Next Article in Journal
Magnetically Aligned and Enriched Pathways of Zeolitic Imidazolate Framework 8 in Matrimid Mixed Matrix Membranes for Enhanced CO2 Permeability
Previous Article in Journal
Effect of the Formation of Ultrathin Selective Layers on the Structure and Performance of Thin-Film Composite Chitosan/PAN Membranes for Pervaporation Dehydration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2/N2 Separation Properties of Polyimide-Based Mixed-Matrix Membranes Comprising UiO-66 with Various Functionalities

1
Singapore Membrane Technology Centre, Nanyang Environment and Water Research Institute, Nanyang Technological University, Singapore 637141, Singapore
2
School of Chemical and Biomedical Engineering, Nanyang Technological University, Singapore 637459, Singapore
3
Department of Chemical and Biomolecular Engineering, Korea Advanced Institute of Science and Technology, Daejeon 34141, Korea
*
Author to whom correspondence should be addressed.
Membranes 2020, 10(7), 154; https://doi.org/10.3390/membranes10070154
Submission received: 5 June 2020 / Revised: 10 July 2020 / Accepted: 16 July 2020 / Published: 17 July 2020
(This article belongs to the Special Issue Metal-Organic Framework Membranes for Molecular Separations)

Abstract

:
Nanocrystalline UiO-66 and its derivatives (containing -NH2, -Br, -(OH)2) were developed via pre-synthetic functionalization and incorporated into a polyimide membrane to develop a mixed-matrix membrane (MMM) for CO2/N2 separation. Incorporation of the non-functionalized UiO-66 nanocrystals into the polyimide membrane successfully improved CO2 permeability, with a slight decrease in CO2/N2 selectivity, owing to its large accessible surface area. The addition of other functional groups further improved the CO2/N2 selectivity of the polymeric membrane, with UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2 demonstrating improvements of 12%, 4%, and 17%, respectively. Further evaluation by solubility–diffusivity analysis revealed that the functionalized UiO-66 in MMMs can effectively increase CO2 diffusivity while suppressing N2 sorption, thus, resulting in improved CO2/N2 selectivity. Such results imply that the structural tuning of UiO-66 by the incorporation of various functional groups is an effective strategy to improve the CO2 separation performance of MMMs.

Graphical Abstract

1. Introduction

Carbon capture, storage, and sequestration (CCS) processes have been heavily researched in recent years as a potentially feasible means to minimize the increase in global CO2 concentration.
CO2 capture accounts for 70% of the total cost of a typical CCS process [1,2,3]. Hence, efficient separation methods are required to remove CO2 for mitigating the impact of its emission into the atmosphere. In comparison with conventional gas separation processes, such as cryogenic distillation, absorption, and adsorption, membrane separation processes offer several competitive advantages, such as a reduced plant footprint and greater energy efficient [4,5,6,7,8,9,10,11]. In particular, polymeric membranes are commonly used in such processes due to their well-established synthesis and ease of commercial availability. However, an inevitable trade-off relationship between permeability and selectivity, as evidenced by the so called Robeson upper bound, has proven to be a major drawback of conventional polymeric membranes [12,13]. Meanwhile, zeolites or metal–organic framework (MOF)-based membranes, which demonstrate high gas selectivity, generally show poorer scalability than polymeric membranes due to their inherent brittleness [14,15,16].
The incorporation of porous fillers into polymer films to form mixed-matrix membranes (MMMs) has been explored with the aim of combining the advantages of polymer and porous fillers. In terms of the choice of porous materials, MOFs have attracted substantial research interest due to properties such as their large accessible surface area and micropore volume [17,18,19,20,21]. MOFs are also capable of pre- or post-synthetic functionalization to promote favorable interaction with target gases such as CO2, which is highly polarizable [22,23,24]. In addition, the presence of organic moieties in MOFs also ensures better compatibility with the polymer matrices than zeolites, which typically require additional compatibilizers to mitigate the interfacial defects that would otherwise be present between the filler and polymer [25,26,27,28,29].
In this work, we demonstrate the potential utility of nanocrystalline UiO-66 and its derivatives (UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2) in a polymeric membrane for CO2/N2 separation. The presence of zirconium at the metal sites in UiO-66 allows very stable coordination bonding with the ligand, resulting in high stability under humid conditions [30,31,32]. Furthermore, the synthesis of UiO-66 nanocrystals can be conducted using a facile, scalable method [33]. The CO2/N2 separation performance was further tuned with the incorporation of ligands with various functional groups, namely amine (-NH2), bromine (-Br), and hydroxyl (-OH), via pre-synthetic functionalization, as these functional groups interact favorably with CO2. In terms of the choice of polymeric membrane, an in-house polyimide, ODPA-TMPDA (the abbreviations will be elaborated in Section 2.1), was used. OPDA-TMPDA ( P C O 2 = 88   barrer ) possesses greater CO2 permeability than commercial polymers such as Matrimid ( P C O 2 = 10   barrer ), polysulfone ( P C O 2 = 5   barrer ), and Ultem ( P C O 2 = 2   barrer ) [34,35,36,37,38]. Moreover, ODPA-TMPDA can be readily synthesized without a monomer purification process [39,40]. In contrast, the synthesis of 6FDA-based polymers (6FDA = 4,4′-(hexafluoroisopropylidene)diphthalic anhydride) (e.g., 6FDA-DAM (DAM = 2,4,6-trimethyl-m-phenylenediamine), P C O 2 = 681   barrer ) and PIM-1 (polymer of intrinsic microporosity-1) ( P C O 2 = 5120   barrer ) typically requires the monomers to be purified before polymer synthesis to allow the production of those polymers with a high average molecular weight in order to develop membrane with high intrinsic CO2 permeability [41,42,43,44]. It should be noted that the CO2 permeabilities indicated above are obtained at the condition that is reasonably close to the measurement condition used in this work (35 °C and 1 bar feed pressure, as stated in Section 2.4.3.). Finally, the effect of different functional groups in UiO-66 on CO2/N2 separation performance was systematically studied.

2. Materials and Methods

2.1. Materials

2-Bromoterephthalic acid, 2-aminoterephthalic acid, 2-5-dihydroxyterephthalic acid, 2,4,6-trimethyl-m-phenylenediamine (TMPDA), 4,4′-oxydiphthalic anhydride (ODPA), acetic anhydride (Ac2O), terephthalic acid, triethylamine (TEA), and zirconium(IV) chloride (ZrCl4) were purchased from Sigma Aldrich (Singapore). Chloroform, dimethylformamide (DMF), methanol, N,N-dimethylacetamide (DMAc), and hydrochloric acid (HCl, 37%) were purchased from VWR (Singapore). All other chemicals were used as received without further purification.

2.2. Synthesis of MOFs (UiO-66 and Its Derivatives) and Polymer (ODPA-TMPDA)

UiO-66 nanocrystals were synthesized by the experimental procedure described as follows, with some modifications [33]. In a round-bottom flask, 1.25 g of ZrCl4, 50 mL of DMF and 10 mL of HCl were loaded, and the resulting mixture was sonicated for at least 20 min to ensure that all reactants were completely dissolved and mixed. In a separate flask, 1.23 g of terephthalic acid and 100 mL of DMF were loaded. The contents of this flask were then poured into the above mixture, followed by an additional sonication for at least 20 min. This was followed by heating at 80 °C for 24 h to ensure an effective formation of UiO-66 particles. The precipitated particles were washed and centrifuged with DMF and methanol to remove the unreacted impurities. Before characterization, the samples were dried in a vacuum at 60 °C overnight. For the synthesis of UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2, the porous materials were prepared by a similar protocol to that described above, and 1.34 g of 2-aminoterephthalic acid, 1.84 g of 2-bromoterephthalic acid, and 1.47 g of 2.5-dihydroxyterephthalic acid were added to the solution. The reaction scheme was summarized in Figure 1. The synthesis of ODPA-TMPDA polymer, on the other hand, is conducted based on the procedure as described elsewhere [45]. The reaction scheme is provided in Figure 2.

2.3. Membrane Fabrication

The membranes were fabricated via a solution-casting method to form a dense membrane film. First, UiO-66 and its derivatives were dispersed in chloroform. The dispersion of porous fillers in the suspension was improved with a sonication horn before the polymer was added into the solution. This step is required for the aggregated nanocrystals of UiO-66 and its derivatives to be dispersed readily in chloroform. To minimize solvent evaporation of chloroform during the sonication process (due to its low boiling point), an ice bath was used. The resulting dope solution was stirred for at least 24 h. Next, membranes were formed by casting on a glass plate, and the thickness was controlled using a casting knife, such that the resulting membrane thickness can be in the range of 50−70 μm. The casting was conducted in a glove bag in an environment filled with chloroform vapor to minimize rapid solvent evaporation, after which the membranes were left undisturbed for 4 to 5 h. The membranes were then annealed at 120 °C in a vacuum oven after ensuring that they had peeled off from the glass plate.

2.4. Characterization

2.4.1. Characterization of Nanocrystalline UiO-66 and Its Derivative

A volumetric gas sorption analyzer (iSorbHP1, Quantachrome, Boynton Beach, FL, USA) was used to determine CO2 and N2 adsorption on nanocrystalline UiO-66 and its derivatives. The samples were activated at 120 °C for 1 day to remove the residual solvents. Measurements of gas adsorption isotherms at 25 and 35 °C were conducted at pressures from 0 to 1 bar, during which a water circulator was used to ensure that the measurement environment remained in isothermal conditions. Considering the shape of the adsorption curve, the single-site Langmuir equation (Equation (1)) was deemed appropriate to fit the isotherm with a sufficiently high R2 value [46,47].
q = q s a t b p 1 + b p
In Equation (1), q, qsat, b, and p are the adsorption quantity (mmol/g), saturation loading (mmol/g), Langmuir constant (bar−1), and pressure (bar), respectively. The CO2/N2 selectivity of UiO-66 and its derivative can be calculated using the ideal adsorbed solution theory (IAST) [48], expressed by Equation (2):
Selectivity = x 1 / x 2 y 1 / y 2
where x1, x2—the mole fraction of the adsorbed phase and y1, y2—the mole fraction of the gas phase. The isosteric heat of adsorption, −Qst for CO2 and N2, was evaluated using the Clausius–Clapeyron equation (Equation 3), with p, T, and q denoting pressure (bar), absolute temperature (K), and amount adsorbed (mmol/g), respectively. An explicit analytical solution for the calculation of −Qst that uses a single-site Langmuir equation has been derived, and it has been observed that −Qst is a weak function of temperature.
Q s t = R T 2 ( ln P T ) q
N2 physisorption isotherm: The porosities of UiO-66 and its derivatives were measured via N2 physisorption analysis (77 K), under the conditions specified above, using a volumetric gas sorption analyzer (NOVATouch LX2, Quantachrome, Boynton Beach, FL, USA).
Powdered X-ray diffraction (PXRD; Advanced D8, Bruker, Billerica, MA, USA): PXRD was used to verify the crystallinity of the powdered samples. The samples were measured at ambient conditions, with 2θ scanned from 5° to 40° (step size of 0.02°) using a Cu-Kα radiation (1.5148 Å) diffractor.
Field-emission scanning electron microscopy (FESEM; JSM6701, JEOL, Akishima, Tokyo, Japan): FESEM was performed to examine the structural morphology of nanocrystalline UiO-66 and its derivatives. The accelerating voltage was set at 5 kV.
Fourier-transformed infrared (FTIR) spectroscopy: FTIR was conducted in the range of 4000−450 cm−1 (resolution of 4 cm−1) to identify the functional groups of the porous materials (IRPrestige-21, Shimadzu, Kyoto, Japan).
Elemental analysis (Vario EL III CHNS Elemental Analyzer, Elementar, Langenselbold, Germany): Elemental analysis was used to determine the elemental composition of nanocrystalline UiO-66 and its derivatives. The Br content in UiO-66-Br was estimated via FESEM equipped with energy-dispersive X-ray (EDX) spectroscopy.
Thermal stability: The thermal stabilities of the nanocrystalline UiO-66 and its derivatives were measured via thermogravimetric analysis (SDT Q600 TGA, TA Instrument, New Castle, DE, USA), under a temperature scan from 40 to 800 °C at the ramping rate of 10 °C/min. Prior to the analysis, the samples are purged at 120 °C under pure nitrogen flow (flow rate set at 100 mL/min) for 8 h to minimize the effect of the residual solvents on the TGA curve.

2.4.2. Characterization of Mixed-Matrix Membranes

FESEM was used to investigate the cross-sectional morphologies of membranes under uniform accelerating voltage conditions. The membranes were fractured before Platinum coating (with the use of liquid nitrogen) to preserve the overall morphologies. FTIR and XRD analyses were conducted with the same settings as above to investigate the properties of the membranes. Similarly, thermogravimetric analysis was used to investigate the thermal stability of the membranes with the same settings as mentioned above. An analytical balance (ME204, Mettler Toledo, Columbus, OH, USA) equipped with a density kit was used to determine the membranes’ densities. This measurement was conducted by computing the difference in the mass of the samples in an auxiliary liquid (ethanol) and air via Archimedes’ principle. The density of the membrane, ρ, can be computed from Equation (4):
ρ = A A B ( ρ L ρ A ) + ρ A
In this equation, A—mass of the membrane sample in air; B—mass of the membrane sample in the auxiliary liquid; ρ L —density of the auxiliary liquid; ρ A —density of air. The calculated value can be accurately determined up to 4 decimal places, based on the precision of the analytical balance.

2.4.3. Gas Permeation Test

A gas permeation test (GTR-11, GTR Tec Corporation, Kyoto, Japan) was carried out using a constant pressure-variable volume system. The gases (CO2/N2 test gas: 80 vol% N2 (99.9995%) and 20 vol% CO2 (99.8%) and helium (99.9995%)) were purchased from Airliquide Singapore Pte. Ltd. The membranes were mounted onto the gas permeation cell with the aid of vacuum grease. To minimize the potential contamination of the vacuum grease onto the desired permeation area (1.77 cm2) in the permeation cell, the membrane that is mounted onto the permeation cell is prepared in such a way that the membrane area is larger than the permeation area. Throughout the measurement, a uniform temperature of 35 °C was maintained with a temperature controller. Using a mass flow controller, the test gas and helium were continuously supplied upstream and downstream of the membrane, respectively. At set time intervals, the downstream permeated gas was swept by helium and sent to the gas chromatograph to calculate the composition of the gas stream. This process continued until it was ensured that the permeated gas’ concentration did not fluctuate substantially over a designated time period. The gas permeability and selectivity were calculated based on the concentration of the permeated gas. The equation of permeability, P, can be written as shown below (5). The parameters q, l, a, p, and t are the concentration (of CO2 and N2 gas calculated from gas chromatography), the membrane thickness, the permeation area, the pressure, and the time, respectively. To ensure sufficient reproducibility of the result, at least three different samples were measured.
P = q l a p t

2.4.4. Gas Adsorption Analysis

The adsorption properties of CO2 and N2 gases in each membrane were measured under identical conditions using a volumetric gas sorption analyzer (iSorbHP1, Quantachrome, Boynton Beach, FL, USA). All membranes were outgassed at 120 °C for 1 day. The CO2 and N2 adsorption isotherms were fitted using the single-site Langmuir equation as elaborated in Equation (1). This was followed by the calculation of the solubility, S, of CO2 and N2 in the respective membranes. In Equation (6), q, ρ, and p are defined as the gas adsorbed per unit membrane mass, the membrane’s density, and the specified pressure, respectively. The diffusivity, D, of gas in the membrane can be calculated using the relationship between permeability and solubility because the solution–diffusion mechanism is the transport mechanism in a dense membrane.
S = q ρ p

2.4.5. Filler Enhancement Index

Equation (7), which evaluates the parameter of filler enhancement index (Findex), was used to calculate the performance of the MMMs. In this expression, Pfilled and Punfilled are defined as the permeability of MMM and pure polymeric membrane; αfilled and αunfilled are the CO2/N2 selectivity of MMM and pure polymeric membrane; η is the enhancement coefficient, which is defined as 2.888 based on the slope of the Robeson upper bound published in 2008. This parameter was defined for CO2/N2 in a 2008 study to assist in quantifying the effectiveness of fillers in MMMs [4].
F i n d e x = ln ( P f i l l e d P u n f i l l e d ) + η ln ( α f i l l e d α u n f i l l e d )

3. Results and Discussion

3.1. Synthesis of Nanocrystalline UiO-66 and Its Derivatives

First, the PXRD patterns were inspected to verify the crystallinity of the UiO-66 nanocrystals (Figure 3a). A comparison of UiO-66 and its derivatives shows that the overall structural crystallinity remained intact, despite the use of different ligands. The peaks observed here are consistent with the diffraction peaks previously reported for UiO-66 [49,50,51,52,53]. Next, the N2 adsorption–desorption isotherm measured at 77 K (Figure 3b) clearly depicted that the UiO-66 nanocrystals possess high N2 sorption at low P/Po. This implies that these nanocrystals possess large micropore volumes, as summarized in Table 1. However, it can be expected that after the incorporation of ligands with different functional groups (via pre-synthetic functionalization), the overall porosity of the framework would decrease substantially. Indeed, the N2 adsorption–desorption measurements confirm the decreases in overall porosity with the various functional groups. As expected, the introduction of difunctional groups causes a sharper decrease in the accessible surface area than modification with monofunctional groups. Comparing amine (-NH2) and bromine (-Br) functional groups, UiO-66-Br suffers a slight decrease in accessible surface area compared with UiO-66-NH2, consistent with the larger atomic size of -Br. Further verification of the presence of -NH2 and -Br functional groups was conducted via elemental analysis and EDX (Figure S1), and the results are summarized in Table S1 (the theoretical values of the elemental composition are supplemented in Table S2). FTIR analysis of UiO-66 and its derivatives demonstrated that the presence of different functional groups did not influence the FTIR spectrum (Figure 3c). The spectrum of UiO-66 is similar to that reported in the literature [54]. A clearly observable O-H stretch as well as N-H stretch at around 3000 and 3300 cm−1, respectively, for UiO-66-(OH)2 and UiO-66-NH2 indicated the unmistakable presence of 2,5-dihydroxyterephthalate and 2-aminoterephthalate in each sample. As for the thermal stability of UiO-66 and its derivatives, thermogravimetric analysis (Figure 3d) demonstrated that the overall structure of UiO-66 and UiO-66-Br remained thermally stable up to 550 and 500 °C, respectively, whereas UiO-66-NH2 and UiO-66-(OH)2 showed a substantial weight loss after 200 °C. Thus, for the subsequent experiments (gas adsorption and membrane annealing), the activation condition of all nanoporous materials were set at 120 °C. In addition, it should be noted that the nanoporous materials that are used in MMM for gas separation should have sufficiently small particle sizes [55]. Thus, the morphologies of UiO-66 and its derivatives were verified via FESEM (Figure 4). Based on the images, the particle sizes generally ranged from 300 to 500 nm, which is sufficiently small for these particles to be used in gas separation.

3.2. CO2 and N2 Adsorption by UiO-66 and Its Derivatives

The properties of UiO-66 and its derivatives were further characterized by the measurement of CO2 and N2 adsorption at 35 °C, with the results summarized in Figure 5. The CO2 and N2 adsorption isotherms at 25 °C were also acquired and are summarized in Figure S2a,b. The fitting parameters for CO2 and N2 adsorption are summarized in Table S3 (25 °C) and Table S4 (35 °C) respectively. In general, UiO-66 and all of its derivatives proved to be able to preferentially adsorb CO2 relative to N2, consistent with the fact that the former gas possesses a higher quadrupole moment (4.3 × 10−26 esu cm2 vs. 1.5 × 10−26 esu cm2) and polarizability (29.11 × 10−25 cm3 vs. 17.4 × 10−25 cm3) [56]. The presence of zirconium at the metal sites and the functional groups in the ligands (-NH2, -Br, -(OH)2) both promoted favorable interaction with CO2. Notably, UiO-66-NH2 and UiO-66-Br showed a clear enhancement of CO2 adsorption capability relative to UiO-66, even though they possessed a smaller surface area. The adsorption capability was poorly correlated with the accessible surface areas, possibly because none of the samples had reached CO2 adsorption saturation at 1 bar. In contrast, although UiO-66-(OH)2 showed reasonably high CO2 adsorption at low partial pressure, due to its low surface area (318 m2/g) relative to the other samples, the slope of the isotherm indicated a faster equilibrium saturation of CO2 adsorption than for the other adsorbents [51,57]. Nevertheless, in terms of the applicability for CO2 adsorption in the field of post-combustion CO2 capture, which is conducted at low partial pressure of CO2 [58], UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2 each demonstrated better CO2 adsorption performance than UiO-66. Each of these observations are supported by the higher isosteric heat of adsorption of CO2 on the functionalized adsorbents (Figure S2c) together with higher IAST CO2/N2 selectivity (Figure S2d), which indicates that the use of ligands with any of the various functional groups in this work is a feasible strategy to improve the CO2 adsorption performance compared with UiO-66.

3.3. Fabrication of Mixed-matrix Membranes

In this study, in-house-made polyimide (ODPA-TMPDA) was used as the polymeric membrane for the gas separation process. Successful synthesis of ODPA-TMPDA was verified from its FTIR spectrum (Figure 6), which exhibits the characteristic asymmetric and symmetric stretching (1770 and 1710 cm−1) of C=O as well as the stretching of C-N (1300 cm−1). The corresponding colors of the functional groups are also indicated in the figure for guidance. The most important evidence of successful formation of ODPA-TMPDA is the band at 3500 cm−1 (i.e., O-H absorption band), corresponding to residual unreacted polyamic acid. As described in the section on synthesis procedure [45], the ODPA-TMPDA polymer was formed in two steps: (1) formation of polyamic acid via condensation reaction and (2) chemical imidization of polyamic acid. This band was not detected in this work, implying the imidization was completely done. The overall features of the spectrum are comparable to those of previously reported spectra [39,40,59].
Hence, MMMs containing UiO-66, UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2 with 10 and 20 wt% loading were developed in this work. The FTIR spectra of the MMMs indicate that the structural properties of the polyimide membrane remained intact (Figure S3). In addition, the XRD patterns of the MMMs indicate that the crystallinity of UiO-66 and its derivatives remained unaffected after annealing (Figure S4). Hence, the cross-sectional morphologies were further investigated by FESEM (Figure 7). The sieve-in-a-cage morphology, which is common in zeolite/polymer MMMs, was not observed in this work [60]. The presence of organic ligands in UiO-66, UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2 improves the polymer/filler compatibility. The use of small particles in this study was also advantageous because it created large interfacial areas between the filler and polymer [41,61]. Thermogravimetric analysis of the pure polymer and MMMs verified that the incorporation of nanoporous materials did not affect the thermal stability of the polymer (Figure S5). The initial drop in weight loss is attributed to the removal of residual solvent present in the sample. For the MMMs, because the thermal stability of the filler is weaker than that of the polymer matrices, a two-stage weight loss was observed. This observation is consistent with a previous study [22].

3.4. Gas Permeation Properties

The gas permeation properties of the membranes were measured at the upstream pressure of 1 bar of CO2 (20/80 mixture) at 35 °C (Table 2). The presence of 20 wt% UiO-66 in the polymeric membrane was found to improve the overall CO2 permeability by 92%, but with a marginal decrease in CO2/N2 selectivity. The large micropore volume of UiO-66 presumably allowed rapid transport of both CO2 and N2 molecules through the MMM with minimal resistance. However, due to the relatively poor CO2/N2 selectivity of UiO-66, it was anticipated that the incorporation of the above-tested derivatives of UiO-66 in MMMs would give better gas separation performance. Indeed, the gas permeation results confirm that UiO-66-NH2, UiO-66-Br, and UiO-66-(OH)2 improved the CO2/N2 selectivity by 12%, 4%, and 17%, respectively, which is consistent with the IAST calculation (Figure S1d), which predicted that UiO-66-(OH)2 would demonstrate the highest improvement in selectivity among the UiO-66 derivatives. Nevertheless, considering the overall performance in terms of both CO2 permeability and CO2/N2 selectivity, 20 wt% UiO-66-Br demonstrates the highest CO2 permeability along with attractive CO2/N2 selectivity.
The solubility–diffusivity of CO2 and N2 in MMMs was then quantitatively analyzed. The solubility of gas in the membranes was determined by measuring the pure-component CO2 and N2 adsorption isotherms at 35 °C, with the data summarized in Figure 8. The fitting parameters for CO2 and N2 adsorption by the membranes are compiled in Table S5. Table 3 summarizes the quantitative solubility–diffusivity data for CO2 and N2 in the membranes. Based on the gas adsorption data (which corresponds to the analysis of solubility), at the point of interest (CO2 at 0.2 bar), the adsorption performance is slightly inferior to the pristine membrane without nanocrystals. Nevertheless, at 0.8 bar of N2 (point of interest in this study), the incorporation of UiO-66 and its derivatives substantially suppressed the N2 solubility, which is consistent with the behavior of other porous materials in membranes, as reported in previous studies [22,40,59]. In contrast, it was found that the diffusivities of both CO2 and N2 dramatically increased upon incorporation of nanocrystalline UiO-66 and its derivatives, which possess large pore volumes, resulting in improved gas permeabilities of MMMs.
Besides, the CO2/N2 separation properties of our membranes are also compared with the literature data, where UiO-66 and its derivative are used as the fillers in the fabrication of MMM. In general, as summarized in the gas permeation data (Table 2) and Table 4, UiO-66 is incapable of improving the CO2/N2 selectivity to a substantial extent, unless rubbery polymer (PEBA) is used, as reported by Shen et al. [51]. This is plausibly attributed to the high chain mobility of the rubbery polymer, which minimize the formation of interfacial nanogaps more effectively than the glassy polymers [61]. Nevertheless, although the performance of UiO-66 in PEBA looks promising, it is noteworthy that membrane should be fabricated on a porous support owing to its poor mechanical stability [62,63,64], thus, limiting potential utility in large scale membrane production. In order to ensure a clear enhancement in CO2/N2 selectivity of glassy polymer membrane, functionalization of UiO-66 to allow favorable CO2 adsorption is deemed necessary.
With reference to the CO2/N2 separation performance reported in Table 4, an empirical metric (Findex) was used to compute the effectiveness of the fillers in MMM, considering the fact that different polymers were used as the matrices in membranes. In this work, the most promising performance was that of UiO-66-Br at 20 wt% loading, with Findex = 0.94, a performance close to the Robeson upper bound for CO2/N2 separation [12], as plotted in Figure S6. In contrast, the value of Findex for UiO-66 was only 0.09 at 10 wt% and 0.55 at 20 wt% loading, lower than for any other fillers investigated in this work. Notably, the Findex of 0.94 for the case of UiO-66-Br is found to be higher than most of the reported literature data, with the exception of the incorporation of azo-UiO-66 at 10 wt% loading in Matrimid membrane [65] and UiO-66-H (20 wt%) and UiO-66-NH2 (10 wt%) in PIM-1 membrane [66]. However, it should be noted that the data in the above studies were obtained from the pure gas permeation test, which does not reflect the practical conditions. For example, the partial pressure of CO2 in the feed is considerably low (<20 vol%) in a typical post-combustion CO2 capture process [58,67], as compared to volume fraction of 50 vol% (of CO2) used in several studies [51,66]. Thus, evidently, the measurement of UiO-66-NH2 (10 wt%) in PIM-1 membrane under the mixed-gas condition has led to a clear 54% decrease in Findex from 1.48 to 0.67, as shown in Table 4. Hence, it is generally important to conduct the measurement under the mixed-gas condition rather than pure gas testing in order to demonstrate a clear illustration of the gas separation performance of porous materials in MMM.

4. Conclusions

Nanocrystalline UiO66 and its derivatives were developed via pre-synthetic functionalization with ZrCl4, and these porous fillers were used to develop MMMs for the analysis of CO2/N2 separation. It was observed that the addition of UiO-66 nanocrystals successfully improved CO2 permeability but with a slight dip in CO2/N2 selectivity. Thus, derivatives of UiO-66 with different functional groups (-NH2, -Br, and -(OH)2) were prepared and evaluated under similar conditions. All of these UiO-66 derivatives were found to improve the CO2/N2 selectivity, which was consistent with the calculated IAST of these porous materials together with the experimental results of a gas permeation test, where the particles were incorporated into an MMM. Further solubility–diffusivity analysis indicated that the addition of UiO-66 derivatives with different functional groups suppressed the solubility of N2, which led, in turn, to enhanced CO2/N2 selectivity. Based on our findings, UiO-66-Br at 20 wt% loading is capable of improving both the CO2 permeability and CO2/N2 selectivity of MMMs for gas separation, with performance close to the reported upper bound. Future efforts can be devoted to realizing a practically attractive performance by employing high-performance polymer matrices (e.g., 6FDA-based polyimides and PIM-1) in MMM fabrication.

Supplementary Materials

The following are available online at https://www.mdpi.com/2077-0375/10/7/154/s1, Figure S1: EDX analysis of UiO-66-Br, Figure S2: (a) CO2 adsorption at 25 °C; (b) N2 adsorption at 25 °C; (c) Isosteric heat of adsorption of CO2 and (d) CO2/N2 IAST selectivity (feed mixture of CO2/N2 = 20/80) for UiO-66, UiO-66-NH2, UiO-66-Br and UiO-66-(OH)2, Figure S3: FTIR spectrum of mixed-matrix membrane, Figure S4: XRD analysis of mixed-matrix membrane, Figure S5: TGA analysis of mixed-matrix membrane, Figure S6: Comparison of the gas permeation data with the upper bound limit for CO2/N2 constructed in 2008. The numerical number indicated in the figure illustrate the value of Findex as described in Section 2.4.5. The data used in this plot is provided in Table 4, Table S1: Elemental analysis of UiO-66 and its derivative, Table S2: Theoretical amount of UiO-66 and its derivative, Table S3: Fitting parameters for CO2 and N2 for UiO-66 and its derivative at 25 °C, Table S4: Fitting parameters for CO2 and N2 for UiO-66 and its derivative at 35 °C, Table S5: Fitting parameters for CO2 and N2 for polymeric and mixed-matrix membranes at 35 °C.

Author Contributions

Original draft preparation, experimental and characterization: C.Y.C., assistance in the membrane characterization: J.L.; Writing—review and editing: T.-H.B. and J.S.; funding acquisition, T.-H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work is funded by Saudi Aramco-KAIST CO2 Management Center.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chuah, C.Y. Microporous Materials with Tailored Structural Properties for Enhanced Gas Separation; Nanyang Technological University: Singapore, 2019. [Google Scholar]
  2. Chuah, C.Y.; Kim, K.; Lee, J.; Koh, D.Y.; Bae, T.H. CO2 absorption using membrane contactors: Recent progress and future perspective. Ind. Eng. Chem. Res. 2020, 59, 6773–6794. [Google Scholar] [CrossRef]
  3. Aaron, D.; Tsouris, C. Separation of CO2 from flue gas: A review. Sep. Sci. Technol. 2005, 40, 321–348. [Google Scholar] [CrossRef]
  4. 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] [PubMed]
  5. Mondal, M.K.; Balsora, H.K.; Varshney, P. Progress and trends in CO2 capture/separation technologies: A review. Energy 2012, 46, 431–441. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Sunarso, J.; Liu, S.; Wang, R. Current status and development of membranes for CO2/CH4 separation: A review. Int. J. Greenh. Gas Con. 2013, 12, 84–107. [Google Scholar] [CrossRef]
  7. Yu, C.H.; Huang, C.H.; Tan, C.S. A review of CO2 capture by absorption and adsorption. Aerosol Air Qual. Res 2012, 12, 745–769. [Google Scholar] [CrossRef] [Green Version]
  8. Yang, H.; Xu, Z.; Fan, M.; Gupta, R.; Slimane, R.B.; Bland, A.E.; Wright, I. Progress in carbon dioxide separation and capture: A review. J. Environ. Sci. 2008, 20, 14–27. [Google Scholar] [CrossRef]
  9. Esposito, E.; Dellamuzia, L.; Moretti, U.; Fuoco, A.; Giorno, L.; Jansen, J.C. Simultaneous production of biomethane and food grade CO2 from biogas: An industrial case study. Energy Environ. Sci. 2019, 12, 281–289. [Google Scholar] [CrossRef]
  10. Sanders, D.F.; Smith, Z.P.; Guo, R.; Robeson, L.M.; McGrath, J.E.; Paul, D.R.; Freeman, B.D. Energy-efficient polymeric gas separation membranes for a sustainable future: A review. Polymer 2013, 54, 4729–4761. [Google Scholar] [CrossRef] [Green Version]
  11. 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, 1906697. [Google Scholar] [CrossRef]
  12. Robeson, L.M. The upper bound revisited. J. Membr. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
  13. Robeson, L.M. Correlation of separation factor versus permeability for polymeric membranes. J. Membr. Sci. 1991, 62, 165–185. [Google Scholar] [CrossRef]
  14. 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]
  15. Li, S.; Falconer, J.L.; Noble, R.D. SAPO-34 membranes for CO2/CH4 separation. J. Membr. Sci. 2004, 241, 121–135. [Google Scholar] [CrossRef]
  16. Pera-Titus, M. Porous inorganic membranes for CO2 capture: Present and prospects. Chem. Rev. 2013, 114, 1413–1492. [Google Scholar] [CrossRef]
  17. Zhang, X.; Chuah, C.Y.; Dong, P.; Cha, Y.H.; Bae, T.H.; Song, M.K. Hierarchically Porous Co-MOF-74 Hollow Nanorods for Enhanced Dynamic CO2 Separation. ACS Appl. Mater. Int. 2018, 10, 43316–43322. [Google Scholar] [CrossRef] [PubMed]
  18. Vinoba, M.; Bhagiyalakshmi, M.; Alqaheem, Y.; Alomair, A.A.; Pérez, A.; Rana, M.S. Recent progress of fillers in mixed matrix membranes for CO2 separation: A review. Sep. Purif. Technol. 2017, 188, 431–450. [Google Scholar] [CrossRef]
  19. Adatoz, E.; Avci, A.K.; Keskin, S. Opportunities and challenges of MOF-based membranes in gas separations. Sep. Purif. Technol. 2015, 152, 207–237. [Google Scholar] [CrossRef]
  20. Lin, Y.S. Metal organic framework membranes for separation applications. Curr. Opin. Chem. Eng. 2015, 8, 21–28. [Google Scholar] [CrossRef]
  21. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.H.; Long, J.R. Carbon dioxide capture in metal–organic frameworks. Chem. Rev. 2012, 112, 724–781. [Google Scholar] [CrossRef]
  22. 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]
  23. McDonald, T.M.; Lee, W.R.; Mason, J.A.; Wiers, B.M.; Hong, C.S.; Long, J.R. Capture of carbon dioxide from air and flue gas in the alkylamine-appended metal–organic framework mmen-Mg2(dobpdc). J. Am. Chem. Soc. 2012, 134, 7056–7065. [Google Scholar] [CrossRef] [PubMed]
  24. Darunte, L.A.; Oetomo, A.D.; Walton, K.S.; Sholl, D.S.; Jones, C.W. Direct air capture of CO2 using amine functionalized MIL-101 (Cr). ACS Sust. Chem. Eng. 2016, 4, 5761–5768. [Google Scholar] [CrossRef]
  25. Gong, H.; Lee, S.S.; Bae, T.H. Mixed-matrix membranes containing inorganically surface-modified 5A zeolite for enhanced CO2/CH4 separation. Micropor. Mesopor. Mater. 2017, 237, 82–89. [Google Scholar] [CrossRef]
  26. Bae, T.H.; Liu, J.; Thompson, J.A.; Koros, W.J.; Jones, C.W.; Nair, S. Solvothermal deposition and characterization of magnesium hydroxide nanostructures on zeolite crystals. Micropor. Mesopor. Mater. 2011, 139, 120–129. [Google Scholar] [CrossRef]
  27. Li, Y.; Chung, T.S.; Cao, C.; Kulprathipanja, S. The effects of polymer chain rigidification, zeolite pore size and pore blockage on polyethersulfone (PES)-zeolite A mixed matrix membranes. J. Membr. Sci. 2005, 260, 45–55. [Google Scholar] [CrossRef]
  28. Li, Y.; Guan, H.M.; Chung, T.S.; Kulprathipanja, S. Effects of novel silane modification of zeolite surface on polymer chain rigidification and partial pore blockage in polyethersulfone (PES)–zeolite A mixed matrix membranes. J. Membr. Sci. 2006, 275, 17–28. [Google Scholar] [CrossRef]
  29. Bae, T.H.; Liu, J.; Lee, J.S.; Koros, W.J.; Jones, C.W.; Nair, S. Facile high-yield solvothermal deposition of inorganic nanostructures on zeolite crystals for mixed matrix membrane fabrication. J. Am. Chem. Soc. 2009, 131, 14662–14663. [Google Scholar] [CrossRef]
  30. Liu, J.; Benin, A.I.; Furtado, A.M.; Jakubczak, P.; Willis, R.R.; LeVan, M.D. Stability effects on CO2 adsorption for the DOBDC series of metal–organic frameworks. Langmuir 2011, 27, 11451–11456. [Google Scholar] [CrossRef]
  31. 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]
  32. Garibay, S.J.; Cohen, S.M. Isoreticular synthesis and modification of frameworks with the UiO-66 topology. Chem. Commun. 2010, 46, 7700–7702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Katz, M.J.; Brown, Z.J.; Colón, Y.J.; Siu, P.W.; Scheidt, K.A.; Snurr, R.Q.; Hupp, J.T.; Farha, O.K. A facile synthesis of UiO-66, UiO-67 and their derivatives. Chem. Commun. 2013, 49, 9449–9451. [Google Scholar] [CrossRef] [PubMed]
  34. Samarasinghe, S.; Chuah, C.Y.; Karahan, H.E.; Sethunga, G.; Bae, T.H. Enhanced O2/N2 Separation of Mixed-Matrix Membrane Filled with Pluronic-Compatibilized Cobalt Phthalocyanine Particles. Membranes 2020, 10, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Dorosti, F.; Omidkhah, M.; Abedini, R. Fabrication and characterization of Matrimid/MIL-53 mixed matrix membrane for CO2/CH4 separation. Chem. Eng. Res. Des. 2014, 92, 2439–2448. [Google Scholar] [CrossRef]
  36. Castro-Munoz, R.; Fila, V.; Martin-Gil, V.; Muller, C. Enhanced CO2 permeability in Matrimid® 5218 mixed matrix membranes for separating binary CO2/CH4 mixtures. Sep. Purif. Technol. 2019, 210, 553–562. [Google Scholar] [CrossRef]
  37. Yong, W.F.; Li, F.Y.; Xiao, Y.C.; Chung, T.S.; Tong, Y.W. High performance PIM-1/Matrimid hollow fiber membranes for CO2/CH4, O2/N2 and CO2/N2 separation. J. Membr. Sci. 2013, 443, 156–169. [Google Scholar] [CrossRef]
  38. Dai, Y.; Johnson, J.; Karvan, O.; Sholl, D.S.; Koros, W. Ultem®/ZIF-8 mixed matrix hollow fiber membranes for CO2/N2 separations. J. Membr. Sci. 2012, 401, 76–82. [Google Scholar] [CrossRef]
  39. Samarasinghe, S.; Chuah, C.Y.; Li, W.; Sethunga, G.; Wang, R.; Bae, T.H. Incorporation of CoIII acetylacetonate and SNW-1 nanoparticles to tailor O2/N2 separation performance of mixed-matrix membrane. Sep. Purif. Technol. 2019, 223, 133–141. [Google Scholar] [CrossRef]
  40. 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]
  41. 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]
  42. Yang, Y.; Goh, K.; Weerachanchai, P.; Bae, T.H. 3D covalent organic framework for morphologically induced high-performance membranes with strong resistance toward physical aging. J. Membr. Sci. 2019, 574, 235–242. [Google Scholar] [CrossRef]
  43. Chuah, C.Y.; Samarasinghe, S.; 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]
  44. Askari, M.; Chung, T.S. Natural gas purification and olefin/paraffin separation using thermal cross-linkable co-polyimide/ZIF-8 mixed matrix membranes. J. Membr. Sci. 2013, 444, 173–183. [Google Scholar] [CrossRef]
  45. Duan, C.; Jie, X.; Liu, D.; Cao, Y.; Yuan, Q. Post-treatment effect on gas separation property of mixed matrix membranes containing metal organic frameworks. J. Membr. Sci. 2014, 466, 92–102. [Google Scholar] [CrossRef]
  46. Chuah, C.Y.; Yang, Y.; Bae, T.H. Hierarchically porous polymers containing triphenylamine for enhanced SF6 separation. Micropor. Mesopor. Mater. 2018, 272, 232–240. [Google Scholar] [CrossRef]
  47. Chuah, C.Y.; Yu, S.; Na, K.; Bae, T.H. Enhanced SF6 recovery by hierarchically structured MFI zeolite. J. Ind. Eng. Chem. 2018, 62, 64–71. [Google Scholar] [CrossRef]
  48. Myers, A.; Prausnitz, J.M. Thermodynamics of mixed-gas adsorption. AIChE J. 1965, 11, 121–127. [Google Scholar] [CrossRef]
  49. Li, J.; Musho, T.; Bright, J.; Wu, N. Functionalization of a Metal–organic Framework Semiconductor for Tuned Band Structure and Catalytic Activity. J. Electrochem. Soc. 2019, 166, H3029–H3034. [Google Scholar] [CrossRef]
  50. Khdhayyer, M.R.; Esposito, E.; Fuoco, A.; Monteleone, M.; Giorno, L.; Jansen, J.C.; Attfield, M.P.; Budd, P.M. Mixed matrix membranes based on UiO-66 MOFs in the polymer of intrinsic microporosity PIM-1. Sep. Purif. Technol. 2017, 173, 304–313. [Google Scholar] [CrossRef]
  51. Shen, J.; Liu, G.; Huang, K.; Li, Q.; Guan, K.; Li, Y.; Jin, W. UiO-66-polyether block amide mixed matrix membranes for CO2 separation. J. Membr. Sci. 2016, 513, 155–165. [Google Scholar] [CrossRef]
  52. Rodrigues, M.A.; de Souza Ribeiro, J.; de Souza Costa, E.; de Miranda, J.L.; Ferraz, H.C. Nanostructured membranes containing UiO-66 (Zr) and MIL-101 (Cr) for O2/N2 and CO2/N2 separation. Sep. Purif. Technol. 2018, 192, 491–500. [Google Scholar] [CrossRef]
  53. Jadhav, N.; Kumar, S.N.; Tanvidkar, P.S.; Kuncharam, B.V.R. Synthesis and characterization of mixed-matrix material of Zirconium based metal organic framework (MOF: UiO-66-NH2) and poly (ether-urethane-urea). Mater. Today Proc. 2020, 28, 734–738. [Google Scholar] [CrossRef]
  54. Han, Y.; Liu, M.; Li, K.; Zuo, Y.; Wei, Y.; Xu, S.; Zhang, G.; Song, C.; Zhang, Z.; Guo, X. Facile synthesis of morphology and size-controlled zirconium metal–organic framework UiO-66: The role of hydrofluoric acid in crystallization. CrystEngComm 2015, 17, 6434–6440. [Google Scholar] [CrossRef]
  55. Bae, T.H.; Lee, J.S.; Qiu, W.; Koros, W.J.; Jones, C.W.; Nair, S. A high-performance gas-separation membrane containing submicrometer-sized metal–organic framework crystals. Angew. Chem. Int. Ed. 2010, 49, 9863–9866. [Google Scholar] [CrossRef]
  56. 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] [PubMed]
  57. Kim, M.B.; Kim, K.M.; Kim, T.H.; Yoon, T.U.; Kim, E.J.; Kim, J.H.; Bae, Y.S. Highly selective adsorption of SF6 over N2 in a bromine-functionalized zirconium-based metal–organic framework. Chem. Eng. J. 2018, 339, 223–229. [Google Scholar] [CrossRef]
  58. 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]
  59. Li, W.; Samarasinghe, S.; Bae, T.H. Enhancing CO2/CH4 separation performance and mechanical strength of mixed-matrix membrane via combined use of graphene oxide and ZIF-8. J. Ind. Eng. Chem. 2018, 67, 156–163. [Google Scholar] [CrossRef]
  60. Goh, P.; Ismail, A.; Sanip, S.; Ng, B.; Aziz, M. Recent advances of inorganic fillers in mixed matrix membrane for gas separation. Sep. Purif. Technol. 2011, 81, 243–264. [Google Scholar] [CrossRef]
  61. Gong, H.; Nguyen, T.H.; Wang, R.; Bae, T.H. Separations of binary mixtures of CO2/CH4 and CO2/N2 with mixed-matrix membranes containing Zn(pyrz)2(SiF6) metal–organic framework. J. Membr. Sci. 2015, 495, 169–175. [Google Scholar] [CrossRef]
  62. Kim, K.; Ingole, P.G.; Kim, J.; Lee, H. Separation performance of PEBAX/PEI hollow fiber composite membrane for SO2/CO2/N2 mixed gas. Chem. Eng. J. 2013, 233, 242–250. [Google Scholar] [CrossRef]
  63. Esposito, E.; Clarizia, G.; Bernardo, P.; Jansen, J.C.; Sedláková, Z.; Izák, P.; Curcio, S.; de Cindio, B.; Tasselli, F. Pebax®/PAN hollow fiber membranes for CO2/CH4 separation. Chem. Eng. Process 2015, 94, 53–61. [Google Scholar] [CrossRef]
  64. 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]
  65. Prasetya, N.; Donose, B.C.; Ladewig, B.P. A new and highly robust light-responsive Azo-UiO-66 for highly selective and low energy post-combustion CO2 capture and its application in a mixed matrix membrane for CO2/N2 separation. J. Mater. Chem. A 2018, 6, 16390–16402. [Google Scholar] [CrossRef]
  66. Ghalei, B.; Sakurai, K.; Kinoshita, Y.; Wakimoto, K.; Isfahani, A.P.; Song, Q.; Doitomi, K.; Furukawa, S.; Hirao, H.; Kusuda, H. Enhanced selectivity in mixed matrix membranes for CO2 capture through efficient dispersion of amine-functionalized MOF nanoparticles. Nat. Energy 2017, 2, 17086. [Google Scholar] [CrossRef]
  67. Mason, J.A.; McDonald, T.M.; Bae, T.H.; Bachman, J.E.; Sumida, K.; Dutton, J.J.; Kaye, S.S.; Long, J.R. Application of a high-throughput analyzer in evaluating solid adsorbents for post-combustion carbon capture via multicomponent adsorption of CO2, N2, and H2O. J. Am. Chem. Soc. 2015, 137, 4787–4803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Su, N.C.; Sun, D.T.; Beavers, C.M.; Britt, D.K.; Queen, W.L.; Urban, J.J. Enhanced permeation arising from dual transport pathways in hybrid polymer–MOF membranes. Energy Environ. Sci. 2016, 9, 922–931. [Google Scholar] [CrossRef]
Figure 1. Synthesis scheme of nanocrystalline UiO-66 and its derivatives.
Figure 1. Synthesis scheme of nanocrystalline UiO-66 and its derivatives.
Membranes 10 00154 g001
Figure 2. Synthesis of ODPA-TMPDA.
Figure 2. Synthesis of ODPA-TMPDA.
Membranes 10 00154 g002
Figure 3. (a) Powdered X-ray diffraction (PXRD) pattern; (b) N2 adsorption–desorption isotherms at 77 K (open and closed symbols describe the adsorption and desorption isotherms); (c) FTIR curves; (d) thermogravimetric curves of UiO-66 and its derivatives.
Figure 3. (a) Powdered X-ray diffraction (PXRD) pattern; (b) N2 adsorption–desorption isotherms at 77 K (open and closed symbols describe the adsorption and desorption isotherms); (c) FTIR curves; (d) thermogravimetric curves of UiO-66 and its derivatives.
Membranes 10 00154 g003
Figure 4. FESEM images of (a) UiO-66; (b) UiO-66-NH2; (c) UiO-66-Br; (d) UiO-66-(OH)2.
Figure 4. FESEM images of (a) UiO-66; (b) UiO-66-NH2; (c) UiO-66-Br; (d) UiO-66-(OH)2.
Membranes 10 00154 g004
Figure 5. (a) CO2 and (b) N2 adsorption by nanocrystalline UiO-66 nanocrystals and its derivatives at 35 °C.
Figure 5. (a) CO2 and (b) N2 adsorption by nanocrystalline UiO-66 nanocrystals and its derivatives at 35 °C.
Membranes 10 00154 g005
Figure 6. FTIR spectrum of ODPA-TMPDA polymer.
Figure 6. FTIR spectrum of ODPA-TMPDA polymer.
Membranes 10 00154 g006
Figure 7. FESEM images of (a) 10 wt% UiO-66; (b) 20 wt% UiO-66; (c) 10 wt% UiO-66-NH2; (d) 20 wt% UiO-66-NH2; (e) 10 wt% UiO-66-Br; (f) 20 wt% UiO-66-Br; (g) 10 wt% UiO-66-(OH)2; (h) 20 wt% UiO-66-(OH)2 in ODPA-TMPDA membrane.
Figure 7. FESEM images of (a) 10 wt% UiO-66; (b) 20 wt% UiO-66; (c) 10 wt% UiO-66-NH2; (d) 20 wt% UiO-66-NH2; (e) 10 wt% UiO-66-Br; (f) 20 wt% UiO-66-Br; (g) 10 wt% UiO-66-(OH)2; (h) 20 wt% UiO-66-(OH)2 in ODPA-TMPDA membrane.
Membranes 10 00154 g007aMembranes 10 00154 g007b
Figure 8. (a) CO2 and (b) N2 isotherm of ODPA-TMPDA membrane and mixed-matrix membrane containing UiO-66, UiO-66-NH2, UiO-66-Br and UiO-66-(OH)2 at 20 wt% loading.
Figure 8. (a) CO2 and (b) N2 isotherm of ODPA-TMPDA membrane and mixed-matrix membrane containing UiO-66, UiO-66-NH2, UiO-66-Br and UiO-66-(OH)2 at 20 wt% loading.
Membranes 10 00154 g008
Table 1. Surface areas and pore volumes of UiO-66 and its derivatives, determined by N2 physisorption at 77 K.
Table 1. Surface areas and pore volumes of UiO-66 and its derivatives, determined by N2 physisorption at 77 K.
SampleSBET (a) (m2/g)SLANG (a) (m2/g)Smicro (b) (m2/g)Vmicro (b) (cc/g)Vtotal (c) (cc/g)
UiO-661733226616620.8090.938
UiO-66-NH21218159911600.5690.682
UiO-66-Br85111178180.3800.440
UiO-66-(OH)23184182930.1360.178
(a) BET surface area and Langmuir surface area (SBET and SLANG) were determined at P/Po = 0.05–0.2; (b) Micropore surface area and micropore volume (Smicro and Vmicro) were determined at P/Po = 0.4–0.6 using t-plot method; (c) Total pore volume (Vtotal) was determined at P/Po = 0.99.
Table 2. Permeation results of the membranes at 35 °C at the feed pressure of 1 bar (or 105 Pa) CO2/N2 mixture (20 vol%/80 vol%) (a).
Table 2. Permeation results of the membranes at 35 °C at the feed pressure of 1 bar (or 105 Pa) CO2/N2 mixture (20 vol%/80 vol%) (a).
MembraneCO2 Permeability (Barrer) (b)CO2/N2 Selectivity
ODPA-TMPDA88 ± 233.1 ± 1.2
10 wt% UiO-66142 ± 529.0 ± 0.4
20 wt% UiO-66169 ± 231.9 ± 0.2
10 wt% UiO-66-NH2129 ± 336.1 ± 0.8
20 wt% UiO-66-NH2142 ± 137.1 ± 2.3
10 wt% UiO-66-Br158 ± 233.7 ± 1.0
20 wt% UiO-66-Br200 ± 434.5 ± 1.9
10 wt% UiO-66-(OH)298 ± 235.2 ± 0.8
20 wt% UiO-66-(OH)2125 ± 438.9 ± 0.9
(a) The overall thickness of the membrane is determined to be ranging from 50 to 70 μm, based on the measurement from a micrometer screw gauge. (b) 1 Barrer = 3.35 × 10 −16 mol-m/m2-s-Pa.
Table 3. Solubility–diffusivity data of CO2 and N2 on membranes measured at 35 °C and total feed pressure of 1 bar. The pressure point is determined to be 0.2 bar for CO2 and 0.8 bar for N2.
Table 3. Solubility–diffusivity data of CO2 and N2 on membranes measured at 35 °C and total feed pressure of 1 bar. The pressure point is determined to be 0.2 bar for CO2 and 0.8 bar for N2.
MembraneDensity (g/cm3)CO2 Solubility (mol/m3.bar)CO2 Diffusivity, × 10−12 (m2/s)N2 Solubility (mol/m3.bar)N2 Diffusivity, × 10−12 (m2/s)
ODPA-TMPDA1.28615261.9632.12.81
20 wt% UiO-661.21011644.9231.55.71
20 wt% UiO-66-NH21.33011724.1128.14.62
20 wt% UiO-66-Br1.35412565.4022.58.72
20 wt% UiO-66-(OH)21.38013853.0629.33.72
Table 4. Summary of gas permeation results of the mixed-matrix membranes that utilize UiO-66 and its derivatives (a).
Table 4. Summary of gas permeation results of the mixed-matrix membranes that utilize UiO-66 and its derivatives (a).
FillerPolymerFiller Loading (wt%) (b)Separation PerformanceFindexYr (Ref.)
Testing ConditionP(CO2) Barrer% En.α (CO2/N2)% En.
Pressure (bar)Temp. (°C)
UiO-66PEBA10- (c)2596.387.056.634.41.4816′ [51]
UiO-66PEBA10- (c, d)25139.717161.145.12.0716′ [51]
UiO-66PSF203351618626.2−11.00.7116′ [68]
UiO-66-NH2PEBA20- (c)2587.068.966.157.01.8216′ [51]
UiO-66-NH2PEBA20- (c, d)25130.215372.271.52.4816′ [51]
UiO-66-refPIM-1204256981128.613.0−19.30.2117′ [66]
UiO-66PIM-123.1125761059.520.7−5.10.3117′ [50]
UiO-66PIM-1 (MeOH treated)23.11259980109.221.6−0.90.7117′ [50]
UiO-66-COOHPIM-123.1125530011.120−8.26−0.1417′ [50]
UiO-66-HPIM-1204252606−14.724.652.81.0717′ [66]
UiO-66-NH2PIM-19.112548100.8322.32.290.0717′ [50]
UiO-66-NH2PIM-1104252869−6.127.570.81.4817′ [66]
UiO-66-NH2PIM-1104 (c)251900−37.82449.00.6717′ [66]
UiO-66-NH2PIM-1 (3-month aging)9.11 (e)2548351.3628.229.40.7517′ [50]
UiO-66PU24--75.295.834.2−12.80.2718′ [52]
UiO-66Matrimid104377.813.029.4−1.40.0818′ [65]
Azo-UiO-66Matrimid104371044.93724.00.9918′ [65]
UiO-66ODPA-TMPDA201 (f)3516992.031.9−3.60.54This work
UiO-66-NH2ODPA-TMPDA201 (f)3514261.337.112.00.81This work
UiO-66-BrODPA-TMPDA201 (f)3520012734.54.230.94This work
UiO-66-(OH)2ODPA-TMPDA201 (f)3512542.038.917.50.81This work
Note: % En.—Percentage enhancement; Findex—Filler enhancement index (in Section 2.4.5); MeOH—methanol; PEBA—polyether block amide; PSF—Polysulfone; PU—Polyurethane ether; azo—azobenzene; UiO-66-ref—UiO-66 particles without water modulation (to make small particles). (a) The performance in the table is reported as pure gas permeation unless stated (indicated in the pressure column); (b) the loading amount is selected to be as close as possible to the amount that is used in this work in order to provide a better comparability with the Findex calculation; (c) CO2/N2 (50/50, in vol%); (d) humid condition; (e) N2/CO2/O2 (80:10:10, in vol%); (f) N2/CO2 (80:20, in vol%).

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/membranes10070154

AMA Style

Chuah CY, 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(7):154. https://doi.org/10.3390/membranes10070154

Chicago/Turabian Style

Chuah, Chong Yang, Junghyun Lee, Juha Song, and Tae-Hyun Bae. 2020. "CO2/N2 Separation Properties of Polyimide-Based Mixed-Matrix Membranes Comprising UiO-66 with Various Functionalities" Membranes 10, no. 7: 154. https://doi.org/10.3390/membranes10070154

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