Next Article in Journal
Desalination Technology in South Korea: A Comprehensive Review of Technology Trends and Future Outlook
Next Article in Special Issue
Membranes for Gas Separation and Purification Processes
Previous Article in Journal
Gas Permeability, Fractional Free Volume and Molecular Kinetic Diameters: The Effect of Thermal Rearrangement on ortho-hydroxy Polyamide Membranes Loaded with a Porous Polymer Network
Previous Article in Special Issue
Density Functional Theory Study of B, N, and Si Doped Penta-Graphene as the Potential Gas Sensors for NH3 Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Recent Advances in Mixed-Matrix Membranes for Light Hydrocarbon (C1–C3) Separation

1
Department of Chemical Engineering, Universiti Teknologi Petronas, Bandar Seri Iskandar, Perak 32610, Malaysia
2
CO2 Research Centre (CO2RES), Institute of Contaminant Management, Universiti Teknologi Petronas, Bandar Seri Iskandar, Perak 32610, Malaysia
3
Department of Chemical and Biomolecular Engineering, Korea Advanced Institute of Science and Technology, Daejeon 34141, Korea
*
Authors to whom correspondence should be addressed.
Membranes 2022, 12(2), 201; https://doi.org/10.3390/membranes12020201
Submission received: 14 January 2022 / Revised: 29 January 2022 / Accepted: 6 February 2022 / Published: 9 February 2022
(This article belongs to the Special Issue Membranes for Gas Separation and Purification Processes)

Abstract

:
Light hydrocarbons, obtained through the petroleum refining process, are used in numerous applications. The separation of the various light hydrocarbons is challenging and expensive due to their similar melting and boiling points. Alternative methods have been investigated to supplement cryogenic distillation, which is energy intensive. Membrane technology, on the other hand, can be an attractive alternative in light hydrocarbon separation as a phase change that is known to be energy-intensive is not required during the separation. In this regard, this study focuses on recent advances in mixed-matrix membranes (MMMs) for light hydrocarbon (C1–C3) separation based on gas permeability and selectivity. Moreover, the future research and development direction of MMMs in light hydrocarbon separation is discussed, considering the low intrinsic gas permeability of polymeric membranes.

1. Introduction

Light hydrocarbons (specifically methane, acetylene, ethylene, ethane, propylene, and propane) are major raw materials in petrochemical industries. In particular, the increasing demand for polyethylene, poly(vinyl chloride), ethane, and polypropylene, which are used in the production of everyday materials [1], is driving the substantial production of ethylene and propylene (~191 and 120 million tons, respectively, worldwide in 2019) [2,3,4]. These two hydrocarbons are typically produced through the steam cracking of hydrocarbon sources, such as naphtha, natural gas, coal, and shale gas [5]. Irrespective of the source, the purities of the extracted ethylene and propylene streams should ideally be at least 99.5% [6,7]. The presence of acetylene (>40 ppm) in the ethylene stream, for instance, hampers ethylene polymerization because acetylene could poison the Ziegler–Natta catalyst [8]. This polymerization process is essential because it nullifies contaminants that may trigger side reactions, which could substantially reduce the molecular weight of the resulting polymers [9].
High-pressure cryogenic distillation, a mature technology, has been adopted in industrial operations for light alkene/alkane separation. For example, effective ethylene/ethane separation requires an operating temperature and pressure of −160 °C and 23 bar, respectively, given their similar melting and boiling points (Table 1) [10]; specifically, the boiling and melting points of ethylene and ethane differ by 14 and 15 °C, respectively. Such energy-intensive operating conditions are required for propylene/propane separation as well, because the melting and boiling points of this pair are even more similar (differing by only 6 and 3 °C, respectively). In particular, 75–85% of the input energy is utilized for the effective production of ethylene and propylene [11]. Thus, alternative technologies for effectively separating light hydrocarbons to isolate specific products are highly desirable.
Various swing adsorption processes have been proposed to alleviate the high energy penalty in conventional (i.e., pressure- or vacuum-based) processes [10,11,12,13,14,15]. These processes enable hydrocarbon separation at ambient temperature by varying the feed pressure; thus, their energy consumption is relatively low. However, swing adsorption typically entails challenges, such as low recovery, depending on the type of adsorbent used [13,16,17,18,19,20,21]. In addition, their adsorption–desorption cycling is energy-intensive, which restricts their widespread adoption for light hydrocarbon separation [22]. To overcome these drawbacks, membrane-based gas separation technology (Figure 1) has been developed for cost-effective and energy-efficient light hydrocarbon separation [23]. This approach eliminates the need for phase change as separation is achieved under ambient conditions. Polymeric membranes are the main drivers of this technology as they can be fabricated in large sizes with reasonably high mechanical stability [24,25]. However, because the gas transport mechanism is dominated by the solution-diffusion mechanism, the performance of polymeric membrane-based gas separation is governed by the inevitable trade-off relationship between permeability and selectivity [26,27]. Moreover, molecular sieve membranes, developed using pure zeolites and metal-organic frameworks (MOFs), have low scalability due to their poor mechanical strength [28,29,30].
Table 1. Properties of light hydrocarbons (C1–C3) [31,32] .
Table 1. Properties of light hydrocarbons (C1–C3) [31,32] .
Light HydrocarbonNormal Boiling Point (°C)Normal Melting Point (°C)Critical Pressure (MPa)Critical Temperature (°C)Kinetic Diameter (Å) 1Van der Waals Diameter (Å) 1Polarizability × 1025 (cm3)Dipole Moment × 1018 (esu cm)Quadrupole Moment × 1026 (esu cm2)
Methane (CH4)−162−1834.6−833.803.2525.900
Acetylene (C2H2)−84−816.2−353.30-33.30-
Ethylene (C2H4)−103−1695.194.163.5942.501.5
Ethane (C2H6)−89−1844.9324.443.7244.300.7
Propylene (C3H6)−48−1854.6914.68 14.0362.60.4-
Propane (C3H8)−42−1884.3974.30 14.1662.90.1-
1 Hybrid molecular dimension scales based on kinetic diameter and van der Waals diameter are required to obtain a clearer understanding of membrane performance.
Mixed-matrix membranes (MMMs), which combine the advantages of polymeric and molecular sieves (adsorbents), have been a focus of research attention. Numerous papers have reviewed the research on membranes for hydrocarbon separation [2,5,33,34,35]; therefore, the present study focuses mainly on the performance of MMMs and composite membranes in separating C1 to C3 hydrocarbons. In general, the utilization of MOFs as potential fillers for light hydrocarbon separation has been the main focus due to their higher flexibility in both pore size control and post-synthetic functionalization compared to conventional porous materials, such as zeolites [23]. In addition, their performance with respect to the upper bound limit (C2H4/C2H6 and C3H6/C3H8) [36,37] is discussed. Finally, insights on the future research and development direction of MMMs in light hydrocarbon separation are shared.

2. C2 Hydrocarbon Separation

Membranes in C2 hydrocarbon separation have been mainly investigated with respect to their potential in C2H4/C2H6 separation (Table 2). However, although polymeric membranes, especially those commercially available, are robust and processable, they do not exhibit high permeabilities in ethylene/ethane separation. For example, Naghsh et al. [38], Davoodi et al. [39], and Jeroen et al. [40] demonstrated that the C2H4 permeabilities of Matrimid®, cellulose acetate (CA), and polyimide (P84) under ambient conditions were merely 0.26, 0.051, and 0.049 barrer, respectively, which is attributable to their low fractional free volume (FFV). C2H4/C2H6 selectivities, on the other hand, were reported to be moderate, with values of 3.25, 2.27, and 4.00, respectively. It is expected that the reported selectivity is comparatively higher than those of polymers with large FFV, such as 6FDA-based polymers and PIM-1, as observed in CO2 separation [24,28,41]. Therefore, these low intrinsic gas permeabilities make the lab-scale (gas chromatography) investigation of such membranes challenging, as it requires specific devices (high-accuracy pressure transducers or thermal conductivity detectors) or conditions (high feed pressure). The advantages of incorporating filler materials into polymeric membranes have been investigated. For instance, Naghsh et al. [38] and Davoodi et al. [39] incorporated amorphous silica nanoparticles into Matrimid® and CA membranes, respectively, and investigated their performance in C2H4/C2H6 separation; the incorporation of 20 wt% silica in these two polymeric membranes was found to increase C2H4 permeability by 58% and 76%, with no substantial change in C2H6 permeability, in comparison with pure polymer matrix. Although silica nanoparticles are not porous, the residual hydroxyl (–OH) functionalities make interactions with polar C2H4 molecules feasible. Furthermore, the incorporation of silica nanoparticles disrupts polymer packing, which increases the FFV and facilitates rapid gas transport.
These advantages notwithstanding, the use of non-porous silica nanoparticles constrains improvements in C2H4/C2H6 separation performance due to their lack of intrinsic porosity, highlighting the need for the incorporation of porous materials into separation membranes. In this regard, MOFs with coordinative unsaturated open metal sites are of particular interest. These metal sites interact favorably with the π-bond in C2H4, thus facilitating C2H4 adsorption. Bachman et al. [42] identified M2dobdc (M = Co, Fe, Mg, Mn, Ni, Zn), which possess reasonable C2H4/C2H6 selectivity (c.a. 8–12 at 50/50 vol/vol C2H4/C2H6 and 35 °C, based on ideal adsorbed solution theory [IAST] calculations), as a feasible adsorbent. Studies on MMMs incorporated with M2dobdc nanocrystals have revealed that the addition of Mg2dobdc and Mn2dobdc, unlike Ni2dobdc and Co2dobdc, creates undesirable interfacial gaps in the polymer matrix, leading to non-selective gas transport [42,43]. Nevertheless, given their high C2H4 adsorption, these M2dobdc-incorporated MMMs have exhibited large increases in C2H4 permeability, making it increasingly likely to surpass the upper bound limit for C2H4/C2H6 separation. The advantages of incorporating MOFs with open metal sites were also demonstrated by Chuah et al. [44], who used HKUST-1 as the adsorbent; specifically, the incorporation of 20 wt% HKUST-1 in 6FDA-TMPDA membranes substantially improved their C2H4 permeability, pushing their performance close to the C2H4/C2H6 upper bound limit (Figure 2).
MMMs have also been applied in acetylene separation. However, adsorbents with high acetylene/ethylene selectivity are rare [6]. MOFs are generally more favorable for acetylene/ethylene separation than conventional adsorbents, such as zeolites and activated carbon, because MOFs offer more systematic control over pore size (through the selection of ligands with varying strut lengths) and are compatible with pre- and post-synthetic functionalization. Qian et al. [45] recently demonstrated the C2H2/C2H4 selectivity of two types of MOF, namely SIFSIX-2-Cu-i and ZRFSIX-2-Ni-i (Figure 3a); this favorable performance is attributable to the presence of MF62− moieties (M = Cu, Ni), which enables strong interactions with C2H2. Furthermore, poly(ionic liquids) have also been used in this manner given their reasonable C2H2 permeability and filler–polymer compatibility [46,47,48]. Specifically, single-gas permeation experiments have shown that 30 wt% SIFSIX-2-Cu-i and ZRFSIX-2-Ni-i in poly(ionic liquids) improve C2H2 permeability by 129% and 100%, respectively (Figure 3b), without a substantial change in C2H2/C2H4 selectivity.
The feasibility of using MMMs for C2H6/CH4 separation has also been explored in research. In contrast to MOFs, which are effective at C2H4/C2H6 separation, the applicability of zeolite adsorbents in this separation is relatively less common than in CO2 separation (e.g., post-combustion or pre-combustion carbon capture) [49,50] due to the comparable polarizability and quadrupole moment of C2H4 and C2H6 molecules (Table 1). Hence, C2H6/CH4 separation was investigated instead due to a more substantial difference in their polarizabilities. Tirouni et al. [51] investigated C2H6/CH4 separation using polyurethane MMMs incorporated with zeolite 4A and ZSM-5. With this zeolite-based approach, the polarizability of C2H6 is higher than that of CH4; accordingly, the permeability of C2H6 is also higher than that of CH4. However, this approach is not promising due to its low C2H6/CH4 selectivity, which in turn is attributable to the poor compatibility between the polymer matrix and the zeolite particles; specifically, the zeolite–polymer configuration has been reported to exhibit the sieve-in-a-cage morphology [52,53,54,55]. These shortcomings can be potentially overcome via post-synthetic treatments, such as the incorporation of silane coupling agents and post-synthetic amine grafting. However, these treatments may jeopardize the effective transport of the desired gas through the membranes [23].
Table 2. Performance of mixed-matrix membranes (MMMs) in C2 hydrocarbon separation.
Table 2. Performance of mixed-matrix membranes (MMMs) in C2 hydrocarbon separation.
MembraneMeasurement ConditionGas Permeability (Barrer)Gas SelectivityRef.
Polymer/SupportFillerLoading (wt%)Pres. (bar)Temp. (°C)C2H2C2H4C2H6CH4C2H2/C2H4C2H4/C2H6
6FDA-DAMNi-gallate20---9125--4.13[56]
6FDA- TMPDACo2(dobdc)33235-27671--3.89[42]
6FDA- TMPDANi2(dobdc)250.7535-426101--4.22[42]
6FDA- TMPDAMg2(dobdc)23235-1140431--2.65[42]
6FDA- TMPDAMn2(dobdc)13235-433188--2.30[42]
6FDA- TMPDAHKUST-120135-18376--2.41[44]
α-aluminaZeolite MFI-925--37 [1]6 [1]--[57]
Cellulose
Acetate
Silica30---0.110.026--4.23[38]
DBzPBI-BuIZIF-8302.735-11241--2.73[58]
MatrimidSilica20330-0.420.07--6[39]
NylonBoron Nitride [2]-2--160 [1]1.9 [1]--84.2[29]
ODPA-
TMPDA
HKUST-120135-164.7--3.40[44]
P1–1 [3]SIFSIX-2-Cu-i301.5-452.1--21.4-[45]
P1–1 [3]ZRFSIX-2-Ni-i301.5-151.3--11.5-[45]
P84HKUST-1205--0.0520.0075--6.93[59]
P84Fe-BTC205--0.030.0052--5.77[40]
P84MIL-53205--0.0960.027--3.56[40]
PolystyreneFullerene1-35-0.580.34--1.71[60]
PolyurethaneZeolite 4A10230--66.653--[51]
PolyurethaneZSM-520230--71.332.3--[51]
PPEESZIF-8301--3.21.0--3.2[61]
PVDFAg-Al/NMA-0.125-45030--15[62]
[1] Permeance in GPU; [2] incorporated with reactive ionic liquid (RIL); [3] P1-1: (ATMA)(BF4)/PEG500 = 1:1 (2-(acryloyloxy)ethyl-trimethylammonium tetrafluoroborate)/poly(ethylene glycol) methyl ether methacrylate; 6FDA = 4,4′-(hexafluoroisoprropylidene)diphthalic anhydride; DAM = 2,4-diaminomesitylene; DBzPBI-BuI = substituted polybenzimidazole; ODPA = 4,4′-oxydiphthalic anhydride; PPEES = poly(1,4-phenylene ether-ether-sulfone); PVDF = polyvinylidene fluoride; TMPDA = 2,4,6-trimethyl-m-phenylenediamine.

3. C3 Hydrocarbon Separation

C3 hydrocarbon separation has primarily been examined for the propylene/propane (C3H6/C3H8) gas pair (Table 3), with few studies examining the C3H8/CH4 gas pair. As discussed in Section 2, commercial polymer membranes do not exhibit intrinsically high C3H6 permeability and C3H6/C3H8 selectivity due to their considerably larger kinetic diameter compared with other common gaseous molecules, such as CO2, N2, C2H4, and C2H6. Naghsh et al. [38] and Davoodi et al. [39] demonstrated that the C2H6 permeabilities of Matrimid® and CA under ambient conditions were as low as 0.09 and 0.046 barrer, respectively, due to their low FFV. Thus, it is highly desirable to use polyimide membranes with high intrinsic C3H6 permeabilities, such as 6FDA-based polymers and rubbery polymers (e.g., polydimethylsiloxane (PDMS]) and cross-linked poly(ethylene oxide) (XLPEO)) to achieve sufficient flux for gas separation.
Regarding other porous materials, the effectiveness of ZIF-8 (ZIF = zeolitic imidazolate framework), which is generally stable toward water and organic solvents (e.g., methanol), in C3H6/C3H8 separation has been investigated. Based on single-crystal X-ray diffraction data, the aperture of ZIF-8 has been reported to be 3.4 Å [63], which is fairly flexible compared with rigid frameworks (e.g., zeolites and zeotype materials). Thus, considering the flexibility of the ligands present in ZIF-8, Zhang et al. [64] estimated (revised) the effective aperture size. According to the adsorption kinetics of C3H6 and C3H8, C3H6 has a shorter equilibration time than C3H8 (Figure 4a), indicating a potential means of size discrimination between C3H6 and C3H8. However, a precise cut-off for the ZIF-8 aperture size cannot be determined. As an alternative, based on the diffusivities of gaseous molecules against pure ZIF-8 membrane (as calculated from the Maxwell model [23,65]), the diffusion selectivity of C3H6/C3H8 has been reported to exceed 100 at 35 °C (Figure 4b); this result indicates the feasibility of using the ZIF-8 aperture size to achieve effective molecular sieving between C3H6 and C3H8. As evident from Table 1, kinetic diameter alone is not satisfactory for characterizing the molecular diffusion for selective C3H6 transport. Thus, as suggested by Zhang et al. [64], a hybrid molecular dimension scale (i.e., a combination of kinetic and van der Waals diameters) is required to effectively explain the feasibility of ZIF-8 to perform effective C3H6/C3H8 separation.
Numerous researchers have investigated the use of ZIF-8 for effective C3H6/C3H8 separation. The first widely known investigation of ZIF-8 as a filler in MMMs for C3H6/C3H8 separation was conducted by Zhang et al. [66]. Based on the gas adsorption isotherms and adsorption kinetics of both C3H6 and C3H8, their sorption capacities and fractional uptake are comparable. The incorporation of 48 wt% ZIF-8 in 6FDA-TMPDA membranes was found to increase the C3H6 permeability and C3H6/C3H8 selectivity by 258% and 150%, respectively. These conflicting results are attributable to the difficulty of accurately characterizing the adsorption kinetics (i.e., determining the mass and heat transfer resistance and particle size) during the equilibrium gas adsorption measurements [12,67].
Efforts have also been made to improve the separation performance of ZIF-8-containing MMMs. Specifically, polymers other than 6FDA-TMPDA [68] have been investigated with the aim of surpassing the upper bound limit for C3H6/C3H8 separation. For example, Ma et al. [69] examined the use of PIM-6FDA-OH, which possesses higher C3H6/C3H8 selectivity (30 vs 12 for 6FDA-TMPDA) due to the hydroxyl (–OH) group. As evidenced by X-ray photon spectroscopy (XPS) data (Figure 5a), hydrogen bonding (N…H–O) between the nitrogen moiety in the ZIF-8 ligand and the –OH group in the polymer improves the compatibility between the filler and the polymer (Figure 5b), in turn improving the C3H6 permeability and C3H6/C3H8 selectivity by 1.086% and 43%, respectively (Figure 5c), at 65 wt% ZIF-8. The incorporation of ZIF-8 in rubbery polymers, such as PDMS and XLPEO, has also been explored given their high C3H6 permeability [70,71], but these compounds have yet to be used in industrial operations due to their low inherent brittleness [24].
Other porous materials have also been explored for C3H6/C3H8 separation. For instance, cobalt-substituted ZIF-8 (ZIF-67) has been proposed to be more efficient in this regard than ZIF-8 because the Co–N bond in ZIF-67 is stiffer than the Zn–N bond in ZIF-8. In particular, due to the higher electronegativity of Co compared with Zn, the generation of a stiffer ionic bond has the potential to restrict the flipping motion of the organic ligands present in the respective MOFs (Figure 6a) [72,73,74]. An et al. [75] demonstrated that the incorporation of 20 wt% ZIF-67 in 6FDA-TMPDA improved the C3H6/C3H8 selectivity by 165%, whereas a comparable ZIF-8 incorporation resulted in an 83% improvement. Furthermore, Oh et al. [76] developed Co2+ and Zn2+ mixed-metal hybrids (ZIF-8-67) to take advantage of the benefits of both fillers. A comparison study of ZIF-8, ZIF-67, and ZIF-8-67 revealed that ZIF-8-67 could improve C3H6 permeability to a much larger extent than could the ZIF-8 and ZIF-67 frameworks (Figure 6b). The use of composite fillers, such as ZIF-67 with porous graphene oxide (GO) [77], has also been explored. The creation of three-dimensional architectures with MOFs and GO has been reported to increase the accessible surface area due to the favorable interactions between the functional groups in GO and the ligands in MOFs [78,79,80].
Frameworks other than those involving ZIFs have also been investigated in previous research. For example, Liu et al. [81] examined the potential of Zr-fum-fcu-MOF crystals in C3H6/C3H8 separation and found that the resulting C3H6/C3H8 sorption selectivity is comparable with that of ZIF-8, and a pore aperture size of 3.3–4.6 Å was found to further improve the C3H6/C3H8 diffusion selectivity. Furthermore, it has been postulated that C3H8 has a comparatively higher energy barrier to pass through Zr-fum-fcu-MOF crystals than does C3H6, due to the C–C bond rotation of C3H8. Similarly, Lee et al. [82] developed defect-engineered MOFs derived from hydrolytically-stable [83] UiO-66 to create additional sorption sites for effective C3H6/C3H8 separation; their defect-engineered UiO-66-based MMM significantly improved C3H6 permeance (1.005%) relative to the control (UiO-66-based MMM, 308%), without substantially jeopardizing the C3H6/C3H8 selectivity.
Table 3. Performance of mixed-matrix membranes (MMMs) in C3 hydrocarbon separation.
Table 3. Performance of mixed-matrix membranes (MMMs) in C3 hydrocarbon separation.
MembraneMeasurement ConditionGas Permeability (barrer)Gas SelectivityRef.
Polymer/
Support
FillerLoading (wt%)Pres. (bar)Temp. (°C)C3H6C3H8CH4C3H6/C3H8
6FDA-
Durene
ZIF-830235492.8-17.5[71]
6FDA-
TMPDA
UiO-662023587.48.9-9.82[82]
6FDA-
TMPDA
UIO-TF3620235236.524.5-9.65[82]
6FDA-
TMPDA
ZIF-672023547.83.7-14.06[77]
6FDA-
TMPDA
ZIF-6720-3534.11.1-31.00[75]
6FDA-
TMPPDA
ZIF-67/GO (ZGO67)2023543.13.1-13.90[77]
6FDA-
TMPDA
ZIF-820-3527.71.8-15.39[75]
6FDA-
TMPDA
ZIF-848235522.4-21.67[69]
6FDA-
TMPDA
ZIF-84823556.21.8-31.22[66]
6FDA-
TMPDA
ZIF-8302351512115.00[68]
6FDA-
TMPDA
ZIF-83023510 [1]1.5 [1]-6.67[68]
6FDA-
TMPDA
ZIF-830235351.7-20.59[71]
6FDA-
TMPDA
ZIF-8–6721.9235653.9-16.67[76]
6FDA-
TMPDA
ZPGO672023555.43.9-14.21[77]
6FDA-
TMPDA
Zr-fum-fcu-MOF1513540.42.8-14.43[81]
6FDA-
TMPDA (with PDMS)
ZIF-8302356 [1]0.4 [1]-15.00[68]
α-aluminaZeolite MFI-925-3006 [2]-[57]
Cellulose
Acetate
silica30--0.0920.018-5.11[38]
Comb
copolymers
AgBF4/MgO nanosheet (3:7)---11.8 [1]0.9 [1]-13.11[84]
DBzPBI-BuIZIF-8302.735110.4-27.50[58]
Ethyl
Cellulose
Fullerene101-103-3.33[85]
MatrimidSilica203300.160.008-20.00[39]
PEBAX® 1657ZIF-83023519578-2.50[71]
PEBAX® 2533ZIF-820235420220-1.91[71]
PDMSACN-N10435-11,0001000-[86]
PDMSACN-O10435-13,0001050-[86]
PDMSSiO2105359800---[87]
PDMSZIF-8202--160 [1]--[70]
PIM-
6FDA-OH
ZIF-865235351.1-31.81[69]
PolyurethaneZeolite 4A10230-7853-[51]
PolyurethaneZSM-520230-71.332.3-[51]
PVAcZIF-840235242-12[71]
XLPEOZIF-842235281.9-14.73[71]
XLPEOZIF-8-IT20-3511.61.8-6.44[88]
XLPEOZIF-8-NC20-3513.81.9-7.26[88]
XLPEOZIF-8-NR20-3516.61.8-9.22[88]
XLPEOZIF-8-OP20-3512.52.2-5.68[88]
XLPEOZIF-8-RD20-3510.21.9-5.37[88]
[1] Permeance in GPU; 6FDA = 4,4′-(hexafluoroisoprropylidene)diphthalic anhydride; CAN = adsorptive carbon nanoparticles; DBzPBI-BuI = substituted polybenzimidazole; IT = interpenetration twin; NC = nanocube; NR = nanorod; OP = octagonal plate; PIM = polymer of intrinsic microporosity; PDMS = polydimethylsiloxane; PVAc = polyvinyl acetate; TF = trifluoroacetic acid; TMPDA = 2,4,6-trimethyle-m-phenylenediamine; XLPEO = crosslinked poly(ethylene oxide).

4. Comparison of Membrane Performance against the Upper Bound

In this study, the hydrocarbon separation performance of membranes is benchmarked against the upper bound curve for C2H4/C2H6 and C3H6/C3H8 separation, based on the data in Table 2 and Table 3, respectively. The parameters used to obtain the upper bound plot (Figure 7) are listed in Table 4. The results reveal that achieving excellent light hydrocarbon separation performance is highly challenging. For example, the trend in Figure 7a indicates that the upper bound limit could easily be surpassed by using polymeric membranes with M2dobdc series fillers, considering their high C2H4 adsorption and C2H4/C2H6 selectivity (based on IAST calculations). However, MOFs with open metal sites generally exhibit poor hydrolytic stability, which may affect their long-term separation performance. The incorporation of deep eutectic solvents (DES) into polymeric membrane [62] has been proven to be feasible in substantially improving C2H4/C2H6 selectivity, the challenges of developing cost-effective and highly stable DES notwithstanding [89,90]. As discussed in Section 3, the use of ZIF-8 as porous material in membranes for C3H6/C3H8 separation has been proven effective, as ZIF-8 substantially improves C3H6/C3H8 selectivity. Thus, the incorporation of ZIF-8 in polymeric membranes has the potential to surpass the C3H6/C3H8 upper bound limit, as indicated in Figure 7b.

5. Conclusions and Future Perspectives

This perspective paper presents an overview of MMMs used for C2H4/C2H6 and C3H6/C3H8 separation. Such light hydrocarbon separations are generally challenging due to the similar polarizabilities, kinetic diameters, and van der Waals diameters of the adsorbates, which make it difficult to realize high selectivity. Moreover, because light hydrocarbon molecules are bulkier than carbon dioxide, nitrogen, and methane (which have been extensively investigated for CO2 separation [23,28,91]), membranes with high FFV are highly desirable for separation. Thus, most relevant studies have investigated high-performance polyimides, namely 6FDA-based polymers and PIMs [44,92,93]. However, these polymers require extensive monomer purification, making their large-scale fabrication difficult [53]. Composite membranes are thus a feasible alternative for light hydrocarbon separation. This can be potentially achieved with the utilization of porous fillers possessing large accessible surface areas that are not limited to zeolites and MOFs, as they can effectively increase the diffusion rates of gas-molecule MMMs.
Future efforts should focus on the investigation of MMMs for acetylene separation. Acetylene is typically produced via the thermal cracking of CH4 and typically contains CO2 as the byproduct. Acetylene is also present as an impurity in C2H4 production, which tends to disrupt the effective polymerization of C2H4. Acetylene separation is largely reliant on the use of adsorbents, but effective membrane-based separation is difficult because of the comparable properties of the adsorbates (Table 1). Nevertheless, the membrane performance of MMMs can be expected to improve through the use of porous materials proven to be effective in previous studies.

Author Contributions

Writing—draft preparation, review and editing, C.Y.C. and T.-H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Universiti Teknologi Petronas (UTP) and National Research Foundation of Korea (NRF), grant funded by the Korean government MSIT (Reference number: 2021M313A1084977).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sholl, D.S.; Lively, R.P. Seven chemical separations to change the world. Nature 2016, 532, 435. [Google Scholar] [CrossRef]
  2. Ren, Y.; Liang, X.; Dou, H.; Ye, C.; Guo, Z.; Wang, J.; Pan, Y.; Wu, H.; Guiver, M.D.; Jiang, Z. Membrane-Based Olefin/Paraffin Separations. Adv. Sci. 2020, 7, 2001398. [Google Scholar] [CrossRef] [PubMed]
  3. Garside, M. Global Production Capacity of Propylene 2018–2030. Available online: https://www.statista.com/statistics/1065879/global-propylene-production-capacity/ (accessed on 3 July 2021).
  4. Garside, M. Global Production Capacity of Ethylene 2014–2024. Available online: https://www.statista.com/statistics/1067372/global-ethylene-production-capacity/#:~:text=The%20production%20capacity%20of%20ubiquitous,to%20283%20million%20metric%20tons (accessed on 3 July 2021).
  5. Chen, X.Y.; Xiao, A.; Rodrigue, D. Polymer-based Membranes for Propylene/Propane Separation. Sep. Purif. Rev. 2021, 51, 1–13. [Google Scholar] [CrossRef]
  6. Lee, J.; Chuah, C.Y.; Kim, J.; Kim, Y.; Ko, N.; Seo, Y.; Kim, K.; Bae, T.H.; Lee, E. Separation of acetylene from carbon dioxide and ethylene by a water-stable microporous metal–organic framework with aligned imidazolium groups inside the channels. Angew. Chem. Int. Ed. 2018, 130, 7995–7999. [Google Scholar] [CrossRef]
  7. Nitzsche, R.; Budzinski, M.; Gröngröft, A. Techno-economic assessment of a wood-based biorefinery concept for the production of polymer-grade ethylene, organosolv lignin and fuel. Bioresour. Technol. 2016, 200, 928–939. [Google Scholar] [CrossRef]
  8. Wen, H.-M.; Li, B.; Wang, H.; Krishna, R.; Chen, B. High acetylene/ethylene separation in a microporous zinc (II) metal–organic framework with low binding energy. Chem. Commun. 2016, 52, 1166–1169. [Google Scholar] [CrossRef] [Green Version]
  9. Drobny, J.G. 4-Fluoroelastomer Monomers. In Fluoroelastomers Handbook, 2nd ed.; Drobny, J.G., Ed.; William Andrew Publishing: Norwich, NY, USA, 2016; pp. 29–40. [Google Scholar] [CrossRef]
  10. Van Miltenburg, A.; Zhu, W.; Kapteijn, F.; Moulijn, J. Adsorptive separation of light olefin/paraffin mixtures. Chem. Eng. Res. Des. 2006, 84, 350–354. [Google Scholar] [CrossRef] [Green Version]
  11. Liang, W.; Zhang, Y.; Wang, X.; Wu, Y.; Zhou, X.; Xiao, J.; Li, Y.; Wang, H.; Li, Z. Asphalt-derived high surface area activated porous carbons for the effective adsorption separation of ethane and ethylene. Chem. Eng. Sci. 2017, 162, 192–202. [Google Scholar] [CrossRef]
  12. Chuah, C.Y.; Lee, Y.; Bae, T.-H. Potential of adsorbents and membranes for SF6 capture and recovery: A review. Chem. Eng. J. 2020, 404, 126577. [Google Scholar] [CrossRef]
  13. Alcántara-Avila, J.R.; Gómez-Castro, F.I.; Segovia-Hernández, J.G.; Sotowa, K.-I.; Horikawa, T. Optimal design of cryogenic distillation columns with side heat pumps for the propylene/propane separation. Chem. Eng. Process. Process Intensif. 2014, 82, 112–122. [Google Scholar] [CrossRef]
  14. Yang, Y.; Chuah, C.Y.; Bae, T.-H. Polyamine-appended porous organic polymers for efficient post-combustion CO2 capture. Chem. Eng. J. 2019, 358, 1227–1234. [Google Scholar] [CrossRef]
  15. Yang, Y.; Chuah, C.Y.; Bae, T.-H. Highly efficient carbon dioxide capture in diethylenetriamine-appended porous organic polymers: Investigation of structural variations of chloromethyl monomers. J. Ind. Eng. Chem. 2020, 88, 207–214. [Google Scholar] [CrossRef]
  16. Lin, R.-B.; Li, L.; Zhou, H.-L.; Wu, H.; He, C.; Li, S.; Krishna, R.; Li, J.; Zhou, W.; Chen, B. Molecular sieving of ethylene from ethane using a rigid metal–organic framework. Nat. Mater. 2018, 17, 1128–1133. [Google Scholar] [CrossRef]
  17. Aguado, S.; Bergeret, G.r.; Daniel, C.; Farrusseng, D. Absolute molecular sieve separation of ethylene/ethane mixtures with silver zeolite A. J. Am. Chem. Soc. 2012, 134, 14635–14637. [Google Scholar] [CrossRef]
  18. Anson, A.; Wang, Y.; Lin, C.; Kuznicki, T.; Kuznicki, S. Adsorption of ethane and ethylene on modified ETS-10. Chem. Eng. Sci. 2008, 63, 4171–4175. [Google Scholar] [CrossRef]
  19. Maghsoudi, H. Comparative study of adsorbents performance in ethylene/ethane separation. Adsorption 2016, 22, 985–992. [Google Scholar] [CrossRef]
  20. Liang, B.; Zhang, X.; Xie, Y.; Lin, R.-B.; Krishna, R.; Cui, H.; Li, Z.; Shi, Y.; Wu, H.; Zhou, W. An ultramicroporous metal–organic framework for high sieving separation of propylene from propane. J. Am. Chem. Soc. 2020, 142, 17795–17801. [Google Scholar] [CrossRef]
  21. Li, J.; Jiang, L.; Chen, S.; Kirchon, A.; Li, B.; Li, Y.; Zhou, H.-C. Metal–organic framework containing planar metal-binding sites: Efficiently and cost-effectively enhancing the kinetic separation of C2H2/C2H4. J. Am. Chem. Soc. 2019, 141, 3807–3811. [Google Scholar] [CrossRef]
  22. Chuah, C.Y.; Li, W.; Yang, Y.; Bae, T.-H. Evaluation of porous adsorbents for CO2 capture under humid conditions: The importance of recyclability. Chem. Eng. J. Adv. 2020, 3, 100021. [Google Scholar] [CrossRef]
  23. Chuah, C.Y.; Goh, K.; Yang, Y.; Gong, H.; Li, W.; Karahan, H.E.; Guiver, M.D.; Wang, R.; Bae, T.-H. Harnessing filler materials for enhancing biogas separation membranes. Chem. Rev. 2018, 118, 8655–8769. [Google Scholar] [CrossRef]
  24. Gong, H.; Chuah, C.Y.; Yang, Y.; Bae, T.-H. High performance composite membranes comprising Zn(pyrz)2(SiF6) nanocrystals for CO2/CH4 separation. J. Ind. Eng. Chem. 2018, 60, 279–285. [Google Scholar] [CrossRef]
  25. Karahan, H.E.; Goh, K.; Zhang, C.; Yang, E.; Yıldırım, C.; Chuah, C.Y.; Ahunbay, M.G.; Lee, J.; Tantekin-Ersolmaz, Ş.B.; Chen, Y. MXene materials for designing advanced separation membranes. Adv. Mater. 2020, 32, 1906697. [Google Scholar] [CrossRef] [PubMed]
  26. Robeson, L.M. The upper bound revisited. J. Membr. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
  27. Robeson, L.M. Correlation of separation factor versus permeability for polymeric membranes. J. Membr. Sci. 1991, 62, 165–185. [Google Scholar] [CrossRef]
  28. Samarasinghe, S.; Chuah, C.Y.; Yang, Y.; Bae, T.-H. Tailoring CO2/CH4 separation properties of mixed-matrix membranes via combined use of two-and three-dimensional metal-organic frameworks. J. Membr. Sci. 2018, 557, 30–37. [Google Scholar] [CrossRef]
  29. Dou, H.; Jiang, B.; Xu, M.; Zhang, Z.; Wen, G.; Peng, F.; Yu, A.; Bai, Z.; Sun, Y.; Zhang, L. Boron nitride membranes with a distinct nanoconfinement effect for efficient ethylene/ethane separation. Angew. Chem. Int. Ed. 2019, 131, 14107–14113. [Google Scholar] [CrossRef]
  30. Rungta, M.; Xu, L.; Koros, W.J. Carbon molecular sieve dense film membranes derived from Matrimid® for ethylene/ethane separation. Carbon 2012, 50, 1488–1502. [Google Scholar] [CrossRef]
  31. Li, J.-R.; Kuppler, R.J.; Zhou, H.-C. Selective gas adsorption and separation in metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1477–1504. [Google Scholar] [CrossRef]
  32. Chuah, C.Y.; Lee, H.; Bae, T.-H. Recent advances of nanoporous adsorbents for light hydrocarbon (C1–C3) separation. Chem. Eng. J. 2022, 430, 132654. [Google Scholar] [CrossRef]
  33. Najari, S.; Saeidi, S.; Gallucci, F.; Drioli, E. Mixed matrix membranes for hydrocarbons separation and recovery: A critical review. Rev. Chem. Eng. 2019, 37, 363–406. [Google Scholar] [CrossRef]
  34. Yang, L.; Qian, S.; Wang, X.; Cui, X.; Chen, B.; Xing, H. Energy-efficient separation alternatives: Metal–organic frameworks and membranes for hydrocarbon separation. Chem. Soc. Rev. 2020, 49, 5359–5406. [Google Scholar] [CrossRef] [PubMed]
  35. Chuah, C.Y.; Jiang, X.; Goh, K.; Wang, R. Recent progress in mixed-matrix membranes for hydrogen separation. Membranes 2021, 11, 666. [Google Scholar] [CrossRef] [PubMed]
  36. Burns, R.L.; Koros, W.J. Defining the challenges for C3H6/C3H8 separation using polymeric membranes. J. Membr. Sci. 2003, 211, 299–309. [Google Scholar] [CrossRef]
  37. Rungta, M.; Zhang, C.; Koros, W.J.; Xu, L. Membrane-based ethylene/ethane separation: The upper bound and beyond. AIChE J. 2013, 59, 3475–3489. [Google Scholar] [CrossRef]
  38. Naghsh, M.; Sadeghi, M.; Moheb, A.; Chenar, M.P.; Mohagheghian, M. Separation of ethylene/ethane and propylene/propane by cellulose acetate–silica nanocomposite membranes. J. Membr. Sci. 2012, 423, 97–106. [Google Scholar] [CrossRef]
  39. Davoodi, S.M.; Sadeghi, M.; Naghsh, M.; Moheb, A. Olefin–paraffin separation performance of polyimide Matrimid®/silica nanocomposite membranes. RSC Adv. 2016, 6, 23746–23759. [Google Scholar] [CrossRef]
  40. Ploegmakers, J.; Japip, S.; Nijmeijer, K. Mixed matrix membranes containing MOFs for ethylene/ethane separation Part A: Membrane preparation and characterization. J. Membr. Sci. 2013, 428, 445–453. [Google Scholar] [CrossRef]
  41. Chuah, C.Y.; Bae, T.-H. Incorporation of Cu3BTC2 nanocrystals to increase the permeability of polymeric membranes in O2/N2 separation. BMC Chem. Eng. 2019, 1, 2. [Google Scholar] [CrossRef]
  42. Bachman, J.E.; Smith, Z.P.; Li, T.; Xu, T.; Long, J.R. Enhanced ethylene separation and plasticization resistance in polymer membranes incorporating metal–organic framework nanocrystals. Nat. Mater. 2016, 15, 845–849. [Google Scholar] [CrossRef]
  43. Bae, T.-H.; Long, J.R. CO2/N2 separations with mixed-matrix membranes containing Mg2(dobdc) nanocrystals. Energy Environ. Sci. 2013, 6, 3565–3569. [Google Scholar] [CrossRef]
  44. Chuah, C.Y.; Samarasinghe, S.A.S.C.; Li, W.; Goh, K.; Bae, T.-H. Leveraging nanocrystal HKUST-1 in mixed-matrix membranes for ethylene/ethane separation. Membranes 2020, 10, 74. [Google Scholar] [CrossRef] [PubMed]
  45. Qian, S.; Xia, L.; Yang, L.; Wang, X.; Suo, X.; Cui, X.; Xing, H. Defect-free mixed-matrix membranes consisting of anion-pillared metal-organic frameworks and poly (ionic liquid)s for separation of acetylene from ethylene. J. Membr. Sci. 2020, 611, 118329. [Google Scholar] [CrossRef]
  46. Tomé, L.C.; Guerreiro, D.C.; Teodoro, R.M.; Alves, V.D.; Marrucho, I.M. Effect of polymer molecular weight on the physical properties and CO2/N2 separation of pyrrolidinium-based poly (ionic liquid) membranes. J. Membr. Sci. 2018, 549, 267–274. [Google Scholar] [CrossRef]
  47. Cowan, M.G.; Masuda, M.; McDanel, W.M.; Kohno, Y.; Gin, D.L.; Noble, R.D. Phosphonium-based poly (Ionic liquid) membranes: The effect of cation alkyl chain length on light gas separation properties and Ionic conductivity. J. Membr. Sci. 2016, 498, 408–413. [Google Scholar] [CrossRef]
  48. Ye, L.; Wan, L.; Tang, J.; Li, Y.; Huang, F. Novolac-based poly(1,2,3-triazolium) s with good ionic conductivity and enhanced CO2 permeation. RSC Adv. 2018, 8, 8552–8557. [Google Scholar] [CrossRef] [Green Version]
  49. Li, W.; Chuah, C.Y.; Kwon, S.; Goh, K.; Wang, R.; Na, K.; Bae, T.-H. Nanosizing zeolite 5A fillers in mixed-matrix carbon molecular sieve membranes to improve gas separation performance. Chem. Eng. J. Adv. 2020, 2, 100016. [Google Scholar] [CrossRef]
  50. Li, W.; Chuah, C.Y.; Bae, T.-H. Hierarchical 5A Zeolite-Containing Carbon Molecular Sieve Membranes for O2/N2 Separation. Membr. J. 2020, 30, 260–268. [Google Scholar] [CrossRef]
  51. Tirouni, I.; Sadeghi, M.; Pakizeh, M. Separation of C3H8 and C2H6 from CH4 in polyurethane–zeolite 4Å and ZSM-5 mixed matrix membranes. Sep. Purif. Technol. 2015, 141, 394–402. [Google Scholar] [CrossRef]
  52. Li, W.; Goh, K.; Chuah, C.Y.; Bae, T.-H. Mixed-matrix carbon molecular sieve membranes using hierarchical zeolite: A simple approach towards high CO2 permeability enhancements. J. Membr. Sci. 2019, 588, 117220. [Google Scholar] [CrossRef]
  53. Chuah, C.Y.; Lee, J.; Bao, Y.; Song, J.; Bae, T.-H. High-performance porous carbon-zeolite mixed-matrix membranes for CO2/N2 separation. J. Membr. Sci. 2021, 622, 119031. [Google Scholar] [CrossRef]
  54. Chuah, C.Y.; Anwar, S.N.B.M.; Weerachanchai, P.; Bae, T.-H.; Goh, K.; Wang, R. Scaling-up defect-free asymmetric hollow fiber membranes to produce oxygen-enriched gas for integration into municipal solid waste gasification process. J. Membr. Sci. 2021, 640, 119787. [Google Scholar] [CrossRef]
  55. Chuah, C.Y.; Goh, K.; Bae, T.-H. Enhanced performance of carbon molecular sieve membranes incorporating zeolite nanocrystals for air separation. Membranes 2021, 11, 489. [Google Scholar] [CrossRef]
  56. Chen, G.; Chen, X.; Pan, Y.; Ji, Y.; Liu, G.; Jin, W. M-gallate MOF/6FDA-polyimide mixed-matrix membranes for C2H4/C2H6 separation. J. Membr. Sci. 2021, 620, 118852. [Google Scholar] [CrossRef]
  57. Min, B.; Yang, S.; Korde, A.; Jones, C.W.; Nair, S. Single-Step Scalable Fabrication of Zeolite MFI Hollow Fiber Membranes for Hydrocarbon Separations. Adv. Mater. Int. 2020, 7, 2000926. [Google Scholar] [CrossRef]
  58. Kunjattu, S.H.; Ashok, V.; Bhaskar, A.; Pandare, K.; Banerjee, R.; Kharul, U.K. ZIF-8@ DBzPBI-BuI composite membranes for olefin/paraffin separation. J. Membr. Sci. 2018, 549, 38–45. [Google Scholar] [CrossRef]
  59. Ploegmakers, J.; Japip, S.; Nijmeijer, K. Mixed matrix membranes containing MOFs for ethylene/ethane separation—Part B: Effect of Cu3BTC2 on membrane transport properties. J. Membr. Sci. 2013, 428, 331–340. [Google Scholar] [CrossRef]
  60. Higuchi, A.; Agatsuma, T.; Uemiya, S.; Kojima, T.; Mizoguchi, K.; Pinnau, I.; Nagai, K.; Freeman, B.D. Preparation and gas permeation of immobilized fullerene membranes. J. Appl. Polym. Sci. 2000, 77, 529–537. [Google Scholar] [CrossRef]
  61. Díaz, K.; López-González, M.; del Castillo, L.F.; Riande, E. Effect of zeolitic imidazolate frameworks on the gas transport performance of ZIF8-poly(1,4-phenylene ether-ether-sulfone) hybrid membranes. J. Membr. Sci. 2011, 383, 206–213. [Google Scholar] [CrossRef]
  62. Jiang, B.; Zhou, J.; Xu, M.; Dou, H.; Zhang, H.; Yang, N.; Zhang, L. Multifunctional ternary deep eutectic solvent-based membranes for the cost-effective ethylene/ethane separation. J. Membr. Sci. 2020, 610, 118243. [Google Scholar] [CrossRef]
  63. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef] [Green Version]
  64. Zhang, C.; Lively, R.P.; Zhang, K.; Johnson, J.R.; Karvan, O.; Koros, W.J. Unexpected molecular sieving properties of zeolitic imidazolate framework-8. J. Phys. Chem. Lett. 2012, 3, 2130–2134. [Google Scholar] [CrossRef] [PubMed]
  65. Vinh-Thang, H.; Kaliaguine, S. Predictive models for mixed-matrix membrane performance: A review. Chem. Rev. 2013, 113, 4980–5028. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, C.; Dai, Y.; Johnson, J.R.; Karvan, O.; Koros, W.J. High performance ZIF-8/6FDA-DAM mixed matrix membrane for propylene/propane separations. J. Membr. Sci. 2012, 389, 34–42. [Google Scholar] [CrossRef]
  67. Chuah, C.Y.; Goh, K.; Bae, T.-H. Hierarchically structured HKUST-1 nanocrystals for enhanced SF6 capture and recovery. J. Phys. Chem. C 2017, 121, 6748–6755. [Google Scholar] [CrossRef]
  68. Zhang, C.; Zhang, K.; Xu, L.; Labreche, Y.; Kraftschik, B.; Koros, W.J. Highly scalable ZIF-based mixed-matrix hollow fiber membranes for advanced hydrocarbon separations. AIChE J. 2014, 60, 2625–2635. [Google Scholar] [CrossRef]
  69. Ma, X.; Swaidan, R.J.; Wang, Y.; Hsiung, C.-e.; Han, Y.; Pinnau, I. Highly compatible hydroxyl-functionalized microporous polyimide-ZIF-8 mixed matrix membranes for energy efficient propylene/propane separation. ACS Appl. Nano Mater. 2018, 1, 3541–3547. [Google Scholar] [CrossRef]
  70. Fang, M.; Wu, C.; Yang, Z.; Wang, T.; Xia, Y.; Li, J. ZIF-8/PDMS mixed matrix membranes for propane/nitrogen mixture separation: Experimental result and permeation model validation. J. Membr. Sci. 2015, 474, 103–113. [Google Scholar] [CrossRef]
  71. Liu, D.; Xiang, L.; Chang, H.; Chen, K.; Wang, C.; Pan, Y.; Li, Y.; Jiang, Z. Rational matching between MOFs and polymers in mixed matrix membranes for propylene/propane separation. Chem. Eng. Sci. 2019, 204, 151–160. [Google Scholar] [CrossRef]
  72. Kwon, H.T.; Jeong, H.-K.; Lee, A.S.; An, H.S.; Lee, J.S. Heteroepitaxially grown zeolitic imidazolate framework membranes with unprecedented propylene/propane separation performances. J. Am. Chem. Soc. 2015, 137, 12304–12311. [Google Scholar] [CrossRef]
  73. Pimentel, B.R.; Parulkar, A.; Zhou, E.k.; Brunelli, N.A.; Lively, R.P. Zeolitic imidazolate frameworks: Next-generation materials for energy-efficient gas separations. ChemSusChem 2014, 7, 3202–3240. [Google Scholar] [CrossRef]
  74. Krokidas, P.; Castier, M.; Moncho, S.; Sredojevic, D.N.; Brothers, E.N.; Kwon, H.T.; Jeong, H.-K.; Lee, J.S.; Economou, I.G. ZIF-67 framework: A promising new candidate for propylene/propane separation. experimental data and molecular simulations. J. Phys. Chem. C 2016, 120, 8116–8124. [Google Scholar] [CrossRef]
  75. An, H.; Park, S.; Kwon, H.T.; Jeong, H.-K.; Lee, J.S. A new superior competitor for exceptional propylene/propane separations: ZIF-67 containing mixed matrix membranes. J. Membr. Sci. 2017, 526, 367–376. [Google Scholar] [CrossRef] [Green Version]
  76. Oh, J.W.; Cho, K.Y.; Kan, M.-Y.; Yu, H.J.; Kang, D.-Y.; Lee, J.S. High-flux mixed matrix membranes containing bimetallic zeolitic imidazole framework-8 for C3H6/C3H8 separation. J. Membr. Sci. 2020, 596, 117735. [Google Scholar] [CrossRef]
  77. Moghadam, F.; Lee, T.H.; Park, I.; Park, H.B. Thermally annealed polyimide-based mixed matrix membrane containing ZIF-67 decorated porous graphene oxide nanosheets with enhanced propylene/propane selectivity. J. Membr. Sci. 2020, 603, 118019. [Google Scholar] [CrossRef]
  78. Chuah, C.Y.; Lee, J.; Bae, T.-H. Graphene-based Membranes for H2 Separation: Recent Progress and Future Perspective. Membranes 2020, 10, 336. [Google Scholar] [CrossRef]
  79. Li, W.; Chuah, C.Y.; Yang, Y.; Bae, T.-H. Nanocomposites formed by in situ growth of NiDOBDC nanoparticles on graphene oxide sheets for enhanced CO2 and H2 storage. Micropor. Mesopor. Mater. 2018, 265, 35–42. [Google Scholar] [CrossRef]
  80. Li, W.; Chuah, C.Y.; Nie, L.; Bae, T.-H. Enhanced CO2/CH4 selectivity and mechanical strength of mixed-matrix membrane incorporated with NiDOBDC/GO composite. J. Ind. Eng. Chem. 2019, 74, 118–125. [Google Scholar] [CrossRef]
  81. Liu, Y.; Chen, Z.; Liu, G.; Belmabkhout, Y.; Adil, K.; Eddaoudi, M.; Koros, W. Conformation-controlled molecular sieving effects for membrane-based propylene/propane separation. Adv. Mater. 2019, 31, 1807513. [Google Scholar] [CrossRef]
  82. Lee, T.H.; Jung, J.G.; Kim, Y.J.; Roh, J.S.; Yoon, H.W.; Ghanem, B.S.; Kim, H.W.; Cho, Y.H.; Pinnau, I.; Park, H.B. Defect Engineering in Metal–Organic Frameworks Towards Advanced Mixed Matrix Membranes for Efficient Propylene/Propane Separation. Angew. Chem. Int. Ed. 2021, 133, 13191–13198. [Google Scholar] [CrossRef]
  83. Chuah, C.Y.; Lee, J.; Song, J.; Bae, T.-H. CO2/N2 separation properties of polyimide-based mixed-matrix membranes comprising UiO-66 with various functionalities. Membranes 2020, 10, 154. [Google Scholar] [CrossRef]
  84. Park, C.H.; Lee, J.H.; Jung, J.P.; Kim, J.H. Mixed matrix membranes based on dual-functional MgO nanosheets for olefin/paraffin separation. J. Membr. Sci. 2017, 533, 48–56. [Google Scholar] [CrossRef]
  85. Sun, H.; Ma, C.; Wang, T.; Xu, Y.; Yuan, B.; Li, P.; Kong, Y. Preparation and characterization of C60-filled ethyl cellulose mixed-matrix membranes for gas separation of propylene/propane. Chem. Eng. Technol. 2014, 37, 611–619. [Google Scholar] [CrossRef]
  86. Heidari, M.; Hosseini, S.S.; Nasrin, M.O.; Ghadimi, A. Synthesis and fabrication of adsorptive carbon nanoparticles (ACNs)/PDMS mixed matrix membranes for efficient CO2/CH4 and C3H8/CH4 separation. Sep. Purif. Technol. 2019, 209, 503–515. [Google Scholar] [CrossRef]
  87. Beltran, A.B.; Nisola, G.M.; Cho, E.; Lee, E.E.D.; Chung, W.-J. Organosilane modified silica/polydimethylsiloxane mixed matrix membranes for enhanced propylene/nitrogen separation. Appl. Surf. Sci. 2011, 258, 337–345. [Google Scholar] [CrossRef]
  88. Yang, F.; Mu, H.; Wang, C.; Xiang, L.; Yao, K.X.; Liu, L.; Yang, Y.; Han, Y.; Li, Y.; Pan, Y. Morphological map of ZIF-8 crystals with five distinctive shapes: Feature of filler in mixed-matrix membranes on C3H6/C3H8 separation. Chem. Mater. 2018, 30, 3467–3473. [Google Scholar] [CrossRef]
  89. Jeong, S.; Kang, S.W. Effect of Ag2O nanoparticles on long-term stable polymer/AgBF4/Al(NO3)3 complex membranes for olefin/paraffin separation. Chem. Eng. J. 2017, 327, 500–504. [Google Scholar] [CrossRef]
  90. Yoon, K.W.; Kang, Y.S.; Kang, S.W. Activated Ag ions and enhanced gas transport by incorporation of KIT-6 for facilitated olefin transport membranes. J. Membr. Sci. 2016, 513, 95–100. [Google Scholar] [CrossRef]
  91. Chuah, C.Y.; Li, W.; Samarasinghe, S.; Sethunga, G.; Bae, T.-H. Enhancing the CO2 separation performance of polymer membranes via the incorporation of amine-functionalized HKUST-1 nanocrystals. Micropor. Mesopor. Mater. 2019, 290, 109680. [Google Scholar] [CrossRef]
  92. Li, P.; Chung, T.-S.; Paul, D. Gas sorption and permeation in PIM-1. J. Membr. Sci. 2013, 432, 50–57. [Google Scholar] [CrossRef]
  93. Luque-Alled, J.M.; Ameen, A.W.; Alberto, M.; Tamaddondar, M.; Foster, A.B.; Budd, P.M.; Vijayaraghavan, A.; Gorgojo, P. Gas separation performance of MMMs containing (PIM-1)-functionalized GO derivatives. J. Membr. Sci. 2021, 623, 118902. [Google Scholar] [CrossRef]
Figure 1. Membranes for gas separation.
Figure 1. Membranes for gas separation.
Membranes 12 00201 g001
Figure 2. Performance of mixed-matrix membranes (MMMs) with 10 and 20 wt% HKUST-1 loading in ODPA-TMPDA and 6FDA-TMPDA membranes; 20 wt% HKUST-1, whose performance is comparable with those of other membranes reported in previous research, has the potential to surpass the upper bound limit of C2H4/C2H6 separation. Reprinted with permission from [44], Creative Commons Attribution 4.0.
Figure 2. Performance of mixed-matrix membranes (MMMs) with 10 and 20 wt% HKUST-1 loading in ODPA-TMPDA and 6FDA-TMPDA membranes; 20 wt% HKUST-1, whose performance is comparable with those of other membranes reported in previous research, has the potential to surpass the upper bound limit of C2H4/C2H6 separation. Reprinted with permission from [44], Creative Commons Attribution 4.0.
Membranes 12 00201 g002
Figure 3. (a) C2H2 and C2H4 adsorption of SIFSIX-2-Cu-i and ZRFSIX-2-Ni-i at 25 °C; (b) C2H2 and C2H4 permeability of MMMs with SIFSIX-2-Cu-i and ZRFSIX-2-Ni-i loading in poly(ionic liquids). Reprinted with permission from [45], copyright 2020, Elsevier.
Figure 3. (a) C2H2 and C2H4 adsorption of SIFSIX-2-Cu-i and ZRFSIX-2-Ni-i at 25 °C; (b) C2H2 and C2H4 permeability of MMMs with SIFSIX-2-Cu-i and ZRFSIX-2-Ni-i loading in poly(ionic liquids). Reprinted with permission from [45], copyright 2020, Elsevier.
Membranes 12 00201 g003
Figure 4. (a) Adsorption kinetics of C3H6 and C3H8 with ZIF-8 at 35 °C; (b) corrected diffusivities and permeability of pure ZIF-8 membrane at 35 °C. Reprinted with permission from [64], copyright 2012, American Chemical Society.
Figure 4. (a) Adsorption kinetics of C3H6 and C3H8 with ZIF-8 at 35 °C; (b) corrected diffusivities and permeability of pure ZIF-8 membrane at 35 °C. Reprinted with permission from [64], copyright 2012, American Chemical Society.
Membranes 12 00201 g004
Figure 5. (a) Effect of incorporating ZIF-8 in PIM-6FDA-OH on the Zn and N signals in XPS spectra. (b) Illustration of the proposed mechanism, based on hydrogen bonding between ZIF-8 and the polymer (PIM-6FDA-OH); (c) C3H6/C3H8 separation performance with varying ZIF-8 loading. Theoretical prediction at 30 wt% (orange), 40 wt% (blue), and 50 wt% (green) ZIF-8 are illustrated for comparison. Reprinted with permission from [69], copyright 2018, American Chemical Society.
Figure 5. (a) Effect of incorporating ZIF-8 in PIM-6FDA-OH on the Zn and N signals in XPS spectra. (b) Illustration of the proposed mechanism, based on hydrogen bonding between ZIF-8 and the polymer (PIM-6FDA-OH); (c) C3H6/C3H8 separation performance with varying ZIF-8 loading. Theoretical prediction at 30 wt% (orange), 40 wt% (blue), and 50 wt% (green) ZIF-8 are illustrated for comparison. Reprinted with permission from [69], copyright 2018, American Chemical Society.
Membranes 12 00201 g005
Figure 6. (a) Illustration of the effect of transport of C3H6 and C3H8 upon passing through the pores of ZIF-67. Reprinted with permission from [75], copyright 2017, Elsevier; (b) effect of incorporating ZIF-67, ZIF-8, and ZIF-8-67 at 21.9 wt% loading in polyimide (PI) on C3H6 permeability, C3H6/C3H8 selectivity, C3H6 diffusivity, and C3H6/C3H8 selectivity. Reprinted with permission from [76], copyright 2020 Elsevier.
Figure 6. (a) Illustration of the effect of transport of C3H6 and C3H8 upon passing through the pores of ZIF-67. Reprinted with permission from [75], copyright 2017, Elsevier; (b) effect of incorporating ZIF-67, ZIF-8, and ZIF-8-67 at 21.9 wt% loading in polyimide (PI) on C3H6 permeability, C3H6/C3H8 selectivity, C3H6 diffusivity, and C3H6/C3H8 selectivity. Reprinted with permission from [76], copyright 2020 Elsevier.
Membranes 12 00201 g006
Figure 7. Performance of MMMs in (a) C2H4/C2H6 and (b) C3H6/C3H8 separation. The data points indicated in Panels (a) and (b) are from Table 2 and Table 3, respectively. The parameters used to generate the upper bound line [36,37] are summarized in Table 4.
Figure 7. Performance of MMMs in (a) C2H4/C2H6 and (b) C3H6/C3H8 separation. The data points indicated in Panels (a) and (b) are from Table 2 and Table 3, respectively. The parameters used to generate the upper bound line [36,37] are summarized in Table 4.
Membranes 12 00201 g007
Table 4. Parameters used to obtain the upper bound for C2H4/C2H6 and C3H6/C3H8 separation.
Table 4. Parameters used to obtain the upper bound for C2H4/C2H6 and C3H6/C3H8 separation.
Upper Bound Curve2003 (2013) [1]
λ[2]β[2]
C2H4/C2H60.147.3
C3H6/C3H80.1720.4
[1] The upper bounds for C2H4/C2H6 and C3H6/C3H8 are based on the data obtained in 2013 and 2003, respectively; [2] λ and β are described as the slope and front factor of the upper bound line, for which the relation between permeability and selectivity can be correlated as follows: α = β / P λ .
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chuah, C.Y.; Bae, T.-H. Recent Advances in Mixed-Matrix Membranes for Light Hydrocarbon (C1–C3) Separation. Membranes 2022, 12, 201. https://doi.org/10.3390/membranes12020201

AMA Style

Chuah CY, Bae T-H. Recent Advances in Mixed-Matrix Membranes for Light Hydrocarbon (C1–C3) Separation. Membranes. 2022; 12(2):201. https://doi.org/10.3390/membranes12020201

Chicago/Turabian Style

Chuah, Chong Yang, and Tae-Hyun Bae. 2022. "Recent Advances in Mixed-Matrix Membranes for Light Hydrocarbon (C1–C3) Separation" Membranes 12, no. 2: 201. https://doi.org/10.3390/membranes12020201

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