organic syntheses. Catalysis surveys from Asia. 2018;22:31–6. doi: 10.1007/s10563-017-9239-9. [DOI] [Google Scholar]
- Kumagai T.. Direct observation and control of hydrogen-bond dynamics using low-temperature scanning tunneling microscopy. Prog. Surf. Sci. 2015;90:239–291. doi: 10.1016/j.progsurf.2015.04.001. [DOI] [Google Scholar]
- Dannenberg J., Rios R.. Theoretical study of the enolic forms of acetylacetone. How strong is the hydrogen bond? J. Phys. Chem. 1994;98:6714–6718. doi: 10.1021/j100078a011. [DOI] [Google Scholar]
- Meany J. E.. Reversible tautomerization of acetylacetone enol. I. Metal ion catalysis. J. Phys. Chem. 1969;73:3421–3423. doi: 10.1021/j100844a046. [DOI] [Google Scholar]
- Belova N. V., Oberhammer H., Trang N. H., Girichev G. V.. Tautomeric properties and gas-phase structure of acetylacetone. Journal of organic chemistry. 2014;79:5412–5419. doi: 10.1021/jo402814c. [DOI] [PubMed] [Google Scholar]
- Rarh, V. Organic synthesis via enolates; Academia, 2008. [Google Scholar]
- Shibata I., Baba A.. Organotin Enolates in Organic Synthesis. A Review. Organic preparations and procedures international. 1994;26:85–100. doi: 10.1080/00304949409458015. [DOI] [Google Scholar]
- Emsley J., Ernst R., Hathaway B., Warren K., Emsley J.. The composition, structure and hydrogen bonding of the β-diketones. Complex chemistry. 1984;57:147–191. doi: 10.1007/BFb0111456. [DOI] [Google Scholar]
- Deng W., Zhang L., Li L., Chen S., Hu C., Zhao Z.-J., Wang T., Gong J.. Crucial role of surface hydroxyls on the activity and stability in electrochemical CO2 reduction. J. Am. Chem. Soc. 2019;141:2911–2915. doi: 10.1021/jacs.8b13786. [DOI] [PubMed] [Google Scholar]
- Xiao W., Zhang J., Wu J.. Recent advances in reactions involving carbon dioxide radical anion. ACS Catal. 2023;13:15991–16011. doi: 10.1021/acscatal.3c04125. [DOI] [Google Scholar]
- Sampson M. D., Kubiak C. P.. Manganese electrocatalysts with bulky bipyridine ligands: Utilizing Lewis acids to promote carbon dioxide reduction at low overpotentials. J. Am. Chem. Soc. 2016;138:1386–1393. doi: 10.1021/jacs.5b12215. [DOI] [PubMed] [Google Scholar]
- Durfy C. S., Zurakowski J. A., Drover M. W.. A Blueprint for Secondary Coordination Sphere Editing: Approaches Toward Lewis-Acid Assisted Carbon Dioxide Co-Activation. ChemSusChem. 2024;17:e202400039. doi: 10.1002/cssc.202400039. [DOI] [PubMed] [Google Scholar]
- Liu W., Sahoo B., Junge K., Beller M.. Cobalt complexes as an emerging class of catalysts for homogeneous hydrogenations. Acc. Chem. Res. 2018;51:1858–1869. doi: 10.1021/acs.accounts.8b00262. [DOI] [PubMed] [Google Scholar]
- Yin X., Moss J. R.. Recent developments in the activation of carbon dioxide by metal complexes. Coord. Chem. Rev. 1999;181:27–59. doi: 10.1016/S0010-8545(98)00171-4. [DOI] [Google Scholar]
- Brittain H. G.. Proton Magnetic Resonance Spectrum of Magnesium Acetylacetonate in Chloroform-d. Evidence for Conformational Equilibrium. Inorganic chemistry. 1975;14(11):2858–2860. doi: 10.1021/ic50153a056. [DOI] [Google Scholar]
- Abraham R. J., Ainger N. J.. Proton Chemical Shifts in NMR. Part 13.1 Proton Chemical Shifts in Ketones and the Magnetic Anisotropy and Electric Field Effect of the Carbonyl Group. Journal of the Chemical Society. 1999;3:441–448. doi: 10.1039/a808908f. [DOI] [Google Scholar]
- Butera V., D’Anna L., Rubino S., Bonsignore R., Spinello A., Terenzi A., Barone G.. How the Metal Ion Affects the 1H NMR Chemical Shift Values of Schiff Base Metal Complexes: Rationalization by DFT Calculations. J. Phys. Chem. A. 2023;127(44):9283–9290. doi: 10.1021/acs.jpca.3c05653. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rahman M. M., Yoon S., Hsan N., Hong I., Kim Y., Kim S., Cho Y., Kumar S., Shin E., Koh J.. Synthesis and Properties of Anionic Waterborne Polyurethane Using Biobased Poly(Trimethylene Ether) Glycol for Breathable Membranes. Journal of the Taiwan Institute of Chemical Engineers. 2025;175:106251–106251. doi: 10.1016/j.jtice.2025.106251. [DOI] [Google Scholar]
- Rozenberg M., Loewenschuss A., Marcus Y.. IR Spectra and Hydration of Short-Chain Polyethyleneglycols. Spectrochimica Acta Part A Molecular and Biomolecular Spectroscopy. 1998;54(12):1819–1826. doi: 10.1016/S1386-1425(98)00062-6. [DOI] [Google Scholar]
- Oruc S., Yılmaz B.ır, Sancak K.. Characterization and Rheological Behavior of Asphalt Binder Modified by a Novel Cyclic Borate Ester Additive. Construction and Building Materials. 2022;348:128673–128673. doi: 10.1016/j.conbuildmat.2022.128673. [DOI] [Google Scholar]
- Ramis G., Busca G.. FTIR Spectra of Adsorbed N-Butylamine. J. Mol. Struct. 1989;193:93–100. doi: 10.1016/0022-2860(89)80124-3. [DOI] [Google Scholar]
- Harris K. R., Newitt P. J., Derlacki Z. J.. Alcohol Tracer Diffusion, Density, NMR and FTIR Studies of Aqueous Ethanol and 2,2,2-Trifluoroethanol Solutions at 25°C. Journal of the Chemical Society, Faraday Transactions. 1998;94(14):1963–1970. doi: 10.1039/a802567c. [DOI] [Google Scholar]
- Ye Q., Meng X., Pang L.. D2O Assisted FTIR Spectroscopic Analysis of Moisture in Edible Oil. Food Chemistry: X. 2023;18:100679. doi: 10.1016/j.fochx.2023.100679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deshmukh S. S., Kovalchuk V. I., Borovkov V. Y., d’Itri J. L.. FTIR Spectroscopic and Reaction Kinetics Study of the Interaction of CF3CFCl2 with γ-al2O3 . J. Phys. Chem. B. 2000;104(6):1277–1284. doi: 10.1021/jp993088f. [DOI] [Google Scholar]
- Chen B., Yang X., Zeng X., Huang Z., Xiao J., Wang J., Zhan G.. Multicomponent Metal Oxides Derived from Mn-BTC Anchoring with Metal Acetylacetonate Complexes as Excellent Catalysts for VOCs and CO Oxidation. Chemical Engineering Journal. 2020;397:125424–125424. doi: 10.1016/j.cej.2020.125424. [DOI] [Google Scholar]
- Dineshkumar P., Sahana R., Rajmohan G., Elangovan A., Naganandhini S. P., Sangeetha T., Arivazhagan G.. Interaction Behaviour of Acetonitrile in the Presence of Acetylacetone: FTIR Studies and DFT Calculations. Phys. Chem. Liq. 2022;60(5):696–706. doi: 10.1080/00319104.2022.2037134. [DOI] [Google Scholar]
- Čeponkus, J. ; Platakytė, R. ; Šablinskas, V. ; Gutierrez Quintanilla, A. . FTIR Study of Acetylacetone, D2-Acetylacetone and Hexafluoroacetylacetone–Water Complexes in Argon and Nitrogen Matrices. Chemija 2018, 29 (1). 10.6001/chemija.v29i1.3639. [DOI] [Google Scholar]
- Dismukes J. P., Jones L. H., Bailar J. C.. THE MEASUREMENT of METAL-LIGAND BOND VIBRATIONS in ACETYLACETONATE COMPLEXES1. J. Phys. Chem. 1961;65(5):792–795. doi: 10.1021/j100823a021. [DOI] [Google Scholar]
- Yang F., Wang B., Jiao Y., Tan H., Wang X.. Density and viscosity of three ionic liquids with 2, 2, 2-trifluoroethanol. J. Chem. Thermodyn. 2023;181:107038. doi: 10.1016/j.jct.2023.107038. [DOI] [Google Scholar]
- Xu Y., Heck K. N., Ayala-Orozco C., Arredondo J. H., Zenor W., Shammai M., Wong M. S.. Heavy oil viscosity reduction at mild temperatures using palladium acetylacetonate. Fuel. 2021;294:120546. doi: 10.1016/j.fuel.2021.120546. [DOI] [Google Scholar]
- Del Sesto R. E., McCleskey T. M., Macomber C., Ott K. C., Koppisch A. T., Baker G. A., Burrell A. K.. Limited thermal stability of imidazolium and pyrrolidinium ionic liquids. Thermochimica acta. 2009;491:118–120. doi: 10.1016/j.tca.2009.02.023. [DOI] [Google Scholar]
- Maton C., De Vos N., Stevens C. V.. Ionic liquid thermal stabilities: decomposition mechanisms and analysis tools. Chem. Soc. Rev. 2013;42:5963–5977. doi: 10.1039/c3cs60071h. [DOI] [PubMed] [Google Scholar]
- Xu C., Cheng Z.. Thermal stability of ionic liquids: Current status and prospects for future development. Processes. 2021;9:337. doi: 10.3390/pr9020337. [DOI] [Google Scholar]
- Lecomte H. A., Liggat J. J.. Degradation mechanism of diethylene glycol units in a terephthalate polymer. Polymer degradation and stability. 2006;91:681–689. doi: 10.1016/j.polymdegradstab.2005.05.028. [DOI] [Google Scholar]
- Nenashev R., Mordvinova N., Zlomanov V., Kuznetsov V.. Thermal decomposition of vanadyl acetylacetonate. Inorg. Mater. 2015;51:891–896. doi: 10.1134/S0020168515090150. [DOI] [Google Scholar]
- Farooq M. Q., Chand D., Odugbesi G. A., Varona M., Mudryk Y., Anderson J. L.. Investigating the effect of ligand and cation on the properties of metal fluorinated acetylacetonate based magnetic ionic liquids. New J. Chem. 2019;43:11334–11341. doi: 10.1039/C9NJ02595B. [DOI] [Google Scholar]
- Halliday, C. Molten Alkali Metal Borates for High Temperature Carbon Capture. Ph.D. thesis, Massachusetts Institute of Technology, 2022. [Google Scholar]
- Hu Y., Li X., Liu J., Li L., Zhang L.. Experimental investigation of CO2 absorption enthalpy in conventional imidazolium ionic liquids. Greenhouse Gases: Science and Technology. 2018;8:713–720. doi: 10.1002/ghg.1777. [DOI] [Google Scholar]
- Bresciani G., Biancalana L., Pampaloni G., Marchetti F.. Recent advances in the chemistry of metal carbamates. Molecules. 2020;25:3603. doi: 10.3390/molecules25163603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Phipps C. A., Zirilli C. D., Duff B. G., Erickson J. D., Karki S., Okolocha C., Mashuta M. S., Buchanan R. M., Grapperhaus C. A.. Enhancing CO2 Capture via Metal–Ligand Cooperativity: Tuning Ligand Basicity and Zn (II) Lewis Acidity. Inorg. Chem. 2024;63:9992–10000. doi: 10.1021/acs.inorgchem.4c01159. [DOI] [PubMed] [Google Scholar]
- Hazra A., Bonakala S., Adalikwu S. A., Balasubramanian S., Maji T. K.. Fluorocarbon-functionalized superhydrophobic metal–organic framework: enhanced CO2 uptake via photoinduced postsynthetic modification. Inorg. Chem. 2021;60:3823–3833. doi: 10.1021/acs.inorgchem.0c03575. [DOI] [PubMed] [Google Scholar]
- McEldrew M., Goodwin Z. A., Molinari N., Kozinsky B., Kornyshev A. A., Bazant M. Z.. Salt-in-ionic-liquid electrolytes: Ion network formation and negative effective charges of alkali metal cations. J. Phys. Chem. B. 2021;125:13752–13766. doi: 10.1021/acs.jpcb.1c05546. [DOI] [PubMed] [Google Scholar]
- Chiappe C., Malvaldi M.. Highly concentrated “solutions” of metal cations in ionic liquids: current status and future challenges. Phys. Chem. Chem. Phys. 2010;12:11191–11196. doi: 10.1039/c001796e. [DOI] [PubMed] [Google Scholar]
- Raeissi S., Peters C. J.. Carbon Dioxide Solubility in the Homologous 1-Alkyl-3-Methylimidazolium Bis(Trifluoromethylsulfonyl)Imide Family†. Journal of Chemical & Engineering Data. 2009;54(2):382–386. doi: 10.1021/je800433r. [DOI] [Google Scholar]
- Kim Y. S., Choi W. Y., Jang J. H., Yoo K.-P., Lee C. S.. Solubility Measurement and Prediction of Carbon Dioxide in Ionic Liquids. Fluid Phase Equilib. 2005;228–229:439–445. doi: 10.1016/j.fluid.2004.09.006. [DOI] [Google Scholar]
- Taylor C. D. L., Klemm A., Al-Mahbobi L., Bradford B. J., Gurkan B., Pentzer E. B.. Ionic Liquid–Glycol Mixtures for Direct Air Capture of CO2: Decreased Viscosity and Mitigation of Evaporation via Encapsulation. ACS Sustainable Chem. Eng. 2024;12(20):7882–7893. doi: 10.1021/acssuschemeng.4c01265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramdin M., de Loos T. W., Vlugt T. J.. State-of-the-art of CO2 capture with ionic liquids. Ind. Eng. Chem. Res. 2012;51:8149–8177. doi: 10.1021/ie3003705. [DOI] [Google Scholar]
- Yuan Y., Teja A. S.. Quantification of Specific Interactions between CO2 and the Carbonyl Group in Polymers via ATR-FTIR Measurements. Journal of Supercritical Fluids. 2011;56(2):208–212. doi: 10.1016/j.supflu.2010.12.010. [DOI] [Google Scholar]
- Paschoal V. H., Faria L. F. O., Ribeiro M. C. C.. Vibrational Spectroscopy of Ionic Liquids. Chem. Rev. 2017;117(10):7053–7112. doi: 10.1021/acs.chemrev.6b00461. [DOI] [PubMed] [Google Scholar]
- Dolati F., Vakili M., Ebrahimi A., Tayyari S. F.. Vibrational Spectra of Tris(Acetylacetonato)Chromium(III) J. Mol. Struct. 2015;1099:340–347. doi: 10.1016/j.molstruc.2015.05.058. [DOI] [Google Scholar]
- Blatchford M. A., Raveendran P., Wallen S. L.. Spectroscopic Studies of Model Carbonyl Compounds in CO2: Evidence for Cooperative C–H···O Interactions. J. Phys. Chem. A. 2003;107(48):10311–10323. doi: 10.1021/jp027208m. [DOI] [Google Scholar]
- Baltrusaitis J., Jensen J. H., Grassian V. H.. FTIR spectroscopy combined with isotope labeling and quantum chemical calculations to investigate adsorbed bicarbonate formation following reaction of carbon dioxide with surface hydroxyl groups on Fe2O3 and Al2O3. J. Phys. Chem. B. 2006;110:12005–12016. doi: 10.1021/jp057437j. [DOI] [PubMed] [Google Scholar]
- Zanatta M., Simon N. M., Dupont J.. The nature of carbon dioxide in bare ionic liquids. ChemSusChem. 2020;13:3101–3109. doi: 10.1002/cssc.202000574. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





