Carbon Dioxide Solubility in Phosphonium-, Ammonium-, Sulfonyl

Mar 20, 2017 - This work provides insight information regarding CO2 solubility for IL–IL mixing effect pressures up to 10 bar and at 298 K. Density ...
2 downloads 0 Views 1MB Size
Article pubs.acs.org/jced

Carbon Dioxide Solubility in Phosphonium‑, Ammonium‑, Sulfonyl‑, and Pyrrolidinium-Based Ionic Liquids and their Mixtures at Moderate Pressures up to 10 bar Tausif Altamash,† Tamara Shabib Haimour,† Mahsa Ali Tarsad,† Baraa Anaya,‡ Moustafa Hussein Ali,‡ Santiago Aparicio,*,§ and Mert Atilhan*,‡ †

Department of Chemical Engineering, Qatar University, P.O. Box 2713, Doha, Qatar Texas A&M University at Qatar, P.O. Box 23874, Doha, Qatar § Department of Chemistry, University of Burgos, 09001 Burgos, Spain ‡

S Supporting Information *

ABSTRACT: Carbon dioxide solubility in four ionic liquids (ILs) of different families with different cationic−anionic groups (tributylmethylphosphonium formate, butyltrimethylammonium bis(trifluoromethyl sulfonyl) imide, 1-methyl-1-propylpyrrolidinium dicyanamide, and 1ethyl-3-methylimidazolium acetate) at temperature of 298 K and a pressure range from vacuum to 10 bar were studied in this work using state of the art gravimetric sorption experiments. This work provides insight information regarding CO2 solubility for IL−IL mixing effect pressures up to 10 bar and at 298 K. Density values were used to calculate molar volume of ionic liquids for further discussions on CO2 solubility-molar volume relationship. Noticeably higher CO2 solubility with IL−IL hybridized systems of different family is opening a new window for research on a molecular level by simulations and intellectually designed ILs. Chemisorption behavior has been observed for the ILs that contain acetate-based anions in the structure and relevant discussion is included in this work.

when for the first time ever the Brennecke Group16−18 presented CO2 solubility in ILs. However, at that time ILs already were considered as a greener liquid due to nonvolatile, noncorrosive, nonflamable behavior while high thermal stability and recyclability were considered as energy saver properties. In this continuation, the same groups published several researches on a similar topic to promote a new generation of technology as well as provided knowledge of facts and factors on molecular level of CO 2−IL.1,16,19,20 Various families of ILs like imidazolium, pyridinium, and pyrrolidinium have been tested to search for the most potential one.21−25 Enhanced CO2 absorption results have been reported on elongation/branching of the alkyl chain or ether linkages on cation core or fluorination on cation and anion core of IL.26 However, the presence of hydroxyl, nitrile, methyl, alkyl, and an ether group on second C position on the cation core impoverished the CO2 solubility of IL interactions.24 CO2 absorption studies on the molecular level suggested that the anion factor plays a stronger role than the cation, while elongation of the alkyl chain supports the theory of free volume or molar volume.27,28 On the other hand, IR spectroscopy determination supports the theory of Lewis acid−base interaction between the CO2 fluorinated anion of an IL.29,30

1. INTRODUCTION Carbon dioxide (CO2) absorption researches have been initiated since scientists realized the rising level of CO2 in atmosphere is one of the factors for global warming. With emission of CO2 mainly from the industries, burning fuels are most common that dissolved CO2 in Earth’s atmosphere, leading to greenhouse effect and acid rain. Scientific reports stated that CO2 concentration increases from 280 to 400 ppm from the preindustrial era to the year 2013 and in May 2016 concentration increased up to 407.70 ppm.1−3 The increase in CO2 level in atmosphere absolutely changes the temperature globally by around 60% alone in comparison to other factors, and human activity is greatly responsible for this high CO2 emission.4,5 It is also true that continuous effort was taken into action seriously by the scientists toward CO2 absorption technologies to develop methodologies.6−10 Very common technology is amine aqueous solutions or pure ones (volatile organic compounds) being used for CO2 capture that absorb CO2 molecules easily in adequate amount.11 Amines or amine solutions are not considered as an environmentally benign candidate due to corrosive behavior, readily oxidative nature, and evaporation of volatile compounds.12,13 Required energy and its increased energy cost are also another considerable issue in the present scenario of the world. Moreover, CO 2 regeneration from amine-based solvent needs high temperature, which is directly related with energy loss.14,15 Because of their growing interest, ILs came into light as an alternative of amines © 2017 American Chemical Society

Received: September 25, 2016 Accepted: March 14, 2017 Published: March 20, 2017 1310

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data

Article

Table 1. Chemicals Used in this Worka,b

a

Standard uncertainties u are ur(density) = 0.003 and u(temperature) = 0.05 K. N/A: not available. bDensities are reported at 0.1 MPa. cReference 43. dReference 46. eReference 47. fReference48

system of a different family opens a new window for research on molecular level by simulations and intellectually designed ILs.

The main aim of the IL-CO2 studies is to replace tradition amine/alkyl amine or organic volatile solvent from ongoing separation technology and research encouraging us to move forward in this direction as a continuation of our previous works.31−35 Application of IL using technology for CO2 capture on an industrial level seems challenging due to higher cost in comparison with present technology as well as low fluidity of ILs. Therefore, viscosity and physical properties can be altered by playing with structural ionic entities (cation or anion or alkyl chain) according to their tasks. Many authors preparing ammonium- or phosphonium-based ILs with lower cost and CO2 solubility data have been reported in recent years.1,36−39 Moreover, unlike CO2 solubility with single IL, solubility of CO2 possibility with IL−IL mixtures of their ratios are also examined. Furthermore, CO2 solubility has been explained in terms of molar volume or free volume of ILs with the help of molecular dynamic simulation, quantum, or simple calculations. The theory of free volume believes loosely held or weak ionic interactions create a space that is favorable for CO 2 solubility.27,40−42 In the continuation of such research progress, we have attempted to measure the CO2 solubility in four ILs of different families with different cationic−anionic groups at a temperature of 298 K and a pressure range from vacuum to 10 bar. On extending our research thoughts, IL−IL molar mixtures were included to provide better insight comparison of CO2 solubility with pure IL system. In addition, measured densities were used to calculated molar volume of ILs to compare the solubility. Noticeably higher CO2 solubility with a IL−IL hybridized

2. EXPERIMENTAL SECTION 2.1. Materials. CO2 gas was received with a purity of ≥99.99% from Buzware Scientific Technical Gases, Qatar. The ILs used for this work are tributylmethylphosphonium formate ([TBMP] [Formate]), butyltrimethylammonium bis(trifluoromethyl sulfonyl) imide ([N1114][NTf2]), 1-methyl-1propylpyrrolidinium dicyanamide ([PMPy][DCA]), and 1ethyl-3-methylimidazolium acetate ([EMIM][Ac]) and were provided by IoLiTec, Germany, Table 1. Degassed ILs and their mixtures of different molar ratios were prepared by weighing using a Mettler electronic with a precision of ±0.0001 g. Experimental density and other details are listed in Table 1. Water content determinations of sample have been performed with Karl Fischer moisture titrator (Model C20). The water content of all neat and mixed ILs samples is reported in Tables 1 and S1 (Supporting Information). All samples were prepared and handled under nitrogen inert atmosphere to avoid moisture contamination during measurements. 2.2. Density Measurements. Density of ILs were measured at different temperatures using Anton Paar DMA 4500M. This densitometer measures the density by oscillating the U-tube sensor principle and only 1 mL of IL is required to get density results. Densitometer calibration was previously performed with water (Milli-Q, Millipore, resistivity 18.2 MΩ cm) as a calibration fluid together with 14-parameters 1311

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data

Article

calibration equation in order to convert the densitometer oscillating period readings into density and the details of the calibration procedure was described previously.44 For densitometer calibration purpose, the reference density values of water were obtained from the fundamental equation of state by Wagner and Pruss (uncertainty lower than ±0.003% in the full pressure and temperature ranges).45 Moreover a further densitometer calibration was performed by using dimethyl sulfoxide (DMSO), which was measured at different temperature three times and compared with literature values. Temperature was measured in the oscillating U-tube to ±0.05 K. The uncertainty of the pressure measurements in densitometer is 0.005 MPa. The combined relative uncertainty of density measurements considering the samples purity was estimated as 0.3%. Calibration results with DMSO are included in Supporting Information Figure S1 and Table S2. 2.3. CO2 Solubility Measurements. The solubility measurements of carbon dioxide in IL were carried out at temperature 298 K and pressure up to 10 bar from vacuum using high-pressure magnetic suspension sorption apparatus (MSA) provided by Rubotherm Präzisionsmesstechnik GmbH. This apparatus is rated up to high pressure and temperature of 350 bar and 373 K, respectively. The installed pressure transducers of Paroscientific, U.S.A. is able to obtain pressure from vacuum and up to 350 bar with a typical uncertainty of 0.01% of the full scale (u(p) ≈ 0.035 bar). The temperature sensor from Minco PRT, U.S.A. has an accuracy of ±0.5 K (u(T) = 0.05 K) for the temperature measurements. There were two different operation positions. In the first operation position, the measuring cell was filled with gas by which MSA detected change in the weight of the loaded sample fitted in the sample container as the pressurized gas was directly absorb in the sample. In the second position, in situ density data of gas were computed with the help of the absorbed amount of gas into the sample. In order to provide a precise data set as well as consistency, each density value was cross-checked further with REFPROP 9.0.46 All the isothermal absorption/desorption results were collected by pressurizing and depressurizing the gas in MSA apparatus. Absorption/ desorption cycle for each IL are available in Supporting Information Figure S2.

Figure 1. Plot of experimental density versus temperature of ILs ([EMIM][Ac], [TBMP][Formate], and [PMPy][DCA]). Water content is reported in Table 1.

promising candidate and [N1114][NTf2] the least efficient. A close scrutiny of Figure 2 shows a marginal CO2 solubility difference exists between [TBMP][Formate] and [EMIM][Ac] at different pressures. However, overall CO2 solubility in ILs results demonstrates the trend of [TBMP][Formate] > [Emim][Ac] > [PMPy][DCA] > [N1114][NTf2] at different pressures. Further comparison of CO2 solubility in said ionic liquids on the basis of their structures was difficult due to unmatched cation/anion combination with each other of ILs. 3.3. Effect of Structure on CO2 Solubility. CO2 solubility trend in said ILs of this work surprising and challenging the theory CO2−philicity toward fluorinated anion and Lewis acid−base interaction concept given by any authors.29,40,49−52 However, the anion plays a primary role for CO2 solubility determination while the cation plays secondary role. Ramdin et al.53 found higher solubilities in the [BMIM][Tf2N] in comparison with 1-butyl-3-methylimidazolium methylsulfate ([BMIM][MeSO4]) and also reported that tributylmethylphosphonium methylsulfate ([TBMP][MeSO4]) had given a better result than [BMIM][MeSO4]. As far as the CO2 solubility effect of the cation concern for [N1114][NTf2], Jacquemin et al.54 measured the solubilities of CO2 and presented them while searching for the influence of the cations; the trend orders were found to be [N4,1,1,1]+ > [C4mim]+ > [C2mim]+ and most probably were due to the flexibility of the [N4,1,1,1]+. From Figure 2, it has been also found small increment in CO2 solubility with [TBMP][Formate] than [EMIM][Ac]. In this case, we were unable to explain on the basis of the cation/ anion effect until reviewing the Yasaka and Kimura37 work, which reveals CO2−formate has less affinity rather than acetate anion. There was another speculation proposed by several authors that the anion is an important entity to capture the CO2 gas but, for instance, the situation may be altered with increased alkyl chain of cation that may diminish the anion factor. So, [TBMP][Formate] that has the three butyl chain with their cation seems to be a more dominating factor over the anion’s CO2 selectivity. A noticeable point shown by Figure 2 is that the acetate anion of [EMIM][Ac] has the potential to keep CO2 the solubility profile close to the three butyl alkyl chain of [TBMP][Formate]. The molar volume has been used to describe the gas solubility of ILs. Lei et al.24 calculated molar volumes for

3. RESULTS AND DISCUSSION 3.1. Density of Ionic Liquids. Experimental densities of [EMIM][Ac], [PMPy][DCA], and [TBMP][Formate] ILs were measured at different temperatures. Density of each pure IL decreases when increasing the temperature as expected, Figure 1 (Table S1). Density data of ILs at 298 K are listed in Table 1 and compared with available literature values. Measured density values were further used to calculate the molar volume of ILs to understand the effect of CO2 solubility−IL molar volume relationship. 3.2. CO2 Solubility of ILs. The solubility performances of CO2 in four different ILs were investigated experimentally at a temperature of 298 K from 0.00 to 10 bar by using MSA. Experimentally measured CO2 solubility data for [EMIM][Ac], [PMPy][DCA], [N1114][NTf2], and [TBMP][Formate] are listed in Table 2 and the values are plotted for graphical representation as Figure 2. The CO2 concentration−pressure relationship for all ILs shown to be linear (Figure 2) within the range of experimental data measured in this study. The concentration−pressure plot reveals CO2 absorption capacity of ILs, where [TBMP][Formate] were found to be the most 1312

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data

Article

Table 2. Solubilities of CO2 Gas in Ionic Liquids at Temperature 298 Ka [TBMP][Formate] + CO2

[N1114][NTf2] + CO2

[PMPy][DCA] + CO2

[EMIM][Ac] + CO2

P/bar

solubility/(mmol/g)

P/bar

solubility/(mmol/g)

P/bar

solubility/(mmol/g)

P/bar

solubility/(mmol/g)

0.06 0.93 1.91 2.92 3.92 4.92 5.92 6.92 7.92 8.77 9.90

0.0040 0.0898 0.1709 0.2494 0.3272 0.4048 0.4831 0.5611 0.6410 0.7044 0.7833

0.06 0.92 1.93 2.93 3.94 4.94 5.94 6.95 7.95 8.94 9.95

0.0040 0.0542 0.1137 0.1734 0.2345 0.2955 0.3567 0.4195 0.4774 0.5455 0.6105

0.06 0.92 1.93 2.94 3.94 4.93 5.95 6.94 7.94 8.95 9.94

0.0044 0.0634 0.1338 0.2049 0.2760 0.3475 0.4217 0.4949 0.5690 0.6446 0.7142

0.06 0.92 1.92 2.93 3.92 4.92 5.92 6.94 7.94 9.94

0.0043 0.0659 0.1416 0.2233 0.3051 0.3870 0.4672 0.5476 0.6256 0.7790

a

Water content for each ionic liquid is reported in Table 1. bStandard uncertainties u are u(P) = 0.035 bar, u(T) = 0.05 K, and ur(CO2 solubility) = 0.005.

Figure 3. Plot of experimentally measured CO2 solubility (mmol/g IL) at pressure of 10.0 bar versus calculated molar volume at pressure of 10.0 bar and temperature of 298 K. Water content was reported in Table 1.

Figure 2. Experimentally determined solubility of CO2 at different pressures for the ILs ([EMIM][Ac], [PMPy][DCA], and [TBMP][Formate], and [N4111NTf2]) at temperature of 298 K. Water content reported in Table 1.

most popular examples referenced on this mechanism was on 1alkyl-3-methlyimidazolium acetate case.55,56 It has been discussed that the anion of the ILs has superior effect on the interactions with CO2 through the Lewis acid−base interactions57 and the strength of the CO2 solubility is not directly related with the strength of the CO2 gas and anion interactions. In the case of high basicity of the anion part of he IL, the anion (acetate in this case) has an effect on stronger interactions with the C-2 proton on the cation part (imidazolium in this case) and leads to abstract the proton to produce an N-heterocyclic carbane and acetic acid.58 The gravimetric experimental adsorption/desorption cycle figures (Figure S2) for mixtures that contain anions other than acetate do not show any hysteresis and show complete desorption when the applied pressure is released. However, acetate-containing IL mixtures show slight hysteresis and do not completely recover when the applied pressure is released. Thus, these observations lead to chemisorption behavior in the case of the presence of acetate anion in the mixtures. However, this behavior has not been strong enough to result in distinct maximum CO2 capture capacity in comparison to [TBMP][Formate]. The comparison of the results for [EMIM][Ac] reported in this work regarding CO2 solubility with available literature data59,60 shows lower

imidazolium, phosphonium, and ammonium-based ILs at 303.2 K, to estimate gas solubility factor affecting due to molar volume correspond to free volume and responsible for higher solubility. In order to develop better understanding role of alkyl chain justified for high CO2 solubility of [TBMP][Formate]. Calculated molar volume of each IL and their CO2 solubilities in mmol/g was demonstrated in Figure 3. A perusal of Figure 3 and structure of IL at the same time in Table 1 demonstrate the CO2 solubility-IL molar volume relationship for [TBMP][Formate] and [N1114][NTf2]. [TBMP][Formate] and [N1114][NTf2] larger molar volume than others but CO2 solubility values found lesser for [TBMP][Formate] than [N1114][NTf2]. Structurally, in [N1114][NTf2] molar volume major part is [NTf2] anion while molar volume of [TBMP][Formate] major part is covered by three butyl alkyl chain of Phosphonium cation which supports the theory of dominancy of long alkyl chain over anion very well for CO2 absorption. Most of the recent studies on ILs and gas sorption show the dominant behavior of physisorption driven gas sorption mechanisms at high pressures. However, in afew occasions chemisorption with ILs has also been reported and one of the 1313

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data

Article

conformation rearrangement of ions takes place thus cavity formation occurs for gas solubility and selectivity. It may be possible that the IL mixture provides some additional cavities to accommodate gas molecules within rearranged ions. However, a systematic study was required to understand by simulation or spectroscopic methods. Furthermore, in recent years Klähn and Seduraman66 explained CO2 absorptivity by studying molecular dynamic simulation in which weaker ionic interaction facilitates to create the space in between where the CO2 molecule accommodates, and Hu et al.67 experimentally investigated the absorption capacity of CO2 with different nitrate ion-contacting ILs that provided the information of how charge densities of ion are responsible for CO2 absorption. Results for x [TBMP][Formate] + (1 − x) [EMIM][Ac] reported in Figure 4 are remarkably complex. First, it should be remarked that the CO2 absorption capacity of neat [TBMP][Formate] and [EMIM][Ac] are almost the same, thus confirming the prevailing role of chemisorption for these ionic liquids, which is equivalent for both anions, [Formate] and [Ac], with a null role of the type of involved cation, [TBMP] or [EMIM]. Upon mixing of [TBMP][Formate] and [EMIM][Ac], CO2 absorption is reinforced, showing a cooperative effect by the simultaneous presence of both anions being capable of chemisorbing CO2 molecules. Nevertheless, the behavior of x [TBMP][Formate] + (1 − x) [EMIM][Ac] reported in Figure 4 shows instead a single maximum of two maxima, one for [TBMP][Formate] rich mixtures and the other one for [EMIM][Ac] mixtures, which is in remarkable contrast with the behavior reported for the other mixtures in Figure 4. Although the available reported results do not allow one to obtain a detailed explanation of this behavior, it may be hypothesized that considering the very different properties of the involved cations, [TBMP] or [EMIM], in terms of size and shape, the available free volume in the mixtures should develop a pivotal role. The addition of [TBMP][Formate] to [EMIM][Ac] should increase the free volume because of the larger size of the [TBMP] cation in comparison with that in [EMIM][Ac], hence increasing CO2 solubility; hwever, this disrupting effect of [TBMP in [EMIM][Ac] should evolve through a maxima on going to [TBMP][Formate] mixtures, in which a new maxima in CO2 solubility should be expected in this case by the disrupting effect of [EMIM] cation in [TBMP][Formate] rich mixtures. Therefore, the behavior for x [TBMP][Formate] + (1 − x) [EMIM][Ac] reported in Figure 4 is a combination of chemisorption reinforcement and of the increaser of available free volume produced by the considered cations, having effect for both mixtures composition regions.

solubilities for the samples used in this work. Additional comparison of data reported in this work with literature was only possible for [N1114][NTf2],61−64 although data was not reported at the same temperature the data reported in this work show at lower CO2 solubilities. A complete study on the chemisorption mechanism of CO2 via several other functionalized ionic liquids shall be investigated in a separate study. 3.4. Effect of IL-IL Mixtures on CO2 Solubility. In addition to the solubility measurements of four pure IL, binary mixtures of ILs by mole fraction were also investigated to explore the mixing effect on the CO2 solubility. The [N1114][NTf2] + [EMIM][Ac], [N1114][NTf2] + [PMPy][DCA], [PMPy][DCA] + [EMIM][Ac], [TBMP][Formate] + [EMIM][Ac], and [TBMP][Formate] + [PMPy] [DCA] hybridized possibilities were tested. All the IL−IL mixtures were measured at 298 K at pressure of 10 bar of molar composition of 0.2, 0.4, 0.6, and 0.8. Experimental values tabulated in Table S3 of Supporting Information and trends for CO2 solubilities at various molar compositions are illustrated in Figure 4. The trends of CO2 concentration−molar composition

Figure 4. Maximum CO2 uptake trend for IL−IL mixtures at various molar mixing ratios at 10.0 bar and 298 K. Water content is reported in Table S2 (Supporting Information).

relationship of IL−IL mixture as [N1114][NTf2] + [EMIM][Ac] (decreases and stable), [N1114][NTf2] + [PMPy][DCA] (decreases from start to end), [PMPy][DCA] + [EMIM][Ac] (increases and then decreases) and mixed behavior for [TBMP][Formate] + [EMIM][Ac], [TBMP][Formate] + [PMPy][DCA]. Among values of all IL−IL mixtures, the maximum CO2 solubility (0.85 mmol CO2/g) shown on small addition (0.2 mole fraction) of [PMPy][DCA] in [EMIM][Ac] was presented in Figure 5. The density was measured at a temperature of 298 K of 0.2 molar composition of [PMPy][DCA] in [EMIM][Ac] where a decreased value of density was received from pure [EMIM][Ac] (1.09783 g.cm−3) to 0.2 molar composition of [PMPy][DCA] in [EMIM][Ac] (1.0830 g.cm−3). It was quite difficult to estimate the reason for higher CO2 solubility in demonstrated hybridized IL ([PMPy][DCA] + [EMIM][Ac]) in comparison to pure ones with the help of experimental results (Figure 5). However, Liu et al.65 revealed interesting facts while working on electrostatic potential within the free volume space of ionic liquid in the presence of gas and they suggested that as a gas dissolved in ionic liquids

4. CONCLUSIONS The CO2 solubility on four different ionic liquids and their mixtures of various molar mixing rates were experimentally investigated by using state-of-the-art gravimetric sorption experiments. The effect of cation in ILs and the effect of the molar volume of the IL were experimentally investigated in detail for the effects on the CO2 solubility on the studied pure IL and IL−IL mixtures. For pure IL measurements, it was found that [TBMP][Formate] has three butyl chains and their cation seems to be a more dominating factor over the anion’s CO2 selectivity. There is a noticeable point reading [EMIM][Ac] IL, where thev acetate anion has the potential to keep close to a three butyl alkyl chain in CO2 solubility profile. On the other hand, in addition to the solubility measurements for 1314

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data

Article

Figure 5. CO2 uptake profile for ILs and IL−IL mixture at pressure of 10.0 bar and temperature of 298 K. Water content is reported in Table S2 (Supporting Information).



pure IL, binary mixtures of ILs by mole fraction were also measured to explore the mixing effect on the solubility. Among values of an all IL−IL mixture, the maximum CO2 solubility (0.85 mmol CO2/g) was shown on small addition (0.2 mole fraction) of [PMPy][DCA] in [EMIM][Ac]. [EMIM][Ac] containing mixtures shows larger solubilities than those expected from a linear evolution of CO2 solubility with composition (dashed lines in Figure 4). [EMIM][Ac] shows the stronger anion−cation interactions among the studied ILs and thus when a second IL is added they disrupt the [EMIM][Ac] H-bonded network and potentially create voids that allow CO2 fitting whereas on the other side liberating some [Ac] anions for interacting with CO2 molecules. On the contrary, when ILs involving large cations such as [TBMP][Formate] are considered, the addition of a second type of ILs should lead to less disruption in the IL structuring considering that the large alkyl chains are able to fit a second component. Moreover, the addition of a second IL would occupy some of the available excess space, volume reordering, decreasing the free volume available and thus leading to lower CO2 solubility. This result encourages us to collect insight information to understand the possible higher solubility on this mixture and other similar mixtures as future work. Evidence of chemisorption behavior by using gravimetric experiments were observed in case of acetate-containing IL mixtures and a complete study on the chemisorption mechanism of CO2 via several other functionalized ionic liquids shall be investigated in a separate study.

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.6b00833. Densimeter calibration information for the density of DMSO (Figure S1 and Table S2), data table for the experimental CO2 solubility data for IL−IL mixtures, gravimetric CO2 adsorption/desoption experimental data on [Ac] anion including IL mixtures (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (M.A.). *E-mail: [email protected] (S.A). ORCID

Santiago Aparicio: 0000-0001-9996-2426 Mert Atilhan: 0000-0001-8270-7904 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This paper was made possible by the support of Qatar National Research Fund, Undergraduate Research Experience Program (UREP 15-131-2-044) and National Priorities Research Program Grant (NPRP 6-330-2-140). The statements made herein are solely the responsibility of the authors. We also acknowledge Ministerio de Economiá y Competitividad (Spain, project CTQ2013-40476-R) and Junta de Castilla y León (Spain, project BU324U14). 1315

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data



Article

(22) Ramdin, M.; de Loos, T. W.; Vlugt, T. J. H. State-of-the-Art of CO2 Capture with Ionic Liquids. Ind. Eng. Chem. Res. 2012, 51, 8149− 8177. (23) Zhang, X.; Zhang, X.; Dong, H.; Zhao, Z.; Zhang, S.; Huang, Y. Carbon capture with ionic liquids: overview and progress. Energy Environ. Sci. 2012, 5, 6668−6681. (24) Lei, Z.; Dai, C.; Chen, B. Gas Solubility in Ionic Liquids. Chem. Rev. 2014, 114, 1289−1326. (25) MacFarlane, D. R.; Tachikawa, N.; Forsyth, M.; Pringle, J. M.; Howlett, P. C.; Elliott, G. D.; Davis, J. H.; Watanabe, M.; Simon, P.; Angell, C. A. Energy applications of ionic liquids. Energy Environ. Sci. 2014, 7, 232−250. (26) Corvo, M. C.; Sardinha, J.; Casimiro, T.; Marin, G.; Seferin, M.; Einloft, S.; Menezes, S. C.; Dupont, J.; Cabrita, E. J. A Rational Approach to CO2 Capture by Imidazolium Ionic Liquids: Tuning CO2 Solubility by Cation Alkyl Branching. ChemSusChem 2015, 8, 1935− 1946. (27) Shannon, M. S.; Tedstone, J. M.; Danielsen, S. P. O.; Hindman, M. S.; Irvin, A. C.; Bara, J. E. Free Volume as the Basis of Gas Solubility and Selectivity in Imidazolium-Based Ionic Liquids. Ind. Eng. Chem. Res. 2012, 51, 5565−5576. (28) Kilaru, P. K.; Scovazzo, P. Correlations of Low-Pressure Carbon Dioxide and Hydrocarbon Solubilities in Imidazolium-, Phosphonium-, and Ammonium-Based Room-Temperature Ionic Liquids. Part 2. Using Activation Energy of Viscosity. Ind. Eng. Chem. Res. 2008, 47, 910−919. (29) Kazarian, S. G.; Briscoe, B. J.; Welton, T. Combining ionic liquids and supercritical fluids: in situ ATR-IR study of CO2 dissolved in two ionic liquids at high pressures. Chem. Commun. 2000, 2047− 2048. (30) Bhargava, B. L.; Balasubramanian, S. Insights into the Structure and Dynamics of a Room-Temperature Ionic Liquid: Ab Initio Molecular Dynamics Simulation Studies of 1-n-Butyl-3-methylimidazolium Hexafluorophosphate ([bmim][PF6]) and the [bmim][PF6]− CO2 Mixture. J. Phys. Chem. B 2007, 111, 4477−4487. (31) Chong, F. K.; Foo, D. C. Y.; Eljack, F. T.; Atilhan, M.; Chemmangattuvalappil, N. G. Ionic liquid design for enhanced carbon dioxide capture by computer-aided molecular design approach. Clean Technol. Environ. Policy 2015, 17, 1301−1312. (32) Garcia, G.; Atilhan, M.; Aparicio, S. Assessment of DFT methods for studying acid gas capture by ionic liquids. Phys. Chem. Chem. Phys. 2015, 17, 26875−26891. (33) Sanz, V.; Alcalde, R.; Atilhan, M.; Aparicio, S. Insights from quantum chemistry into piperazine-based ionic liquids and their behavior with regard to CO2. J. Mol. Model. 2014, 20, 1−14. (34) Aparicio, S.; Atilhan, M. A Computational Study on Choline Benzoate and Choline Salicylate Ionic Liquids in the Pure State and After CO2 Adsorption. J. Phys. Chem. B 2012, 116, 9171−9185. (35) Aparicio, S.; Atilhan, M. On the Properties of CO2 and Flue Gas at the Piperazinium-Based Ionic Liquids Interface: A Molecular Dynamics Study. J. Phys. Chem. C 2013, 117, 15061−15074. (36) Firaha, D. S.; Kirchner, B. CO2 Absorption in the Protic Ionic Liquid Ethylammonium Nitrate. J. Chem. Eng. Data 2014, 59, 3098− 3104. (37) Yasaka, Y.; Kimura, Y. Effect of Temperature and Water Concentration on CO2 Absorption by Tetrabutylphosphonium Formate Ionic Liquid. J. Chem. Eng. Data 2016, 61, 837−845. (38) 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. (39) Carvalho, P. J.; Á lvarez, V. H.; Marrucho, I. M.; Aznar, M.; Coutinho, J. A. P. High carbon dioxide solubilities in trihexyltetradecylphosphonium-based ionic liquids. J. Supercrit. Fluids 2010, 52, 258−265. (40) Babarao, R.; Dai, S.; Jiang, D.-E. Understanding the High Solubility of CO2 in an Ionic Liquid with the Tetracyanoborate Anion. J. Phys. Chem. B 2011, 115, 9789−9794.

REFERENCES

(1) Seo, S.; Simoni, L. D.; Ma, M.; DeSilva, M. A.; Huang, Y.; Stadtherr, M. A.; Brennecke, J. F. Phase-Change Ionic Liquids for Postcombustion CO2 Capture. Energy Fuels 2014, 28, 5968−5977. (2) Jones, N. Troubling milestone for CO2. Nat. Geosci. 2013, 6, 589−589. (3) ESRL Global Monitoring Division - Global Greenhouse Gas Reference Network. Recent monthly CO2 Mauna Loa. http://www. esrl.noaa.gov/gmd/ccgg/trends/ (Access date: November 30, 2016 and data updated on November 09, 2016 at site). (4) Baiocchi, G.; Minx, J. C. Understanding Changes in the UK’s CO2 Emissions: A Global Perspective. Environ. Sci. Technol. 2010, 44, 1177−1184. (5) Huber, M.; Knutti, R. Anthropogenic and natural warming inferred from changes in Earth/’s energy balance. Nat. Geosci. 2011, 5, 31−36. (6) Pan, M.; Wang, C. Advances in CO2 Capture, Sequestration, and Conversion. ACS Symp. Ser. 2015, 1194, 341−369. (7) Kenarsari, S. D.; Yang, D.; Jiang, G.; Zhang, S.; Wang, J.; Russell, A. G.; Wei, Q.; Fan, M. Review of recent advances in carbon dioxide separation and capture. RSC Adv. 2013, 3, 22739−22773. (8) Espinal, L.; Poster, D. L.; Wong-Ng, W.; Allen, A. J.; Green, M. L. Measurement, Standards, and Data Needs for CO2 Capture Materials: A Critical Review. Environ. Sci. Technol. 2013, 47, 11960−11975. (9) Li, B.; Duan, Y.; Luebke, D.; Morreale, B. Advances in CO2 capture technology: A patent review. Appl. Energy 2013, 102, 1439− 1447. (10) Ben-Mansour, R.; Habib, M. A.; Bamidele, O. E.; Basha, M.; Qasem, N. A. A.; Peedikakkal, A.; Laoui, T.; Ali, M. Carbon capture by physical adsorption: Materials, experimental investigations and numerical modeling and simulations − A review. Appl. Energy 2016, 161, 225−255. (11) Dutcher, B.; Fan, M.; Russell, A. G. Amine-Based CO2 Capture Technology Development from the Beginning of 2013A Review. ACS Appl. Mater. Interfaces 2015, 7, 2137−2148. (12) Gonzalez-Miquel, M.; Bedia, J.; Abrusci, C.; Palomar, J.; Rodriguez, F. Anion Effects on Kinetics and Thermodynamics of CO2 Absorption in Ionic Liquids. J. Phys. Chem. B 2013, 117, 3398−3406. (13) Saiwan, C.; Supap, T.; Idem, R. O.; Tontiwachwuthikul, P. Part 3: Corrosion and prevention in post-combustion CO2 capture systems. Carbon Manage. 2011, 2, 659−675. (14) Haszeldine, R. S. Carbon Capture and Storage: How Green Can Black Be? Science 2009, 325, 1647−1652. (15) Papatryfon, X. L.; Heliopoulos, N. S.; Molchan, I. S.; Zubeir, L. F.; Bezemer, N. D.; Arfanis, M. K.; Kontos, A. G.; Likodimos, V.; Iliev, B. CO2 Capture Efficiency, Corrosion Properties, and Ecotoxicity Evaluation of Amine Solutions Involving Newly Synthesized Ionic Liquids. Ind. Eng. Chem. Res. 2014, 53, 12083−12102. (16) Anderson, J. L.; Dixon, J. K.; Brennecke, J. F. Solubility of CO2, CH4, C2H6, C2H4, O2, and N2 in 1-Hexyl-3-methylpyridinium Bis(trifluoromethylsulfonyl)imide: Comparison to Other Ionic Liquids. Acc. Chem. Res. 2007, 40, 1208−1216. (17) Muldoon, M. J.; Aki, S. N. V. K.; Anderson, J. L.; Dixon, J. K.; Brennecke, J. F. Improving Carbon Dioxide Solubility in Ionic Liquids. J. Phys. Chem. B 2007, 111, 9001−9009. (18) Aki, S. N. V. K.; Mellein, B. R.; Saurer, E. M.; Brennecke, J. F. High-Pressure Phase Behavior of Carbon Dioxide with ImidazoliumBased Ionic Liquids. J. Phys. Chem. B 2004, 108, 20355−20365. (19) 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. (20) Mejía, I.; Stanley, K.; Canales, R.; Brennecke, J. F. On the HighPressure Solubilities of Carbon Dioxide in Several Ionic Liquids. J. Chem. Eng. Data 2013, 58, 2642−2653. (21) Karadas, F.; Atilhan, M.; Aparicio, S. Review on the Use of Ionic Liquids (ILs) as Alternative Fluids for CO2 Capture and Natural Gas Sweetening. Energy Fuels 2010, 24, 5817−5828. 1316

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317

Journal of Chemical & Engineering Data

Article

(41) Zhang, X.; Liu, X.; Yao, X.; Zhang, S. Microscopic Structure, Interaction, and Properties of a Guanidinium-Based Ionic Liquid and Its Mixture with CO2. Ind. Eng. Chem. Res. 2011, 50, 8323−8332. (42) Jeffrey Horne, W.; Shannon, M. S.; Bara, J. E. Correlating fractional free volume to CO2 selectivity in [Rmim][Tf2N] ionic liquids. J. Chem. Thermodyn. 2014, 77, 190−196. (43) Liu, H.; Maginn, E.; Visser, A. E.; Bridges, N. J.; Fox, E. B. Thermal and Transport Properties of Six Ionic Liquids: An Experimental and Molecular Dynamics Study. Ind. Eng. Chem. Res. 2012, 51, 7242−7254. (44) Dávila, M. J.; Alcalde, R.; Atilhan, M.; Aparicio, S. PρT measurements and derived properties of liquid 1-alkanols. J. Chem. Thermodyn. 2012, 47, 241−259. (45) García, B.; Alcalde, R.; Aparicio, S.; Dávila, M. J.; Leal, J. M. Modeling the PVTx Behavior of the N-Methylpyrrolidinone/Water Mixed Solvent. Ind. Eng. Chem. Res. 2004, 43, 3205−3215. (46) Kilaru, P.; Baker, G. A.; Scovazzo, P. Density and Surface Tension Measurements of Imidazolium-, Quaternary Phosphonium-, and Ammonium-Based Room-Temperature Ionic Liquids: Data and Correlations. J. Chem. Eng. Data 2007, 52, 2306−2314. (47) Wandschneider, A.; Lehmann, J. K.; Heintz, A. Surface Tension and Density of Pure Ionic Liquids and Some Binary Mixtures with 1Propanol and 1-Butanol. J. Chem. Eng. Data 2008, 53, 596−599. (48) Quijada-Maldonado, E.; van der Boogaart, S.; Lijbers, J. H.; Meindersma, G. W.; de Haan, A. B. Experimental densities, dynamic viscosities and surface tensions of the ionic liquids series 1-ethyl-3methylimidazolium acetate and dicyanamide and their binary and ternary mixtures with water and ethanol at T = (298.15 to 343.15 K). J. Chem. Thermodyn. 2012, 51, 51−58. (49) Lemmon, E. W.; Huber, M. L.; McLinden, M. O. NIST Standard Reference Database 23: Reference Fluid Thermodynamic and Transport Properties-REFPROP, Version 9.1; Standard Reference Data; NIST: Gaithersburg, MD, 2013. (50) Anthony, J. L.; Anderson, J. L.; Maginn, E. J.; Brennecke, J. F. Anion Effects on Gas Solubility in Ionic Liquids. J. Phys. Chem. B 2005, 109, 6366−6374. (51) Carvalho, P. J.; Á lvarez, V. H.; Marrucho, I. M.; Aznar, M.; Coutinho, J. A. P. High pressure phase behavior of carbon dioxide in 1butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide and 1butyl-3-methylimidazolium dicyanamide ionic liquids. J. Supercrit. Fluids 2009, 50, 105−111. (52) Cadena, C.; Anthony, J. L.; Shah, J. K.; Morrow, T. I.; Brennecke, J. F.; Maginn, E. J. Why Is CO2 So Soluble in ImidazoliumBased Ionic Liquids? J. Am. Chem. Soc. 2004, 126, 5300−5308. (53) Ramdin, M.; Vlugt, T. J. H.; de Loos, T. W. Solubility of CO2 in the Ionic Liquids [TBMN][MeSO4] and [TBMP][MeSO4]. J. Chem. Eng. Data 2012, 57, 2275−2280. (54) Jacquemin, J.; Husson, P.; Majer, V.; Costa Gomes, M. F. Influence of the Cation on the Solubility of CO2 and H2 in Ionic Liquids Based on the Bis(trifluoromethylsulfonyl)imide Anion. J. Solution Chem. 2007, 36, 967−979. (55) Besnard, M.; Cabaço, M. I.; Chávez, F. V.; Pinaud, N.; Sebastião, P. J.; Coutinho, J. A. P.; Danten, Y. On the spontaneous carboxylation of 1-butyl-3-methylimidazolium acetate by carbon dioxide. Chem. Commun. 2012, 48, 1245−1247. (56) Gurau, G.; Rodgigues, H.; Kelley, S.; 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. (57) Kazarian, S. G.; Briscoe, B. J.; Welton, T. Combining ionic liquids and supercritical fluids: in situ ATR-IR study of CO2 dissolved in two ionic liquids at high pressures. Chem. Commun. 2000, 2047− 2048. (58) Seo, S.; DeSilva, M. A.; Brennecke, J. F. Physical Properties and CO2 Reaction Pathway of 1-Ethyl-3-Methylimidazolium Ionic Liquids with Aprotic Heterocyclic Anions. J. Phys. Chem. B 2014, 118, 14870− 14879.

(59) Pinto, A. M.; Rodriguez, H.; Arce, A.; Soto, A. Combined physical and chemical absorption of carbon dioxide in a mixture of ionic liquids. J. Chem. Thermodyn. 2014, 77, 197−205. (60) Shiflett, M. B.; Yokozeki, A. Phase Behavior of Carbon Dioxide in Ionic Liquids: [emim][Acetate], [emim][Trifluoroacetate], and [emim][Acetate] + [emim][Trifluoroacetate] Mixtures. J. Chem. Eng. Data 2009, 54, 108−114. (61) Jacquemin, J.; Husson, P.; Majer, V.; Gomes, M. F. C. Influence of the Cation on the Solubility of CO2 and H2 in Ionic Liquids Based on the Bis(trifluoromethylsulfonyl)imide Anion. J. Solution Chem. 2007, 36, 967−979. (62) Nam, S. G.; Lee, B. C. Solubility of carbon dioxide in ammonium-based ionic liquids: Butyltrimethylammonium bis(trifluoromethylsulfonyl)imide and methyltrioctylammonium bis(trifluoromethylsulfonyl)imide. Korean J. Chem. Eng. 2013, 30, 474− 481. (63) Deng, Y.; Husson, P.; Delort, A.-M.; Besse-Hoggan, P.; Sancelme, M.; Gomes, M. F. C. Influence of an Oxygen Functionalization on the Physicochemical Properties of Ionic Liquids: Density, Viscosity, and Carbon Dioxide Solubility as a Function of Temperature. J. Chem. Eng. Data 2011, 56, 4194−4202. (64) Muldoon, M. J.; Aki, S. N. V. K.; Anderson, J. L.; Dixon, J. K.; Brennecke, J. F. Improving Carbon Dioxide Solubility in Ionic Liquids. J. Phys. Chem. B 2007, 111, 9001−9009. (65) Liu, H.; Zhang, Z.; Bara, J. E.; Turner, C. H. Electrostatic Potential within the Free Volume Space of Imidazole-Based Solvents: Insights into Gas Absorption Selectivity. J. Phys. Chem. B 2014, 118, 255−264. (66) Klähn, M.; Seduraman, A. What Determines CO2 Solubility in Ionic Liquids? A Molecular Simulation Study. J. Phys. Chem. B 2015, 119, 10066−10078. (67) Hu, P.; Zhang, R.; Liu, Z.; Liu, H.; Xu, C.; Meng, X.; Liang, M.; Liang, S. Absorption Performance and Mechanism of CO2 in Aqueous Solutions of Amine-Based Ionic Liquids. Energy Fuels 2015, 29, 6019− 6024.

1317

DOI: 10.1021/acs.jced.6b00833 J. Chem. Eng. Data 2017, 62, 1310−1317