Next Article in Journal
The Application of the Nanofiltration Membrane NF270 for Separation of Fermentation Broths
Next Article in Special Issue
Surface Treatment by Physical Irradiation for Antifouling, Chlorine-Resistant RO Membranes
Previous Article in Journal
The Re-Localization of Proteins to or Away from Membranes as an Effective Strategy for Regulating Stress Tolerance in Plants
Previous Article in Special Issue
The Structure Stability of Metal Diffusion Membrane-Filters in the Processes of Hydrogen Absorption/Desorption
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ionic Liquids-Polymer of Intrinsic Microporosity (PIMs) Blend Membranes for CO2 Separation

by
Giuseppe Ferraro
1,
Carmela Astorino
1,2,
Mattia Bartoli
1,
Alberto Martis
1,3,
Stefania Lettieri
1,3,*,
Candido Fabrizio Pirri
1,3 and
Sergio Bocchini
1,3,*
1
Center for Sustainable Future Technologies (CSFT), Istituto Italiano di Tecnologia (IIT), Via Livorno 60, 10144 Turin, Italy
2
Dipartimento di Chimica Generale ed Organica Applicata, Università di Torino, Corso Massimo D’Azeglio 48, 10125 Turin, Italy
3
Department of Applied Science and Technology, Politecnico di Torino, Corso Duca Degli Abruzzi 24, 10129 Turin, Italy
*
Authors to whom correspondence should be addressed.
Membranes 2022, 12(12), 1262; https://doi.org/10.3390/membranes12121262
Submission received: 13 November 2022 / Revised: 3 December 2022 / Accepted: 6 December 2022 / Published: 13 December 2022
(This article belongs to the Special Issue Membrane Materials and Processes for Liquid and Gas Separation)

Abstract

:
Membranes with high CO2 solubility are essential for developing a separation technology with low carbon footprint. To this end, physical blend membranes of [BMIM][Ac] and [BMIM][Succ] as Ionic Liquids (ILs) and PIM-1 as the polymer were prepared trying to combine the high permeability properties of PIM-1 with the high CO2 solubility of the chosen ILs. Membranes with a PIM-1/[BMIM][Ac] 4/1 ratio nearly double their CO2 solubility at 0.8 bar (0.86 cm3 (STP)/cm3 cmHg), while other ratios still maintain similar solubilities to PIM-1 (0.47 cm3 (STP)/cm3 cmHg). Moreover, CO2 permeability of PIM-1/[BMIM][Ac] blended membranes were between 1050 and 2090 Barrer for 2/1 and 10/1 ratio, lower than PIM-1 membrane, but still highly permeable. The here presented self-standing and mechanically resistant blend membranes have yet a lower permeability compared to PIM-1 yet an improved CO2 solubility, which eventually will translate in higher CO2/N2 selectivity. These promising preliminary results will allow us to select and optimize the best performing PIM-1/ILs blends to develop outstanding membranes for an improved gas separation technology.

1. Introduction

To decrease the anthropogenic climate change due to global greenhouse gas (GHG) emissions [1], which is mostly caused by carbon dioxide (CO2) [2], and to meet the long-term goals of the Paris Agreement (Adoption of the Paris Agreement FCCC/CP/2015/L.9/Rev.1 (UNFCCC, 2015) [3], one of the most promising carbon mitigation strategies is the Carbon Capture and Storage (CCS) [4] being an effective method for the recovery of CO2 from power plant flue gas (CO2/N2) [5], syngas in hydrogen production (CO2/H2), natural gas, and biogas (CO2/CH4) [6]. Membranes have been recently identified as one of the most promising and emerging technologies for the separation processes in carbon sequestration, as they are cost-effective and not require thermal driving force during their use, reducing the energy needs of 90% compared to the cryogenic distillation process which is an energy-intensive separation method (10–15% of total energy consumption) [6]. Therefore, the development of membranes with high CO2 solubility is essential for developing a separation technology with a low carbon footprint.
Selectivity (S) and permeability (P) of the membrane are the main factors that oversee the efficiency of the membrane in gas separation technology. The selectivity is the ability of the membrane to physically, chemically, or by pore size exclusion, separate one species from a mixture, while the permeability is affected by pore size and other surface characteristics of the membranes. The microporosity of the polymer can be finely tuned to enhance their gas diffusion and thus permeability. An example was proposed by Budd and McKeown in 2004 [7] by reporting on a polymer of intrinsic microporosity (PIM) named PIM-1, which microporosity (pore size < 2 nm) is created simply because the highly rigid and contorted molecular structure cannot fill space efficiently when packed. PIMs create self-standing membranes with high surface area (600–900 m2 g−1). In general, all polymers contain some void space or free volume, however, as in the case of PIM, the voids need to be interconnected, enabling the free movement of gas molecules, generating high gas adsorption.
Membranes prepared from PIM-1 show a large area of free volume producing high permeability for many gases [8], especially for CO2 (4000 to 10,000 Barrer) being highly soluble in this polymer [5]. Therefore, PIMs have been largely used in gas separation technology, as they are able to reach high values not only in terms of permeability, but also in selectivity for O2/N2, CO2/CH4 and CO2/N2 [9,10]. PIMs are also polymers that are easy to synthesize, which makes them great candidates for industrial scale-up. However, there is a trade-off between selectivity and permeability, meaning that polymers with higher permeability will generally be less selective and vice versa, as demonstrated by Robeson [11,12]. This is a massive drawback hardly overcoming. This happen mainly because the solubility selectivity is modest. Thus, the improved performance of PIMs over Robeson’s upper bound is linked to the increase of solubilities while not decreasing the permeability [8].
One of the strategies to obtain polymer-based composite materials with improved gas transport properties beyond the Robeson’s upper bound is to combine polymers with fillers (being solid or liquids), resulting in a synergy being able to positively enhance their properties. Although the physical blending approach can have some drawback, such as miscibility and homogeneity of the blend components, it is still one of the most cost effective and fast alternatives to enhance the physical properties of pristine PIMs polymers [13,14,15,16,17]. In particular, an attracting material highly soluble to CO2 and largely studied for CCS [18], are ionic liquids (ILs). ILs have also attracted attentions as a green solvent compared to traditional ones, besides being chemically and thermally stable, having a negligible vapor pressure and for their ease and scalable synthesis [19,20]. ILs could absorb CO2 either physically or chemically [20]. Chemical sorption is always favorite at lower pressure and thus effective ILs that can interact chemically with CO2 were chosen. One of the best ILs for CO2 absorption are imidazolium based ionic liquids having great CO2 absorption properties because the CO2 is believed to form a complex with the acidic C-2 position in the imidazolium ring [21]. The 1-Butyl-3-methylimidazolium (BMIM) cation, due to its longer side chain, is expected to better interact with polymer matrix and thus decrease leaching of the IL during operation. Therefore, 1-Butyl-3-methylimidazolium acetate [BMIM][Ac] and 1-Butyl-3-methylimidazolium succinimidate [BMIM][Succ] were selected as ILs. In particular, succinimidate as an anion is theoretically able to absorb a high amount of CO2 (>1 mol CO2/mol IL). We estimate by preliminary absorption test on pure ILs [BMIM][Ac] and [BMIM][Succ] a CO2 uptake of 7.6 wt% and 10.9 wt%, respectively. However, it is worth highlighting that the capture in [BMIM][Succ], according to literature, should be at least equimolar, however the molar efficiency considerably lower than 1. Most probably, the high viscosity of [BMIM][Succ] hindered gas diffusion and equilibrium condition was not achieved, despite the high solubility of CO2 of IL.
The aim of this work was to prepare physical blend membranes of [BMIM][Ac] and [BMIM][Succ] as ILs, and PIM-1 as the polymer, to combine the high permeability properties of PIM-1 with the high CO2 solubility of the chosen ILs. These preliminary results will allow the selection of the best performing blended membrane for a future investigation on the effect of ILs on the gas selectivity (CO2/N2) of the blends. We expected the novel membranes to have a lower permeability compared to PIM-1 alone membranes, in trade of an improved CO2 solubility and eventually improved CO2/N2 selectivity.

2. Materials and Methods

2.1. Chemicals

All solvents and reagents were used as received without further purification. 2,3,5,6-Tetrafluoroterephtalonitrile were purchased from FluoroChem (Hadfield, UK). ILs [BMIM][Ac] and [BMIM][Succ] were received from Iolitec (Heilbronn, Germany). All the other reagents and solvents were purchased from Merck-Sigma-Aldrich (St. Louis, MO, USA).

2.2. Synthesis of PIM-1

PIM-1 (Figure 1a) was synthetized following a previously reported procedure [7]. Briefly, a mixture of anhydrous K2CO3 (5.53 g, 0.04 mol, FW = 138.205), 3,3,3′,3′-Tetramethyl-1,1′-spirobiindane-5,5′,6,6′-tetraol (3 g, 0.008813 mol, FW = 340.41), and 2,3,5,6-Tetrafluoroterephthalonitrile (1.76 g, 0.008813 mol, FW = 200,096) in dry DMF (60 mL) was stirred at 65 °C for 72 h under N2 atmosphere. The crude was then poured in 400 mL of DI water and the precipitated washed by centrifugation several times in water and methanol. After drying under vacuum overnight a bright yellow solid was obtained in a quantitative yield.

2.3. Membrane Preparation

PIM-1/ILs membranes were prepared by a solvent casting method. PIM-1 was first dissolved in chloroform and stirred for 15 min at 40 °C. Subsequently, IL was added in the solution and stirred for 15 min at 40 °C. PIM-1 and ILs films were prepared with a total 2 wt% polymer concentration in chloroform. The blend ratio PIM-1/IL contents varies from 10/1, 4/1, and 2/1 where IL is [BMIM][Ac] or [BMIM][Succ] (Figure 1a). The blended solution was cast onto a leveled nylon substrate at ambient temperature. The polymer films were formed after the evaporation of the solvent (Figure 1b,c). The resultant membranes were dried at 40 °C under vacuum for at least 16 h to remove the remained traces of solvent. The membranes were labeled as “PIM-1/IL ratio composition”, for example, PIM-1/[BMIM][Ac] 10/1. The thicknesses of the casted films were around 1 µm, while the densities were between 0.979 and 1.955 (g/cm3) (data reported in the Supplementary Materials, Table S1).

2.4. Physicochemical Characterization of Membranes

Thermal degradation analysis (TGA) was performed using a thermogravimetric analyzer NETZSCH TG 209 F3 Tarsus (Selb, Germany). An amount of 10–15 mg of the sample was burned in an alumina pan under airflow (20 mL min−1) with N2 as a protective gas (20 mL min−1) in a temperature range of 30–800 °C (heating rate 10 °C min−1). The TGA spectra and degradation weight losses are reported in the Supplementary Materials (Figure S1, Table S2). Attenuated total reflection infrared spectroscopy (ATR-IR) was employed to characterize the membranes. Measurements were carried out on a Bruker Tensor II Fourier transform spectrophotometer (Billerica, MA, USA). The spectra were acquired by accumulating 64 scans (64 for the background spectrum) in 4000–600 cm−1 range with a resolution of 2 cm−1. The Tg was defined as the midpoint of the heat capacity change observed in the DSC thermogram. FESEM measurements were performed on a Zeiss SupraTM 25 (Oberkochen, Germany). Pure CO2 absorption-desorption measurements were performed using a Surface Measurement System, Dynamic Vapor Sorption (DVS) instrument. The pure CO2 isotherms were carried out at a constant temperature of 40 °C while increasing the pressure up to 0.8 bar (20% of p/p0 increase at each step). The equilibrium criterion condition for each step was chosen as dm/dt = 0.001% min−1. Gas chromatograph (µGC, Inficon Fusion®; Bad Ragaz, Switzerland), equipped with two columns (a 10 m Rt-Molsieve 5A and an 8 m Rt-Q-Bond) and micro thermal conductivity detectors Zeiss SupraTM 25 (Oberkochen, Germany). The thickness of the membranes was measured using a Digimatic indicator (IDC-112B-5) (Mitutoyo, Kawasaki, Japan) with an accuracy of 1 µm. The thickness of the membranes recorded is an average value obtained from at least 25 different points of the membrane.

2.5. Solubility Measurements

Gas permeation characteristics of PIM-1 based blends were studies, especially its selectivity was evaluated being, with the permeability, one of the main factors that indicate their overall efficiency in gas separation technology. The measurements were performed using a Dynamic Vapor Sorption (DVS) instrument following the method previously described [22]. If the diffusion process obeys Fick’s law and the down-stream pressure is much less than the up-stream pressure, the permeability P of the membranes, expressed in Barrer (10−10 cm3 (STP) cm/(cm2 s cmHg)) [23], is given by (1)
PA = DA × SA,
where DA is the average diffusion coefficient, and SA is the solubility of penetrant A in the polymer. Solubility measurements were performed at for four partial gas pressure (0.2, 0.4, 0.6, 0.8 Bar) at 40 °C. The experiments were performed at least two times for each sample at the different pressure values.

2.6. Permeability Measurement

The experimental set-up used is shown in Figure 2. The flow cell consists of two gas chambers separated by the membrane the exposed area is 10 cm2. The chambers were kept at 25 °C and at 1 absolute bar. In the first chamber 20 mL min−1 CO2 was fluxed, while in the second chamber, a flux of 50 mL min−1 was used to convey CO2 to the gas chromatograph (µGC, Inficon Fusion®), equipped with two columns (a 10 m Rt-Molsieve 5A and an 8 m Rt-Q-Bond). Micro thermal conductivity detectors monitored the signal of N2 and CO2. The permeability was calculated through the evaluation of CO2 flux from first chamber to the second chamber divided by membrane surface and multiplied by membrane thickness. The experiments were performed at least two times for each sample at the different pressure values. The results were reported in Barrer (Section 3.3).

3. Results and Discussion

3.1. Physico-Chemical Properties

3.1.1. Attenuated Total Reflectance (ATR)

ATR was performed on blended membranes and bare PIM-1 membrane to prove the successful introduction of ILs within the PIM-1 -based films. All spectra are shown in Figure 3. In the PIM-1 membrane spectrum, the typical peaks of the polymer are present. In particular, the C-O-C stretching peak of the aromatic ethers and the CN stretching of the nitrile group at 2240 cm−1 are present [24]. The imidazolium in-plane ring modes present in both ILs are located at (∼1560 cm−1) for [BMIM][Ac] and [BMIM][Succ] [25]. In Figure 3a, the presence of [BMIM][Succ] in the membrane is confirmed from the presence of the peak at 1558 cm−1 referring to the carbonyl stretching peak of succinimidate anion of [BMIM][Succ]. The appearances of characteristic bands for [SUC]-anion at 1695 cm−1 and 1064 cm−1 are attributed to C=O and C–N [26]. As expected, the increase of the typical ILs peaks intensities is directly linked with the increase of ILs % in the blend.
The imidazolium vibration partially superimposes with the asymmetric stretching (1570 cm−1) of the ATR spectra reported in Figure 3b of the acetate anion of [BMIM][Ac] [27]. The acetate bending is instead at 634 cm−1. The presence of both absorption in PIM-1/[BMIM][Ac] blends are shown, confirming the successful incorporation ILs in PIM-1 polymer.

3.1.2. FESEM

In Figure 4, the FESEM capture of representative of membrane with and without the addition of ILs. PIM-1 showed a smooth surface without appreciable porous network with cracks due to the intrinsically brittleness of the polymeric film as shown in Figure 4a highlight. As reported in Figure 4c, PIM-1 cross section shows a thick ranging from 13 up to 20 µm confirming the absence of diffuse pores. The PIM-1/ILs showed a very different topology characterized by porous with an average diameter size of up to 1.5-µm (Figure 4b). Furthermore, the cross section for the sample shown in Figure 4d shows the presence of small channels inside the film bulk and average thickness comparable with neat PIM-1. The FESEM analysis clearly enlightens the porous formation as a consequence of the addition of ILs without altering the thickness of the specimens. Additional FESEM capture of the here presented membranes are reported in the Supplementary Materials Figure S2.

3.2. Pure CO2 Absorption

The solubility of CO2 in PIM-1 is usually quite high due to the nitrile groups that may be assumed to enhance both the strength of intermolecular forces and, because of their lateral position, the free volume, giving rise to even higher apparent solubilities. The high apparent solubility of gases in PIMs thus it is attributed mainly to their microporous character, which provides a high capacity for gas [10].
Once [BMIM][Succ] is introduced, the CO2 solubility in the blended membranes decrease in a first instance to the following order: 0 > 4/1 > 10/1 > 2/1 (Figure 5a, Table 1).
Again, the blend with the 4/1 of IL has the higher solubility compared to the other blended membranes; we speculate that there is a trade-off between the beneficial effect of ILs in enhancing the solubility and their “negative” limiting the free volume. Indeed, the ILs are filling the empty voids of the PIM-1, thus a decreased diffusion of gas molecules is expected. The reason of this decrease is due to the change in the mechanism that the gas molecules have undergone when passing through a membrane in the presence or absence of the ILs. In fact, the gas molecules mechanism changes from a molecular sieving mechanism [28] to an adsorption and diffusion into a liquid phase one when the ILs is added. Moreover, the high viscosity of [BMIM][Succ] could further limit the diffusion of CO2 in the membrane. Thus, despite the high solubility of CO2 in [BMIM][Succ] that could reach an uptake of 10.9 wt% at CO2 pressure of 1 bar, the inclusion of this ILs into the membrane decreases the total solubility (Figure 5a, Table 1).
On the contrary, the solubility coefficient of CO2 is found to increase significantly with the increment of [BMIM][Ac] from 10/1 to 2/1, compared to pure PIM-1 membrane, with the higher increase in solubility for PIM-1/[BMIM][Ac] 4/1 reaching and surpassing the PIM-1 for high pressures (Figure 5b, Table 1). This result was expected since [BMIM][Ac] has a high solubility specifically for CO2 [29].

3.3. Flow Cell Measurements

The permeability is given from the product of diffusion coefficient for the solubility (Section 2.5. Solubility measurements). Diffusion is expected to decrease with the increase of ILs content being diffusion into the empty microporosity much easier than in the presence of a liquid. In fact, the newly synthesized PIM-1/ILs show lower CO2 permeability coefficients for all blends compared to the pure PIM-1. The addition of [BMIM][Succ] onto PIM-1 generates a decrease in permeability an order of magnitude lower compared to PIM-1 (Table 2), where the permeability increase proportionally to the ILs content.
As the CO2 solubility also plays a role in the permeability values, it is not surprising that the introduction of [BMIM][Succ] into the polymer plays a general “negative” effect, as the solubilities of the membranes containing this ILs are lower than PIM-1 (Table 1). Instead, the overall “negative” effect on the permeability coefficient of [BMIM][Ac] is limited, as the high CO2 solubility (Table 1) of these blends positively contribute to create membranes with great permeability. The optimization of this membrane will need to consider the fine equilibrium between gas diffusion, which is affected by the presence of ILs into the empty voids of PIMs, and the CO2 solubility, which can be enhanced by increasing the amount of ILs. However, there is a limit in the amount of ILs that can be introduced into the blends, as a loss of mechanical properties or ILs bleaching can be envisage.
For comparison, Table 3 shows CO2 permeability values of other PIM-1/ILs membranes published till now. Compared with relevant literature data, the PIM-1/ILs membranes developed in this work show great if not better CO2 permeability.

4. Conclusions

Self-standing and mechanically resistant blend membranes of the ILs [BMIM][Succ] and [BMIM][Ac] and a polymer of intrinsic microporosity (PIM-1) were successfully prepared. The ILs were chosen for their high CO2 solubility, while PIM-1 was chosen for its high gas diffusion coefficient. The ILs used gave no problems of miscibility with PIM-1 and the films obtained have, on qualitative analysis, the necessary properties to be used in gas separation processes. The introduction of ILs changes the properties and morphology of PIM-1, probably partially occupying the microporosities that underlie the gas separation process of this class of polymers. The change in these properties is obviously related to the concentration of the ILs, and in some cases, allows the solubility of CO2 in the composite to be increased. To this end, we demonstrate the positive effect of ILs in enhancing the CO2 solubility, as membranes with PIM-1/[BMIM][Ac] 4/1 ratio nearly doubled their CO2 solubility at 0.8 bar compared to PIM-1 alone, being still highly permeable. However, this was not the case for [BMIM][Succ], as its presence caused a “negative” effect on S and permeability of the blends. In general, the presence of ILs into the membrane still decreases the permeability, however an improved CO2 S is still possible by selecting the finest ILs and by tuning its loading. The here reported preliminary studies will allow us to optimize blended membranes and demonstrated as PIM-1/IL composite plays an important role in improving the gas solubility and probably the overall gas separation performance of PIM-1 membranes.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/membranes12121262/s1, Table S1: Thickness and density values of PIM-1/ILs membrane, Figure S1: Thermal degradation of pure PIM-1 and PIM-1/ILs blended membranes by means of Thermo Gravimetric Analysis (TGA). (A) PIM-1 [BMIM][Ac] 10/1, 4/1, 2/1; (B) PIM-1 [BMIM][Succ] 10/1, 4/1, 2/1, Table S2: Degradation weight losses of PIM-1 and PIM-1/ILs membranes and ILs calculated by Thermo Gravimetric Analysis (TGA), Figure S2: FESEM captures of (a) PIM-1 membranes loaded with BMIM Ac ((b) 10/1, (c) 4/1, (d) 2/1) and of BMIM Succ ((e) 10/1, (f) 4/1, (g) 2/1).

Author Contributions

G.F. and C.A. have worked together and contributed equally to this work. Conceptualization, S.B., S.L. and G.F.; methodology, S.B., S.L. and G.F.; investigation, G.F., C.A., M.B. and A.M.; writing—original draft preparation, S.B. and S.L.; review and editing, S.B. and S.L.; supervision, S.B., S.L. and G.F; funding acquisition, C.F.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministero dello Sviluppo Economico (MISE) and Ministero della Transizione Ecologica (MITE).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lamb, W.F.; Wiedmann, T.; Pongratz, J.; Andrew, R.; Crippa, M.; Olivier, J.G.J.; Wiedenhofer, D.; Mattioli, G.; Khourdajie, A.A.; House, J.; et al. A Review of Trends and Drivers of Greenhouse Gas Emissions by Sector from 1990 to 2018. Environ. Res. Lett. 2021, 16, 073005. [Google Scholar] [CrossRef]
  2. 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. [Google Scholar] [CrossRef]
  3. Grassi, G.; House, J.; Kurz, W.A.; Cescatti, A.; Houghton, R.A.; Peters, G.P.; Sanz, M.J.; Viñas, R.A.; Alkama, R.; Arneth, A.; et al. Reconciling Global-Model Estimates and Country Reporting of Anthropogenic Forest CO2 Sinks. Nat. Clim. Chang. 2018, 8, 914–920. [Google Scholar] [CrossRef]
  4. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon Capture and Storage (CCS): The Way Forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef] [Green Version]
  5. Powell, C.E.; Qiao, G.G. Polymeric CO2/N2 Gas Separation Membranes for the Capture of Carbon Dioxide from Power Plant Flue Gases. J. Memb. Sci. 2006, 279, 1–49. [Google Scholar] [CrossRef]
  6. Sidhikku Kandath Valappil, R.; Ghasem, N.; Al-Marzouqi, M. Current and Future Trends in Polymer Membrane-Based Gas Separation Technology: A Comprehensive Review. J. Ind. Eng. Chem. 2021, 98, 103–129. [Google Scholar] [CrossRef]
  7. Budd, P.M.; Elabas, E.S.; Ghanem, B.S.; Makhseed, S.; McKeown, N.B.; Msayib, K.J.; Tattershall, C.E.; Wang, D. Solution-Processed, Organophilic Membrane Derived from a Polymer of Intrinsic Microporosity. Adv. Mater. 2004, 16, 456–459. [Google Scholar] [CrossRef]
  8. Budd, P.M.; Msayib, K.J.; Tattershall, C.E.; Ghanem, B.S.; Reynolds, K.J.; McKeown, N.B.; Fritsch, D. Gas Separation Membranes from Polymers of Intrinsic Microporosity. J. Memb. Sci. 2005, 251, 263–269. [Google Scholar] [CrossRef]
  9. Budd, P.M.; McKeown, N.B.; Ghanem, B.S.; Msayib, K.J.; Fritsch, D.; Starannikova, L.; Belov, N.; Sanfirova, O.; Yampolskii, Y.; Shantarovich, V. Gas Permeation Parameters and Other Physicochemical Properties of a Polymer of Intrinsic Microporosity: Polybenzodioxane PIM-1. J. Memb. Sci. 2008, 325, 851–860. [Google Scholar] [CrossRef]
  10. Bengtson, G.; Neumann, S.; Filiz, V. Membranes of Polymers of Intrinsic Microporosity (PIM-1) Modified by Poly(Ethylene Glycol). Membranes 2017, 7, 28. [Google Scholar] [CrossRef]
  11. Robeson, L.M. Correlation of Separation Factor versus Permeability for Polymeric Membranes. J. Memb. Sci. 1991, 62, 165–185. [Google Scholar] [CrossRef]
  12. Robeson, L.M. The Upper Bound Revisited. J. Memb. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
  13. Yong, W.F.; Li, F.Y.; Xiao, Y.C.; Li, P.; Pramoda, K.P.; Tong, Y.W.; Chung, T.S. Molecular Engineering of PIM-1/Matrimid Blend Membranes for Gas Separation. J. Memb. Sci. 2012, 407–408, 47–57. [Google Scholar] [CrossRef]
  14. Halder, K.; Khan, M.M.; Grünauer, J.; Shishatskiy, S.; Abetz, C.; Filiz, V.; Abetz, V. Blend Membranes of Ionic Liquid and Polymers of Intrinsic Microporosity with Improved Gas Separation Characteristics. J. Memb. Sci. 2017, 539, 368–382. [Google Scholar] [CrossRef]
  15. Vázquez, M.I.; Romero, V.; Fontàs, C.; Anticó, E.; Benavente, J. Polymer Inclusion Membranes (PIMs) with the Ionic Liquid (IL) Aliquat 336 as Extractant: Effect of Base Polymer and IL Concentration on Their Physical-Chemical and Elastic Characteristics. J. Memb. Sci. 2014, 455, 312–319. [Google Scholar] [CrossRef]
  16. Maziarz, K.M.; Monaco, H.L.; Shen, F.; Ratnam, M. Complete Mapping of Divergent Amino Acids Responsible for Differential Ligand Binding of Folate Receptors alpha and beta. J. Biol. Chem. 1999, 274, 11086–11091. [Google Scholar] [CrossRef] [Green Version]
  17. Chen, W.; Zhang, Z.; Yang, C.; Liu, J.; Shen, H.; Yang, K.; Wang, Z. PIM-Based Mixed-Matrix Membranes Containing MOF-801/Ionic Liquid Nanocomposites for Enhanced CO2 Separation Performance. J. Memb. Sci. 2021, 636, 119581. [Google Scholar] [CrossRef]
  18. Bocchini, S.; Castro, C.; Cocuzza, M.; Ferrero, S.; Latini, G.; Martis, A.; Pirri, F.; Scaltrito, L.; Rocca, V.; Verga, F.; et al. The Virtuous CO2 Circle or the Three Cs: Capture, Cache, and Convert. J. Nanomater. 2017, 2017, 6594151. [Google Scholar] [CrossRef] [Green Version]
  19. Davarpanah, E.; Hernández, S.; Latini, G.; Pirri, C.F.; Bocchini, S. Enhanced CO2 Absorption in Organic Solutions of Biobased Ionic Liquids. Adv. Sustain. Syst. 2020, 4, 1–8. [Google Scholar] [CrossRef] [Green Version]
  20. Latini, G.; Signorile, M.; Rosso, F.; Fin, A.; d’Amora, M.; Giordani, S.; Pirri, F.; Crocellà, V.; Bordiga, S.; Bocchini, S. Efficient and Reversible CO2 Capture in Bio-Based Ionic Liquids Solutions. J. CO2 Util. 2022, 55, 101815. [Google Scholar] [CrossRef]
  21. Hussan, S.K.P.; Thayyil, M.S.; Rajan, V.K.; Antony, A. The Interplay between Charge Transport and CO2 Capturing Mechanism in [EMIM][SCN] Ionic Liquid: A Broadband Dielectric Study. J. Phys. Chem. B 2019, 123, 6618–6626. [Google Scholar] [CrossRef]
  22. Maruccia, E.; Lourenço, M.A.O.; Priamushko, T.; Bartoli, M.; Bocchini, S.; Pirri, F.C.; Saracco, G.; Kleitz, F.; Gerbaldi, C. Nanocast Nitrogen-Containing Ordered Mesoporous Carbons from Glucosamine for Selective CO2 Capture. Mater. Today Sustain. 2022, 17, 100089. [Google Scholar] [CrossRef]
  23. Stern, S.A. The “Barrer” Permeability Unit. J. Polym. Sci. Part A-2 Polym. Phys. 1968, 6, 1933–1934. [Google Scholar] [CrossRef]
  24. Hao, L.; Liao, K.S.; Chung, T.S. Photo-Oxidative PIM-1 Based Mixed Matrix Membranes with Superior Gas Separation Performance. J. Mater. Chem. A 2015, 3, 17273–17281. [Google Scholar] [CrossRef]
  25. Cha, S.; Ao, M.; Sung, W.; Moon, B.; Ahlström, B.; Johansson, P.; Ouchi, Y.; Kim, D. Structures of Ionic Liquid-Water Mixtures Investigated by IR and NMR Spectroscopy. Phys. Chem. Chem. Phys. 2014, 16, 9591–9601. [Google Scholar] [CrossRef]
  26. Wang, H.; Zhu, J.; Tan, L.; Zhou, M.; Zhang, S. Encapsulated Ionic Liquids for CO2 Capture. Mater. Chem. Phys. 2020, 251, 122982. [Google Scholar] [CrossRef]
  27. Besnard, M.; Cabaço, M.I.; Chávez, F.V.; Pinaud, N.; Sebastião, P.J.; Coutinho, J.A.P.; Dantena, Y. On the Spontaneous Carboxylation of 1-Butyl-3-Methylimidazolium Acetate by Carbon Dioxide. Chem. Commun. 2012, 48, 1245–1247. [Google Scholar] [CrossRef]
  28. Zhang, J.; Schott, J.A.; Mahurin, S.M.; Dai, S. Porous Structure Design of Polymeric Membranes for Gas Separation. Small Methods 2017, 1, 1–7. [Google Scholar] [CrossRef]
  29. 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]
  30. Guiver, M.D.; Yahia, M.; Dal-Cin, M.M.; Robertson, G.P.; Garakani, S.S.; Du, N.; Tavajohi, N. Gas Transport in a Polymer of Intrinsic Microporosity (PIM-1) Substituted with Pseudo-Ionic Liquid Tetrazole-Type Structures. Macromolecules 2020, 53, 8951–8959. [Google Scholar] [CrossRef]
Figure 1. (a) Molecular structures of PIM-1(1), 1-Butyl-3-methylimidazolium succinimidate [BMIM][Succ] (2) and 1 -Butyl-3-methylimidazolium acetate [BMIM][Ac] (3); (b) Self-standing PIM-1 membrane; (c) Self-standing PIM-1/[BMIM][Succ] 2/1 membrane.
Figure 1. (a) Molecular structures of PIM-1(1), 1-Butyl-3-methylimidazolium succinimidate [BMIM][Succ] (2) and 1 -Butyl-3-methylimidazolium acetate [BMIM][Ac] (3); (b) Self-standing PIM-1 membrane; (c) Self-standing PIM-1/[BMIM][Succ] 2/1 membrane.
Membranes 12 01262 g001
Figure 2. Experimental set up consisting of gas chromatograph, cell, and mass flow.
Figure 2. Experimental set up consisting of gas chromatograph, cell, and mass flow.
Membranes 12 01262 g002
Figure 3. Attenuated total reflectance (ATR) spectra of blended PIM-1/ILs membranes compared to PIM-1: (a) [BMIM][Succ] spectra; (b) [BMIM][Ac] spectra.
Figure 3. Attenuated total reflectance (ATR) spectra of blended PIM-1/ILs membranes compared to PIM-1: (a) [BMIM][Succ] spectra; (b) [BMIM][Ac] spectra.
Membranes 12 01262 g003
Figure 4. FESEM capture of surface (a) PIM-1, (b) PIM-1/[BMIM][Ac] 10/1 and their cross sections (c,d), respectively.
Figure 4. FESEM capture of surface (a) PIM-1, (b) PIM-1/[BMIM][Ac] 10/1 and their cross sections (c,d), respectively.
Membranes 12 01262 g004
Figure 5. CO2 solubility of blended membranes at four different CO2 pressure of 0.2, 0.4, 0.6, 0.8 bar of [BMIM][Succ] (a) and [BMIM][Ac] (b).
Figure 5. CO2 solubility of blended membranes at four different CO2 pressure of 0.2, 0.4, 0.6, 0.8 bar of [BMIM][Succ] (a) and [BMIM][Ac] (b).
Membranes 12 01262 g005
Table 1. CO2 Solubility coefficients at 0.8 bar pressure derived from the direct sorption experiments.
Table 1. CO2 Solubility coefficients at 0.8 bar pressure derived from the direct sorption experiments.
MembraneCO2 Solubility 1
PIM-10.47
PIM-1/[BMIM][Succ] 10/1 0.22
PIM-1/[BMIM][Succ] 4/10.33
PIM-1/[BMIM][Succ] 2/10.18
PIM-1/[BMIM][Ac] 10/10.32
PIM-1/[BMIM][Ac] 4/10.86
PIM-1/[BMIM][Ac] 2/10.69
1 cm3 (STP)/cm3 cmHg.
Table 2. CO2 Permeability at 1 bar pressure derived from flow cell experiments.
Table 2. CO2 Permeability at 1 bar pressure derived from flow cell experiments.
MembraneCO2 Permeability
(Barrer 1)
PIM-15177
PIM-1/[BMIM][Succ] 10/1275
PIM-1/[BMIM][Succ] 4/1304
PIM-1/[BMIM][Succ] 2/1421
PIM-1/[BMIM][Ac] 10/12090
PIM-1/[BMIM][Ac] 4/1171
PIM-1/[BMIM][Ac] 2/11053
1 1 Barrer = 10−10 cm3 (STP) cm/(cm2 s cmHg)).
Table 3. Comparison of the experimental data with other literature based on PIM-based gas separation membranes.
Table 3. Comparison of the experimental data with other literature based on PIM-based gas separation membranes.
MembraneCO2 Permeability
(Barrer 1)
CO2 Permeability
PIM-1 Ref. (Barrer 1)
Ref.
PIM-15167 (@1 Bar, 40 °C)5167This work
PIM-1/[BMIM][Ac] 10/12090 (@1 Bar, 40 °C)5167This work
PIM-1/[BMIM][Ac] 2/11053 (@1 Bar, 40 °C)5167This work
IL@MOF/PIM−5 %9420 (@4 Bar, 35 °C)4110[17]
PIM-1 -IL1 (pseudo-IL tetrazole-type)1043 (@6.86 Bar, 25 °C)7340[30]
PIM-1/Matrimid (95:5)3355 (@3.5 Bar, 35 °C)3815[13]
PIM-1/Matrimid (50:50)155 (@3.5 Bar, 35 °C)3815[13]
PIM-1 + 5 wt%[C2mim][Tf2N]6650 (@Bar 2, 30 °C)7440[14]
1 1 Barrer = 10−10 cm3 (STP) cm (cm−2 s−1 cmHg−1); 2 unknown.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ferraro, G.; Astorino, C.; Bartoli, M.; Martis, A.; Lettieri, S.; Pirri, C.F.; Bocchini, S. Ionic Liquids-Polymer of Intrinsic Microporosity (PIMs) Blend Membranes for CO2 Separation. Membranes 2022, 12, 1262. https://doi.org/10.3390/membranes12121262

AMA Style

Ferraro G, Astorino C, Bartoli M, Martis A, Lettieri S, Pirri CF, Bocchini S. Ionic Liquids-Polymer of Intrinsic Microporosity (PIMs) Blend Membranes for CO2 Separation. Membranes. 2022; 12(12):1262. https://doi.org/10.3390/membranes12121262

Chicago/Turabian Style

Ferraro, Giuseppe, Carmela Astorino, Mattia Bartoli, Alberto Martis, Stefania Lettieri, Candido Fabrizio Pirri, and Sergio Bocchini. 2022. "Ionic Liquids-Polymer of Intrinsic Microporosity (PIMs) Blend Membranes for CO2 Separation" Membranes 12, no. 12: 1262. https://doi.org/10.3390/membranes12121262

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

Article Metrics

Back to TopTop