Next Article in Journal
Anticancer Potential of Raddeanin A, a Natural Triterpenoid Isolated from Anemone raddeana Regel
Next Article in Special Issue
Bio-Based Solvents and Gasoline Components from Renewable 2,3-Butanediol and 1,2-Propanediol: Synthesis and Characterization
Previous Article in Journal
Small Dimension—Big Impact! Nanoparticle-Enhanced Non-Invasive and Intravascular Molecular Imaging of Atherosclerosis In Vivo
Previous Article in Special Issue
Studies on the Solubility of Terephthalic Acid in Ionic Liquids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The CO2 Absorption in Flue Gas Using Mixed Ionic Liquids

College of Chemistry and Chemical Engineering, Inner Mongolia University, Hohhot 010021, China
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(5), 1034; https://doi.org/10.3390/molecules25051034
Submission received: 22 January 2020 / Revised: 18 February 2020 / Accepted: 22 February 2020 / Published: 25 February 2020

Abstract

:
Because of the appealing properties, ionic liquids (ILs) are believed to be promising alternatives for the CO2 absorption in the flue gas. Several ILs, such as [NH2emim][BF4], [C4mim][OAc], and [NH2emim[OAc], have been used to capture CO2 of the simulated flue gas in this work. The structural changes of the ILs before and after absorption were also investigated by quantum chemical methods, FTIR, and NMR technologies. However, the experimental results and theoretical calculation showed that the flue gas component SO2 would significantly weaken the CO2 absorption performance of the ILs. SO2 was more likely to react with the active sites of the ILs than CO2. To improve the absorption capacity, the ionic liquid (IL) mixture [C4mim][OAc]/ [NH2emim][BF4] were employed for the CO2 absorption of the flue gas. It is found that the CO2 absorption capacity would be increased by about 25%, even in the presence of SO2. The calculation results suggested that CO2 could not compete with SO2 for reacting with the IL during the absorption process. Nevertheless, SO2 might be first captured by the [NH2emim][BF4] of the IL mixture, and then the [C4mim][OAc] ionic liquid could absorb more CO2 without the interference of SO2.

1. Introduction

The CO2 absorption of flue gas is an important process for reducing greenhouse gas [1]. To date, most flue absorptions are performed by using amine solvents. However, conventional absorption methods usually have some disadvantages: High equipment corrosion rate, high absorbent make-up rate due to the amine degradation by SO2, NO2, and O2 in the flue gas, and high energy consumption during the regeneration process [2]. In the last decades, ionic liquids (ILs) have been used in many fields. Multifunctional ionic liquids are easily prepared, and the vapor pressure of ionic liquids can be neglected [3]. The other attractive properties of ILs include: High thermal stability, large electrochemical window, and high dissolve ability of compounds [4]. Blanchard et al. [5] have reported that certain ILs can considerably dissolve CO2 gas. Since then, ILs for CO2 capture have attracted much attention. For example, multi-N-containing ionic liquids can absorb much more SO2/CO2 in the flue gas than that of the limestone solvent [6]. Shiflett et al. [7,8] have found that the imidazolium-based ionic liquid [C4mim][OAc] can markedly reduce the energy losses of CO2 absorption comparing with those of the commercial monoethanolamine solvent. Guanidinium salt ILs (e.g., [TMG][L]) and functional ILs (e.g., [NH2p-bmim][BF4] and 2-(2-hydroxyethoxy) ammonium acetate) all show high efficiency for CO2 and SO2 capture [9,10].
The typical flue gas from coal-burning usually contains about 15 vol% CO2, 10 vol% H2O, and more than 2 vol% SO2 [11]. Apart from CO2, the effects of SO2 on the flue gas absorption should be taken into account [12]. Most researchers consider that water has a little influence on the CO2 capture, but the effects of SO2 would not be neglected. For the impurities of the flue gas, it is found that the ILs are more likely to absorb SO2 than CO2. Specifically, the N element of the ionic liquid (IL) prefers to capture the SO2 molecules, and then CO2 molecules are repelled by the captured SO2 [13]. Thus, the CO2 absorption capacity of ionic liquids would be rapidly decreased over several cycles. Moreover, almost all of the ILs would exhibit much higher SO2 absorption capacity than CO2 due to the higher solubility of SO2 in ILs. For instance, pure CO2 solubility in the guanidinium-based ILs was only 0.4 mol/mol, while the SO2 solubility in these ILs was as high as 2.5 mol/mol under the same conditions [10]. For an extreme case, the absorption capacities of pure SO2 and CO2 in the azole-based ILs (e.g., [P66614][Im]) were 3.5 mol/mol and 0.1 mol/mol, respectively [14]. Although SO2 has stronger interactions with ILs than CO2 does, the actual partial pressure of SO2 in the flue gas is very low [15]. It is well accepted that the partial pressure of SO2 is about two orders of magnitude lower than that of CO2, and low SO2 partial pressure usually leads to low absorption capacity for SO2. That is, a lot of energy will be consumed to remove SO2 and CO2 from the flue gas, if we used a two-step absorption process (absorbing SO2 at first, and then CO2).
The effects of single IL on the CO2 capture have been widely discussed in recent years [2,7,16,17,18,19]. However, there are few reports on the IL mixtures for CO2 absorption, especially on the CO2 capture from the flue gas. Therefore, if the IL mixture was employed, the CO2 absorption capacity of the mixed ILs might be greater than that of the single ionic liquid. Specifically, if the SO2 of flue gas was absorbed by one IL of the mixture, the negative influence of SO2 on the whole IL mixture might be significantly decreased, and then more CO2 could be captured.
In this work, we want to use IL mixtures to remove CO2 and SO2 from the flue gas. The amine-functional ionic liquid [NH2emim][BF4] and the imidazolium-based ionic liquid [C4mim][OAc] were synthesized at first. Subsequently, the two ILs were mixed to investigate the CO2 and SO2 absorption. Here, [NH2emim][BF4] was mainly used to absorb SO2 in the flue gas, while the [C4mim][OAc] ionic liquid was employed to capture CO2. In order to more clearly study the actual flue gas, the CO2 absorption performance in the IL mixture was measured at the simulated flue gas with 15 vol% CO2 and 2 vol% SO2. The effect of SO2 on the IL mixture and the interaction between CO2 and SO2 in the IL absorption were also studied. Furthermore, the absorption mechanism at the molecule level was investigated by the quantum chemical calculation and the instrumental analysis.

2. Results and Discussion

2.1. CO2 and SO2 Absorption Performance of the Single ILs

When the simulated flue gas (15% CO2/85% N2) only contains the CO2 impurity, one mole [C4mim][OAc] could absorb 0.298 mol CO2 (Table 1). However, when SO2 was mixed in the simulated flue gas (15% CO2/2% SO2/N2), the CO2 absorption capacity of [C4mim][OAc] was reduced to 0.204 CO2/mol IL. The result shows that SO2 has a negative effect on the IL absorbing CO2. The single IL [NH2emim][OAc] and [NH2emim][BF4] were also used to capture the CO2 of flue gas. Without the interference of SO2, the CO2 absorption capacities of [NH2emim][OAc] and [NH2emim][BF4] were 0.291 mol CO2/mol IL and 0.290 mol CO2/mol IL, respectively. However, like the [C4mim][OAc] ionic liquid, the CO2 absorption capacities under exposure to 2% SO2 would be markedly decreased to 0.171 mol CO2/mol [NH2emim][OAc] and 0.180 mol CO2/mol [NH2emim][BF4], respectively. In short, the single ionic liquids all exhibit the CO2 absorption capacities, but this capacity would be greatly weakened by the interference of SO2.
The researchers believed that CO2 and SO2 of the flue gas would be absorbed simultaneously [13,20]. The [C4mim][OAc] and [NH2emim][BF4] absorption results supported this conclusion. For example, an experiment was carried out of outlet SO2 concentration vs. time to investigate the SO2 absorption performance. In this study (Figure 1a), the simulated flue gas contained 15% CO2, 2% SO2, and 83% N2. The concentrations of CO2 and SO2 of the outlet stream were simultaneously detected with time. It was found that [C4mim][OAc] and [NH2emim][BF4] all can capture CO2, but the outlet SO2 concentration was hardly detected before 40 min (Figure 1a). It indicates that SO2 could be completely absorbed by [C4mim][OAc] and [NH2emim][BF4] during the absorption process. Additionally, an extreme case was investigated in which the simulated flue gas contained 80% CO2, 2% SO2, and 18% N2. However, SO2 was also not found at the gas stream of the outlet before 15 min. Compared with CO2, SO2 has higher dipole moments and molecular polarity, which often results in the strong affinity of SO2 with ionic liquids [21,22].
The presence of SO2 in flue gas usually leads to a competitive and negative influence on the separation of CO2. Figure 2 shows the CO2 absorption performance of [C4mim][OAc] at the atmosphere of 15% CO2/85% N2 and 15% CO2/2% SO2/83% N2, respectively. After 6 regeneration cycles, 1 mol [C4mim][OAc] could absorb 0.255 mol CO2, although the IL absorption capacity slightly decreased. In contrast, the absorption capacity of 1 mol [C4mim][OAc] was only 0.10 mol CO2 after the same number of cycles. In addition, the net SO2 absorption experiment (2% SO2/98% N2) showed that [NH2emim][BF4] has high SO2 absorption capacity (0.35 mol SO2/mol IL), while 1 mol [C4mim][OAc] only absorbs 0.18 mol SO2 (Figure 1b). These results agreed well with previous studies [23,24,25].
FT-IR can investigate the interaction between IL and CO2/SO2 [26]. The spectra of [C4mim][OAc] showed the changes after 15% CO2 and 2% SO2 absorption, respectively (Figure 3a−c). However, the [C4mim][OAc] spectrum had minimal changes when it was used to remove pure CO2. Although the appearance of the carbonyl band at 1720 cm−1 shows that the acetate anion might be partly converted into the acetate acid [27], Shiflett considered that the amount of such a chemical reaction was minor and reversible [24]. Thus, the other reactions between CO2 and the cation species in this work might not be detected within the wavenumber of 800−1600 cm−1, as Figure 3a,c shown. Similarly, the FT-IR spectra of [NH2emim][BF4] almost have no change before and after CO2 absorption (Figure 3d,f), indicating that the chemical reaction between the [NH2emim] cation and CO2 was not enough to be detected by FT-IR.
In contrast, the chemical reaction between SO2 and [C4mim][OAc] was much stronger than that of [C4mim][OAc]−CO2. Even if there was only a small amount of SO2 (2%) in the flue gas, the FT-IR spectrum still shows marked changes (Figure 3b). Compared with the spectrum of fresh [C4mim][OAc], the peak intensity at 1580 and 1371 cm−1 decreased significantly as the [C4mim][OAc] absorbing SO2. Meanwhile, the peaks at 1720, 1321, 1254, 1144, and 950 cm−1 newly appeared in the spectrum. The bands at 1321 and 1144 cm−1 can be attributed to the stretching of SO2 absorbed by the ionic liquid [26]. After SO2 absorption, the new peak at 1720 cm−1 shows the formation of a carbonyl group, which also indicates that most of the acetate ions were no longer associated with [C4mim] cations. The intense band at 950 cm−1 should be assigned to the vibrational mode of SO32− or S2O52−. It once again suggests that the interaction between the SO2 and [C4mim][OAc] ionic liquid was strong. Similarly, the peak intensity at 885 and 1543 cm−1 changed markedly when the [NH2emim][BF4] absorbed SO2. Particularly, two new peaks appeared at 968 and1367 cm−1, which can be attributed to the interaction between the N elements of the IL and SO2 [6].

2.2. CO2/SO2 Absorption Properties in IL Mixtures

Due to the influence of SO2, the CO2 absorption capacity of single ionic liquids was greatly decreased. If [NH2emim][BF4] and [C4mim][OAc] were simultaneously utilized, more CO2 (with SO2) in the flue gas might be captured. The CO2 absorption capacities of IL mixtures at different mole fractions of [C4mim][OAc] or [NH2emim][BF4] were displayed in Figure 4. Here, X is defined as the molar ratio of [NH2emim][BF4] to the IL mixture ([C4mim][OAc]/[NH2emim][BF4]). Compared with the CO2 absorption capacity of the single IL, the IL mixture could remove more CO2. This result may be due to the presence of [NH2emim][BF4]. The small amount of SO2 might be absorbed by [NH2emim][BF4] at first. Without the interference of SO2, the CO2 absorption capacity of the IL mixture was significantly enhanced. It was also found that the absorption of CO2 did not increase significantly with the increase of the X value. When X was 0.3, the CO2 absorption capacity of the [C4mim][OAc]/[NH2emim][BF4] mixture reached up to the maximum. Similarly, when the IL mixture [C4mim][OAc]/[NH2emim][OAc] was used to remove CO2/SO2 of the flue gas, the poison effect of SO2 on [C4mim][OAc] was also greatly reduced. As Figure 4 shows, the CO2 absorption of [C4mim][OAc]/[NH2emim][OAc] was about 0.4 mol CO2/mol IL. Compared to the single [C4mim][OAc], the absorption capacity of the IL mixture was improved.
The 1H NMR spectrum (Figure 5) shows that SO2 would interact with [NH2emim][BF4] and [NH2emim][OAc]. The NMR data of fresh [NH2emim][BF4] and fresh [NH2emim][OAc] are listed as follows:
Fresh [NH2emim][BF4], δ = 7.721 (s, 1H, unsaturated C−H in the imidazole ring, with N connected to the left and right), 7.641 (d, 1H, unsaturated C−H in the imidazole ring), 7.636 (d, 1H, unsaturated C−H in the imidazole ring), 4.295 (s, 3H, H3C−N ring), 4.039 (t, 2H, H2C−N ring), 3.112 (m, 2H, N−CH2−C−N ring), and 1.878 (t, 2H, NH2).
Fresh [NH2emim][OAc], δ = 12.751 (s, 1H, unsaturated C−H in the imidazole ring, with N connected to the left and right), 7.865 (d, 1H, unsaturated C−H in the imidazole ring), 7.703 (d, 1H, unsaturated C−H in the imidazole ring), 3.577 (s, 3H, H3C−N ring), 3.749 (t, 2H, H2C−N ring), 2.693 (m, 2H, N−CH2−C−N ring), 1.973 (t, 2H, NH2) and 1.651 (s, 3H, CH3 in OAc-).
In comparison to the 1H NMR spectrum of the fresh [NH2emim][BF4] (Figure 5a), new resonance peaks at 8.10 ppm were found after SO2 absorption, which indicates the formation of SN [28]. According to this result, it was considered that the interaction between SO2 and [NH2emim][BF4] should mainly occur at the N element of the [NH2emim] cation. For the case of the [NH2emim][OAc] (Figure 5b), a typical peak of −COOH in the 1H NMR spectrum moved from 12.75 to 11.83 ppm, and a new resonance peak was observed at 7.52 ppm after SO2 absorption. These results suggest that the interaction between [NH2emim][OAc] and SO2 had occurred [25]. That is, the interaction between [OAc] and SO2 leads to the moving from 12.75 ppm to 11.83, while the reaction of [NH2emim] and SO2 makes the new peak 7.52 ppm appearance.
In order to further investigate the effects of SO2 on the CO2 absorption capacity of ionic liquids, the CO2 absorption performance of fresh IL and after SO2-saturated IL are illustrated in Figure 6. Specifically, fresh [C4mim][OAc] and fresh [NH2emim][BF4] ionic liquids were used to absorb SO2 at first. When the IL was saturated by SO2, the CO2 absorption performance of the IL was investigated. It was found that the SO2-saturated [C4mim][OAc] did not have the ability to absorb CO2. The concentration of CO2 at the outlet was almost equal to that at the inlet. While fresh [C4mim][OAc] can absorb CO2 even after 60 min. The similar results were also observed using fresh [C2mim][OAc] and SO2 saturated [C2mim][OAc] to absorb CO2 [24]. Shiflet et al. considered that the interaction between the [OAc] anion and CO2 plays an important role in the CO2 removal of [OAc]−based ionic liquids [8]. However, the presence of SO2 makes a great impact on the CO2 absorption of [C4mim][OAc]. In contrast, when [NH2emim][BF4] was saturated by SO2, the [NH2emim][BF4] still had the CO2 absorption capacity. As Figure 6b shows, SO2-saturated [NH2emim][BF4] could capture about 2−7% CO2 during the absorption process.

2.3. Quantum Chemical Calculation on the Interaction of IL Mixture with CO2/SO2

The absorption capacity of CO2 in the IL mixtures was higher than that of the single ionic liquid. This may be related to the interactions between ILs and CO2/SO2 molecules. Thus, the interaction of the [C4mim][OAc] anion and [NH2emim][BF4] with CO2/SO2 was deeply investigated through quantum chemical calculation, which might be helpful to understand well the roles of CO2 and SO2 in the IL absorption. In this work, the structure of [NH2emim][BF4] and [C4mim][OAc] was optimized on the basis of DFT-D3 calculation at first. The configuration of the IL with the lowest energy was considered as the optimized structure. Additionally, the structures of [C4mim][OAc] and [NH2emim][BF4] with CO2 and SO2 were also investigated (Figure 7). The structural parameters for the IL−CO2/SO2 complexes are listed in Table 2.
In general, the CO2/SO2 gas molecules around the anions and cations were related to the absorption reaction. As Figure 7a shows, there was a strong interaction between the N atom and the S atom of SO2. Due to the complexation of SO2N, the average angle of SO2 was 116.0°. Compared to 119.5° of the pure SO2 molecule, the bending degree of O=S=O was increased. Similarly, [NH2emim][BF4] also leads to an impact on the CO2 structure. The angle of CO2 was bent from 180° to 166°, and the bond length of C−O was extended from 1.16 Å to 1.19 Å. For the cases of [C4mim][OAc], the interaction between the O and the C atom of carbon dioxide was also strong due to the negatively charged oxygen (O atom) in the [OAc] anion. The curvature of CO2 could be increased by the interaction of the [OAc] anion. The average angle of CO2 was bent to 146°, and the bond length of C−O was elongated to 1.21 Å. The configurations in Figure 7 also suggest that [NH2emim] cation and [OAc] anion are the active sites for CO2/SO2 absorption.
To some extent, the interaction energy and absorption enthalpy might reflect the absorption capacity of the ILs. It was found that the NH2emim] cation and [OAc] anion were the main active sites for the absorption of CO2 and SO2. In order to save the calculation cost and reduce the interference of other ions, herein, only the thermodynamic data of [OAc]−CO2, [OAc]−SO2, CO2−[NH2emim], and SO2−[NH2emim] complexes were compared (Table 3). It was found that the interaction energy and absorption enthalpy of [OAc]−CO2−SO2 complex were less than the sum of the energy and the enthalpy for [OAc]−CO2 and [OAc]−SO2, suggesting that CO2 and SO2 would competitively react with [OAc] anion.
In contrast, the interaction energy and absorption enthalpy of [NH2emim]−CO2−SO2 complex were approximately equal to the sum of those for the [NH2emim]−CO2 and [NH2emim]−SO2 complexes, indicating that the competitive reaction between [NH2emim]−CO2 and [NH2emim]−SO2 was not obvious. The absorption reaction might also lead to a change in charge distribution. It was found that the amount of net charge transfer from CO2 to [NH2emim] in [NH2emim]−CO2−SO2 complex was almost equal to that of [NH2emim]−CO2, suggesting that [NH2emim] had strong interactions with either SO2 or CO2. However, due to the impact of SO2, the net charge transfer from [OAc] to CO2 was significantly reduced from −0.510 in the [OAc]−CO2 complex to −0.035 in [OAc]−CO2−SO2 complex, respectively, which might account for the decrease of CO2 absorption capacity for [C4mim][OAc] in Table 1.
Table 4 collects the thermochemical data of [NH2emim][BF4]−CO2 and [NH2emim][BF4]−SO2 complexes. In order to consider the solvent effect of ionic liquids, the continuum universal solvation model (SMD) was used in the calculation. Based on the SMD model, the interaction energies of [NH2emim][BF4]−CO2 and [NH2emim][BF4]−SO2 system were −16.8 and −76.3 kJ/mol, respectively. They were lower than those in the gas phase (−19.2 and −85.4 kJ/mol). Notably, the difference between the interaction energy ([NH2emim][BF4]−CO2 and [NH2emim][BF4]−SO2) in the gas phase (59.5 kJ/mol) was very consistent with the energy difference (66.2 kJ/mol) using the SMD model. In the liquid phase, the interaction energy between [NH2emim][BF4] and SO2 was slightly greater than that of [NH2emim][BF4]−CO2, suggesting that [NH2emim][BF4] tends to react with SO2 rather than with CO2 during the absorption process. Similarly, this phenomenon could also be observed in the thermodynamic data of the [C4mim][OAc]−CO2 and [C4mim][OAc]−SO2 complexes. In short, SO2 was more active than CO2 in the reaction with ionic liquids, and the [NH2emim][BF4] may be more likely to absorb SO2.
The interaction between IL mixture ([C4mim][OAc]/[NH2emim][BF4]) and CO2/SO2 has also been investigated by the quantum chemistry calculation. The optimized structure is displayed in Figure 8. In the mixed ionic liquids, it is found that SO2 was close to [NH2emim][BF4], while the CO2 molecule was near the [C4mim][OAc] ionic liquid. Specifically, SO2 would have interacted with the N atom on the [NH2emim] cation, and CO2 was more likely to react with the [OAc] anion. This result might explain why the IL mixture can more effectively absorb the CO2 of flue gas. Because of the existence of [NH2emim][BF4], SO2 may be first captured by [NH2emim]. Without the interference of SO2, the [C4mim][OAc] ionic liquid then could absorb more CO2.

3. Materials and Methods

3.1. Materials

The simulated flue gas was obtained by pure gas CO2, SO2, and N2 (purity of >99.99 wt%). They were all purchased from Beifen (China) Gas Technology Company. 1-butyl-3-methylimidazolium tetrafluoroborate ([C4mim][BF4], 99 wt%) was obtained from Sigma-Aldrich Chemical Co., but the ionic liquid 1-butyl-3-methylimidazolium acetate ([C4mim][OAc], 99 wt%) were purchased from Lanzhou Greenchem ILs, LICP, CAS, China. Additionally, [NH2emim][BF4] and [NH2emim][OAc] have been synthesized by ourselves in this work. The used materials were as follows: 1-methylimidazole (C4H6N2), 2-bromoethylamine hydrobromide (C2H7Br2N), sodium acetate anhydrous (CH3COONa), 1-methylimidazole (C4H6N2), acetic acid (CH3COOH), and sodium borate (NaBF4). They were all provided by Sinopharm Chemical Reagent Co., Ltd., China, with purity over 98 wt%.

3.2. Ionic Liquid Preparation

In this work, the ionic liquids [NH2emim][BF4] and [NH2emim][OAc] were prepared by ourselves according to the method used in the literature [17,18,29]. First, the [NH2emim] cation was prepared by the reaction of 1-methylimidazole and 2-bromoethylamine hydrobromide under reflux for 12 h. Second, the [NH2emim]-based IL was simply synthesized by ion exchange with NaBF4 or NaOAc/CH3COOH in ethanol, and then the ethanol was removed in vacuum. The structures of the ILs were confirmed by proton nuclear magnetic resonance (1H NMR, Bruker WB400 AMX spectrometer, Billerica, MA, USA). Here, deuterated chloroform (CDCl3) was used as a solvent, and tetramethylsilane (TMS) was employed as an internal standard for 1H NMR measurement.

3.3. CO2 and SO2 Absorption

As Figure 9 shows, CO2 and SO2 absorption experiments were performed in a 30 mL reactor immersed with a water-bath temperature controller. The temperatures were controlled at 293 K for absorption and 353 K for desorption, respectively. The simulated flue gas was a mixture of N2, CO2, and SO2 in accordance with a certain proportion. As a typical absorption process, 10 mL IL or IL mixtures were added to the reactor at first. Subsequently, 15 vol% CO2, 2 vol% SO2 and 83 vol% N2 were mixed in storage. The intake speed of the mixed gas was controlled at 60 mL/min, and the absorption pressure was controlled at 101.3 kPa. The concentrations of CO2 and SO2 were analyzed by a gas analyzer (MRU NOVA2000) at the outlet. To investigate the IL regeneration, CO2 or SO2 saturated IL was also loaded in the reactor. Desorption was performed at 353 K under a pure N2 gas atmosphere for 30 min.
The amount of absorbed CO2 or SO2 was calculated by the following equation.
A gas = M IL ρ gas Q t 1 t 2 ( C 0 C gas ( t ) ) d t m IL M gas
where Agas is the molar amount of CO2 or SO2 in the ionic liquid; Q is the flow rate of the gas stream; C0 and Cgas are the CO2 or SO2 concentrations at the inlet and the outlet streams, respectively; t1 refers to the beginning time of the absorption process; when the CO2 and SO2 concentration at the outlet stream returns to the initial concentration, the time is t2; MIL and Mgas are the molecular weight of IL and CO2 (or SO2), respectively; mIL is the weight of the ILs, and ρgas is the density of CO2 or SO2.
After the IL was saturated by CO2 and SO2, the complex structure was investigated by the FTIR and NMR technologies. The FTIR spectra of the samples were analyzed on an FTIR spectrometer (PerkinElmer, Frontier 2500). In addition, the structure changes of the [NH2emim]-based IL after absorption were also detected by the NMR spectrometer (Bruker WB400 AMX, 300 MHz) using chloroform-d (CDCl3) as a solvent and tetramethylsilane (TMS) as an internal standard.

3.4. Theoretical Calculation

A quantum chemical calculation was used in this work to study the interaction between the ionic liquid and CO2 with SO2. All calculations were carried out by the Gaussian 16 program [30]. For IL calculations, Li et al. [31] suggested that the density function of the Minnesota family [32] (e.g., M06-2X) with a diffusion function basis set (e.g., 6-311++G(d,p)) might give reasonable results. If dispersion-corrected density functionals (e.g., gd3bj, DFT-D3) were used, more reliable results could be obtained [33]. Therefore, the geometry optimization and frequency analysis of all ILs and IL mixtures were performed at the M06-2X/6-311++G(d,p) level and correction with Grimme’s method. In order to calculate the interaction energy of the IL complexes, the basis set superposition error (BSSE) method was employed to correct the energy results [34]. The effect of the solvent should be taken into consideration in the theoretical calculation of the ionic liquids. It was found that the SMD solvation model proposed by Truhlar et al. can be used for the IL calculation very well [35,36]. Thus, the density functional theory (M06-2X and dispersion-corrected method) with the SMD model was also used to calculate the interaction energy of IL−CO2/SO2.

4. Conclusions

The CO2 and SO2 absorption of the flue gas in ionic liquids were investigated by the experimental method and theoretical calculation. The single ionic liquids, such as [NH2emim][BF4] and [C4mim][OAc], all showed good CO2 absorption performance for the simulated flue gas without SO2 interference. However, SO2 was more likely to react with the active sites of the ILs. When SO2 was in the flue gas, the CO2 absorption capacity of the single ionic liquid would be significantly inhibited. It was found that the interference of SO2 on the CO2 absorption performance might be markedly reduced by using the ionic liquid mixtures. The CO2 absorption capacity of the IL mixture [C4mim][OAc]/[NH2emim][BF4] was about 0.4 mol CO2/mol IL even at an atmosphere of 15% CO2/2%SO2/83% N2, which was greater than that of single [C4mim][OAc] (0.204 mol CO2/mol IL). There was a competitive relationship between CO2 and SO2 during the absorption process. The single ILs prefer to capture SO2 rather than remove CO2, due to the stronger interaction energy of SO2 and the ILs. The experimental and calculated results suggested that the [OAc] anion and [NH2emim] cation are the main active sites for CO2 and SO2 absorption. A lower absorption enthalpy of the IL−SO2 or IL−CO2 system usually means low absorption capacity. Thus, for the IL mixture [C4mim][OAc]/[NH2emim][BF4], the quantum calculation results indicated that [NH2emim][BF4] might be more likely to absorb SO2 of the flue gas and CO2 was easily removed by the [C4mim][OAc].

Author Contributions

Funding acquisition, Y.L.; investigation, G.W., Y.L., and G.L.; methodology, Y.L.; data curation, X.P.; writing—original draft, G.W., and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC were funded by the National Natural Science Foundation of China (grant numbers: 21766021 and 21266015).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aghaie, M.; Rezaei, N.; Zendehboudi, S. A systematic review on CO2 capture with ionic liquids: Current status and future prospects. Renew. Sustain. Energy Rev. 2018, 96, 502–525. [Google Scholar] [CrossRef]
  2. Yu, C.H.; Huang, C.H.; Tan, C.S. A review of CO2 capture by absorption and adsorption. Aerosol Air Qual. Res. 2012, 12, 745–769. [Google Scholar] [CrossRef] [Green Version]
  3. Wang, B.; Qin, L.; Mu, T.; Xue, Z.; Gao, G. Are ionic liquids chemically stable? Chem. Rev. 2017, 117, 7113–7131. [Google Scholar] [CrossRef]
  4. Zhang, S.; Sun, N.; He, X.; Lu, X.; Zhang, X. Physical properties of ionic liquids: Database and evaluation. J. Phys. Chem. Ref. Data 2006, 35, 1475–1517. [Google Scholar] [CrossRef]
  5. 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]
  6. Wang, C.M.; Cui, G.K.; Luo, X.Y.; Xu, Y.J.; Li, H.R.; Dai, S. Highly efficient and reversible SO2 capture by tunable azole-based ionic liquids through multiple-site chemical absorption. J. Am. Chem. Soc. 2011, 133, 11916–11919. [Google Scholar] [CrossRef]
  7. Shiflett, M.B.; Yokozeki, A. Separation of carbon dioxide and sulfur dioxide using room-temperature ionic liquid [bmim][MeSO4]. Energy Fuels 2010, 24, 1001–1008. [Google Scholar] [CrossRef]
  8. Shiflett, M.B.; Yokozeki, A. Solubilities and diffusivities of carbon dioxide in ionic liquids: Bmim PF6 and bmim BF4. Ind. Eng. Chem. Res. 2005, 44, 4453–4464. [Google Scholar] [CrossRef]
  9. Zhang, X.; Jiang, K.; Liu, Z.; Yao, X.; Liu, X.; Zeng, S.; Dong, K.; Zhang, S. Insight into the performance of acid gas in ionic liquids by molecular simulation. Ind. Eng. Chem. Res. 2019, 58, 1443–1453. [Google Scholar] [CrossRef]
  10. Shang, Y.; Li, H.P.; Zhang, S.J.; Xu, H.; Wang, Z.X.; Zhang, L.; Zhang, J.M. Guanidinium-based ionic liquids for sulfur dioxide sorption. Chem. Eng. J. 2011, 175, 324–329. [Google Scholar] [CrossRef]
  11. Izgorodina, E.I.; Hodgson, J.L.; Weis, D.C.; Pas, S.J.; MacFarlane, D.R. Physical absorption of CO2 in protic and aprotic ionic liquids: An interaction perspective. J. Phys. Chem. B 2015, 119, 11748–11759. [Google Scholar] [CrossRef] [PubMed]
  12. Karousos, D.S.; Kouvelos, E.; Sapalidis, A.; Pohako-Esko, K.; Bahlmann, M.; Schulz, P.S.; Wasserscheid, P.; Siranidi, E.; Vangeli, O.; Falaras, P.; et al. Novel inverse supported ionic liquid absorbents for acidic gas removal from flue gas. Ind. Eng. Chem. Res. 2016, 55, 5748–5762. [Google Scholar] [CrossRef]
  13. Li, X.S.; Zhang, L.Q.; Zheng, Y.; Zheng, C.G. Effect of SO2 on CO2 absorption in flue gas by ionic liquid 1-ethyl-3-methylimidazolium acetate. Ind. Eng. Chem. Res. 2015, 54, 8569–8578. [Google Scholar] [CrossRef]
  14. 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]
  15. Lei, Z.G.; Dai, C.N.; Chen, B.H. Gas solubility in ionic liquids. Chem. Rev. 2014, 114, 1289–1326. [Google Scholar] [CrossRef] [PubMed]
  16. 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. Ind. Eng. Chem. Res. 2011, 50, 111–118. [Google Scholar] [CrossRef]
  17. 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]
  18. Wang, M.; Zhang, L.Q.; Gao, L.X.; Pi, K.W.; Zhang, J.Y.; Zheng, C.G. Improvement of the CO2 absorption performance using ionic liquid [NH2emim][BF4] and [Emim][BF4]/[Bmim][BF4] mixtures. Energy Fuels 2013, 27, 461–466. [Google Scholar] [CrossRef]
  19. Damas, G.B.; Dias, A.B.A.; Costa, L.T. A quantum chemistry study for ionic liquids applied to gas capture and separation. J. Phys. Chem. B 2014, 118, 9046–9064. [Google Scholar] [CrossRef]
  20. Amarasekara, A.S. Acidic Ionic Liquids. Chem. Rev. 2016, 116, 6133–6183. [Google Scholar] [CrossRef]
  21. Rezaei, F.; Jones, C.W. Stability of supported amine adsorbents to SO2 and NOx in postcombustion CO2 capture. 1. Single-component adsorption. Ind. Eng. Chem. Res. 2013, 52, 12192–12201. [Google Scholar] [CrossRef]
  22. Hallenbeck, A.P.; Kitchin, J.R. Effects of O2 and SO2 on the capture capacity of a primary-amine based polymeric CO2 sorbent. Ind. Eng. Chem. Res. 2013, 52, 10788–10794. [Google Scholar] [CrossRef]
  23. Shiflett, M.B.; Yokozeki, A. Chemical absorption of sulfur dioxide in room-temperature ionic liquids. Ind. Eng. Chem. Res. 2010, 49, 1370–1377. [Google Scholar] [CrossRef]
  24. Shiflett, M.B.; Drew, D.W.; Cantini, R.A.; Yokozeki, A. Carbon dioxide capture using ionic liquid 1-butyl-3-methylimidazolium acetate. Energy Fuels 2010, 24, 5781–5789. [Google Scholar] [CrossRef]
  25. Li, W.; Liu, Y.; Wang, L.; Gao, G. Using ionic liquid mixtures to improve the so2 absorption performance in flue gas. Energy Fuels 2017, 31, 1771–1777. [Google Scholar] [CrossRef]
  26. Lee, K.Y.; Kim, H.S.; Kim, C.S.; Jung, K.D. Behaviors of SO2 absorption in BMIm OAc as an absorbent to recover SO2 in thermochemical processes to produce hydrogen. Int. J. Hydrog. Energy 2010, 35, 10173–10178. [Google Scholar] [CrossRef]
  27. 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] [PubMed]
  28. Zhai, L.Z.; Zhong, Q.; He, C.; Wang, J. Hydroxyl ammonium ionic liquids synthesized by water-bath microwave: Synthesis and desulfurization. J. Hazard. Mater. 2010, 177, 807–813. [Google Scholar] [CrossRef]
  29. Kurnia, K.A.; Mutalib, M.I.A.; Ariwahjoedi, B. Estimation of physicochemical properties of ionic liquids [H2N-C2mim][BF4] and [H2N-C3mim][BF4]. J. Chem. Eng. Data 2011, 56, 2557–2562. [Google Scholar] [CrossRef]
  30. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 16 (Revision A.01); Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  31. Li, H.P.; Chang, Y.H.; Zhu, W.S.; Jiang, W.; Zhang, M.; Xia, J.X.; Yin, S.; Li, H.M. A DFT Study of the Extractive Desulfurization Mechanism by BMIM (+) AlCl4 (-) Ionic Liquid. J. Phys. Chem. B 2015, 119, 5995–6009. [Google Scholar] [CrossRef]
  32. Zhao, Y.; Truhlar, D.G. A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101–194118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Grimme, S.; Hujo, W.; Kirchner, B. Performance of dispersion-corrected density functional theory for the interactions in ionic liquids. Phys. Chem. Chem. Phys. 2012, 14, 4875–4883. [Google Scholar] [CrossRef] [PubMed]
  34. Boys, S.F.; Bernardi, F. The Calculations of Small Molecular Interaction by the Difference of Separate Total Energies. Some Procedures with Reduced Error. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  35. Bernales, V.S.; Marenich, A.V.; Contreras, R.; Cramer, C.J.; Truhlar, D.G. Quantum Mechanical Continuum Solvation Models for Ionic Liquids. J. Phys. Chem. B 2012, 116, 9122–9129. [Google Scholar] [CrossRef] [PubMed]
  36. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
Sample Availability: Samples of the ionic liquids are available from the authors.
Figure 1. SO2 absorption performance of [C4mim][OAc] and [NH2emim][BF4]: (a) Outlet SO2 concentration vs. time at the atmosphere of 15% CO2, 2% SO2, and 83% N2; (b) SO2 absorption capacity at the atmosphere of 2% SO2/98% N2.
Figure 1. SO2 absorption performance of [C4mim][OAc] and [NH2emim][BF4]: (a) Outlet SO2 concentration vs. time at the atmosphere of 15% CO2, 2% SO2, and 83% N2; (b) SO2 absorption capacity at the atmosphere of 2% SO2/98% N2.
Molecules 25 01034 g001
Figure 2. CO2 absorption capacity of [C4mim][OAc] during 6 regeneration cycles at the atmosphere of 15% CO2/85% N2 and 15% CO2/2% SO2/83% N2.
Figure 2. CO2 absorption capacity of [C4mim][OAc] during 6 regeneration cycles at the atmosphere of 15% CO2/85% N2 and 15% CO2/2% SO2/83% N2.
Molecules 25 01034 g002
Figure 3. FT-IR spectra of ionic liquids (ILs): (a) Fresh [C4mim][OAc]; (b) [C4mim][OAc] after SO2 absorption (2% SO2/98% N2); (c) [C4mim][OAc] after CO2 absorption (15% CO2/85% N2); (d) Fresh [NH2emim][BF4]; (e) [NH2emim][BF4] after SO2 absorption (2% SO2/98% N2); (f) [NH2emim][BF4] after CO2 absorption (15% CO2/85% N2).
Figure 3. FT-IR spectra of ionic liquids (ILs): (a) Fresh [C4mim][OAc]; (b) [C4mim][OAc] after SO2 absorption (2% SO2/98% N2); (c) [C4mim][OAc] after CO2 absorption (15% CO2/85% N2); (d) Fresh [NH2emim][BF4]; (e) [NH2emim][BF4] after SO2 absorption (2% SO2/98% N2); (f) [NH2emim][BF4] after CO2 absorption (15% CO2/85% N2).
Molecules 25 01034 g003
Figure 4. CO2 absorption capacity of ionic liquid (IL) mixtures at different X: Flue atmosphere 15% CO2/2% SO2/83% N2; absorption temperature, 293 K.
Figure 4. CO2 absorption capacity of ionic liquid (IL) mixtures at different X: Flue atmosphere 15% CO2/2% SO2/83% N2; absorption temperature, 293 K.
Molecules 25 01034 g004
Figure 5. 1H NMR spectra of ILs: (a) [NH2emim][BF4]; (b) [NH2emim][OAc].
Figure 5. 1H NMR spectra of ILs: (a) [NH2emim][BF4]; (b) [NH2emim][OAc].
Molecules 25 01034 g005
Figure 6. CO2 absorption performance of fresh IL and after SO2 saturated IL: (a) [C4mim][OAc]; (b) [NH2emim][BF4].
Figure 6. CO2 absorption performance of fresh IL and after SO2 saturated IL: (a) [C4mim][OAc]; (b) [NH2emim][BF4].
Molecules 25 01034 g006
Figure 7. Optimized structure of the complexes of IL with CO2/SO2: (a) [NH2emim][BF4]−SO2; (b) [NH2emim][BF4]−CO2; (c) [C4mim][OAc]−SO2; (d) [C4mim][OAc]−CO2.
Figure 7. Optimized structure of the complexes of IL with CO2/SO2: (a) [NH2emim][BF4]−SO2; (b) [NH2emim][BF4]−CO2; (c) [C4mim][OAc]−SO2; (d) [C4mim][OAc]−CO2.
Molecules 25 01034 g007
Figure 8. Optimized structure of the IL mixture and CO2/SO2.
Figure 8. Optimized structure of the IL mixture and CO2/SO2.
Molecules 25 01034 g008
Figure 9. Schematic diagram of the apparatus used for absorption and desorption.
Figure 9. Schematic diagram of the apparatus used for absorption and desorption.
Molecules 25 01034 g009
Table 1. Summary of CO2 absorption capacity by single ionic liquids.
Table 1. Summary of CO2 absorption capacity by single ionic liquids.
Ionic LiquidsT, KGasAbsorption Capacity, mol CO2/mol IL
[C4mim][OAc]29315% CO2/85% N20.298
[C4mim][OAc]29315% CO2/2% SO2/83% N20.204
[NH2emim][BF4]29315% CO2/85% N20.290
[NH2emim][BF4]29315% CO2/2% SO2/83% N20.180
[NH2emim][OAc]29315% CO2/85% N20.291
[NH2emim][OAc]29315% CO2/2% SO2/83% N20.171
Table 2. Structural parameters for the complexes.
Table 2. Structural parameters for the complexes.
Structural ParametersCO2−[NH2emim][BF4]SO2−[NH2emim][BF4][C4mim][OAc]−CO2[C4mim][OAc]−SO2
C−O, Å1.191.21
∠O−C−O, °166146
S−O, Å1.461.49
∠O−S−O, °113.5112.6
Table 3. Thermochemical parameters and charge transfer of the ion−CO2/SO2 complexes.
Table 3. Thermochemical parameters and charge transfer of the ion−CO2/SO2 complexes.
OAc−CO2OAc−SO2OAc−CO2−SO2
ΔE, kJ/mol−40.7−113.4−140.0
ΔH, kJ/mol−46.5−125.1−151.4
ΔG, kJ/mol−1.6−70.8−72.1
net charge transfer, e−0.510−0.382−0.035(CO2)/−0.316(SO2)
CO2−[NH2emim]SO2−[NH2emim]SO2−CO2−[NH2emim]
ΔE, kJ/mol−33.8−123.9−156.8
ΔH, kJ/mol−36.3−126.7−160.1
ΔG, kJ/mol−7.2−55.1−61.9
net charge transfer, e−0.312−0.399−0.308(CO2)/−0.334(SO2)
Table 4. Thermochemical parameters for the IL−CO2/SO2 complexes a.
Table 4. Thermochemical parameters for the IL−CO2/SO2 complexes a.
[C4mim][OAc]−CO2[C4mim][OAc]−SO2CO2−[NH2emim][BF4]SO2−[NH2emim][BF4]
ΔE, kJ/mol−26.4 (−21.9)−80.1 (−70.5)−19.2 (−16.8)−85.4 (−76.3)
ΔH, kJ/mol−30.5−93.2−29.1−91.5
ΔG, kJ/mol−2.7−40.68.1−37.2
a Values of brackets were calculated by the continuum universal solvation model (SMD).

Share and Cite

MDPI and ACS Style

Wu, G.; Liu, Y.; Liu, G.; Pang, X. The CO2 Absorption in Flue Gas Using Mixed Ionic Liquids. Molecules 2020, 25, 1034. https://doi.org/10.3390/molecules25051034

AMA Style

Wu G, Liu Y, Liu G, Pang X. The CO2 Absorption in Flue Gas Using Mixed Ionic Liquids. Molecules. 2020; 25(5):1034. https://doi.org/10.3390/molecules25051034

Chicago/Turabian Style

Wu, Guoqing, Ying Liu, Guangliang Liu, and Xiaoying Pang. 2020. "The CO2 Absorption in Flue Gas Using Mixed Ionic Liquids" Molecules 25, no. 5: 1034. https://doi.org/10.3390/molecules25051034

Article Metrics

Back to TopTop