Rational Design of Metal-Organic-Framework-Based Membranes: Comparison
Please note this is a comparison between Version 2 by Peter Tang and Version 1 by Yutian Duan.

Metal-organic framework (MOF) materials have been widely investigated and have been found to have enormous potential in membrane separation due to their uniform pore size and high designability. Notably, pure MOF films and MOF mixed matrix membranes (MMMs) are the core of the “next generation” MOF materials.

  • metal organic framework (MOF)
  • MOF-based membranes
  • gas separation
  • liquid separation

1. Introduction

Separation of chemical mixtures is the most energy-consuming process in the chemical industry [1,2][1][2]. Traditional separation methods, such as thermally driven phase change separation processes including distillation and regeneration of adsorbents after adsorption and separation [3], consume enormous resources, and the energy consumption of distillation accounts for about 50% of total industrial separation [4]. However, membrane separation technology conducted through molecular size and chemical affinity, not limited to thermal-driven force and adsorbent, has the advantages of energy saving, environmental protection, small floor space, recovery of valuable chemicals, and so on [5,6,7,8,9][5][6][7][8][9]. Thus, membrane separation has improved energy efficiency, making it a promising alternative to traditional separation technology [7]. For example, replacing distillation with membrane separation technology in the petrochemical field could save 80% of energy [10] and reduce global energy consumption by 8%. In particular, the membranes consist of inorganic, polymer, and mixed matrix membranes, of which the inorganic membrane has been successfully applied to the separation of CO2 and CH4 [11]. However, the inorganic membrane is difficult to realize in large-scale industrial applications due to its high capital cost. Meanwhile, although the polymeric membrane is easy to process at a low cost, there still exist problems such as aging, poor thermal stability, and mechanical strength, such as easy plasticization under high pressure during separation and purification. In addition, it is worth noting that there is also an inherent trade-off between the selectivity and permeability of polymer membranes [12], viz., increasing membrane permeability may decrease selectivity, and vice versa.
Later, researchers proposed the concept of metal-organic framework (MOF), which is a class of porous crystalline materials formed by the self-assembly of metal ions (or metal clusters) and organic ligands through covalent and coordination bonds [13]. Compared with commonly used microporous materials for separation, such as activated carbon and zeolite, MOF has a high specific surface area (i.e., 500–7000 m2/g) with good thermal/physical stability [14], and the structure, pore size, and functions of MOF could be delicately tailored by metal ions, organic ligands, etc. [15[15][16][17],16,17], thus it is widely used for gas storage, adsorption separation, chemical sensing, catalysis, biological applications, and so on [17,18,19,20][17][18][19][20]. As early as 2016, MOF adsorbents were used for the storage and subsequent release of 1-methylcyclopropene (1-MCP) to extend the shelf life of fruits and vegetables [21]. MOF membrane separation is one of the most effective methods to maximize the utilization of the potential of MOF-based materials [22], which is expected to serve as a substitute to solve many thorny problems (e.g., stability problems of polymers) for membrane separation.
MOF-based composites include pure MOF membranes and MOF-based mixed matrix membranes (MMMs), both of which are considered the next generation of MOF materials [4]. To date, numerous high-performance pure MOF membranes and MMMs have been prepared for gas separation (CO2, H2, olefin/paraffin, etc.) and liquid separation (water purification, organic solvent separation, chiral resolution, etc.), according to the difference in molecular size and chemical affinity, while the sorption-diffusion model has been employed to describe the permeation phenomenon [23,24][23][24].
Pure MOF membranes and MMMs are the next generation of MOF materials, which could be used for gas and liquid separation as an energy-saving method [30][25], in place of conventional energy-consuming techniques such as distillation. As a widely known burgeoning material, pure MOF membranes are typically assembled on a porous substrate (e.g., α-alumina and silica), which nevertheless still has restrictive issues of framework flexibility [25][26], defects [26][27], orientation [27][28], and so on [4]. As another important membrane, MMMs have a two-phase composition by incorporating MOF as fillers into the polymer matrix; however, there are problems such as MOF dispersity [31][29], polymer plasticization, aging [32][30], interfacial compatibility [33][31], and so on [34][32]. Therefore, the composition of MOF and matrix and their compatibility with each other are crucial for their function and efficiency [21].

2. Properties of MOFs

MOF is normally composed of metal nodes and organic linkers. Compared to conventional materials such as silica gel and zeolites, MOFs have outstanding water absorption capacity, recyclability, ease of reconstruction, and unique heat transfer properties [35][33]. Typical structures of MOF [14] used for membrane preparation are zeolitic imidazolate frameworks (ZIFs), UiO-66 (i.e., Zr6O4(OH)4(benzene-1,4-dicarboxylato)6), HKUST-1 (i.e., Cu3(1,3,5-benzenetricarboxylate)2), MIL-53, and MIL-101. Due to the diversity of metal nodes and organic linkers in MOF, ~70,000 MOF materials have been synthesized in the past 10 years [36][34], and the pore size, structure (e.g., slit, tubular, spherical, cylindrical, etc.) [37][35], and functions formed are multifarious [38][36]. Therefore, gas or liquid molecules could be separated according to the difference in size and adsorption affinity [39,40][37][38]. Note that the size of the MOF could be precisely tailored by isoreticular approaches, framework interpenetration, pore space partition, and other methods. For example, Eddaoudi et al. [41][39] used an organic linker with different lengths to replace the original MOF linker, wherein the replaced organic linker has the same connectivity and geometry. Herein, the framework structure remained unchanged, but the aperture changed from 3.8 Å to 23.8 Å. In addition to changing the organic linker, the isoreticular approach could also replace metal ions. For example, Shekhah et al. [42][40] replaced the metal ion of SIFSIX-3-M MOF from Zn to Cu, with the pore size changing from 3.84 Å to 3.50 Å. However, the interaction between the pore surface and molecules could be enhanced with the decrease in pore size; therefore, small molecules could also pass through pore voids and thus be sieved. The improvement of MOF separation performance could be achieved not only through the precise control of apertures but also through the functionalization of MOF. Herein, functionalization could render MOF selectively recognize gas or liquid molecules by selectively enhancing binding affinity between them. Functionalized sites include organic linkers [43][41] and unsaturated metal sites (i.e., open metal sites, OMSs) [44][42], of which organic linkers sites could further introduce polar groups (e.g., –NH2, –NO2, –OH, –Br, and so on) and realize the selective sieving of MOF by intermolecular interactions, such as H-binding.

3. Pure MOF Membranes

Pure MOF membranes are prepared on porous substrates by some commonly used preparation methods, including in situ growth, seeded growth, layer-by-layer (LbL) techniques, and vapor phase growth [45][43]. The main factors affecting the separation performance of pure MOF membranes are mainly composed of framework flexibility and grain boundary structure [46][44], wherein the framework flexibility allows larger molecules to pass through [25][26], while the poor grain boundary structure could produce defects acting as a nonselective pore to affect the separation performance of membranes [26][27]. Note that recent reports have demonstrated that defects can help enhance the separation performance of pure MOF membranes rather than cause side effects through defect engineering due to increased porosity and OMSs [47][45]. In addition, since pure MOF membranes are polycrystalline materials, grain orientation can also affect the separation performance of the membranes [48][46].

4. Mixed-Matrix Membranes

The MMM is composed of MOF particle filler and polymer substrate, wherein the commonly used polymer substrate includes poly(methyl methacrylate) (PMMA), PI, poly(vinylidene fluoride) (PVDF), polysulfone(PSF), polyethersulfone (PES), etc. [34][32]. Due to the fact that permeance and selectivity are inversely correlated with highly permeable membranes lacking selectivity and vice versa, the separation performance of polymeric membranes is limited, which could be improved to exceed the Robeson upper bound limits by adding MOF particles to the polymer matrix. However, MMM’s capability for gas and liquid separation is dented due to the aggregation of MOF particles [31][29], the plasticization and aging of polymer [32][30], and the interfacial compatibility [33][31] between MOF and matrix. Therefore, first and foremost, the aforementioned concerns need to be properly addressed in order to prepare high-performance MMMs.

5. Properties of MOF Membranes

5.1. Stability

Stability is a prerequisite for material performance, and many efforts have been made to improve the stability of MOF-based materials against chemical changes induced by active liquids and/or gases through exquisite structural design and functionalization [4,72][4][47]. Notably, ZIFs consisting of imidazolate ligands and metal nodes possess prominent stability due to their zeolite-like structure. The stability of other water-stable MOFs, such as MIL-53 (Al3+), UiO-66 (Zr4+), MIL-100, and MIL-101 (Cr3+), can be enhanced by using highly charged metals to form strong metal-linker bonds, which presented outstanding water stability when used for membrane separation [21]. For example, Hu et al. [73][48] used the MIL-53 membrane for dehydration of the azeotrope of ethyl acetate (EA) aqueous solution by pervaporation. The MIL-53 membrane maintained high stability during evaporation for over 200 h, and hydroxyl groups on the MIL-53 surface formed hydrogen bonds with water molecules, promoting water transport with a flow rate of 454 g m−2 h−1. Liu et al. [74][49] used the UiO-66 membrane for water softening. Note that the Uio-66 membrane could maintain excellent stability and exhibited good multivalent rejection effects lasting ~170 h in different saline solutions (i.e., Ca2+, Mg2+, and Al3+). Among them, the ZIF-8/(TA-Zn2+)2/PES membrane displayed almost no changes in both pure water permeance and Na2SO4 rejection after long-term filtration for 100 h, indicating remarkable stability. Moreover, when Jian et al. [75][50] prepared Al-MOF membranes (two-dimensional (2D) monolayer aluminum tetra-(4-carboxyphenyl) porphyrin framework), no Al3+ dissolution was detected after immersion in water for one month, which revealed good water stability as well. In liquid separation, MOFs are generally stable, although metal ions (e.g., Zn2+) could release from the matrix into the surrounding medium [19,76][19][51]. Notably, the hydrolysis of MOFs (e.g., ZIF) in aqueous media under ambient conditions was associated with the release of metal ions (e.g., Zn2+) that feature a well-established acid sensitivity [19,76][19][51]. Over time, the release rate first increases and then decreases, resulting in long-term sustained release [77][52]. Whereas MOF stability decreases with an accelerated rate of degradation in a weakly acidic environment (e.g., pH = 5.5), wherein metal ions (e.g., Zn2+) could be released at a higher rate than in neutral conditions [76,77][51][52]. For example, Schnabel et al. [78][53] prepared the Zn-MOF-74, comprising Zn(II) ions and 2,5-dihydroxybenzene-1,4-dicarboxylate ligand, wherein the detected amount of Zn2+ was merely ~0.1% during the whole time period after being tested for 160 h in neutral aqueous conditions (pH = 7.4) and ~0.8% at pH 6.0. Note that with the slow degradation of MOF materials, the long-term release of metal ions could be achieved, which greatly reduces the toxicity of metal ions and prolongs the function time [79,80,81][54][55][56]. Notably, the stability of MOF membranes could be affected by humidity and corrosive and acidic substances when used for gas separation. For example, moisture could cause the MOF OMS to be inactive. Meanwhile, H2S, SOx, and NOx can disrupt weak ligand-metal linkages in MOF. In addition, heavy hydrocarbons (such as heptane or toluene) can damage pores or active sites of MOF membranes, thus affecting stability [4]. Herein, the stability of MOFs could be improved by various methods, such as structural changes and functional modifications. For example, when ZIF-8 with a sodalite (SOD) topology is exposed to SO2 under high humidity, it is unstable with a damaged structure. However, the stability of ZIF with RHO topology was improved, showing stability for both SO2 and CO2 even under highly humid conditions [4]. Moreover, MIL-125(Ti)-NH2, obtained after amine functionalization of MIL-125(Ti), exhibited high H2S stability [82][57]. Furthermore, the stability of UiO-66 MOF-based MMMs could also be improved through a defect-engineered strategy, where performance remained unchanged even after industrial separation of C3H6/C3H8 under harsh conditions (i.e., 50 °C and 5 bar) for 14 days, displaying outstanding stability [71][58].

5.2. Lifetime

Note that MOFs (e.g., ZIF) demonstrated remarkable chemical resistance and thermal stability (e.g., up to 550 °C in N2 for ZIF) against solvents because of the strong interactions between core metal ions and ligands [83][59], which could greatly extend the lifetime of MOF-based materials in both liquid and gas separations for practical applications [27,84][28][60]. For example, ZIF-8, which is made up of imidazolate anions forming a tetrahedral joint with ZnN4, has ideal stability (~7 days) in phosphate buffer (pH = 7.4) at 37 °C. Meanwhile, ZIF-8 and ZIF-11 could maintain their crystalline structures in water at 50 °C for 7 days [85][61]. Liu et al. [86][62] reported a series of continuous UiO-66 polycrystalline membranes on pre-structured yttria-stabilized zirconia hollow fiber supports. UiO-66 membranes as prepared could maintain ideal separation performance during a 300 h stability test and remain robust even under harsh conditions (e.g., boiling benzene, boiling water, and sulphuric acid). Deng et al. [87][63] fabricated a novel superwetting HKUST-1 membrane by seed-mediated growth. Notably, the permeation fluxes in both cases (NaCl and pH effect) decreased with increasing time, and the flux was still high at 7 days after washing the membrane immersed in an acid-based solution, which demonstrated the outstanding stability of the HKUST-1 membrane in acidic, basic, and salty environments. Furthermore, several continuous and dense ZIF-8 polycrystalline membranes were fabricated and tested for seawater desalination [88][64]. The obtained ZIF-8 membranes showed high ion rejection of >99% and excellent water flux of ~6 kg m−2 h−1 at 25 °C. Notably, these ZIF-8 membranes could maintain good separation performance even during a stability test of up to 7 days. Furthermore, related studies have shown that MOF membranes also exhibit exceptional robustness and a long-lasting lifetime for gas separation [27,84][28][60]. For example, Zhou et al. [89][65] pioneered the electrochemical synthesis of a series of defect-free face-centered cubic (fcu)-MOF polycrystalline membranes with molecular sieving capacities for hydrocarbon separations, including butane/isobutane (nC4/iC4) and C3H6/C3H8 mixtures. Thanks to the rigid structure of the framework, the optimized membrane demonstrated outstanding and stable nitrogen rejection performance even after a continuous permeation test over 150 days. Later, they reported the rational design of various MMMs with high loading content (~60 wt%) of (001)-oriented AlFFIVE-1-Ni nanosheets for efficient CO2/H2S/CH4 separation [90][66]. Notably, the optimized MMM could maintain outstanding separation capacity even after a long-lasting stability test of more than 30 days. This work distinctly highlights the promising potential of utilizing MOF membranes with unprecedented separation abilities. Hou et al. [50][67] proposed a mixed-linker strategy to prepare a series of ZIF-7x-8 hybrid membranes using Zn2+ cations and 2-methylimidazole mixtures as ligands via the FCDS approach (x represents the molar percentage of benzimidazole). Moreover, these hybrid membranes could maintain separation performance even after a 180 h temperature swing separation test, consolidating their structural stability and robustness. Moreover, Sabetghadam et al. [91][68] prepared a MOF/PIMAT membrane, resulting in both a substantial enhancement of CO2/N2 selectivity and CO2 permeability under both dry and humid conditions while greatly reducing aging. The MMM obtained exceeds the 2008 Robeson upper bound limit and reaches the economic target zone for post-combustion CO2 capture even after 510 days of aging.

5.3. Environmental Friendliness and Biocompatibility

Environmentally friendly and biocompatible MOFs have already played a crucial and prerequisite role in practical applications, including bio-MOFs or biomimetic MOFs [92,93][69][70]. Notably, MOFs based on amino acids and their derivatives are widely recognized for their environmental friendliness and biocompatibility. For example, Huang et al. [94][71] prepared “green” MOFs with aspartic acid derivatives (i.e., spartic acid, succinic acid, fumaric acid, and malic acid) as organic ligands and Zr(IV) as central ions as materials, which are ideal eco-friendly candidates for phosphorus and arsenic(V) removal in complex real water bodies. Moreover, some metals are essential trace metal elements for humans and are not considered toxic for specific applications. For example, the recommended daily calcium intake is ~1 g, which is considered highly biocompatible. Therefore, Ca-MOFs are biocompatible materials with environment-friendly properties that are green, safe, and sustainable [95][72]. Overall, the factors affecting the biocompatibility of MOFs include coordination metals, organic ligands, particle size, morphology, shape, surface charge, and so on [92][69]. For coordination metals, the median lethal dose of different metal elements and chemical formulas (counter ion and oxidation states) varies. In particular, organic ligands can be divided into exogenous and endogenous ligands, where exogenous ligands can enhance their biocompatibility through functionalization, while endogenous ligands have essentially good biocompatibility. Therefore, in order to improve MOF compatibility, endogenous metal ions and ligands should be used as much as possible during the synthesis stage. In addition, relevant studies have shown that MOFs have good biocompatibility when particle size is below 200 nm with a hydrophilic surface and stability at physiological pH [92][69].

5.4. Fouling

Notably, fouling refers to the deposition and accumulation of undesired materials on the surface or inside of a membrane (such as dissolved particles, organic macromolecules, inorganic macromolecules, and biological micro-organisms) [96][73], which presents the greatest challenge to a more widespread use of membrane separation, a potentially energy-efficient and cost-effective separation procedure, in a wide range of industrial sectors [97][74]. Note that reverse osmosis is currently one of the most successful membrane-based water purification processes in industrial seawater desalination, yet it remains critically affected by membrane fouling [98][75]. Thus, the pressure would be increased to maintain the high flux, further resulting in an increase in energy requirements [96][73]. However, Prince et al. developed a self-cleaning PANCMA-PEI-Ag modified PES membrane; (PANCMA = poly(acrylonitrile-co-maleic acid); PEI = polyethyleneimine) that afforded a delicate solution to fouling in membrane separation processes. Wherein the flux drop for the novel membrane is lower (16.3% of the initial flux) during long-term experiments with protein solution. Moreover, the novel membrane continues to exhibit inhibition of microbes even after 1320 min of protein filtration, which opens up a promising MOF membrane-based solution for biofouling in wastewater purification [99][76].

References

  1. Sholl, D.S.; Lively, R.P. Seven chemical separations to change the world. Nature 2016, 532, 435–437.
  2. Semiat, R. Energy issues in desalination processes. Environ. Sci. Technol. 2008, 42, 8193–8201.
  3. Gascon, J.; Kapteijn, F. Metal-organic framework membranes—High potential, bright future? Angew. Chem. Int. Ed. 2010, 49, 1530–1532.
  4. Qian, Q.; Asinger, P.A.; Lee, M.J.; Han, G.; Rodriguez, K.M.; Lin, S.; Benedetti, F.M.; Wu, A.X.; Chi, W.S.; Smith, Z.P. MOF-based membranes for gas separations. Chem. Rev. 2020, 120, 8161–8266.
  5. Eyal, A.M.; Bressler, E. Industrial separation of carboxylic and amino acids by liquid membranes: Applicability, process considerations, and potential advantage. Biotechnol. Bioeng. 1993, 41, 287–295.
  6. Brunetti, A.; Scura, F.; Barbieri, G.; Drioli, E. Membrane technologies for CO2 separation. J. Membr. Sci. 2010, 359, 115–125.
  7. Baker, R.W. Future directions of membrane gas separation technology. Ind. Eng. Chem. Res. 2002, 41, 1393–1411.
  8. Khosravikia, M.; Rahbar-Kelishami, A. A simulation study of an applied approach to enhance drug recovery through electromembrane extraction. J. Mol. Liq. 2022, 358, 119210.
  9. Marson, G.V.; Belleville, M.P.; Lacour, S.; Hubinger, M.D. Membrane fractionation of protein hydrolysates from by-products: Recovery of valuable compounds from spent yeasts. Membranes 2021, 11, 23.
  10. Wang, H.; Wang, M.; Liang, X.; Yuan, J.; Yang, H.; Wang, S.; Ren, Y.; Wu, H.; Pan, F.; Jiang, Z. Organic molecular sieve membranes for chemical separations. Chem. Soc. Rev. 2021, 50, 5468–5516.
  11. Li, H.; Haas-Santo, K.; Schygulla, U.; Dittmeyer, R. Inorganic microporous membranes for H2 and CO2 separation—Review of experimental and modeling progress. Chem. Eng. Sci. 2015, 127, 401–417.
  12. Robeson, L.M. Correlation of separation factor versus permeability for polymeric membranes. J. Membr. Sci. 1991, 62, 165–185.
  13. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705–714.
  14. Lin, R.B.; Xiang, S.; Zhou, W.; Chen, B. Microporous metal-organic framework materials for gas separation. Chem 2020, 6, 337–363.
  15. Cohen, S.M. Postsynthetic methods for the functionalization of metal-organic frameworks. Chem. Rev. 2012, 112, 970–1000.
  16. Ahmadijokani, F.; Tajahmadi, S.; Bahi, A.; Molavi, H.; Rezakazemi, M.; Ko, F.; Aminabhavi, T.M.; Arjmand, M. Ethylenediamine-functionalized Zr-based MOF for efficient removal of heavy metal ions from water. Chemosphere 2021, 264, 128466.
  17. Islamoglu, T.; Goswami, S.; Li, Z.; Howarth, A.J.; Farha, O.K.; Hupp, J.T. Postsynthetic tuning of metal-organic frameworks for targeted applications. Acc. Chem. Res. 2017, 50, 805–813.
  18. Zhang, M.; Qiao, R.; Hu, J. Engineering metal–organic frameworks (MOFs) for controlled delivery of physiological gaseous transmitters. Nanomaterials 2020, 10, 1134.
  19. Wu, M.X.; Yang, Y.W. Metal–organic framework (MOF)-based drug/cargo delivery and cancer therapy. Adv. Mater. 2017, 29, 1606134.
  20. Rabiee, N.; Atarod, M.; Tavakolizadeh, M.; Asgari, S.; Rezaei, M.; Akhavan, O.; Pourjavadi, A.; Jouyandeh, M.; Lima, E.C.; Mashhadzadeh, A.H.; et al. Green metal-organic frameworks (MOFs) for biomedical applications. Microporous Mesoporous Mater. 2022, 335, 111670.
  21. Freund, R.; Zaremba, O.; Arnauts, G.; Ameloot, R.; Skorupskii, G.; Dincă, M.; Bavykina, A.; Gascon, J.; Ejsmont, A.; Goscianska, J.; et al. The Current status of MOF and COF applications. Angew. Chem. Int. Ed. 2021, 60, 23975–24001.
  22. Huang, K.; Wang, B.; Guo, S.; Li, K. Micropatterned ultrathin MOF membranes with enhanced molecular sieving property. Angew. Chem. Int. Ed. 2018, 57, 13892–13896.
  23. Graham, T. On the absorption and dialytic separation of gases by colloid septa Part I.—Action of a septum of caoutchouc. J. Membr. Sci. 1995, 100, 27–31.
  24. Wijmans, J.G.; Baker, R.W. The solution-diffusion model: A review. J. Membr. Sci. 1995, 107, 1–21.
  25. Hosseini Monjezi, B.; Kutonova, K.; Tsotsalas, M.; Henke, S.; Knebel, A. Current trends in metal–organic and covalent organic framework membrane materials. Angew. Chem. Int. Ed. 2021, 60, 15153–15164.
  26. 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.
  27. Wu, W.; Li, Z.; Chen, Y.; Li, W. Polydopamine-modified metal-organic framework membrane with enhanced selectivity for carbon capture. Environ. Sci. Technol. 2019, 53, 3764–3772.
  28. Cheng, Y.; Datta, S.J.; Zhou, S.; Jia, J.; Shekhah, O.; Eddaoudi, M. Advances in metal-organic framework-based membranes. Chem. Soc. Rev. 2022, 51, 8300–8350.
  29. Zhu, H.; Wang, L.; Jie, X.; Liu, D.; Cao, Y. Improved interfacial affinity and CO2 separation performance of asymmetric mixed matrix membranes by incorporating postmodified MIL-53(Al). ACS Appl. Mater. Interfaces 2016, 8, 22696–22704.
  30. Husna, A.; Hossain, I.; Choi, O.; Lee, S.M.; Kim, T.H. Efficient CO2 separation using a PIM-PI-functionalized UiO-66 MOF incorporated mixed matrix membrane in a PIM-PI-1 polymer. Macromol. Mater. Eng. 2021, 306, 2100298.
  31. Qian, Q.; Wu, A.X.; Chi, W.S.; Asinger, P.A.; Lin, S.; Hypsher, A.; Smith, Z.P. Mixed-matrix membranes formed from imide-functionalized UiO-66-NH2 for improved interfacial compatibility. ACS Appl. Mater. Interfaces 2019, 11, 31257–31269.
  32. Kalaj, M.; Bentz, K.C.; Ayala, S.; Palomba, J.M.; Barcus, K.S.; Katayama, Y.; Cohen, S.M. MOF-polymer hybrid materials: From simple composites to tailored architectures. Chem. Rev. 2020, 120, 8267–8302.
  33. Mohseni, M.M.; Jouyandeh, M.; Sajadi, S.M.; Hejna, A.; Habibzadeh, S.; Mohaddespour, A.; Rabiee, N.; Daneshgar, H.; Akhavan, O.; Asadnia, M.; et al. Metal-organic frameworks (MOF) based heat transfer: A comprehensive review. Chem. Eng. J. 2022, 449, 137700.
  34. Moghadam, P.Z.; Li, A.; Wiggin, S.B.; Tao, A.; Maloney, A.G.P.; Wood, P.A.; Ward, S.C.; Fairen-Jimenez, D. Development of a cambridge structural database subset: A collection of metal-organic frameworks for past, present, and future. Chem. Mater. 2017, 29, 2618–2625.
  35. Stock, N.; Biswas, S. Synthesis of metal-organic frameworks (MOFs): Routes to various MOF topologies, morphologies, and composites. Chem. Rev. 2012, 112, 933–969.
  36. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444.
  37. Millward, A.R.; Yaghi, O.M. Metal-organic frameworks with exceptionally high capacity for storage of carbon dioxide at room temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999.
  38. Yu, S.; Pang, H.; Huang, S.; Tang, H.; Wang, S.; Qiu, M.; Chen, Z.; Yang, H.; Song, G.; Fu, D.; et al. Recent advances in metal-organic framework membranes for water treatment: A review. Sci. Total Environ. 2021, 800, 149662.
  39. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O.M. Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage. Science 2002, 295, 469–472.
  40. Shekhah, O.; Belmabkhout, Y.; Chen, Z.; Guillerm, V.; Cairns, A.; Adil, K.; Eddaoudi, M. Made-to-order metal-organic frameworks for trace carbon dioxide removal and air capture. Nat. Commun. 2014, 5, 4228.
  41. Shah, M.; McCarthy, M.C.; Sachdeva, S.; Lee, A.K.; Jeong, H.K. Current status of metal-organic framework membranes for gas separations: Promises and challenges. Ind. Eng. Chem. Res. 2012, 51, 2179–2199.
  42. Trickett, C.A.; Helal, A.; Al-Maythalony, B.A.; Yamani, Z.H.; Cordova, K.E.; Yaghi, O.M. The chemistry of metal-organic frameworks for CO2 capture, regeneration and conversion. Nat. Rev. Mater. 2017, 2, 17045.
  43. Deng, Y.; Wu, Y.; Chen, G.; Zheng, X.; Dai, M.; Peng, C. Metal-organic framework membranes: Recent development in the synthesis strategies and their application in oil-water separation. Chem. Eng. J. 2021, 405, 127004.
  44. Li, H.; Li, L.; Lin, R.B.; Zhou, W.; Zhang, Z.; Xiang, S.; Chen, B. Porous metal-organic frameworks for gas storage and separation: Status and challenges. EnergyChem 2019, 1, 100006.
  45. Hou, J.; Sutrisna, P.D.; Wang, T.; Gao, S.; Li, Q.; Zhou, C.; Sun, S.; Yang, H.C.; Wei, F.; Ruggiero, M.T.; et al. Unraveling the interfacial structure-performance correlation of flexible metal-organic framework membranes on polymeric substrates. ACS Appl. Mater. Interfaces 2019, 11, 5570–5577.
  46. Pham, T.C.T.; Nguyen, T.H.; Yoon, K.B. Gel-free secondary growth of uniformly oriented silica MFI zeolite films and application for xylene separation. Angew. Chem. Int. Ed. 2013, 52, 8693–8698.
  47. Wang, C.; Liu, X.; Demir, N.K.; Chen, J.P.; Li, K. Applications of water stable metal-organic frameworks. Chem. Soc. Rev. 2016, 45, 5107–5134.
  48. Hu, Y.; Dong, X.; Nan, J.; Jin, W.; Ren, X.; Xu, N.; Lee, Y.M. Metal-organic framework membranes fabricated via reactive seeding. Chem. Commun. 2011, 47, 737–739.
  49. Liu, X.; Demir, N.K.; Wu, Z.; Li, K. Highly water-stable zirconium metal-organic framework UiO-66 membranes supported on alumina hollow fibers for desalination. J. Am. Chem. Soc. 2015, 137, 6999–7002.
  50. Jian, M.; Qiu, R.; Xia, Y.; Lu, J.; Chen, Y.; Gu, Q.; Liu, R.; Hu, C.; Qu, J.; Wang, H.; et al. Ultrathin water-stable metal-organic framework membranes for ion separation. Sci. Adv. 2020, 6, eaay3998.
  51. Zhao, Q.; Gong, Z.; Wang, J.; Fu, L.; Zhang, J.; Wang, C.; Miron, R.J.; Yuan, Q.; Zhang, Y. A zinc- and calcium-rich lysosomal nanoreactor rescues monocyte/macrophage dysfunction under sepsis. Adv. Sci. 2023, 10, 2205097.
  52. Chen, Y.; Cai, J.; Liu, D.; Liu, S.; Lei, D.; Zheng, L.; Wei, Q.; Gao, M. Zinc-based metal organic framework with antibacterial and anti-inflammatory properties for promoting wound healing. Regen. Biomater. 2022, 9, rbac019.
  53. Schnabel, J.; Ettlinger, R.; Bunzen, H. Zn-MOF-74 as pH-responsive drug-delivery system of arsenic trioxide. ChemNanoMat 2020, 6, 1229–1236.
  54. Velásquez-Hernández, M.D.J.; Ricco, R.; Carraro, F.; Limpoco, F.T.; Linares-Moreau, M.; Leitner, E.; Wiltsche, H.; Rattenberger, J.; Schröttner, H.; Frühwirt, P.; et al. Degradation of ZIF-8 in phosphate buffered saline media. CrystEngComm 2019, 21, 4538–4544.
  55. Zhang, S.; Pei, X.; Gao, H.; Chen, S.; Wang, J. Metal-organic framework-based nanomaterials for biomedical applications. Chin. Chem. Lett. 2020, 31, 1060–1070.
  56. Fan, G.; Bao, M.; Zheng, X.; Hong, L.; Zhan, J.; Chen, Z.; Qu, F. Growth inhibition of harmful cyanobacteria by nanocrystalline Cu-MOF-74: Efficiency and its mechanisms. J. Hazard. Mater. 2019, 367, 529–538.
  57. Vaesen, S.; Guillerm, V.; Yang, Q.; Wiersum, A.D.; Marszalek, B.; Gil, B.; Vimont, A.; Daturi, M.; Devic, T.; Llewellyn, P.L.; et al. A robust amino-functionalized titanium(IV) based MOF for improved separation of acid gases. Chem. Commun. 2013, 49, 10082–10084.
  58. 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, 60, 13081–13088.
  59. Moharramnejad, M.; Ehsani, A.; salmani, S.; shahi, M.; Malekshah, R.E.; Robatjazi, Z.S.; Parsimehr, H. Zinc-based metal-organic frameworks: Synthesis and recent progress in biomedical application. J. Inorg. Organomet. Polym. Mater. 2022, 32, 3339–3354.
  60. Dechnik, J.; Gascon, J.; Doonan, C.J.; Janiak, C.; Sumby, C.J. Mixed-Matrix Membranes. Angew. Chem. Int. Ed. 2017, 56, 9292–9310.
  61. Beg, S.; Rahman, M.; Jain, A.; Saini, S.; Midoux, P.; Pichon, C.; Ahmad, F.J.; Akhter, S. Nanoporous metal organic frameworks as hybrid polymer–metal composites for drug delivery and biomedical applications. Drug Discov. Today 2017, 22, 625–637.
  62. Liu, X.; Wang, C.; Wang, B.; Li, K. Novel organic-dehydration membranes prepared from zirconium metal-organic frameworks. Adv. Funct. Mater. 2017, 27, 1604311.
  63. Deng, Y.; Dai, M.; Wu, Y.; Ali, I.; Zhao, J.; Li, S.; Peng, C. High-efficient novel super-wetting HKUST-1 membrane for oil-water separation: Development, characterization and performance. J. Clean. Prod. 2022, 333, 130109.
  64. Zhu, Y.; Gupta, K.M.; Liu, Q.; Jiang, J.; Caro, J.; Huang, A. Synthesis and seawater desalination of molecular sieving zeolitic imidazolate framework membranes. Desalination 2016, 385, 75–82.
  65. Zhou, S.; Shekhah, O.; Jia, J.; Czaban-Jóźwiak, J.; Bhatt, P.M.; Ramírez, A.; Gascon, J.; Eddaoudi, M. Electrochemical synthesis of continuous metal–organic framework membranes for separation of hydrocarbons. Nat. Energy 2021, 6, 882–891.
  66. Datta, S.J.; Mayoral, A.; Bettahalli, N.M.S.; Bhatt, P.M.; Karunakaran, M.; Carja, I.D.; Fan, D.; Mileo, P.G.M.; Semino, R.; Maurin, G.; et al. Rational design of mixed-matrix metal-organic framework membranes for molecular separations. Science 2022, 376, 1080–1087.
  67. Hou, Q.; Wu, Y.; Zhou, S.; Wei, Y.; Caro, J.; Wang, H. Ultra-tuning of the aperture size in stiffened ZIF-8_Cm frameworks with mixed-linker strategy for enhanced CO2/CH4 separation. Angew. Chem. Int. Ed. 2019, 58, 327–331.
  68. Sabetghadam, A.; Liu, X.; Orsi, A.F.; Lozinska, M.M.; Johnson, T.; Jansen, K.M.B.; Wright, P.A.; Carta, M.; McKeown, N.B.; Kapteijn, F.; et al. Towards high performance metal–organic framework–microporous polymer mixed matrix membranes: Addressing compatibility and limiting aging by polymer doping. Chem. Eur. J. 2018, 24, 12796–12800.
  69. Singh, N.; Qutub, S.; Khashab, N.M. Biocompatibility and biodegradability of metal organic frameworks for biomedical applications. J. Mater. Chem. B 2021, 9, 5925–5934.
  70. Tamames-Tabar, C.; Cunha, D.; Imbuluzqueta, E.; Ragon, F.; Serre, C.; Blanco-Prieto, M.J.; Horcajada, P. Cytotoxicity of nanoscaled metal-organic frameworks. J. Mater. Chem. B 2014, 2, 262–271.
  71. Huang, Q.; Jiang, X.; Xiong, J.; Zhou, Q.; Zhu, Y.; Xie, Q.; Wang, S.; Yang, X.; Jiang, F. Aspartic acid derivative-based MOFs: A promising green material for simultaneous removal of phosphorus and arsenic(V) in contaminated spring water. J. Water Process Eng. 2023, 52, 103547.
  72. Sukatis, F.F.; Wee, S.Y.; Aris, A.Z. Potential of biocompatible calcium-based metal-organic frameworks for the removal of endocrine-disrupting compounds in aqueous environments. Water Res. 2022, 218, 118406.
  73. Alsawaftah, N.; Abuwatfa, W.; Darwish, N.; Husseini, G. A comprehensive review on membrane fouling: Mathematical modelling, prediction, diagnosis, and mitigation. Water 2021, 13, 1327.
  74. She, Q.; Wang, R.; Fane, A.G.; Tang, C.Y. Membrane fouling in osmotically driven membrane processes: A review. J. Membr. Sci. 2016, 499, 201–233.
  75. Nthunya, L.N.; Bopape, M.F.; Mahlangu, O.T.; Mamba, B.B.; Van der Bruggen, B.; Quist-Jensen, C.A.; Richards, H. Fouling, performance and cost analysis of membrane-based water desalination technologies: A critical review. J. Environ. Manag. 2022, 301, 113922.
  76. Prince, J.A.; Bhuvana, S.; Anbharasi, V.; Ayyanar, N.; Boodhoo, K.V.K.; Singh, G. Self-cleaning metal organic framework (MOF) based ultra filtration membranes—A solution to bio-fouling in membrane separation processes. Sci. Rep. 2014, 4, 6555.
More
Video Production Service