Next Article in Journal
Novel Module-Based Membrane Reactor Design Approach for Improved Operability Performance
Previous Article in Journal
A Unified, One Fluid Model for the Drag of Fluid and Solid Dispersals by Permeate Flux towards a Membrane Surface
Previous Article in Special Issue
Recent Progress in the Engineering of Polymeric Membranes for CO2 Capture from Flue Gas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Recent Developments in High-Performance Membranes for CO2 Separation

1
National Energy Technology Laboratory, U.S. Department of Energy, 626 Cochrans Mill Road, Pittsburgh, PA 15236, USA
2
Oak Ridge Institute for Science and Education, Pittsburgh, PA 15236, USA
3
Leidos Research Support Team, 626 Cochrans Mill Road, P.O. Box 10940, Pittsburgh, PA 15236, USA
*
Authors to whom correspondence should be addressed.
Membranes 2021, 11(2), 156; https://doi.org/10.3390/membranes11020156
Submission received: 27 January 2021 / Revised: 19 February 2021 / Accepted: 19 February 2021 / Published: 23 February 2021
(This article belongs to the Special Issue Advanced Membranes for Carbon Capture)

Abstract

:
In this perspective article, we provide a detailed outlook on recent developments of high-performance membranes used in CO2 separation applications. A wide range of membrane materials including polymers of intrinsic microporosity, thermally rearranged polymers, metal–organic framework membranes, poly ionic liquid membranes, and facilitated transport membranes were surveyed from the recent literature. In addition, mixed matrix and polymer blend membranes were covered. The CO2 separation performance, as well as other membrane properties such as film flexibility, processibility, aging, and plasticization, were analyzed.

1. Introduction

Despite the remarkable renewable energy capacities being operated around the world, global CO2 concentrations continue to rise. A set of transformational carbon capture, utilization, and storage (CCUS) technologies are urgently needed that can prevent CO2 from entering the atmosphere, use the captured CO2 for downstream applications, and safely and permanently store it deep under the ground [1]. Membrane gas separation technology has been believed to be one of the most promising technologies to replace the traditional technologies such as amine scrubbing, due to its small footprint, simplicity, and high energy efficiency. To overcome the challenges in lowering carbon capture cost or profitable carbon capture, a membrane material will need to be on or above the Robeson Upper Bound, i.e., in the high permeability/moderate selectivity regime [2]. The state-of-the-art membranes utilized in industry are heavily dependent on conventional membrane materials such as cellulose acetate, polysulfone, polydimethylsiloxane, polyethylene oxide, etc. Applications of these gas separation membranes are clustered around hydrogen recovery and nitrogen separation, methane reforming, and vapor recovery [3]. On the other hand, current CO2 separation membranes have been relatively less preferred and did not surpass the cost efficiency of conventional CO2 separation technologies such as solvents and sorbents. Recent studies suggested that the CO2 permeability for a potential membrane candidate should be at least two orders of magnitude larger than the permeability of conventional membranes to break through [4]. The high CO2 permeability should be also complemented with adequate selectivity over other gases, such as nitrogen, in gas mixtures. Therefore, there is a clear need for more advanced membrane materials to be used CO2 separation. Recently, a growing amount of interest has been concentrated in advanced membrane materials to achieve cost-efficient CO2 separation. A large body of literature has been accumulated in the past decade covering these high-performance membranes. Accordingly, several new classes of advanced membrane materials have been introduced. Membranes include polymers of intrinsic micro porosity (PIMs) [5], thermally rearranged (TR) polymers [6], polyethylene oxide (PEO)-based copolymers [7], poly ionic liquids [8] and facilitated transport membranes [9]. In addition, mixed matrix and polymer blend membranes have been introduced by mixing membranes with versatile types of filler materials and polymers [10]. Although most of the membranes are polymeric materials, more recently, promising results have been also reported for inorganic membranes using metal–organic frameworks [11].

2. Discussion

Facilitated transport membranes are an emerging solution that combines the reversible reaction between CO2 and the carrier with the intrinsic kinetic properties of a membrane process. On the feed interface of the membrane, CO2 reacts with a free amine and forms a carbamate complex, which diffuses through the membrane into the permeate interface of the membrane (Figure 1). With the presence of inert sweeping gas or vacuum on the permeate side, CO2 dissociates from the complex because of the near-zero CO2 partial pressure on the permeate side. Meanwhile, the non-reactive gases, such as N2, H2, and CH4, can only transport through the membrane via a physical solution–diffusion process, in which the gas is adsorbed by the membrane on the feed interface and then diffuses through the membrane along its concentration gradient to the permeate interface where the desorption occurs. Most non-facilitated polymeric membranes separate gases following the solution–diffusion mechanism, based on the differences in solubility and/or diffusivity of gas pairs. A variety of approaches are being explored and studied to improve the permeability and selectivity of facilitated transport membranes, such as tuning the chain flexibility of the hydrogel polymer, increasing the carrier density inside the membrane, etc. This article focuses on the discussion of two highly effective approaches, which represent the most promising future direction of the technology: (1) improving the carrier efficiency by tuning the molecular structure and (2) constructing the ultrafast CO2-selective transport channel.
As a result of the specific reaction between CO2 and the amine carrier, facilitated transport membranes usually possess high selectivity. However, the amine is also the origin of a hydrogen bonding formation that results in a reduction in the diffusivity of the CO2–amine complex. To disrupt the hydrogen-bonding network, a favorable carrier molecular structure is needed. The Hideto Matsuyama research group at Kobe University investigated a variety of amino acid ionic liquid-type carriers with different molecular structures and found that a cyclic amino acid ionic liquid had the smallest viscosity increase upon CO2 absorption owing to the ring structure that limits the available hydrogen atoms to form hydrogen bonding [12]. They later used the prolinate ionic liquid carrier and achieved an ultra-high CO2 permeability of 52,000 Barrer and CO2/N2 selectivity of 8100 at 30 °C, 70% relative humidity (RH) and 0.1 KPa of CO2 partial pressure, showing great promise for the direct capture of CO2 from the atmosphere [13].
The above finding of cyclic amino acid ionic liquid also reveals a fact that the steric hindrance of an amino group can interfere with the hydrogen bonding formation. What is more, according to the CO2–amine chemistry, a sterically hindered amine can augment the CO2-loading capacity because of the difficulty of forming a carbamate complex, thus changing the stoichiometry of the overall CO2–amine reaction. By attaching a bulky group to the amine, scientists can manipulate the steric hindrance of the amine carriers. Ho et al. for the first time introduced the amine steric hindrance effect previously applied on solvent absorption into the solid facilitated transport membrane via the attachment of an isopropyl group to the amino groups of polyallylamine [14]. Considering that polyallylamine is less favorable than polyvinylamine in terms of film-forming ability and CO2 separation performance, Tong and Ho later modified polyvinylamine with methyl groups and significantly improved the CO2/H2 separation performance by three times at 102 °C [15]. In the Ho group’s most recent work, they deprotonated a series of α-aminoacids with different alkyl or hydroxyethyl substituents by 2-(1-piperazinyl)ethylamine, leading to the nonvolatile amine carriers with different degrees of steric hindrance. A remarkably high CO2 permeance of 435 GPU (gas permeation units) along with unprecedented CO2/H2 selectivity of over 500 was achieved at 107 °C and 0.4 atm of CO2 partial pressure, which is well above the Robeson Upper Bound, which is an empirical limit describing the trade-off between permeability and selectivity of polymer membranes.
In addition to the efforts of tuning the amine structure to achieve better separation performance, it is noteworthy that the construction of ultrafast CO2 transport channels in facilitated transport membranes significantly advances this field as well. This is a relatively unexplored area in the facilitated transport membrane literature. Since the porous fillers often come with the drawback of low selectivity, they can negate the most attractive advantage of facilitated transport membrane (i.e., the high selectivity). Another limitation is the poor compatibility of mostly hydrophobic porous fillers with hydrophilic amine materials. Recently though, there have been a few advances in this area. Decoration of those metal–organic frameworks (MOFs) and covalent organic frameworks (COFs) with amine-containing functional groups can achieve preferable CO2 adsorption and good compatibility with the polyamine matrix at the same time. The Zhi Wang research group at Tianjin University put forward a strategy for in situ synthesized polyethyleneimine grafted zeolitic imidazolate frameworks ZIF-8. The redundancy of amino groups in polyethyleneimine (PEI) enabled it to coordinatively react with Zn2+ similar to imidazole, thus grafting it into the crystal structure of ZIF-8. Then, the PEI-g-ZIF-8, with an average pore size of 7.53 Å (CO2 kinetic diameter: 3.3 Å), was dispersed into polyvinylamine and formed a gas transport channel-embedded facilitated transport membrane. By this approach, they obtained CO2 permeance of 1990 GPU and CO2/N2 selectivity of 79.9 for a typical flue gas carbon capture scenario [16]. The idea of grafting polyamine onto organic frameworks was further extended by the Wang group with the development of a COF/polymer hybrid membrane by penetrating polyvinylamine into the oversized pore (1.8 nm) of a two-dimensional COF constructed of triformylbenzene and diaminobenzene. As a result of the reaction between the amino groups of polyvinylamine and the pendant formyl groups on the COF pore walls, part of polyvinylamine was chemically immobilized onto the COF, resulting in a membrane with well-dispersed COFs in the polyvinylamine matrix with good interfacial compatibility. The amino environment of the COF pore walls favored the adsorption–diffusion of CO2 through the pores while excluding other gases such as N2, CH4, and H2 [17]. Recently, a gravity-induced interface self-assembly approach was reported by the Wang group to construct an ultrafast transport channel in facilitated transport membranes [18]. In this work, they innovatively leveraged the different affinities of modified MIL-101(Cr) and polyvinylamine onto the membrane substrate and formed well-dispersed modified MIL-101 nanoparticles in the polyvinylamine membrane during the process of nanoparticle sediment and water evaporation. The membrane demonstrated CO2 permeance of 823 GPU and CO2/N2 selectivity of 242 at 0.5 MPa.
Ionic liquids (IL) as nonvolatile materials have received increasing interest in its applications for advanced gas separation membranes, including CO2 separation from N2 and methane. Recently, Bara et al. developed a new type of poly(IL)-IL membranes via the photopolymerization of acrylated IL monomer with free ILs trapped in the crosslinking network, and they achieved excellent CO2/CH4 selectivity of 119 at 20 °C [19]. It should be noted that the CO2 permeability of 20 Barrer limits its commercial application though the performance is above the 2008 Robeson Upper Bound. The membrane composition was further optimized including using a variety of free ILs with different cations and anions and the CO2 permeability increased to 40 Barrer [20]. An advantage of the application of non-amine IL membranes for flue gas carbon capture is the tolerance to contaminants such as sulfur oxides (SOx). By integrating compatible ionic liquids with poly(ionic liquids), membrane materials with synergistic properties can be obtained. Mittenthal et al. synthesized new ionic polyimides with amenability to thermal processing and film-forming ability. The ionic polyimides were able to spontaneously absorb ILs into their structures, indicating a potential orientation of the polymer chain and self-assembly of the ionic polymer and the ILs [21]. Further reducing the thickness is needed in the future to achieve high permeance. With respect to this point, Noble et al. developed the two-step coating technique for thin film fabrication, based on the property of poly(IL) to spontaneously absorb free ILs, as also found in the above-mentioned work of Mittenthal. With the thin film fabrication technique, they successfully achieved unprecedented CO2 permeance of 6000 GPU and CO2/N2 of 22 [22].
Polymers of intrinsic micro porosity (PIMs) are a new class of porous organic polymers (POPs), which possess high micropore and surface area compared to conventional polymers [23]. The polymer structure of PIMs consists of rigid and contorted monomers creating unprecedented micropores (0.7–1.1 nm) within the polymer chains. In general, the CO2 transport mechanism in PIMs depends on solution–diffusions, considering the pore size distribution. Accordingly, many studies have aimed to exploit these unique pores of PIMs to separate the CO2 as well as other gases such as nitrogen, hydrogen, methane, ethylene, etc. [24]. The gas separation performance of PIM membranes has been unprecedented compared to other rigid polymers, such as matrimid, and over 100 times higher CO2 permeability was achieved by PIMs. The Robeson Upper Bound plot, which depicts the trade-off between CO2 permeability and selectivity, has been well surpassed by numerous PIM-based membranes reported in the literature (Figure 2) [25]. For example, ultra-high CO2 permeable PIMs (PIM-TMN-Trip) have been reported by positioning triptycene in the backbone of the polymer [26]. The paddle-wheel-like structure of triptycene provides a free volume desired for high CO2 permeability.
Different types of bulky groups, including bridged-bicyclic ring systems and Troger’s Base, have also been considered to increase the free volume in PIMs. Polymers: PIM-EA-TB, PIM-TMN-SBI and PIM-Trip-TB showed very high CO2 permeability performance (>10,000 barrer) [27]. However, there are some drawbacks of PIMs that need to be addressed. First, the high CO2 permeability usually results in low to moderate selectivity over N2, CH4, and H2. In order to increase the selectivity, PIMs have been functionalized via polar groups such as tetrazole, triazine, amine, amide, etc. [28]. For example, Du et al. improved the CO2/N2 selectivity of PIM-1 over two folds by tetrazole functionalization (TZPIM) [29]. Nonetheless, the selectivity enhancement was accompanied by a great decrease in CO2 permeability in TZPIM, which is a typical case for membranes. Another strategy of boosting CO2 selectivity is to blend PIMs with highly CO2 selective polymers such as polyethylene glycol, matrimid, torlon, polysulfone, etc. The blending polymer usually exhibits higher CO2/N2 selectivity compared to PIMs. Similar to the functionalized PIMs, the selectivity enhancement has been reported with the following loss in CO2 permeability [5]. More recently, polyphosphazenes with ether side chains were blended with PIM-1 [30]. Improved CO2/N2 selectivity with high CO2 permeability was reported. Moreover, these blend membranes offered a solution to the brittle nature of PIM membrane films, which is another drawback reported for PIMs. The last major problem encountered in PIMs, as well as other glassy membranes, is aging [31]. The high CO2 permeability of freshly cast PIM membranes decreases dramatically over time. Structurally, PIMs yield a high free volume mostly from the inefficient packing of their polymer chains. Over time, the polymer chains relax; in other words, they reach the thermodynamic equilibrium that results in a great decrease in the free volume. Numerous studies have managed to impede the aging in PIM-based membranes. Aging-resistant membranes were fabricated by mixing or crosslinking PIMs with porous fillers such as MOFs and POPs [22,32]. Traditionally, adding a filler in membranes causes a selectivity drop in membranes, and most of the PIM-based mixed matrix membranes (MMMs) have followed this trend. However, several studies overcame the trade-off by maintaining high CO2 selectivity with significant CO2 permeability [3]. While some membranes achieved this by crosslinking [33], others added a third component (another polymer) to create blend polymer-based MMMs [34,35]. More importantly, some of these membranes showed great improvement in decreasing the aging. At this point, it should be noted that virtually all these membranes are thick films (<30 microns), and the aging phenomena in membranes should be investigated by considering practical applications, where membrane thickness is below one micron. It is well experienced that as the thickness of the membranes goes down, the aging in membranes accelerates. For example, PIM-1 was fabricated into a thick MMM by using a high surface area POP-type filler (PAF-1) [36]. Polymer chains of PIM-1 have been reported to be intercalated in the large pore of PAF-1, impeding the chain relaxation of PIM and thus the aging. However, Bakhtin et al. systematically compared the thickness effect in membranes and found that PIM/PAF-1 membranes also suffer from aging when fabricated in thin films [37].
Thermally rearranged (TR) polymers have emerged as another class of advanced gas separation membranes [38]. Following PIMs, TR polymers can also be categorized as a porous organic polymer given their high porosity [39]. The pores and free volume in TR polymers are created through thermal treatment: TR polymers are synthesized by the molecule rearrangement of polyimide and/or polyamides with ortho-functional groups (-OH, NH2 etc.) at high temperatures (400–600 °C). Consequently, a heterocyclic ring closure replaces imide and amide groups, creating more rigid polymers such as polybenzoxazoles and polybenzimidazoles. Similar to PIMs, rigid and bulky functional groups provide a free volume for CO2 molecules to permeate. In addition, heterocyclic rings such as benzoimidazole and benzoxazole bestow higher interaction with CO2 molecules compared to less polar gases [40]. Due to the microporous nature of TR polymers, the solution–diffusion mechanism directs the CO2 permeation. Hence, the conversion of polyimides into polybenzoxazole favors both the solubility and diffusivity of CO2 in the membranes. Recent studies have focused on further increasing this free volume in TR polymers. There are two major synthetic strategies that have been used: (i) introducing bulky functional groups into polymer chains, and (ii) crosslinking the polymer chains to adjust the pore opening and free volume. For example, TR polymers prepared from the crosslinking of polyimide with 1,4-butylene glycol evinced advanced CO2 permeability (8000 barer) with relatively high CO2/CH4 (17) selectivity. These reached high CO2/CH4 selectivity and at the same time maintained the high permeability rarely encountered in membranes. While TR polymers offer other desired membrane properties such as high chemical and thermal stability, resistance to plasticization, and tunable free volume, they also suffer from poor mechanical properties and competitive sorption between water and CO2. The high temperature used in TR polymer preparation rigidifies the polymer chains, and the membranes become very brittle. On the other hand, other polar gases such as water can compete with CO2 over interacting with the functional groups of TR polymers. This competition usually results in a CO2 permeability drop. Membranes fabricated from TR polymers also show aging when cast in thin films, following the fate of other porous polymeric membranes. Unlike PIMs, MMM fabrication based on TR polymers is very difficult, because the thermally crosslinked polymer chains are not soluble in common solvents.
Metal–organic frameworks have been a frontrunner material to be used as a filler particle in mixed matrix membranes. MOFs consist of crystalline two and three-dimensional structures with permanent porosity and unprecedented surface area. Pores of MOFs are easily adjustable and functionable, which make them a strong candidate to be used in various gas separation and gas capture applications. More recently, MOFs have been also considered as a gas separation membrane [41]. Pure MOF membranes have been fabricated on a support by several methods including direct growth, secondary growth using MOF seeds, and in situ growth [42]. In general, the pore-sized distribution of MOFs is in the microporous region (<2 nm). Therefore, the CO2 transport mechanism relies mostly on the solubility of CO2, which is similar to PIM and TR polymer-based membranes. Therefore, the pore structure of MOFs has been adjusted to increase CO2 permeability. A well-known class of MOFs, zeolitic imidazolium frameworks (ZIFs), has been the most studied MOF for the fabrication of pure MOF membranes. The main reason is the facile synthesis of ZIFs under mild conditions (room temperature, water solvent, etc.). For example, ZIF-8 membranes grown on hollow fibers (ZIF-8-HF) showed high CO2/N2 selectivity (52) [43]. However, these membranes suffered from low CO2 permeabilities along with other ZIF-based MOF membranes. The continuous flow membrane fabrication on a fiber showed the feasibility of large-scale production, which is usually a hurdle for MOF membrane fabrication. There are also other types of MOFs such as HKUST-1 and SIFSIX-3-Ni, which exceeded the CO2 permeability of ZIF membranes [44]. MOF membranes stand as an advanced gas separation membrane candidate. However, there is still big room to improve MOF membranes, as they mostly suffer from the trade-off between CO2 permeability and selectivity. In addition, other important membrane properties such as mechanical strength, MOF penetration into the support, and defects should be taken into account.
In addition to the glassy polymers mentioned so far, several new high-performance membranes have been also fabricated by using rubbery polymers. Remarkably, rubbery polymers do not suffer the physical aging in thin films that commonly exist in glassy polymers. The high flexibility of the rubbery polymer chains accounts for the large gas diffusivity. Ether oxygen has a strong affinity to CO2, endowing the ether oxygen-containing rubbery material with great potential to surpass the Robeson Upper Bound. In 2019, a group of ether oxygen-containing oligomers with the O/C ratio of 0.71, which is much greater than the 0.5 in polyethylene oxide, was reported by Lin’s group of the University of Buffalo [45]. This work has enlightened a new direction of molecular design for CO2-philic rubbery membranes. Another type of emerging rubbery polymers used in CO2 separation is polyphosphazenes [46]. Polyphosphazenes consist of a phosphazene backbone, which provides both flexibility and tunable chemical structure by the substitution of CO2 attractive groups such as ether. More importantly, polyphosphazene membranes do not show crystal domains as in the other ether-containing polymers, such as polyethylene glycol, which limit the CO2 permeability. Although these polymers have been offered for gas separation before, it was only very recently that high CO2 separation performance was achieved by crosslinking [47] and blending with PIMs [30].

3. Conclusions

Membrane technology has become a competitive alternative technology for carbon capture to the conventional amine scrubbing technology. Recent developments in facilitated transport, porous polymer, MOF-based and PEO-based membranes demonstrated great potential for membrane technology to achieve highly energy-efficient and cost-effective carbon capture. Future challenges to improve the membrane-based carbon capture lie in the thin film composite (TFC) membrane and its module fabrication as well as rational process design. The MOF-based membrane in particular faces challenges in TFC fabrication and membrane scaling up, including MOF material scaling up. Some successful TFC examples have been discussed in this review. More TFC study is expected to convert the high permeability of MOF-based bulk membranes into the excellent permeance of TFC in the near future. A rational process design is helpful to fully leverage the performance characteristics of different membrane systems. For instance, a facilitated transport membrane process does not need a moisture removal process prior to the membrane separator and is more suitable with permeate vacuum instead of air sweep. Last but not the least, in the application of carbon capture, the presence of contaminants should also be taken into consideration. In this regard, field testing should be included as an important part of the development loop of membrane technology.

Funding

This work was funded by the Department of Energy, National Energy Technology Laboratory, an agency of the United States Government, through a support contract with Leidos Research Support Team (LRST). Neither the United States Government nor any agency thereof, nor any of their employees, nor LRST, nor any of their employees, makes any warranty, expressed or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This work was performed in support of the US Department of Energy’s Fossil Energy Carbon Capture Research Program. The Research was executed through the NETL Research and Innovation Center’s Transformational Carbon Capture Field Work Proposal. Research performed by Leidos Research Support Team staff was conducted under the RSS contract 89243318CFE000003.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rivero, J.R.; Panagakos, G.; Lieber, A.; Hornbostel, K. Hollow fiber membrane contactors for post-combustion carbon capture: A review of modeling approaches. Membranes 2020, 10, 382. [Google Scholar] [CrossRef] [PubMed]
  2. Kárászová, M.; Zach, B.; Petrusová, Z.; Červenka, V.; Bobák, M.; Šyc, M.; Izák, P. Post-Combustion Carbon Capture by Membrane Separation, Review. Sep. Purif. Technol. 2020, 238, 116448. [Google Scholar] [CrossRef]
  3. Galizia, M.; Chi, W.S.; Smith, Z.P.; Merkel, T.C.; Baker, R.W.; Freeman, B.D. 50th anniversary perspective: Polymers and mixed matrix membranes for gas and vapor separation: A review and prospective opportunities. Macromolecules 2017, 50, 7809–7843. [Google Scholar] [CrossRef]
  4. Liang, C.Z.; Chung, T.-S.; Lai, J.-Y. A review of polymeric composite membranes for gas separation and energy production. Prog. Polym. Sci. 2019, 97, 101141. [Google Scholar] [CrossRef]
  5. Ma, C.; Urban, J.J. Polymers of Intrinsic Microporosity (PIMs) gas separation membranes: A mini review. Proc. Nat. Res. Soc. 2018, 2, 02002. [Google Scholar] [CrossRef]
  6. Lee, W.H.; Seong, J.G.; Hu, X.; Lee, Y.M. Recent progress in microporous polymers from thermally rearranged polymers and polymers of intrinsic microporosity for membrane gas separation: Pushing performance limits and revisiting trade-off lines. J. Polym. Sci. 2020, 58, 2450–2466. [Google Scholar] [CrossRef]
  7. Wang, M.; Zhao, J.; Wang, X.; Liu, A.; Gleason, K.K. Recent progress on submicron gas-selective polymeric membranes. J. Mater. Chem. A 2017, 5, 8860–8886. [Google Scholar] [CrossRef]
  8. Dai, Z.; Noble, R.D.; Gin, D.L.; Zhang, X.; Deng, L. Combination of ionic liquids with membrane technology: A new approach for CO2 separation. J. Membr. Sci. 2016, 497, 1–20. [Google Scholar] [CrossRef]
  9. Han, Y.; Ho, W.S.W. Recent advances in polymeric facilitated transport membranes for carbon dioxide separation and hydrogen purification. J. Polym. Sci. 2020, 58, 2435–2449. [Google Scholar] [CrossRef]
  10. Vinoba, M.; Bhagiyalakshmi, M.; Alqaheem, Y.; Alomair, A.A.; Pérez, A.; Rana, M.S. Recent progress of fillers in mixed matrix membranes for CO2 separation: A review. Sep. Purif. Technol. 2017, 188, 431–450. [Google Scholar] [CrossRef]
  11. Qian, Q.; Asinger, P.A.; Lee, M.J.; Han, G.; Rodriguez, M.K.; Lin, S.; Benedetti, F.M.; Wu, A.X.; Chi, W.S.; Smith, Z.P. MOF-based membranes for gas separations. Chem. Rev. 2020, 120, 8161–8266. [Google Scholar] [CrossRef] [PubMed]
  12. Kamio, E.; Matsuki, T.; Kasahara, S.; Matsuyama, H. The effect of chemical structures of cyclic amino acid type ionic liquids as CO2 carriers on facilitated transport membrane performances. Sep. Sci. Technol. 2017, 52, 209–220. [Google Scholar] [CrossRef]
  13. Moghadam, F.; Kamio, E.; Matsuyama, H. High CO2 separation performance of amino acid ionic liquid-based double network ion gel membranes in low CO2 concentration gas mixtures under humid conditions. J. Membr. Sci. 2017, 525, 290–297. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Winston Ho, W.S. Steric hindrance effect on amine demonstrated in solid polymer membranes for CO2 transport. J. Membr. Sci. 2012, 415–416, 132–138. [Google Scholar] [CrossRef]
  15. Tong, Z.; Ho, W.S.W. New sterically hindered polyvinylamine membranes for CO2 separation and capture. J. Membr. Sci. 2017, 543, 202–211. [Google Scholar] [CrossRef]
  16. Gao, Y.; Qiao, Z.; Zhao, S.; Wang, Z.; Wang, J. In Situ synthesis of polymer grafted ZIFs and application in mixed matrix membrane for CO2 separation. J. Mater. Chem. A 2018, 6, 3151–3161. [Google Scholar] [CrossRef]
  17. Cao, X.; Wang, Z.; Qiao, Z.; Zhao, S.; Wang, J. Penetrated COF channels: Amino environment and suitable size for CO2 preferential adsorption and transport in mixed matrix membranes. ACS Appl. Mater. Interfaces 2019, 11, 5306–5315. [Google Scholar] [CrossRef]
  18. Wang, B.; Qiao, Z.; Xu, J.; Wang, J.; Liu, X.; Zhao, S.; Wang, Z.; Guiver, M.D. Unobstructed ultrathin gas transport channels in composite membranes by interfacial self-assembly. Adv. Mater. 2020, 32, 1907701. [Google Scholar] [CrossRef] [PubMed]
  19. Kammakakam, I.; Bara, J.E.; Jackson, E.M.; Lertxundi, J.; Mecerreyes, D.; Tome, L.C. Tailored CO2-philic anionic poly(ionic liquid) composite membranes: Synthesis, characterization, and gas transport properties. ACS Sustain. Chem. Eng. 2020, 8, 5954–5965. [Google Scholar] [CrossRef]
  20. Kammakakam, I.; Bara, J.E.; Jackson, E.M. Dual anion-cation crosslinked poly(ionic liquid) composite membranes for enhanced CO2 separation. ACS Appl. Polym. Mater. 2020, 2, 5067–5076. [Google Scholar] [CrossRef]
  21. Mittenthal, M.S.; Flowers, B.S.; Bara, J.E.; Whitley, J.W.; Spear, S.K.; Roveda, J.D.; Wallace, D.A.; Shannon, M.S.; Holler, R.; Martens, R.; et al. Ionic polyimides: Hybrid polymer architectures and composites with ionic liquids for advanced gas separation membranes. Ind. Eng. Chem. Res. 2017, 56, 5055–5069. [Google Scholar] [CrossRef]
  22. Zhou, J.; Mok, M.M.; Cowan, M.G.; McDanel, W.M.; Carlisle, T.K.; Gin, D.L.; Noble, R.D. High-permeance room-temperature ionic-liquid-based membranes for CO2/N2 separation. Ind. Eng. Chem. Res. 2014, 53, 20064–20067. [Google Scholar] [CrossRef]
  23. McKeown, N.B.; Budd, P.M. Polymers of Intrinsic Microporosity (PIMs): Organic materials for membrane separations, heterogeneous catalysis and hydrogen storage. Chem. Soc. Rev. 2006, 35, 675. [Google Scholar] [CrossRef]
  24. McKeown, N.B. Polymers of Intrinsic Microporosity (PIMs). Polymer 2020, 202, 122736. [Google Scholar] [CrossRef]
  25. Robeson, L.M. The upper bound revisited. J. Membr. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
  26. Yin, Y.; Guiver, M.D. Ultrapermeable membranes. Nat. Mater. 2017, 16, 880–881. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, J.; Hou, X.; Park, H.B.; Lin, H. High-performance polymers for membrane CO2 /N2 separation. Chem. Eur. J. 2016, 22, 15980–15990. [Google Scholar] [CrossRef] [PubMed]
  28. Yanaranop, P.; Santoso, B.; Etzion, R.; Jin, J. Facile conversion of nitrile to amide on Polymers of Intrinsic Microporosity (PIM-1). Polymer 2016, 98, 244–251. [Google Scholar] [CrossRef]
  29. Du, N.; Park, H.B.; Robertson, G.P.; Dal-Cin, M.M.; Visser, T.; Scoles, L.; Guiver, M.D. Polymer nanosieve membranes for CO2-capture applications. Nat. Mater. 2011, 10, 372–375. [Google Scholar] [CrossRef]
  30. Sekizkardes, A.K.; Kusuma, V.A.; McNally, J.S.; Gidley, D.W.; Resnik, K.; Venna, S.R.; Hopkinson, D. Microporous polymeric composite membranes with advanced film properties: Pore intercalation yields excellent CO2 separation performance. J. Mater. Chem. A 2018, 6, 22472–22477. [Google Scholar] [CrossRef]
  31. Low, Z.-X.; Budd, P.M.; McKeown, N.B.; Patterson, D.A. Gas permeation properties, physical aging, and its mitigation in high free volume glassy polymers. Chem. Rev. 2018, 118, 5871–5911. [Google Scholar] [CrossRef]
  32. Sekizkardes, A.K.; Kusuma, V.A.; Dahe, G.; Roth, E.A.; Hill, L.J.; Marti, A.; Macala, M.; Venna, S.R.; Hopkinson, D. Separation of carbon dioxide from flue gas by mixed matrix membranes using dual phase microporous polymeric constituents. Chem. Commun. 2016, 52, 11768–11771. [Google Scholar] [CrossRef]
  33. Yu, G.; Zou, X.; Sun, L.; Liu, B.; Wang, Z.; Zhang, P.; Zhu, G. Constructing connected paths between UiO-66 and PIM-1 to improve membrane CO2 separation with crystal-like gas selectivity. Adv. Mater. 2019, 31, 1806853. [Google Scholar] [CrossRef] [PubMed]
  34. Elsaidi, S.K.; Venna, S.; Sekizkardes, A.K.; Steckel, J.A.; Mohamed, M.H.; Baker, J.; Baltrus, J.; Hopkinson, D. Custom formulation of multicomponent mixed-matrix membranes for efficient post-combustion carbon capture. Cell Rep. Phys. Sci. 2020, 1, 100113. [Google Scholar] [CrossRef]
  35. Muldoon, P.F.; Venna, S.R.; Gidley, D.W.; Baker, J.S.; Zhu, L.; Tong, Z.; Xiang, F.; Hopkinson, D.P.; Yi, S.; Sekizkardes, A.K.; et al. Mixed matrix membranes from a microporous polymer blend and nanosized metal–organic frameworks with exceptional CO2/N2 separation performance. ACS Mater. Lett. 2020, 2, 821–828. [Google Scholar] [CrossRef]
  36. Lau, C.H.; Nguyen, P.T.; Hill, M.R.; Thornton, A.W.; Konstas, K.; Doherty, C.M.; Mulder, R.J.; Bourgeois, L.; Liu, A.C.Y.; Sprouster, D.J.; et al. Ending aging in super glassy polymer membranes. Angew. Chem. Int. Ed. 2014, 53, 5322–5326. [Google Scholar] [CrossRef] [Green Version]
  37. Bakhtin, D.S.; Kulikov, L.A.; Legkov, S.A.; Khotimskiy, V.S.; Levin, I.S.; Borisov, I.L.; Maksimov, A.L.; Volkov, V.V.; Karakhanov, E.A.; Volkov, A.V. Aging of thin-film composite membranes based on PTMSP loaded with porous aromatic frameworks. J. Membr. Sci. 2018, 554, 211–220. [Google Scholar] [CrossRef]
  38. Park, H.B.; Jung, C.H.; Lee, Y.M.; Hill, A.J.; Pas, S.J.; Mudie, S.T.; Van Wagner, E.; Freeman, B.D.; Cookson, D.J. Polymers with cavities tuned for fast selective transport of small molecules and ions. Science 2007, 318, 254–258. [Google Scholar] [CrossRef]
  39. Wang, S.; Li, X.; Wu, H.; Tian, Z.; Xin, Q.; He, G.; Peng, D.; Chen, S.; Yin, Y.; Jiang, Z.; et al. Advances in high permeability polymer-based membrane materials for CO2 separations. Energy Environ. Sci. 2016, 9, 1863–1890. [Google Scholar] [CrossRef]
  40. Rabbani, M.G.; Islamoglu, T.; El-Kaderi, H.M. Benzothiazole- and benzoxazole-linked porous polymers for carbon dioxide storage and separation. J. Mater. Chem. A 2017, 5, 258–265. [Google Scholar] [CrossRef]
  41. Hu, Y.; Dong, X.; Nan, J.; Jin, W.; Ren, X.; Xu, N.; Lee, Y.M. Metal–organic framework membranes fabricated via reactive seeding. Chem. Commun. 2011, 47, 737–739. [Google Scholar] [CrossRef]
  42. Li, W.; Yang, Z.; Zhang, G.; Fan, Z.; Meng, Q.; Shen, C.; Gao, C. Stiff metal–organic framework–polyacrylonitrile hollow fiber composite membranes with high gas permeability. J. Mater. Chem. A 2014, 2, 2110–2118. [Google Scholar] [CrossRef]
  43. Marti, A.M.; Wickramanayake, W.; Dahe, G.; Sekizkardes, A.; Bank, T.L.; Hopkinson, D.P.; Venna, S.R. Continuous flow processing of ZIF-8 membranes on polymeric porous hollow fiber supports for CO2 capture. ACS Appl. Mater. Interfaces 2017, 9, 5678–5682. [Google Scholar] [CrossRef] [PubMed]
  44. Elsaidi, S.K.; Venna, S.R.; Mohamed, M.H.; Gipple, M.J.; Hopkinson, D.P. Dual-layer MOF composite membranes with tuned interface interaction for postcombustion carbon dioxide separation. Cell Rep. Phys. Sci. 2020, 1, 100059. [Google Scholar] [CrossRef]
  45. Liu, J.; Zhang, G.; Clark, K.; Lin, H. Maximizing ether oxygen content in polymers for membrane CO2 removal from natural gas. ACS Appl. Mater. Interfaces 2019, 11, 10933–10940. [Google Scholar] [CrossRef] [PubMed]
  46. Orme, C.J.; Klaehn, J.R.; Harrup, M.K.; Luther, T.A.; Peterson, E.S.; Stewart, F.F. Gas permeability in rubbery polyphosphazene membranes. J. Membr. Sci. 2006, 280, 175–184. [Google Scholar] [CrossRef]
  47. Kusuma, V.A.; McNally, J.S.; Baker, J.S.; Tong, Z.; Zhu, L.; Orme, C.J.; Stewart, F.F.; Hopkinson, D.P. Cross-linked polyphosphazene blends as robust CO2 separation membranes. ACS Appl. Mater. Interfaces 2020, 12, 30787–30795. [Google Scholar] [CrossRef]
Figure 1. Schematic of CO2 and amine carrier interaction and an ultrafast CO2-selective transport channel inside a facilitated transport membrane.
Figure 1. Schematic of CO2 and amine carrier interaction and an ultrafast CO2-selective transport channel inside a facilitated transport membrane.
Membranes 11 00156 g001
Figure 2. Robeson Upper Bound 2008 plot depicting CO2 permeability and CO2/N2 selectivity of gas separation membranes based on polyvinylamine PVAm, metal–organic frameworks (MOFs), polyethylene oxide (PEO), thermally rearranged (TR) polymers, polymers of intrinsic micro porosity (PIMs), and conventional membranes [15].
Figure 2. Robeson Upper Bound 2008 plot depicting CO2 permeability and CO2/N2 selectivity of gas separation membranes based on polyvinylamine PVAm, metal–organic frameworks (MOFs), polyethylene oxide (PEO), thermally rearranged (TR) polymers, polymers of intrinsic micro porosity (PIMs), and conventional membranes [15].
Membranes 11 00156 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tong, Z.; Sekizkardes, A.K. Recent Developments in High-Performance Membranes for CO2 Separation. Membranes 2021, 11, 156. https://doi.org/10.3390/membranes11020156

AMA Style

Tong Z, Sekizkardes AK. Recent Developments in High-Performance Membranes for CO2 Separation. Membranes. 2021; 11(2):156. https://doi.org/10.3390/membranes11020156

Chicago/Turabian Style

Tong, Zi, and Ali K. Sekizkardes. 2021. "Recent Developments in High-Performance Membranes for CO2 Separation" Membranes 11, no. 2: 156. https://doi.org/10.3390/membranes11020156

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