Zeolitic imidazolate framework

Last updated
The structure of a zeolitic imidazolate framework is made through three-dimensional assembly of metal(imidazolate)4 tetrahedra. Structure of zeolitic imidazolate frameworks.png
The structure of a zeolitic imidazolate framework is made through three-dimensional assembly of metal(imidazolate)4 tetrahedra.

Zeolitic imidazolate frameworks (ZIFs) are a class of metal-organic frameworks (MOFs) that are topologically isomorphic with zeolites. [1] ZIF glasses can be synthesized by the melt-quench method, and the first melt-quenched ZIF glass was firstly made and reported by Bennett et al. back in 2015. [2] ZIFs are composed of tetrahedrally-coordinated transition metal ions (e.g. Fe, Co, Zn) connected by imidazolate linkers. Since the metal-imidazole-metal angle is similar to the 145° Si-O-Si angle in zeolites, ZIFs have zeolite-like topologies. [3] As of 2010, 105 ZIF topologies have been reported in the literature. [4] [5] Due to their robust porosity, resistance to thermal changes, and chemical stability, ZIFs are being investigated for applications such as carbon dioxide capture. [6]

Contents

ZIF glasses are a newly discovered type of material that has been garnering increasing interest in recent years, with around 13 different ZIFs, including ZIF-4, ZIF-62, and ZIF-76, being successfully prepared in their glassy state. In traditional materials science, glasses can be divided into three major families: inorganic, organic, and metallic. The chemical bonds that make up the structure of members of each family are mixed ionic/covalent bonds, covalent bonds, and metallic bonds, respectively. ZIF glasses, on the other hand, are an organic-inorganic coordinated glass discovered only recently, and have a completely different structure than the three traditional glass families. They thus represent a fourth type of glass. [7]

Glassy structure

The structure of melt-quenched ZIF glasses maintains a certain amount of short-range order, although the chemical configuration and coordination environments, after melting, lose long-range order completely. From a microscopic view, the linkages between metal nodes and organic ligands (e.g., Zn-N linkages) partially break at high temperature and the resulting undercoordinated metal ions have the potential to link with other neighboring organic ligands for exchange.

One notable discovery regarding the structure of ZIF glass was made by Rasmus et al. [7] Before this research was published, the short-range structural order at the scale of the cation-ligand units remained unknown given the limitations of the analytical techniques available. The short-range structural disorder of the tetrahedral ligand environment around metal nodes in the ZIF glass was detected for the first time by performing zinc-67 nuclear magnetic resonance. This finding clearly showed that ZIF glasses are structurally very different from the other known glass types, overturning the traditional view that a glass structure has short-range order and long-range disorder, providing a broader view of what qualifies as a glass.

Synthesis

ZIFs are mainly prepared by solvothermal or hydrothermal techniques. Crystals slowly grow from a heated solution of a hydrated metal salt, an ImH (imidazole with acidic proton), a solvent, and base. [8] Functionalized ImH linkers allow for control of ZIF structure. [9] This process is ideal for generating monocrystalline materials for single-crystal X-ray diffraction. [10] [11] A wide range of solvents, bases, and conditions have been explored, with an eye towards improving crystal functionality, morphology, and dispersity. [12] Prototypically, an amide solvent such as N,N-dimethylformamide (DMF) is used. The heat applied decomposes the amide solvent to generate amines, which in turn generate the imidazolate from the imidazole species. Methanol, [13] [14] ethanol, [15] isopropanol, [16] and water [17] [18] [19] have also been explored as alternative solvents for ZIF formation but require bases such as pyridine, [20] TEA, [21] sodium formate, [22] and NaOH. [23] Polymers such as poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene oxide), [24] polyvinylpyrrolidone, [25] and poly-(diallyldimethylammonium chloride) [26] have been found to act as crystal dispersants, imparting particle-size and morphology control.

Due to their promising material properties, significant interest lies in economical large-scale production methods. Sonochemical synthesis, which allows nucleation reactions to proceed rapidly through acoustic generation of localized heat and pressure, has been explored as a way to shorten synthesis times. [27] [28] As with the case of zeolites, microwave-assisted synthesis has also been of interest for the rapid synthesis of ZIFs. [29] [30] Both methods have been shown to reduce reaction times from days to hours, or from hours to minutes. Solvent-free methods, such as ball-milling or chemical vapor deposition, have also been described to produce high-quality ZIF-8. [31] [32] Chemical vapor deposition is of particular promise due to the high degree of uniformity and aspect ratio control it can offer, and its ability to be integrated into traditional lithographic workflows for functional thin films (e.g. microelectronics). Environmentally-friendly synthesis based on supercritical carbon dioxide (scCO2) have been also reported as a feasible procedure for the preparation of ZIF-8 at an industrial scale. [33] Working under stoichiometric conditions, ZIF-8 could be obtained in 10 hours and does not require the use of ligand excess, additives, organic solvents or cleaning steps.

Using the traditional melt-quench of metals or sintering of ceramics would cause the collapse of MOF structure as its thermal decomposing temperature is lower than its melting temperature. Moreover, the amorphous form of MOF can be achieved through pressurization or heating, but its network feature would be significantly broken during the amorphization process. Bennett et al found certain members from MOF family (ZIF-4, etc.) can be made into a glassy state. [2] Those carefully selected ZIF crystals are able to form a glassy solid after heating and cooling in an argon atmosphere. Moreover, the melting range can be tuned by their network topologies.

Applications

The crystal form of ZIF, or MOF in general, is known for its porosity, but is difficult to mass-produce and incorporate in actual applications due to unavoidable intercrystalline defects. [34] There are several interesting characters about ZIF glasses addressing those challenges to potentially realize promised applications achievable. The first intriguing one is that ZIF glass maintains the porous structure as its crystalline form after melt-quench process, which means it can be applied for applications such as gas separation and storage. The glassy form would also offer unique opportunities for easy processability and mass production. Last but not least, besides pure ZIF glass, composites based on it by tuning the composition and structure has the distinct advantage of a broad design space.

Applications to carbon capture

ZIFs exhibit some properties relevant to carbon dioxide capture, [35] while commercial technology still centers around amine solvents. [36]

Zeolites are known to have tunable pores – ranging between 3-12 Angstroms – which allows them to separate carbon dioxide. Because a molecule is about 5.4 Angstroms in length, zeolites with a pore size of 4-5 Angstroms can be well-suited for carbon dioxide capture. However, other factors also need to be considered when determining how effective zeolites will be at carbon dioxide capture. The first is basicity, which can be created by doing an alkali metal cation exchange. The second is the Si/Al ratio which impacts the cation exchange capacity. To get a higher adsorption capacity, there must be a lower Si/Al ratio in order to increase the cation exchange capacity.

ZIFs 68, 69, 70, 78, 81, 82, 95, and 100 have been found to have very high uptake capacity, meaning that they can store a lot of carbon dioxide, though their affinity to it is not always strong. Of those, 68, 69, and 70 show high affinities for carbon dioxide, evidenced by their adsorption isotherms, which show steep uptakes at low pressures. One liter of ZIF can hold 83 liters of CO2. This could also be useful for pressure-swing adsorption. [37]

Gas separation

ZIF-62 was made into a glassy membrane on the nanoporous alumina support for gas separation for the first time by Yuhan et al in 2020. [38] The vitrification process effectively eliminates grain boundaries formation within the glass, and the molecular sieving ability of such membrane is significantly improved. The value of the ideal selectivities of several gas pairs, e.g. CO2/N2, are much higher than Knudsen selectivities, and the excellent performance of the ZIF-62 glass membrane not only far exceeds the Robeson upper bound, but also exceeds most of other pure polycrystalline MOF materials reported so far.

Other separation applications

Much ZIF research focuses on the separation of hydrogen and carbon dioxide because a well-studied ZIF, ZIF-8, has a very high separation factor for hydrogen and carbon dioxide mixtures. It is also very good for the separation of hydrocarbon mixtures, like the following:

In addition to gas separations, ZIF’s have the potential to separate components of biofuels, specifically, water and ethanol. Of all of the ZIF’s that have been tested, ZIF-8 shows high selectivity. ZIF’s have also shown potential in separating other alcohols, like propanol and butanol, from water. Typically, water and ethanol (or other alcohols) are separated using distillation, however ZIF’s offer a potential lower-energy separation option. [39]

Catalysis

ZIF’s also have great potential as heterogeneous catalysts; ZIF-8 has been shown to act as good catalysts for the transesterification of vegetable oils, the Friedel-Crafts acylation reaction between benzoyl chloride and anisole, and for the formation of carbonates. ZIF-8 nanoparticles can also be used to enhance the performance in the Knoevenagel condensation reaction between benzaldehyde and malononitrile. [40] ZIF’s have also been shown to work well in oxidation and epoxidation reactions; ZIF-9 has been shown to catalyze the aerobic oxidation of tetralin and the oxidation of many other small molecules. It can also catalyze reactions to produce hydrogen at room temperature, specifically the dehydrogenation of dimethylamine borane and Na BH4 hydrolysis.

The table below gives a more comprehensive list of ZIF’s that can act as catalysts for different organic reactions. [4]

ZIF MaterialAdditional MaterialsReaction (s) Catalyzed
ZIF-8gold nanoparticlesOxidation of CO

Oxidation of aldehyde groups

ZIF-8gold and silver core shell nanoparticlesReduction of 4-nitrophenol
ZIF-8gold, silver, and platinum nanoparticlesOxidation of CO

Hydrogenation of n-hexene

ZIF-8platinum nanoparticlesHydrogenation of alkene
ZIF-8platinum and titanium dioxide nanotubesDegradation of phenol
ZIF-8palladium nanoparticlesAminocarbonylation
ZIF-8iridium nanoparticlesHydrogenation of cyclohexene and phenylacetene
ZIF-8ruthenium nanoparticlesAsymmetric hydrogenation of acetophonone
ZIF-8iron oxide microspheresKnoevenagel condensation
ZIF-8Zn2GeO4 nanorodsConversion of CO2
ZIF-65Molybdenum OxideDegradation of methyl orange and orange II dyes

Sensing and electronic devices

ZIF’s are also good candidates for chemical sensors because of their tunable adsorbance properties. ZIF-8 exhibits sensitivity when exposed to the vapor of ethanol and water mixtures, and this response is dependent on the concentration of ethanol in the mixture. [41] Additionally, ZIF’s are attractive materials for matrices for biosensors, like electrochemical biosensors, for in-vivo electrochemical measurements. They also have potential applications as luminescent probes for the detection of metal ions and small molecules. ZIF-8 luminescence is highly sensitive to , and ions as well as acetone. ZIF nanoparticles can also sense fluorescently tagged single stranded pieces of DNA. [41]

Drug delivery

Because ZIF’s are porous, chemically stable, thermally stable, and tunable, they are potentially a platform for drug delivery and controlled drug release. ZIF-8 is very stable in water and aqueous sodium hydroxide solutions but decompose quickly in acidic solutions, indicating a pH sensitivity that could aid in the development of ZIF-based drug-release platforms. [41]

Comparing ZIFs with other compounds

ZIFs vs MOFs

While ZIFs are a subset of the MOF hybrids that combine organic and metal frameworks to create hybrid microporous and crystalline structures, they are much more restricted in their structure. Similar to MOFs, most ZIF properties are largely dependent on the properties of the metal clusters, ligands, and synthesis conditions in which they were created. [42]

Most ZIF alterations up to this point have involved changing the linkers bridging O2 anions and imidazolate-based ligands [36] - or combining two types of linkers to change bond angles or pore size due to limitations in synthesizing methods and production. [43] A large portion of changing linkers included adding functional groups with various polarities and symmetries to the imidazolate ligands to alter the ZIFs carbon dioxide adsorption ability without changing the transitional-metal cations. [44] Compare this to MOFs, which have a much larger degree of variety in the types of their building units.

Despite these similarities with other MOFs, ZIFs have significant properties that distinguish these structures as uniquely applicable to carbon capture processes. Because ZIFs tend to resemble the crystalline framework of zeolites, their thermal and chemical stability are higher than those of other MOFs, allowing them to work at a wider range in temperatures, making them suitable to chemical processes. [42]

Perhaps the most important difference is the ZIFs' hydrophobic properties and water stability. A main issue with zeolites and MOFs, to a certain extent, was their adsorption of water along with CO2. Water vapor is often found in carbon-rich exhaust gases, and MOFs would absorb the water, lowering the amount of CO2 required to reach saturation. [42] MOFs are also less stable in moist and oxygen-rich environments due to metal-oxygen bonds performing hydrolysis. ZIFs, however, have nearly identical performance in dry vs humid conditions, showing much higher CO2 selectivity over water, allowing the adsorbent to store more carbon before saturation is reached. [43]

ZIFs vs commercially available products

Even in comparison with other materials, the ZIFs most attractive quality is still its hydrophobic properties. When compared to ZIFs in dry conditions, activated carbon was nearly identical with its uptake capacity. [43] However, once the conditions were changed to wet, the activated carbon’s uptake was halved. When this saturation and regeneration tests were run at these conditions, ZIFs also showed minimal to no structural degradation, a good indication of the adsorbent’s re-usability. [43]

However, ZIFs tend to be expensive to synthesize. MOFs require synthesis methods with long reaction periods, high pressures, and high temperatures, which aren’t methods that are easy to scale-up. [42] ZIFs do tend to be more affordable than commercially available non-ZIF MOFs.

When combined with polymer-sorbent materials, research determined that hybrid polymer-ZIF sorbent membranes no longer following the upper bound of the Robeson plot, which is a plot of selectivity as a function of permeation for membrane gas separation. [36]

See also

Related Research Articles

Artificial photosynthesis is a chemical process that biomimics the natural process of photosynthesis. The term artificial photosynthesis is used loosely, refer to any scheme for capturing and storing energy from sunlight by producing a fuel, specifically a solar fuel. An advantage of artificial photosynthesis is that the solar energy can be immediately converted and stored. By contrast, using photovoltaic cells, sunlight is converted into electricity and then converted again into chemical energy for storage, with some necessary losses of energy associated with the second conversion. The byproducts of these reactions are environmentally friendly. Artificially photosynthesized fuel would be a carbon-neutral source of energy, which could be used for transportation or homes. The economics of artificial photosynthesis are not competitive.

<span class="mw-page-title-main">Membrane gas separation</span> Technology for splitting specific gases out of mixtures

Gas mixtures can be effectively separated by synthetic membranes made from polymers such as polyamide or cellulose acetate, or from ceramic materials.

<span class="mw-page-title-main">Hydrogen storage</span> Methods of storing hydrogen for later use

Several methods exist for storing hydrogen. These include mechanical approaches such as using high pressures and low temperatures, or employing chemical compounds that release H2 upon demand. While large amounts of hydrogen are produced by various industries, it is mostly consumed at the site of production, notably for the synthesis of ammonia. For many years hydrogen has been stored as compressed gas or cryogenic liquid, and transported as such in cylinders, tubes, and cryogenic tanks for use in industry or as propellant in space programs. Interest in using hydrogen for on-board storage of energy in zero-emissions vehicles is motivating the development of new methods of storage, more adapted to this new application. The overarching challenge is the very low boiling point of H2: it boils around 20.268 K (−252.882 °C or −423.188 °F). Achieving such low temperatures requires expending significant energy.

<span class="mw-page-title-main">Metal–organic framework</span> Class of chemical substance

Metal–organic frameworks (MOFs) are a class of porous polymers consisting of metal clusters coordinated to organic ligands to form one-, two-, or three-dimensional structures. The organic ligands included are sometimes referred to as "struts" or "linkers", one example being 1,4-benzenedicarboxylic acid (BDC).

<span class="mw-page-title-main">Omar M. Yaghi</span> Chemist

Omar M. Yaghi is the James and Neeltje Tretter Chair Professor of Chemistry at the University of California, Berkeley, an affiliate scientist at Lawrence Berkeley National Laboratory, the Founding Director of the Berkeley Global Science Institute, and an elected member of the US National Academy of Sciences as well as the German National Academy of Sciences Leopoldina.

<span class="mw-page-title-main">Electrocatalyst</span> Catalyst participating in electrochemical reactions

An electrocatalyst is a catalyst that participates in electrochemical reactions. Electrocatalysts are a specific form of catalysts that function at electrode surfaces or, most commonly, may be the electrode surface itself. An electrocatalyst can be heterogeneous such as a platinized electrode. Homogeneous electrocatalysts, which are soluble, assist in transferring electrons between the electrode and reactants, and/or facilitate an intermediate chemical transformation described by an overall half reaction. Major challenges in electrocatalysts focus on fuel cells.

Covalent organic frameworks (COFs) are a class of porous polymers that form two- or three-dimensional structures through reactions between organic precursors resulting in strong, covalent bonds to afford porous, stable, and crystalline materials. COFs emerged as a field from the overarching domain of organic materials as researchers optimized both synthetic control and precursor selection. These improvements to coordination chemistry enabled non-porous and amorphous organic materials such as organic polymers to advance into the construction of porous, crystalline materials with rigid structures that granted exceptional material stability in a wide range of solvents and conditions. Through the development of reticular chemistry, precise synthetic control was achieved and resulted in ordered, nano-porous structures with highly preferential structural orientation and properties which could be synergistically enhanced and amplified. With judicious selection of COF secondary building units (SBUs), or precursors, the final structure could be predetermined, and modified with exceptional control enabling fine-tuning of emergent properties. This level of control facilitates the COF material to be designed, synthesized, and utilized in various applications, many times with metrics on scale or surpassing that of the current state-of-the-art approaches.

<span class="mw-page-title-main">Solvothermal synthesis</span>

Solvothermal synthesis is a method of producing chemical compounds, in which a solvent containing reagents is put under high pressure and temperature in an autoclave. Many substances dissolve better in the same solvent in such conditions than at standard conditions, enabling reactions that would not otherwise occur and leading to new compounds or polymorphs. Solvothermal synthesis is very similar to the hydrothermal route; both are typically conducted in a stainless steel autoclave. The only difference being that the precursor solution is usually non-aqueous.

<span class="mw-page-title-main">Periodic graph (crystallography)</span>

In crystallography, a periodic graph or crystal net is a three-dimensional periodic graph, i.e., a three-dimensional Euclidean graph whose vertices or nodes are points in three-dimensional Euclidean space, and whose edges are line segments connecting pairs of vertices, periodic in three linearly independent axial directions. There is usually an implicit assumption that the set of vertices are uniformly discrete, i.e., that there is a fixed minimum distance between any two vertices. The vertices may represent positions of atoms or complexes or clusters of atoms such as single-metal ions, molecular building blocks, or secondary building units, while each edge represents a chemical bond or a polymeric ligand.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

<span class="mw-page-title-main">Two-dimensional polymer</span>

A two-dimensional polymer (2DP) is a sheet-like monomolecular macromolecule consisting of laterally connected repeat units with end groups along all edges. This recent definition of 2DP is based on Hermann Staudinger's polymer concept from the 1920s. According to this, covalent long chain molecules ("Makromoleküle") do exist and are composed of a sequence of linearly connected repeat units and end groups at both termini.

<span class="mw-page-title-main">Imidazolate</span> Ion

Imidazolate (C3H3N
2
) is the conjugate base of imidazole. It is a nucleophile and a strong base. The free anion has C2v symmetry. Imidazole has a pKa of 14.05, so the deprotonation of imidazole (C3H3N2H) requires a strong base.

Niveen M. Khashab is a Lebanese chemist and an associate Professor of chemical Sciences and engineering at King Abdullah University of Science and Technology in Saudi Arabia since 2009. She is a laureate of the 2017 L'Oréal-UNESCO Awards for Women in Science "for her contributions to innovative smart hybrid materials aimed at drug delivery and for developing new techniques to monitor intracellular antioxidant activity." She is also a fellow of the Royal Chemical Society, and a member of the American Chemical Society.

Some metal-organic frameworks (MOF) display large structural changes as a response to external stimuli, and such modifications of their structure can, in turn, lead to drastic changes in their physical and chemical properties. Such stimuli-responsive MOFs are generally referred to as a flexible metal-organic frameworks. They can also be called dynamic metal-organic framework, stimuli-responsive MOFs, multi-functional MOFs, or soft porous crystals.

<span class="mw-page-title-main">MIL-53</span> Chemical compound

MIL-53 belongs to the class of metal-organic framework (MOF) materials. The first synthesis and the name was established by the group of Gérard Férey in 2002. The MIL-53 structure consists of inorganic [M-OH] chains, which are connected to four neighboring inorganic chains by therephthalate-based linker molecules. Each metal center is octahedrally coordinated by six oxygen atoms. Four of these oxygen atoms originate from four different carboxylate groups and the remaining two oxygen atoms belong to two different μ-OH moieties, which bridge neighboring metal centers. The resulting framework structure contains one-dimensional diamond-shaped pores. Many research group have investigated the flexibility of the MIL-53 structure. This flexible behavior, during which the pore cross-section changes reversibly, was termed 'breathing.effect' and describes the ability of the MIL-53 framework to respond to external stimuli.

<span class="mw-page-title-main">HKUST-1</span>

HKUST-1, which is also called MOF-199, is a material in the class of metal-organic frameworks (MOFs). Metal-organic frameworks are crystalline materials, in which metals are linked by ligands to form repeating coordination motives extending in three dimensions. The HKUST-1 framework is built up of dimeric metal units, which are connected by benzene-1,3,5-tricarboxylate linker molecules. The paddlewheel unit is the commonly used structural motif to describe the coordination environment of the metal centers and also called secondary building unit (SBU) of the HKUST-1 structure. The paddlewheel is built up of four benzene-1,3,5-tricarboxylate linkers molecules, which bridge two metal centers. One water molecules is coordinated to each of the two metal centers at the axial position of the paddlewheel unit in the hydrated state, which is usually found if the material is handled in air. After an activation process, these water molecules can be removed and the coordination site at the metal atoms is left unoccupied. This unoccupied coordination site is called coordinatively unsaturated site (CUS) and can be accessed by other molecules.

Tomislav Friščić holds the Leverhulme International Professorship and Chair in Green and Sustainable chemistry at the University of Birmingham. His research focus is at the interface of green chemistry and materials science, developing solvent-free chemistry and mechanochemistry for the cleaner, efficient synthesis of molecules and materials, including organic solids such as pharmaceutical cocrystals, coordination polymers and Metal-Organic Frameworks (MOFs), and a wide range of organic targets such as active pharmaceutical ingredients. He is a Fellow of the Royal Society of Chemistry (RSC), member of the College of New Scholars, Artists and Scientists of the Royal Society of Canada and a corresponding member of the Croatian Academy of Sciences and Arts. He has served on the Editorial Board of CrystEngComm, the Early Career Board of the ACS journal ACS Sustainable Chemistry & Engineering, and was an Associate Editor for the journal Molecular Crystals & Liquid Crystals as well as for the journal Synthesis. He was a Topic Editor and Social Media Editor, and is currently a member of the Editorial Advisory Board of the journal Crystal Growth & Design published by the American Chemical Society (ACS). He famously has a dog named Zizi.

<span class="mw-page-title-main">Wendy Lee Queen</span> American chemist and material scientist

Wendy Lee Queen is an American chemist and material scientist. Her research interest focus on development design and production of hybrid organic/inorganic materials at the intersection of chemistry, chemical engineering and material sciences. As of 2020 she is a tenure-track assistant professor at the École polytechnique fédérale de Lausanne (EPFL) in Switzerland, where she directs the Laboratory for Functional Inorganic Materials.

<span class="mw-page-title-main">Transition metal imidazole complex</span>

A transition metal imidazole complex is a coordination complex that has one or more imidazole ligands. Complexes of imidazole itself are of little practical importance. In contrast, imidazole derivatives, especially histidine, are pervasive ligands in biology where they bind metal cofactors.

Carboxylate–based metal–organic frameworks are metal–organic frameworks that are based on organic molecules comprising carboxylate functional groups.

References

  1. Bennett, Thomas D.; Yue, Yuanzheng; Li, Peng; Qiao, Ang; Tao, Haizheng; Greaves, Neville G.; Richards, Tom; Lampronti, Giulio I.; Redfern, Simon A. T.; Blanc, Frédéric; Farha, Omar K. (2016-03-16). "Melt-Quenched Glasses of Metal–Organic Frameworks". Journal of the American Chemical Society. 138 (10): 3484–3492. doi: 10.1021/jacs.5b13220 . hdl: 2160/43170 . ISSN   0002-7863. PMID   26885940. S2CID   30519423.
  2. 1 2 Bennett, Thomas D.; Tan, Jin-Chong; Yue, Yuanzheng; Baxter, Emma; Ducati, Caterina; Terrill, Nick J.; Yeung, Hamish H. -M.; Zhou, Zhongfu; Chen, Wenlin; Henke, Sebastian; Cheetham, Anthony K. (November 2015). "Hybrid glasses from strong and fragile metal-organic framework liquids". Nature Communications. 6 (1): 8079. arXiv: 1409.3980 . Bibcode:2015NatCo...6.8079B. doi:10.1038/ncomms9079. ISSN   2041-1723. PMC   4560802 . PMID   26314784.
  3. Park, KS; et al. (2006). "Exceptional chemical and thermal stability of zeolitic imidazolate frameworks" (PDF). PNAS. 103 (27): 10186–10191. Bibcode:2006PNAS..10310186P. doi: 10.1073/pnas.0602439103 . PMC   1502432 . PMID   16798880.
  4. 1 2 Phan, A.; Doonan, C. J.; Uribe-Romo, F. J.; et al. (2010). "Synthesis, Structure, and Carbon Dioxide Capture Properties of Zeolitic Imidazolate Frameworks". Acc. Chem. Res. 43 (1): 58–67. doi:10.1021/ar900116g. PMID   19877580.
  5. Zhang, J.-P.; Zhang, Y.-B.; Lin, J.-B.; Chen, X.-M. (2012). "Metal Azolate Frameworks: From Crystal Engineering to Functional Materials". Chem. Rev. 112 (2): 1001–1033. doi:10.1021/cr200139g. PMID   21939178.
  6. Yaghi, Omar M. (January 2010). "Synthesis, Structure, and Carbon Dioxide Capture Properties of Zeolitic Imidazolate Frameworks" (PDF). Accounts of Chemical Research. 43 (1): 58–67. doi:10.1021/ar900116g. PMID   19877580.
  7. 1 2 Madsen, Rasmus S. K.; Qiao, Ang; Sen, Jishnu; Hung, Ivan; Chen, Kuizhi; Gan, Zhehong; Sen, Sabyasachi; Yue, Yuanzheng (2020-03-27). "Ultrahigh-field 67 Zn NMR reveals short-range disorder in zeolitic imidazolate framework glasses". Science. 367 (6485): 1473–1476. Bibcode:2020Sci...367.1473M. doi:10.1126/science.aaz0251. ISSN   0036-8075. PMC   7325427 . PMID   32217725.
  8. Park, Kyo Sung; Ni, Zheng; Côté, Adrien P.; et al. (2006-07-05). "Exceptional chemical and thermal stability of zeolitic imidazolate frameworks". Proceedings of the National Academy of Sciences. 103 (27): 10186–10191. Bibcode:2006PNAS..10310186P. doi: 10.1073/pnas.0602439103 . ISSN   0027-8424. PMC   1502432 . PMID   16798880.
  9. Hayashi, Hideki; Côté, Adrien P.; Furukawa, Hiroyasu; et al. (2007-07-01). "Zeolite A imidazolate frameworks". Nature Materials. 6 (7): 501–506. Bibcode:2007NatMa...6..501H. doi:10.1038/nmat1927. ISSN   1476-1122. PMID   17529969.
  10. Banerjee, Rahul; Phan, Anh; Wang, Bo; et al. (2008-02-15). "High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture". Science. 319 (5865): 939–943. Bibcode:2008Sci...319..939B. doi:10.1126/science.1152516. ISSN   0036-8075. PMID   18276887. S2CID   22210227.
  11. Wang, Bo; Côté, Adrien P.; Furukawa, Hiroyasu; et al. (2008-05-08). "Colossal cages in zeolitic imidazolate frameworks as selective carbon dioxide reservoirs". Nature. 453 (7192): 207–211. Bibcode:2008Natur.453..207W. doi: 10.1038/nature06900 . ISSN   0028-0836. PMID   18464739.
  12. Madhav, Dharmjeet; Malankowska, Magdalena; Coronas, Joaquin (2020-11-06). "Synthesis of nanoparticles of zeolitic imidazolate framework ZIF-94 using inorganic deprotonators". New Journal of Chemistry. 44 (46): 20449–20457. doi:10.1039/D0NJ04402D. ISSN   1144-0546. S2CID   229232268.
  13. Huang, Xiao-Chun; Lin, Yan-Yong; Zhang, Jie-Peng; Chen, Xiao-Ming (2006-02-27). "Ligand-Directed Strategy for Zeolite-Type Metal–Organic Frameworks: Zinc(II) Imidazolates with Unusual Zeolitic Topologies". Angewandte Chemie International Edition. 45 (10): 1557–1559. doi:10.1002/anie.200503778. ISSN   1521-3773. PMID   16440383.
  14. Cravillon, Janosch; Münzer, Simon; Lohmeier, Sven-Jare; et al. (2009-04-28). "Rapid Room-Temperature Synthesis and Characterization of Nanocrystals of a Prototypical Zeolitic Imidazolate Framework". Chemistry of Materials. 21 (8): 1410–1412. doi:10.1021/cm900166h. ISSN   0897-4756.
  15. He, Ming; Yao, Jianfeng; Li, Lunxi; et al. (2013-10-01). "Synthesis of Zeolitic Imidazolate Framework-7 in a Water/Ethanol Mixture and Its Ethanol-Induced Reversible Phase Transition". ChemPlusChem. 78 (10): 1222–1225. doi:10.1002/cplu.201300193. ISSN   2192-6506. PMID   31986784.
  16. Bennett, Thomas D.; Saines, Paul J.; Keen, David A.; et al. (2013-05-27). "Ball-Milling-Induced Amorphization of Zeolitic Imidazolate Frameworks (ZIFs) for the Irreversible Trapping of Iodine". Chemistry – A European Journal. 19 (22): 7049–7055. doi:10.1002/chem.201300216. ISSN   1521-3765. PMID   23576441.
  17. Pan, Yichang; Liu, Yunyang; Zeng, Gaofeng; et al. (2011-02-01). "Rapid synthesis of zeolitic imidazolate framework-8 (ZIF-8) nanocrystals in an aqueous system". Chemical Communications. 47 (7): 2071–3. doi:10.1039/C0CC05002D. ISSN   1364-548X. PMID   21206942.
  18. Tanaka, Shunsuke; Kida, Koji; Okita, Muneyuki; et al. (2012-10-05). "Size-controlled Synthesis of Zeolitic Imidazolate Framework-8 (ZIF-8) Crystals in an Aqueous System at Room Temperature". Chemistry Letters. 41 (10): 1337–1339. doi: 10.1246/cl.2012.1337 . ISSN   0366-7022.
  19. Kida, Koji; Okita, Muneyuki; Fujita, Kosuke; et al. (2013-02-07). "Formation of high crystalline ZIF-8 in an aqueous solution". CrystEngComm. 15 (9): 1794. doi:10.1039/C2CE26847G. ISSN   1466-8033.
  20. Yang, Tingxu; Chung, Tai-Shung (2013-04-23). "Room-temperature synthesis of ZIF-90 nanocrystals and the derived nano-composite membranes for hydrogen separation". Journal of Materials Chemistry A. 1 (19): 6081. doi:10.1039/C3TA10928C. ISSN   2050-7496.
  21. "Solvothermal synthesis of mixed-ligand metal–organic framework ZIF-78 with controllable size and morphology". ResearchGate. Retrieved 2017-05-01.
  22. Cravillon, Janosch; Schröder, Christian A.; Bux, Helge; et al. (2011-12-12). "Formate modulated solvothermal synthesis of ZIF-8 investigated using time-resolved in situ X-ray diffraction and scanning electron microscopy". CrystEngComm. 14 (2): 492–498. doi:10.1039/C1CE06002C. ISSN   1466-8033.
  23. Peralta, David; Chaplais, Gérald; Simon-Masseron, Angélique; Barthelet, Karin; Pirngruber, Gerhard D. (2012-05-01). "Synthesis and adsorption properties of ZIF-76 isomorphs" (PDF). Microporous and Mesoporous Materials. 153: 1–7. doi:10.1016/j.micromeso.2011.12.009.
  24. Yao, Jianfeng; He, Ming; Wang, Kun; et al. (2013-04-16). "High-yield synthesis of zeolitic imidazolate frameworks from stoichiometric metal and ligand precursor aqueous solutions at room temperature". CrystEngComm. 15 (18): 3601. doi:10.1039/C3CE27093A. ISSN   1466-8033.
  25. Shieh, Fa-Kuen; Wang, Shao-Chun; Leo, Sin-Yen; Wu, Kevin C.-W. (2013-08-19). "Water-Based Synthesis of Zeolitic Imidazolate Framework-90 (ZIF-90) with a Controllable Particle Size". Chemistry – A European Journal. 19 (34): 11139–11142. doi:10.1002/chem.201301560. ISSN   1521-3765. PMID   23832867.
  26. Nune, Satish K.; Thallapally, Praveen K.; Dohnalkova, Alice; et al. (2010-06-29). "Synthesis and properties of nano zeolitic imidazolate frameworks". Chemical Communications. 46 (27): 4878–80. doi:10.1039/C002088E. ISSN   1364-548X. PMID   20585703.
  27. Seoane, Beatriz; Zamaro, Juan M.; Tellez, Carlos; Coronas, Joaquin (2012-04-02). "Sonocrystallization of zeolitic imidazolate frameworks (ZIF-7, ZIF-8, ZIF-11 and ZIF-20)". CrystEngComm. 14 (9): 3103. doi:10.1039/C2CE06382D. hdl: 11336/53044 . ISSN   1466-8033.
  28. Cho, Hye-Young; Kim, Jun; Kim, Se-Na; Ahn, Wha-Seung (2013-03-15). "High yield 1-L scale synthesis of ZIF-8 via a sonochemical route". Microporous and Mesoporous Materials. 169: 180–184. doi:10.1016/j.micromeso.2012.11.012.
  29. Bux, Helge; Liang, Fangyi; Li, Yanshuo; et al. (2009). "Zeolitic Imidazolate Framework Membrane with Molecular Sieving Properties by Microwave-Assisted Solvothermal Synthesis". Journal of the American Chemical Society. 131 (44): 16000–16001. doi:10.1021/ja907359t. PMID   19842668.
  30. Hillman, Febrian; Zimmerman, John M.; Paek, Seung-Min; et al. (2017-03-28). "Rapid microwave-assisted synthesis of hybrid zeolitic–imidazolate frameworks with mixed metals and mixed linkers". Journal of Materials Chemistry A. 5 (13): 6090–6099. doi:10.1039/C6TA11170J. ISSN   2050-7496.
  31. Bennett, Thomas D.; Cao, Shuai; Tan, Jin Chong; et al. (2011). "Facile Mechanosynthesis of Amorphous Zeolitic Imidazolate Frameworks". Journal of the American Chemical Society. 133 (37): 14546–14549. doi:10.1021/ja206082s. PMID   21848328.
  32. Stassen, Ivo; Styles, Mark; Grenci, Gianluca; et al. (2016-03-01). "Chemical vapour deposition of zeolitic imidazolate framework thin films". Nature Materials. 15 (3): 304–310. Bibcode:2016NatMa..15..304S. doi:10.1038/nmat4509. ISSN   1476-1122. PMID   26657328.
  33. López-Domínguez, Pedro; López-Periago, Ana M.; Fernández-Porras, Francisco J.; et al. (2017-03-01). "Supercritical CO2 for the synthesis of nanometric ZIF-8 and loading with hyperbranched aminopolymers. Applications in CO2 capture". Journal of CO2 Utilization. 18: 147–155. doi:10.1016/j.jcou.2017.01.019.
  34. Pera-Titus, Marc (2014-01-22). "Porous Inorganic Membranes for CO2 Capture: Present and Prospects". Chemical Reviews. 114 (2): 1413–1492. doi:10.1021/cr400237k. ISSN   0009-2665. PMID   24299113.
  35. Venna, Surendar R.; Carreon, Moises A. (2010-01-13). "Highly Permeable Zeolite Imidazolate Framework-8 Membranes for CO2/CH4 Separation". Journal of the American Chemical Society. 132 (1): 76–78. doi:10.1021/ja909263x. ISSN   0002-7863. PMID   20014839.
  36. 1 2 3 Smit, Bernard; Reimer, Jeffrey A.; Oldenburg, Curtis M.; Bourg, Ian C. (2014). Introduction to Carbon Capture and Sequestration (1 ed.). Hackensack, NJ: Imperial College Press. ISBN   978-1-78326-328-8.
  37. Phan, Anh; Doonan, Christian J.; Uribe-Romo, Fernando J.; et al. (2010-01-19). "Synthesis, structure, and carbon dioxide capture properties of zeolitic imidazolate frameworks". Accounts of Chemical Research. 43 (1): 58–67. doi:10.1021/ar900116g. ISSN   1520-4898. PMID   19877580.
  38. Wang, Yuhan; Jin, Hua; Ma, Qiang; Mo, Kai; Mao, Haizhuo; Feldhoff, Armin; Cao, Xingzhong; Li, Yanshuo; Pan, Fusheng; Jiang, Zhongyi (2020-03-09). "A MOF Glass Membrane for Gas Separation". Angewandte Chemie. 132 (11): 4395–4399. Bibcode:2020AngCh.132.4395W. doi:10.1002/ange.201915807. ISSN   0044-8249. S2CID   226676528.
  39. Zhang, Kang; Nalaparaju, Anjaiah; Chen, Yifei; Jiang, Jianwen (2014-04-23). "Biofuel purification in zeolitic imidazolate frameworks: the significant role of functional groups". Physical Chemistry Chemical Physics. 16 (20): 9643–55. Bibcode:2014PCCP...16.9643Z. doi:10.1039/C4CP00739E. ISSN   1463-9084. PMID   24727907.
  40. Guan, Yebin; Shi, Juanjuan; Xia, Ming; et al. (2017-11-30). "Monodispersed ZIF-8 particles with enhanced performance for CO2 adsorption and heterogeneous catalysis". Applied Surface Science. 423: 349–353. Bibcode:2017ApSS..423..349G. doi:10.1016/j.apsusc.2017.06.183.
  41. 1 2 3 Chen, Binling; Yang, Zhuxian; Zhu, Yanqiu; Xia, Yongde (2014-09-23). "Zeolitic imidazolate framework materials: recent progress in synthesis and applications". Journal of Materials Chemistry A. 2 (40): 16811–16831. doi:10.1039/C4TA02984D. ISSN   2050-7496.
  42. 1 2 3 4 Basnayake, Sajani A.; Su, Jie; Zou, Xiadong; Balkus, Kenneth J. (2015-02-04). "Carbonate-Based Zeolitic Imidazolate Frame for Highly Selective CO2 Capture". Inorganic Chemistry. 54 (4): 1816–1821. doi:10.1021/ic5027174. PMID   25650775.
  43. 1 2 3 4 Nguyen, Nhung T. T.; Lo, Tien N. H.; Kim, Jaheon (2016-04-04). "Mixed-Metal Zeolitic Imidazolate Frameworks and their Selective Capture of Wet Carbon Dioxide over Methane" (PDF). Inorganic Chemistry. 55 (12): 6201–6207. doi:10.1021/acs.inorgchem.6b00814. PMID   27248714.
  44. Wang, Sibo; Wang, Xinchen (2015-12-08). "Imidazolium Ionic Liquids, Imidazolylidene Heterocyclic Carbenes, and Zeolitic Imidazolate Frameworks for CO2 Capture and Photochemical Reduction". Angewandte Chemie. 55 (7): 2308–2320. doi:10.1002/anie.201507145. PMID   26683833.