Next Article in Journal
Effects of Critical Operation and Cleaning Parameters on Performances and Economic Benefits of Biogas Slurry Concentration by Forward Osmosis Membrane
Next Article in Special Issue
Separation of n-Butanol from Aqueous Solutions via Pervaporation Using PDMS/ZIF-8 Mixed-Matrix Membranes of Different Particle Sizes
Previous Article in Journal
The Interplay between TRPM7 and MagT1 in Maintaining Endothelial Magnesium Homeostasis
Previous Article in Special Issue
Investigation on the Performance of CO2 Absorption in Ceramic Hollow-Fiber Gas/Liquid Membrane Contactors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polymer-Infiltrated Metal–Organic Frameworks for Thin-Film Composite Mixed-Matrix Membranes with High Gas Separation Properties

1
Department of Chemical and Biomolecular Engineering, Yonsei University, 50 Yonsei-ro, Seodaemun-gu, Seoul 03722, Republic of Korea
2
Pacific Northwest National Laboratory, 902 Battelle Boulevard, Richland, WA 99352, USA
3
Hydrogen Research Department, Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Membranes 2023, 13(3), 287; https://doi.org/10.3390/membranes13030287
Submission received: 30 January 2023 / Revised: 23 February 2023 / Accepted: 24 February 2023 / Published: 28 February 2023

Abstract

:
Thin-film composite mixed-matrix membranes (TFC-MMMs) have potential applications in practical gas separation processes because of their high permeance (gas flux) and gas selectivity. In this study, we fabricated a high-performance TFC-MMM based on a rubbery comb copolymer, i.e., poly(2-[3-(2H-benzotriazol-2-yl)-4-hydroxyphenyl] ethyl methacrylate)-co-poly(oxyethylene methacrylate) (PBE), and metal–organic framework MOF-808 nanoparticles. The rubbery copolymer penetrates through the pores of MOF-808, thereby tuning the pore size. In addition, the rubbery copolymer forms a defect-free interfacial morphology with polymer-infiltrated MOF-808 nanoparticles. Consequently, TFC-MMMs (thickness = 350 nm) can be successfully prepared even with a high loading of MOF-808. As polymer-infiltrated MOF is incorporated into the polymer matrix, the PBE/MOF-808 membrane exhibits a significantly higher CO2 permeance (1069 GPU) and CO2/N2 selectivity (52.7) than that of the pristine PBE membrane (CO2 permeance = 431 GPU and CO2/N2 selectivity = 36.2). Therefore, the approach considered in this study is suitable for fabricating high-performance thin-film composite membranes via polymer infiltration into MOF pores.

1. Introduction

Anthropogenic CO2 emissions significantly accelerate global warming; therefore, CO2 removal technologies have been extensively investigated [1,2]. In recent years, membrane separation has demonstrated considerable potential as an energy-efficient and environmentally friendly method, particularly for CO2 capture from flue gas [3,4]. Polymeric membranes are favorable because of their low cost and high scalability but are generally limited by the trade-off between permeability and selectivity [5]. Mixed-matrix membranes (MMMs), comprising functional porous fillers in the polymer matrix, have emerged as potential candidates for high-performance membranes to enhance the gas separation performance of polymeric membranes [6,7,8].
Metal–organic frameworks (MOFs) are considered next-generation porous materials suitable for gas separation and capture because of their high surface area, tunable porosity, tailorable pore size, and functionality through the combination of various organic linkers and metal clusters [9,10]. Additionally, the ability to control the particle size of MOFs have a great advantage for the MMMs fabrication [10,11]. The interface morphology between the MOF particle and the polymeric chains is critical to the design and fabrication of high-performance MMMs [12]. Insufficient interfacial adhesion of the polymer on the MOF particle creates interfacial voids and considerably decreases the membrane selectivity [6,13]. Additionally, the pore blockage resulting from the polymer penetration should be avoided because gases cannot permeate through the pore-blocked MOFs in the membranes [14,15]. However, the pore size of the MOFs can be tuned with adequate pore blockage. As reported by Ban et al. [16], a zeolitic imidazolate framework (ZIF-8) with enhanced selectivity was obtained by incorporating an ionic liquid into ZIF-8 nanoparticles. Consequently, the effective cage size of ZIF-8 decreased, enabling the size-sieving of CO2 against N2 or CH4. Moreover, partial polymer penetration into the MOF typically facilitates the formation of an intimate interface between the fillers and the polymer matrix, thereby preventing defects between them [17,18].
In terms of the industrial applications of the membranes, the primary challenge involves improving the gas permeance (flux) of the membrane rather than the permeability (material property regardless of the thickness) because a substantial gas flux needs to be processed through the membrane [19,20]. Therefore, thin-film composite (TFC) membranes comprising submicron-thick selective dense layers on highly porous supports have been recently examined for scaling up gas separation membranes [21]. However, for preparing thin-film composite mixed-matrix membranes (TFC-MMMs), the interfacial morphology control between the MOFs and the polymer matrix must be effectively designed because the MOF particle size is relatively large in a submicron-thick selective layer [22].
MOF-808 is a porous framework composed of hexa-zirconium, which exhibits higher adsorption properties for CO2 than for other gases, such as N2, in addition to sufficient chemical/thermal stability. Plonka et al. [23] investigated the interaction between the MOF-808 structure and CO2. The results indicate that CO2 that enters the MOF structure is preferentially located inside tetrahedral cages owing to the residual electrons of organic linkers. MOF-808 has two different types of pore sizes with tetrahedral cages (size = 4.8 Å) and adamantine-shaped cages (diameter = 18.4 Å) [24,25]. Although the large pore size (which is higher than the kinetic diameter of N2 or CH4) of MOF-808 ensures an increased permeability in MMMs, the increase in selectivity is generally limited. Kulak et al. [26] investigated the CO2 separation performances of MMMs based on MOF-808 and co-polyimide under different conditions. The permeance increased with the incorporation of MOF-808; however, the separation factor was not considerably different from that of the pristine polymer. Vankelecom group [27,28] reported improved separation factors for MOF-808-based MMMs by the functionalization of MOF-808. The smaller pores (size = 4.8 Å) of MOF-808 can serve as selective pore channels by size-sieving if the pore diameter is tuned and lies between the kinetic diameters of CO2 (3.3 Å) and N2 (3.64 Å).
In this study, we fabricated high-performance TFC-MMMs based on the polymer-infiltrated MOF-808. The flexible and rubbery copolymer, poly(2-[3-(2H-benzotriazol-2-yl)-4-hydroxyphenyl] ethyl methacrylate)-co-poly(oxyethylene methacrylate) (PBE), was used for the polymer matrix. The rubbery PBE copolymer penetrated into the pores of MOF-808 through a simple blending, which was characterized using various techniques. The PBE infiltrating into the pore of MOF-808 not only improved the close interfacial contact with the polymer matrix but also reduced the effective pore size of PBE/MOF-808 membranes by controlling the pore size of MOF-808, enhancing the CO2/N2 selectivity. The gas separation performance of the TFC-MMMs was evaluated at 1 atm and 30 °C.

2. Materials and Methods

2.1. Materials

For synthesizing the copolymer, 2-[3-(2H-Benzotriazol-2-yl)-4-hydroxyphenyl]ethyl methacrylate (BEM, Mw = 323.35 g mol−1), and poly(oxyethylene methacrylate) (POEM, poly(ethylene glycol) methyl ether methacrylate, Mn = 500 g mol−1) were obtained from Sigma–Aldrich (St. Louis, MO, USA). Poly(1-(trimethylsilyl-1-propyne)) (PTMSP) was purchased from Gelest Inc. (Morrisville, PA, USA). N,N-Dimethylformamide (DMF), tetrahydrofuran (THF), isopropyl alcohol (IPA), n-hexane, cyclohexane, and methanol were purchased from J.T. Baker (Avantor, Radnor, PA, USA). Azobisisobutyronitrile (AIBN, Mw = 164.21 g mol−1). For synthesizing MOF-808, the ZrOCl2∙8H2O, trimesic acid, DMF, and formic acid were purchased from Acros Organics (Geel, Belgium). All solvents and chemicals were ACS reagent grade and used as received.

2.2. Synthesis of MOF-808

MOF-808 powder was synthesized and purified according to the published procedure [25,29]. ZrOCl2∙8H2O (4.85 g) and trimesic acid (H3BTC; 1.05 g) were completely dissolved in DMF/formic acid (225 mL/225 mL), transferred into a 1000 mL screw-capped jar, and sonicated for 30 min. The clear mixed solution was heated at 130 °C for 2 d and subsequently cooled to room temperature. The resulting white powder was collected via filtration and washed thrice with 150 mL of DMF. The sample was then immersed in 100 mL of DMF for 3 d, replacing fresh DMF daily. The DMF-immersed resulting powder was filtered under vacuum and soaked in 100 mL of acetone for 3 d with replacing fresh acetone daily. The final purified product was collected via centrifugation and subsequently evacuated at room temperature for 24 h.

2.3. Synthesis of the PBE Comb Copolymer

The PBE comb copolymer was synthesized via free-radical polymerization, according to our previously reported procedure [30]. Initially, 2 g of BEM was dissolved in 30 mL of DMF at room temperature for 1 h. Subsequently, 8 g of POEM and 0.002 g of AIBN were added to the solution, which was purged with N2 gas for 30 min. The polymerization reaction was performed at 90 °C for 20 h. The resultant polymer solution was precipitated in an n-hexane/IPA mixture (8:2). Precipitation was performed thrice to purify the polymer, with the resultant rubbery polymer dried in a vacuum oven to completely remove the residual solvent.

2.4. Preparation of TFC-MMMs

Various amounts of MOF-808 powders (0.02, 0.04, 0.06, 0.08, and 0.1 g) were dispersed in 0.2 g of MeOH solution under vigorous stirring and sonication for 1 h, respectively. Simultaneously, 0.2 g of PBE copolymer was dissolved in 1 g of MeOH under vigorous stirring at room temperature. Subsequently, the MOF-808 solution was added to the polymer solution and stirred for 24 h until the solution was well-dispersed. To prevent the selective polymer solution from penetrating the porous support, an RK control coater (Control RK Print-Coat Instruments Ltd., UK) was used to coat 0.5 wt% of PTMSP/cyclohexane solution on the microporous polysulfone (Psf) support layer as a gutter layer. After drying for 3 h, the PBE/MOF-808 solution was directly coated on the PTMSP/Psf membranes. The resulting membranes, with different ratios of MOF-808 particles, were labeled as 10%, 20%, 30%, 40%, and 50%. The prepared membranes were then dried overnight at room temperature and consecutively vacuum-dried overnight at 80 °C.

2.5. Characterization

The Brunauer–Emmett–Teller (BET) surface area from N2 adsorption–desorption isotherms at 77 K of MOF-808 and PBE@MOF-808 were determined using a 3Flex (Micromeritics Instruments, Norcross, GA, USA). Before the N2 adsorption–desorption measurement, approximately 100 mg of the samples were degassed under vacuum at 150 °C for 24 h. The chemical interactions between PBE copolymers and MOF-808 membranes were investigated using Fourier transform infrared spectroscopy (FT-IR). The morphologies of the membranes and particles were examined using a field-emission scanning electron microscope (FE-SEM, JSM-6701F, JEOL Ltd., Tokyo, Japan) and transmission electron microscope (TEM, Libra 120, Zeiss, operated at 120 kV). The thermal stability of the materials was investigated using thermogravimetric analysis (TGA) through a DTA/TGA analyzer (TA instruments, New Castle, DE, USA) at a heating rate of 10 °C/min under a nitrogen environment. The chain morphology of the polymers was analyzed using differential scanning calorimetry (DSC, (DSC8000, Perkin Elmer, Waltham, MA, USA)) at a scan rate of 10 °C/min under a nitrogen atmosphere. The crystalline structure of the membranes was characterized using X-ray diffraction (XRD, RINT2000 wide-angle goniometer, Rigaku, Tokyo, Japan) with a Cu cathode operated at 40 kV and 300 mA.

2.6. Gas Permeation Measurements

The pure gas permeance and separation properties were determined using a constant pressure/variable volume apparatus (Airrane Co., Ltd., Cheongju, Korea), according to our previously reported method [22]. The area of the membrane was approximately 10.2 cm2. All experiments were carried out at 30 °C, 1 bar. Each membrane was tested five times, and the average estimated error in gas permeance was approximately 5%. The gas permeance of the membrane was expressed in gas permeation units (GPU) (1 GPU = 10−6 cm3 (STP) cm−2 s−1 cmHg−1).

3. Results

3.1. Characterization of MOF-808 Nanoparticles and PBE/MOF-808 MMMs

The PBE comb copolymer consists of rigid hydrophobic BEM chains and rubbery hydrophilic POEM chains (Figure 1a). We found that the PBE copolymer is suitable as a thin-film composite membrane for the CO2 separation [30]. The rubbery POEM groups are known to have excellent CO2 selectivity owing to the dipole–quadrupole interaction and a high CO2 permeance, owing to the amorphous state of the polymer chains, whereas the rigid PBEM chains can inhibit N2 permeance and membrane plasticization. The rubbery state of the PBE copolymer can be easily penetrated into the pores of MOF-808 and improve the interfacial morphology between the PBE copolymer and MOF-808 [18]. Particularly, the PBE copolymer is alcohol-soluble, which facilitates the fabrication of TFC membranes because the solution does not dissolve the porous polymer support substrates such as Psf [31].
We synthesized the MOF-808 nanoparticles, and octahedral particles of sizes 500–600 nm were observed from the SEM and TEM images (Figure 1a,b). Figure S1a,b presented the N2 isotherm at 77 K and the pore size distribution for the degassed MOF-808, respectively. Based on the N2 isotherm, the BET surface area of MOF-808 was calculated as 2070 m2 g−1; the high surface area of MOF-808 indicates that it is sufficient to improve the permeability of MMMs by incorporating polymer into MOF-808. Using a density functional theory method, the pore size distributions were evaluated as 7.3 and 18.6 Å; the smaller pore was slightly different from the theoretical data (4.8 Å). The obtained results are consistent with those in the literature, indicating the successful synthesis of the MOF particles [32,33]. Moreover, the high CO2 adsorption properties and CO2 affinity of MOF-808 nanoparticles were confirmed (Figure S1c).
The FT-IR spectra of MOF-808, PBE comb copolymer, and PBE/MOF-808 MMMs with various MOF-808 contents were obtained (Figure 2a) to investigate the chemical interaction between MOF-808 filler and the PBE copolymer matrix. Pristine PBE copolymer has strong absorption bands at 1726 and 1100 cm−1, assigned to the stretching vibrations of carbonyl (C=O) and ether (C–O–C) functional groups, respectively [34]. PBE/MOF-808 MMMs maintained the original absorption band positions of PBE copolymer after incorporation with MOF-808, revealing that only a weak chemical interaction occurs between the polymer matrix and fillers. Strong chemical interactions between polymer matrix and fillers can densify and rigidify polymer chains and adversely affect the gas permeation [15].
The crystalline structures of the PBE copolymer, MOF-808, and PBE/MOF-808 MMMs were characterized using XRD patterns (Figure 2b). All the samples clearly showed isostructural to the pristine MOF-808. The PBE copolymer exhibits a broad peak at approximately 20.0° without a sharp crystalline peak, implying that the polymer structure is amorphous. In contrast to other poly(ethylene oxide) (PEO)-based polymers, the amorphous nature of the PBE copolymer is observed owing to the short chains of ethylene oxide units in the copolymer. The sharp crystalline peaks of MOF-808 were maintained after incorporation with the PBE polymer matrix, and their intensities gradually increased with the increasing MOF-808 content. As corroborated by the FT-IR spectra, the PBE copolymer matrix and MOF-808 filler do not sufficiently interact to affect the crystalline structure of MOF-808.
The chain morphologies of the PBE copolymer and PBE/MOF-808 MMMs were investigated through DSC analysis (Figure 2c). The PBE copolymer exhibits a glass transition temperature (Tg) at −53.7 °C. No sharp endothermic peaks attributed to polymer melting are observed, thus demonstrating the crystalline-free amorphous nature of the PBE copolymer, consistent with the XRD results. In terms of gas permeation, an amorphous structure is considered desirable because hard crystallite in a polymer matrix can act as a barrier to gas molecules and decrease gas permeability. Tg of PBE/MOF-808 MMMs gradually increases from −52.7 °C to −47 °C with the increasing MOF-808 content. The increase in Tg is generally induced by suppressing the polymer chain mobility and flexibility. As confirmed by the FT-IR spectra, only weak chemical interactions are observed between the PBE copolymer matrix and MOF-808 fillers. Thus, the change in Tg possibly originates from the physical penetration of polymer chains into the MOF-808 structure. The polymer chains could penetrate the pores of the MOF-808 structure and exhibit moderate flexibility and mobility.
We investigated the surface area and pore size of washed MOF-808 nanoparticles (Figure 3) to further verify the pore infiltration of polymer chains. The washed MOF-808 particles were prepared as follows: first, MOF-808 particles were blended with PBE solution, similar to the preparation of PBE/MOF-808 MMMs; then, MOF-808 particles were collected from the solution via centrifugation. The resultant particles were washed with simultaneous sonication and centrifuged thrice with MeOH three times to remove the residual polymer on the surface of the MOF-808 particles. Not only the surface area of the MOF-808 mixed with the PBE copolymer decreased from 2070 to 696 m2 g−1 but also the pore volume decreased. Particularly, two peaks at 7.3 and 18.6 Å, corresponding to the small and large pores, respectively, are observed for pristine MOF-808. However, MOF-808 mixed with the PBE copolymer presents peaks from the pore size distribution at approximately 3.9 and 15.9 Å, which are lower than those of the pristine MOF. These results imply the penetration of the polymer chain into the pores of MOF-808. The PBE copolymer chains are entangled in the pores of MOF-808 filler; therefore, they can adhere to the MOF structure even after extensive washing with polymer-dissolving solvents. Considering that the theoretical pore sizes of MOF-808 are 4.8 and 18.4 Å, the effective pore size of the polymer-infiltrated MOF would be lower than the obtained data [32]. The pore-tuned MOF-808 can act as a CO2-philic nanocage in the membrane matrix, which selectively facilitates gas molecules. Furthermore, the PBE@MOF-808 sample exhibits an improved CO2 adsorption property compared to the pristine MOF-808. This is a result of the quadrupole-dipole interaction of the CO2 molecule with the ethylene oxide group of the PBE chain and shows that the PBE chain is well entangled through the pores of MOF-808. Furthermore, the increase in adsorption is related to the decreased pore size of the MOF-808. The adsorption strength can be increased because the target molecule can interact with the other walls of the pore without being affected by only one part of the pore [35]. Therefore, the membrane performance can be improved by incorporating the polymer-infiltrated MOF-808 into the polymer matrix [36].
We investigated the thermal stability of the PBE comb copolymer, MOF-808 particles, and PBE/MOF-808 MMMs through TGA (Figure S2). The pristine PBE copolymer demonstrates excellent thermal stability up to 200 °C. For MOF-808 particles, 0–200 °C is associated with the volatility of the solvent contained in the adsorbent. Generally, the solvent trapped in MOF-808 nanoparticles is evaporated when the temperature increases and the weight is partially decreased. The weight loss in the range of 200–400 °C occurs because the formic acid in MOF-808 is removed. Finally, the weight loss occurring between 400 and 600 °C is considered thermal decomposition owing to the loss of the MOF-808 linker. After incorporating MOF-808 particles, all membranes demonstrate sufficient thermal stability to be applied as CO2 separation membranes under post-combustion process conditions [37].

3.2. Preparation and Gas Separation Performance of TFC-MMMs

Thin-film structures are highly desirable for practical applications in membrane processes because the gas flux is one of the most critical factors in the process. Hence, the thickness of the selective layer should be minimized to maximize the permeance (flux) of the membranes. However, due to the thin selective layer, reducing defects in TFC-MMMs is much more difficult than in thick and dense membranes. We fabricated TFC-MMMs via a simple bar-coating method on a porous Psf support. The morphology of PBE/MOF-808 TFC-MMMs was confirmed through cross-sectional SEM images, with the mixture of PBE copolymer and MOF-808 particles adequately coated on the Psf supporting layer without penetration into the porous support (Figure 4a). MOF-808 particles were well-dispersed on the Psf supporting layer and encapsulated by the PBE comb copolymer. The PBE copolymer acts as a mechanically strong matrix to form a thin selective layer and as a binder between MOF-808 particles and the Psf supporting layer to prevent structural defects. Moreover, the boundary between the MOF-808 particles and the PBE polymer layer is uncertain, implying their considerable compatibility (Figure 4b). The defects from the fabrication of TFC-MMMs are not observed because of the intimate contact between the polymer matrix and the polymer-infiltrated MOF-808 nanoparticles. Moreover, although the selective layer thickness (approximately 350 nm) is less than the MOF-808 nanoparticle size, the polymer covers the surface of the nanofillers without defects. As shown in Figure 4c, partial penetration of polymer chains into the MOF-808 particles enhances interfacial contact, enabling the successful fabrication of TFC-MMMs despite the high loading of MOF particles in the membranes.
The pure gas permeance of membranes was characterized using a constant pressure/variable volume apparatus at 30 °C. The CO2 and N2 permeance and CO2/N2 selectivity of PBE/MOF-808 TFC-MMM with various MOF-808 contents are presented in Figure 5 and Table 1. As the PBE copolymer contains CO2-philic polar groups, such as ether oxygen in POEM chains and tertiary amine groups in triazole in its structure, the pristine PBE membrane exhibits moderate gas separation performance (CO2 permeance = 431 GPU and CO2/N2 selectivity = 36.2). With the increasing MOF-808 content in the TFC-MMMs, both CO2 permeance and CO2/N2 selectivity of the membranes are rapidly enhanced. The PBE/MOF-808 40% membrane presents the optimal gas separation performance (CO2 permeance = 1069 GPU and CO2/N2 selectivity = 52.7). The PBE/MOF-808 40% membrane exhibits a 2.5-fold increase in CO2 permeance and a 45% increase in CO2/N2 selectivity compared with that of the pristine PBE membrane. Both CO2 and N2 permeance increased with the incorporation of MOF-808. The significant increase in permeance of the TFC-MMMs indicates that the gases effectively diffuse through the large pores of the MOF-808 (adamantine-shaped cages with the size of 15.6 Å), and the polymer impregnation does not result in pore blockage or dead MOF pores. Although pristine MOF-808 is known to have an exceedingly large pore size that inhibits the size-sieving of the CO2 molecules, the CO2-philic polymer-encapsulated MOF-808 particles serve as selective nanocages for CO2 molecules. The high CO2-affinity of the nanoparticles can accelerate the CO2 transport relative to that of N2 into the MOF-808 nanoparticle pores. Moreover, pores partially covered with the polymer matrix can selectively sieve the gas molecules. The cumulative effect of the polymer-infiltrated MOF-808 considerably enhances the gas selectivity of the membranes. However, both permeance and selectivity are decreased when MOF-808 is added in excess of 50%. The excess MOF-808 particles are not uniformly dispersed in the polymer matrix and are aggregated by themselves, as shown in the surface SEM image of the membranes (Figure S3). Due to the aggregation, not all incorporated MOF particles are activated to function as selective diffusion pathways for gas transport. As a result, both permeability and selectivity decrease, similar to the membranes with low loading of MOF-808. A MOF-808 membrane without the PBE copolymer was also prepared on a PTMSP-coated Psf support using the same method. The neat MOF-808 membrane showed poor gas separation performance with a CO2 permeance of 12,000 GPU and CO2/N2 selectivity of 1.9, indicating the importance of the PBE copolymer matrix to minimize structural defects.
The gas separation performance of the PBE/MOF-808 membranes is compared with those of other TFC-MMMs reported in the literature (Figure 6 and Table 1) [38,39,40,41,42,43]. The gas separation performance of the TFC-MMMs in our study demonstrates high selectivity owing to the selective nanochannels originating from polymer-entrapped MOFs. Moreover, the permeance of the PBE/MOF-808 40% membrane satisfies the target performance required for membranes to be used in practical post-combustion processes (CO2 permeance >1000 GPU and CO2/N2 selectivity >30) [44]. Using a rubbery and flexible PBE comb copolymer, polymer-infiltrated MOF-808 nanoparticles can be prepared through simple blending. Polymer infiltration into the MOF particles induces intimate contact with the polymer matrix, enabling the successful fabrication of thin-film membranes and the formation of effective nanochannels to selectively transport CO2. Thus, an efficient approach to fabricating high-performance thin-film gas separation membranes is presented.

4. Conclusions

In this study, we fabricated high-performance thin-film membranes by incorporating MOF-808 into the PBE comb copolymer. The flexible and rubbery PBE copolymer was used for the matrix of the membrane, which has moderate gas separation performance and excellent film-forming ability. The flexible PBE chain was partially infiltrated into the pores of MOF-808 through a simple mixing method, and the infiltrated PBE increased the gas separation selectivity by improving the interfacial contact with the polymer matrix and reducing the effective pores of MOF-808. Owing to the intimate contact between the MOF nanoparticles and the PBE polymer matrix, defect-free TFC-MMMs (thickness = 350 nm) were successfully fabricated with a high loading of MOF-808 nanoparticles. The penetration of the polymer matrix could remarkably improve the selectivity and interfacial compatibility of the MOFs. When the MOF-808 content in the membrane was increased up to 40%, the permeance and selectivity continuously increased, and PBE/MOF-808 40% membrane demonstrated the optimal performance with a CO2 permeance of 1069 GPU and CO2/N2 selectivity of 52.7, considerably higher than that of the pristine PBE membrane (CO2 permeance = 431 GPU and CO2/N2 selectivity = 36.2). The resultant performance satisfies the criteria for practical applications of CO2 capture in post-combustion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/membranes13030287/s1, Figure S1: (a) BET surface area, (b) pore size distribution from N2 isotherm at 77K, and (c) CO2 and N2 adsorption isotherms of MOF-808 at 303K, Figure S2: TGA curves of neat PBE and PBE/MOF-808 MMMs with different MOF-808 contents, Figure S3: Surface SEM images of neat PBE and PBE/MOF-808 TFC-MMMs with different MOF-808 contents.

Author Contributions

Conceptualization, J.H.L. and J.H.K.; methodology, H.J.M. and M.-B.K.; software, M.-B.K. and P.K.T.; validation, Y.-S.B. and J.H.L.; formal analysis, H.J.M.; investigation, H.J.M. and M.-B.K.; resources, Y.-S.B.; data curation, J.H.L.; writing—original draft preparation, H.J.M. and M.-B.K.; writing—review and editing, J.H.L. and J.H.K.; visualization, J.H.L.; supervision, J.H.K.; project administration J.H.K.; funding acquisition, J.H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Research Foundation of Korea (NRF) grants funded by the Korean government (MSIT) (2020K1A4A7A02095371) and the Korean Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government (MOTIE) (20214000000090, Fostering human resources training in advanced hydrogen energy industry).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Keith, D.W. Why capture CO2 from the atmosphere? Science 2009, 325, 1654–1655. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, S.; Feron, P.H.; Deng, L.; Favre, E.; Chabanon, E.; Yan, S.; Hou, J.; Chen, V.; Qi, H. Status and progress of membrane contactors in post-combustion carbon capture: A state-of-the-art review of new developments. J. Membr. Sci. 2016, 511, 180–206. [Google Scholar] [CrossRef]
  3. Min, H.J.; Kang, M.; Lee, C.S.; Kim, J.H. Dual-functional interconnected pebble-like structures in highly crystalline poly (ethylene oxide) membranes for CO2 separation. Sep. Purif. Technol. 2021, 263, 118363. [Google Scholar] [CrossRef]
  4. Lv, X.; Li, X.; Huang, L.; Ding, S.; Lv, Y.; Zhang, J. Tailoring physical and chemical microenvironments by polyether-amine in blended membranes for efficient CO2 separation. Korean J. Chem. Eng. 2022, 39, 475–483. [Google Scholar] [CrossRef]
  5. 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]
  6. Aroon, M.; Ismail, A.; Matsuura, T.; Montazer-Rahmati, M. Performance studies of mixed matrix membranes for gas separation: A review. Sep. Purif. Technol. 2010, 75, 229–242. [Google Scholar] [CrossRef]
  7. 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]
  8. Kim, B.J.; Kang, S.W. Composites of poly (vinyl pyrrolidone) and polarized Ag nanoparticles for CO2 separation. Korean J. Chem. Eng. 2022, 39, 2542–2547. [Google Scholar] [CrossRef]
  9. Ji, Y.; Song, Y.; Huang, Y.; Zhu, H.; Yue, C.; Liu, F.; Zhao, J. One-Step Synthesis of Ultrathin Zeolitic Imidazole Framework-8 (ZIF-8) Membrane on Unmodified Porous Support via Electrophoretic Deposition. Membranes 2022, 12, 1062. [Google Scholar] [CrossRef]
  10. Usman, M.; Khan, M.Y.; Anjum, T.; Khan, A.L.; Hoque, B.; Helal, A.; Hakeem, A.S.; Al-Maythalony, B.A. Controlled Covalent Functionalization of ZIF-90 for Selective CO2 Capture & Separation. Membranes 2022, 12, 1055. [Google Scholar] [CrossRef]
  11. Sun, J.; Li, Q.; Chen, G.; Duan, J.; Liu, G.; Jin, W. MOF-801 incorporated PEBA mixed-matrix composite membranes for CO2 capture. Sep. Purif. Technol. 2019, 217, 229–239. [Google Scholar] [CrossRef]
  12. Kim, N.U.; Park, B.J.; Lee, J.H.; Kim, J.H. High-performance ultrathin mixed-matrix membranes based on an adhesive PGMA-co-POEM comb-like copolymer for CO2 capture. J. Mater. Chem. A 2019, 7, 14723–14731. [Google Scholar] [CrossRef]
  13. Venna, S.R.; Lartey, M.; Li, T.; Spore, A.; Kumar, S.; Nulwala, H.B.; Luebke, D.R.; Rosi, N.L.; Albenze, E. Fabrication of MMMs with improved gas separation properties using externally-functionalized MOF particles. J. Mater. Chem. A 2015, 3, 5014–5022. [Google Scholar] [CrossRef]
  14. Mannan, H.A.; Idris, A.; Nasir, R.; Mukhtar, H.; Qadir, D.; Suleman, H.; Basit, A. Interfacial Tailoring of Polyether Sulfone-Modified Silica Mixed Matrix Membranes for CO2 Separation. Membranes 2022, 12, 1129. [Google Scholar] [CrossRef] [PubMed]
  15. Lin, R.; Hernandez, B.V.; Ge, L.; Zhu, Z. Metal organic framework based mixed matrix membranes: An overview on filler/polymer interfaces. J. Mater. Chem. A 2018, 6, 293–312. [Google Scholar] [CrossRef]
  16. Ban, Y.; Li, Z.; Li, Y.; Peng, Y.; Jin, H.; Jiao, W.; Guo, A.; Wang, P.; Yang, Q.; Zhong, C. Confinement of ionic liquids in nanocages: Tailoring the molecular sieving properties of ZIF-8 for membrane-based CO2 capture. Angew. Chem. Int. Ed. 2015, 54, 15483–15487. [Google Scholar] [CrossRef] [PubMed]
  17. Duan, P.; Moreton, J.C.; Tavares, S.R.; Semino, R.; Maurin, G.; Cohen, S.M.; Schmidt-Rohr, K. Polymer infiltration into metal–organic frameworks in mixed-matrix membranes detected in situ by NMR. J. Am. Chem. Soc. 2019, 141, 7589–7595. [Google Scholar] [CrossRef]
  18. Cheng, Y.; Ying, Y.; Japip, S.; Jiang, S.D.; Chung, T.S.; Zhang, S.; Zhao, D. Advanced porous materials in mixed matrix membranes. Adv. Mater. 2018, 30, 1802401. [Google Scholar] [CrossRef]
  19. Baker, R.W.; Low, B.T. Gas separation membrane materials: A perspective. Macromolecules 2014, 47, 6999–7013. [Google Scholar] [CrossRef]
  20. Park, H.B.; Kamcev, J.; Robeson, L.M.; Elimelech, M.; Freeman, B.D. Maximizing the right stuff: The trade-off between membrane permeability and selectivity. Science 2017, 356, eaab0530. [Google Scholar] [CrossRef] [Green Version]
  21. Xie, K.; Fu, Q.; Qiao, G.G.; Webley, P.A. Recent progress on fabrication methods of polymeric thin film gas separation membranes for CO2 capture. J. Membr. Sci. 2019, 572, 38–60. [Google Scholar] [CrossRef]
  22. Lee, C.S.; Kang, M.; Kim, K.C.; Kim, J.H. In-situ formation of asymmetric thin-film, mixed-matrix membranes with ZIF-8 in dual-functional imidazole-based comb copolymer for high-performance CO2 capture. J. Membr. Sci. 2022, 642, 119913. [Google Scholar] [CrossRef]
  23. Plonka, A.M.; Grissom, T.G.; Musaev, D.G.; Balboa, A.; Gordon, W.O.; Collins-Wildman, D.L.; Ghose, S.K.; Tian, Y.; Ebrahim, A.M.; Mitchell, M.B. Effect of carbon dioxide on the degradation of chemical warfare agent simulant in the presence of Zr metal organic framework MOF-808. Chem. Mater. 2019, 31, 9904–9914. [Google Scholar] [CrossRef]
  24. Kim, Y.K.; Cha, S.I.; Hong, S.H.; Jeong, Y.J. A new hybrid architecture consisting of highly mesoporous CNT/carbon nanofibers from starch. J. Mater. Chem. 2012, 22, 20554. [Google Scholar] [CrossRef]
  25. Furukawa, H.; Gandara, F.; Zhang, Y.-B.; Jiang, J.; Queen, W.L.; Hudson, M.R.; Yaghi, O.M. Water adsorption in porous metal–organic frameworks and related materials. J. Am. Chem. Soc. 2014, 136, 4369–4381. [Google Scholar] [CrossRef] [PubMed]
  26. Kulak, H.; Thür, R.; Vankelecom, I.F. MOF/Polymer Mixed-Matrix Membranes Preparation: Effect of Main Synthesis Parameters on CO2/CH4 Separation Performance. Membranes 2022, 12, 425. [Google Scholar] [CrossRef] [PubMed]
  27. Thür, R.; Van Havere, D.; Van Velthoven, N.; Smolders, S.; Lamaire, A.; Wieme, J.; Van Speybroeck, V.; De Vos, D.; Vankelecom, I.F. Correlating MOF-808 parameters with mixed-matrix membrane (MMM) CO2 permeation for a more rational MMM development. J. Mater. Chem. A 2021, 9, 12782–12796. [Google Scholar] [CrossRef]
  28. Thür, R.; Van Velthoven, N.; Lemmens, V.; Bastin, M.; Smolders, S.; De Vos, D.; Vankelecom, I.F. Modulator-mediated functionalization of MOF-808 as a platform tool to create high-performance mixed-matrix membranes. ACS Appl. Mater. Interfaces 2019, 11, 44792–44801. [Google Scholar] [CrossRef]
  29. Wang, L.; Chen, P.; Dong, X.; Zhang, W.; Zhao, S.; Xiao, S.; Ouyang, Y. Porous MOF-808@ PVDF beads for removal of iodine from gas streams. RSC Adv. 2020, 10, 44679–44687. [Google Scholar] [CrossRef]
  30. Lee, J.H.; Jung, J.P.; Jang, E.; Lee, K.B.; Kang, Y.S.; Kim, J.H. CO2-philic PBEM-g-POEM comb copolymer membranes: Synthesis, characterization and CO2/N2 separation. J. Membr. Sci. 2016, 502, 191–201. [Google Scholar] [CrossRef]
  31. Lee, C.S.; Kim, N.U.; Park, J.T.; Kim, J.H. Imidazole-functionalized hydrophilic rubbery comb copolymers: Microphase-separation and good gas separation properties. Sep. Purif. Technol. 2020, 242, 116780. [Google Scholar] [CrossRef]
  32. Rojas-Buzo, S.; García-García, P.; Corma, A. Zr-MOF-808@ MCM-41 catalyzed phosgene-free synthesis of polyurethane precursors. Catal. Sci. Technol. 2019, 9, 146–156. [Google Scholar] [CrossRef]
  33. Hardian, R.; Dissegna, S.; Ullrich, A.; Llewellyn, P.L.; Coulet, M.V.; Fischer, R.A. Tuning the properties of MOF-808 via defect engineering and metal nanoparticle encapsulation. Chem. Eur. J. 2021, 27, 6804–6814. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, N.U.; Park, B.J.; Choi, Y.; Lee, K.B.; Kim, J.H. High-performance self-cross-linked PGP–POEM comb copolymer membranes for CO2 capture. Macromolecules 2017, 50, 8938–8947. [Google Scholar] [CrossRef]
  35. Kim, M.-B.; Yoon, T.-U.; Hong, D.-Y.; Kim, S.-Y.; Lee, S.-J.; Kim, S.-I.; Lee, S.-K.; Chang, J.-S.; Bae, Y.-S. High SF6/N2 selectivity in a hydrothermally stable zirconium-based metal–organic framework. Chem. Eng. J. 2015, 276, 315–321. [Google Scholar] [CrossRef]
  36. Qian, J.; Song, E.; Lian, H.; Jiang, J.; Wang, C.; Pan, Y. High-performance ZIF-302 mixed-matrix membranes for efficient CO2 capture. Korean J. Chem. Eng. 2022, 39, 1020–1027. [Google Scholar] [CrossRef]
  37. 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]
  38. Choi, O.; Hossain, I.; Jeong, I.; Park, C.-H.; Kim, Y.; Kim, T.-H. Modified Graphene Oxide-Incorporated Thin-Film Composite Hollow Fiber Membranes through Interface Polymerization on Hydrophilic Substrate for CO2 Separation. Membranes 2021, 11, 650. [Google Scholar] [CrossRef]
  39. Min, H.J.; Kang, M.; Bae, Y.-S.; Blom, R.; Grande, C.A.; Kim, J.H. Thin-film composite mixed-matrix membrane with irregular micron-sized UTSA-16 for outstanding gas separation performance. J. Membr. Sci. 2022, 669, 121295. [Google Scholar] [CrossRef]
  40. Martínez-Izquierdo, L.; Téllez, C.; Coronas, J. Highly stable Pebax® Renew® thin-film nanocomposite membranes with metal organic framework ZIF-94 and ionic liquid [Bmim][BF4] for CO2 capture. J. Mater. Chem. A 2022, 10, 18822–18833. [Google Scholar] [CrossRef]
  41. Xie, K.; Fu, Q.; Kim, J.; Lu, H.; He, Y.; Zhao, Q.; Scofield, J.; Webley, P.A.; Qiao, G.G. Increasing both selectivity and permeability of mixed-matrix membranes: Sealing the external surface of porous MOF nanoparticles. J. Membr. Sci. 2017, 535, 350–356. [Google Scholar] [CrossRef]
  42. Kang, M.; Kim, T.-H.; Han, H.H.; Min, H.J.; Bae, Y.-S.; Kim, J.H. Submicron-thick, mixed-matrix membranes with metal-organic frameworks for CO2 separation: MIL-140C vs. UiO-67. J. Membr. Sci. 2022, 659, 120788. [Google Scholar] [CrossRef]
  43. Sutrisna, P.D.; Hou, J.; Zulkifli, M.Y.; Li, H.; Zhang, Y.; Liang, W.; D’Alessandro, D.M.; Chen, V. Surface functionalized UiO-66/Pebax-based ultrathin composite hollow fiber gas separation membranes. J. Mater. Chem. A 2018, 6, 918–931. [Google Scholar] [CrossRef]
  44. Merkel, T.C.; Lin, H.; Wei, X.; Baker, R. Power plant post-combustion carbon dioxide capture: An opportunity for membranes. J. Membr. Sci. 2010, 359, 126–139. [Google Scholar] [CrossRef]
Figure 1. (a) Chemical structure of the PBE comb copolymer. (b) SEM and (c) TEM images of the synthesized MOF-808 nanoparticles.
Figure 1. (a) Chemical structure of the PBE comb copolymer. (b) SEM and (c) TEM images of the synthesized MOF-808 nanoparticles.
Membranes 13 00287 g001
Figure 2. (a) FT-IR spectra and (b) XRD patterns of the PBE comb copolymer, MOF-808 nanoparticles, and PBE/MOF MMMs with different ratios. (c) DSC graphs of the PBE copolymer and PBE/MOF MMMs.
Figure 2. (a) FT-IR spectra and (b) XRD patterns of the PBE comb copolymer, MOF-808 nanoparticles, and PBE/MOF MMMs with different ratios. (c) DSC graphs of the PBE copolymer and PBE/MOF MMMs.
Membranes 13 00287 g002
Figure 3. (a) BET isotherms. (b) pore size distributions. (c) CO2 adsorption isotherm of pristine MOF-808 and washed MOF-808 after being mixed with PBE copolymer (PBE@MOF-808).
Figure 3. (a) BET isotherms. (b) pore size distributions. (c) CO2 adsorption isotherm of pristine MOF-808 and washed MOF-808 after being mixed with PBE copolymer (PBE@MOF-808).
Membranes 13 00287 g003
Figure 4. (a,b) Cross-sectional SEM images of PBE/MOF-808 40% TFC-MMM. (c) Schematic of TFC-MMM based on PBE-infiltrated MOF-808.
Figure 4. (a,b) Cross-sectional SEM images of PBE/MOF-808 40% TFC-MMM. (c) Schematic of TFC-MMM based on PBE-infiltrated MOF-808.
Membranes 13 00287 g004
Figure 5. (a) CO2 and N2 permeance of TFC-MMMs with varying MOF-808 content. (b) CO2 permeance and CO2/N2 selectivity of TFC-MMMs with increasing MOF-808 content.
Figure 5. (a) CO2 and N2 permeance of TFC-MMMs with varying MOF-808 content. (b) CO2 permeance and CO2/N2 selectivity of TFC-MMMs with increasing MOF-808 content.
Membranes 13 00287 g005
Figure 6. CO2 permeance versus CO2/N2 selectivity for PBE and PBE/MOF-808 membranes and other TFC-MMMs reported in the literature (details in Table 1).
Figure 6. CO2 permeance versus CO2/N2 selectivity for PBE and PBE/MOF-808 membranes and other TFC-MMMs reported in the literature (details in Table 1).
Membranes 13 00287 g006
Table 1. CO2 separation performance of PBE/MOF-808 membranes and previously reported TFC-MMMs.
Table 1. CO2 separation performance of PBE/MOF-808 membranes and previously reported TFC-MMMs.
SampleCO2 Permeance
(GPU)
CO2/N2 SelectivityCondition
(T (°C)/P (bar))
Reference
PBE43136.230 °C, 1 barThis study
MOF-808 10%53543.3
MOF-808 20%62344.0
MOF-808 30%85351.2
MOF-808 40%1069.052.7
MOF-808 50%67746.4
PEI-TMC/mGO736025 °C, 0.25 bar[38]
PTO/UTSA-16107041.030 °C, 1 bar[39]
Pebax/IL/ZIF-948192535 °C, 3 bar[40]
Polyactive/P@MOF212602235 °C, 3 bar[41]
PGO/MIL-140C17683830 °C, 1 bar[42]
Pebax/UiO-66-NH23385725 °C, 2 bar[43]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Min, H.J.; Kim, M.-B.; Bae, Y.-S.; Thallapally, P.K.; Lee, J.H.; Kim, J.H. Polymer-Infiltrated Metal–Organic Frameworks for Thin-Film Composite Mixed-Matrix Membranes with High Gas Separation Properties. Membranes 2023, 13, 287. https://doi.org/10.3390/membranes13030287

AMA Style

Min HJ, Kim M-B, Bae Y-S, Thallapally PK, Lee JH, Kim JH. Polymer-Infiltrated Metal–Organic Frameworks for Thin-Film Composite Mixed-Matrix Membranes with High Gas Separation Properties. Membranes. 2023; 13(3):287. https://doi.org/10.3390/membranes13030287

Chicago/Turabian Style

Min, Hyo Jun, Min-Bum Kim, Youn-Sang Bae, Praveen K. Thallapally, Jae Hun Lee, and Jong Hak Kim. 2023. "Polymer-Infiltrated Metal–Organic Frameworks for Thin-Film Composite Mixed-Matrix Membranes with High Gas Separation Properties" Membranes 13, no. 3: 287. https://doi.org/10.3390/membranes13030287

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