fbpx
Wikipedia

Metal–organic framework

Metal–organic frameworks (MOFs) are a class of porous polymers consisting of metal clusters (also known as SBUs) 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).

Synthesis of the MIL-101 MOF. Each green octahedron consists of one Cr atom in the center and six oxygen atoms (red balls) at the corners.
Electron micrograph of a MIL-101 crystal showing its supertetrahedra

More formally, a metal–organic framework is a potentially porous extended structure made from metal ions and organic linkers. An extended structure is a structure whose sub-units occur in a constant ratio and are arranged in a repeating pattern. MOFs are a subclass of coordination networks, which is a coordination compound extending, through repeating coordination entities, in one dimension, but with cross-links between two or more individual chains, loops, or spiro-links, or a coordination compound extending through repeating coordination entities in two or three dimensions. Coordination networks including MOFs further belong to coordination polymers, which is a coordination compound with repeating coordination entities extending in one, two, or three dimensions.[1] Most of the MOFs reported in the literature are crystalline compounds, but there are also amorphous MOFs,[2] and other disordered phases.[3]

In most cases for MOFs, the pores are stable during the elimination of the guest molecules (often solvents) and could be refilled with other compounds. Because of this property, MOFs are of interest for the storage of gases such as hydrogen and carbon dioxide. Other possible applications of MOFs are in gas purification, in gas separation, in water remediation,[4] in catalysis, as conducting solids and as supercapacitors.[5]

The synthesis and properties of MOFs constitute the primary focus of the discipline called reticular chemistry (from Latin reticulum, "small net").[6] In contrast to MOFs, covalent organic frameworks (COFs) are made entirely from light elements (H, B, C, N, and O) with extended structures.[7]

Structure edit

MOFs are composed of two main components: an inorganic metal cluster (often referred to as a secondary-building unit or SBU) and an organic molecule called a linker. For this reason, the materials are often referred to as hybrid organic-inorganic materials.[1] The organic units are typically mono-, di-, tri-, or tetravalent ligands.[8] The choice of metal and linker dictates the structure and hence properties of the MOF. For example, the metal's coordination preference influences the size and shape of pores by dictating how many ligands can bind to the metal, and in which orientation.

Classification of hybrid materials based on dimensionality[9]
Dimensionality of Inorganic
0 1 2 3
Dimensionality
of Organic
0 Molecular Complexes Hybrid Inorganic Chains Hybrid Inorganic Layers 3-D Inorganic Hybrids
1 Chain Coordination Polymers Mixed Inorganic-Organic Layers Mixed Inorganic-Organic 3-D Framework
2 Layered Coordination Polymer Mixed Inorganic-Organic 3-D Framework
3 3-D Coordination Polymers

To describe and organize the structures of MOFs, a system of nomenclature has been developed. Subunits of a MOF, called secondary building units (SBUs), can be described by topologies common to several structures. Each topology, also called a net, is assigned a symbol, consisting of three lower-case letters in bold. MOF-5, for example, has a pcu net.

Attached to the SBUs are bridging ligands. For MOFs, typical bridging ligands are di- and tricarboxylic acids. These ligands typically have rigid backbones. Examples are benzene-1,4-dicarboxylic acid (BDC or terephthalic acid), biphenyl-4,4-dicarboxylic acid (BPDC), and the tricarboxylic acid trimesic acid.

 
SBUs are often derived from basic zinc acetate structure, the acetates being replaced by rigid di- and tricarboxylates.

Synthesis edit

General synthesis edit

The study of MOFs has roots in coordination chemistry and solid-state inorganic chemistry, but it developed into a new field. In addition, MOFs are constructed from bridging organic ligands that remain intact throughout the synthesis.[10] Zeolite synthesis often makes use of a "template". Templates are ions that influence the structure of the growing inorganic framework. Typical templating ions are quaternary ammonium cations, which are removed later. In MOFs, the framework is templated by the SBU (secondary building unit) and the organic ligands.[11][12] A templating approach that is useful for MOFs intended for gas storage is the use of metal-binding solvents such as N,N-diethylformamide and water. In these cases, metal sites are exposed when the solvent is evacuated, allowing hydrogen to bind at these sites.[13]

Four developments were particularly important in advancing the chemistry of MOFs.[14] (1) The geometric principle of construction where metal-containing units were kept in rigid shapes. Early MOFs contained single atoms linked to ditopic coordinating linkers. The approach not only led to the identification of a small number of preferred topologies that could be targeted in designed synthesis, but was the central point to achieve a permanent porosity. (2) The use of the isoreticular principle where the size and the nature of a structure changes without changing its topology led to MOFs with ultrahigh porosity and unusually large pore openings. (3) Post- synthetic modification of MOFs increased their functionality by reacting organic units and metal-organic complexes with linkers. (4) Multifunctional MOFs incorporated multiple functionalities in a single framework.

Since ligands in MOFs typically bind reversibly, the slow growth of crystals often allows defects to be redissolved, resulting in a material with millimeter-scale crystals and a near-equilibrium defect density. Solvothermal synthesis is useful for growing crystals suitable to structure determination, because crystals grow over the course of hours to days. However, the use of MOFs as storage materials for consumer products demands an immense scale-up of their synthesis. Scale-up of MOFs has not been widely studied, though several groups have demonstrated that microwaves can be used to nucleate MOF crystals rapidly from solution.[15][16] This technique, termed "microwave-assisted solvothermal synthesis", is widely used in the zeolite literature,[10] and produces micron-scale crystals in a matter of seconds to minutes,[15][16] in yields similar to the slow growth methods.

Some MOFs, such as the mesoporous MIL-100(Fe),[17] can be obtained under mild conditions at room temperature and in green solvents (water, ethanol) through scalable synthesis methods.

A solvent-free synthesis of a range of crystalline MOFs has been described.[18] Usually the metal acetate and the organic proligand are mixed and ground up with a ball mill. Cu3(BTC)2 can be quickly synthesised in this way in quantitative yield. In the case of Cu3(BTC)2 the morphology of the solvent free synthesised product was the same as the industrially made Basolite C300. It is thought that localised melting of the components due to the high collision energy in the ball mill may assist the reaction. The formation of acetic acid as a by-product in the reactions in the ball mill may also help in the reaction having a solvent effect[19] in the ball mill. It has been shown that the addition of small quantities of ethanol for the mechanochemical synthesis of Cu3(BTC)2 significantly reduces the amounts of structural defects in the obtained material.[20]

A recent advancement in the solvent-free preparation of MOF films and composites is their synthesis by chemical vapor deposition. This process, MOF-CVD,[21] was first demonstrated for ZIF-8 and consists of two steps. In a first step, metal oxide precursor layers are deposited. In the second step, these precursor layers are exposed to sublimed ligand molecules, that induce a phase transformation to the MOF crystal lattice. Formation of water during this reaction plays a crucial role in directing the transformation. This process was successfully scaled up to an integrated cleanroom process, conforming to industrial microfabrication standards.[22]

Numerous methods have been reported for the growth of MOFs as oriented thin films. However, these methods are suitable only for the synthesis of a small number of MOF topologies. One such example being the vapor-assisted conversion (VAC) which can be used for the thin film synthesis of several UiO-type MOFs.[23]

High-throughput synthesis edit

High-throughput (HT) methods are a part of combinatorial chemistry and a tool for increasing efficiency. There are two synthetic strategies within the HT-methods: In the combinatorial approach, all reactions take place in one vessel, which leads to product mixtures. In the parallel synthesis, the reactions take place in different vessels. Furthermore, a distinction is made between thin films and solvent-based methods.[24]

Solvothermal synthesis can be carried out conventionally in a teflon reactor in a convection oven or in glass reactors in a microwave oven (high-throughput microwave synthesis). The use of a microwave oven changes, in part dramatically, the reaction parameters.

In addition to solvothermal synthesis, there have been advances in using supercritical fluid as a solvent in a continuous flow reactor. Supercritical water was first used in 2012 to synthesize copper and nickel-based MOFs in just seconds.[25] In 2020, supercritical carbon dioxide was used in a continuous flow reactor along the same time scale as the supercritical water-based method, but the lower critical point of carbon dioxide allowed for the synthesis of the zirconium-based MOF UiO-66.[26]

High-throughput solvothermal synthesis edit

In high-throughput solvothermal synthesis, a solvothermal reactor with (e.g.) 24 cavities for teflon reactors is used. Such a reactor is sometimes referred to as a multiclav. The reactor block or reactor insert is made of stainless steel and contains 24 reaction chambers, which are arranged in four rows. With the miniaturized teflon reactors, volumes of up to 2 mL can be used. The reactor block is sealed in a stainless steel autoclave; for this purpose, the filled reactors are inserted into the bottom of the reactor, the teflon reactors are sealed with two teflon films and the reactor top side is put on. The autoclave is then closed in a hydraulic press. The sealed solvothermal reactor can then be subjected to a temperature-time program. The reusable teflon film serves to withstand the mechanical stress, while the disposable teflon film seals the reaction vessels. After the reaction, the products can be isolated and washed in parallel in a vacuum filter device. On the filter paper, the products are then present separately in a so-called sample library and can subsequently be characterized by automated X-ray powder diffraction. The informations obtained are then used to plan further syntheses.[27]

Pseudomorphic replication edit

Pseudomorphic mineral replacement events occur whenever a mineral phase comes into contact with a fluid with which it is out of equilibrium. Re-equilibration will tend to take place to reduce the free energy and transform the initial phase into a more thermodynamically stable phase, involving dissolution and reprecipitation subprocesses.[28][29]

Inspired by such geological processes, MOF thin films can be grown through the combination of atomic layer deposition (ALD) of aluminum oxide onto a suitable substrate (e.g. FTO) and subsequent solvothermal microwave synthesis. The aluminum oxide layer serves both as an architecture-directing agent and as a metal source for the backbone of the MOF structure.[30] The construction of the porous 3D metal-organic framework takes place during the microwave synthesis, when the atomic layer deposited substrate is exposed to a solution of the requisite linker in a DMF/H2O 3:1 mixture (v/v) at elevated temperature. Analogous, Kornienko and coworkers described in 2015 the synthesis of a cobalt-porphyrin MOF (Al2(OH)2TCPP-Co; TCPP-H2=4,4,4″,4‴-(porphyrin-5,10,15,20-tetrayl)tetrabenzoate), the first MOF catalyst constructed for the electrocatalytic conversion of aqueous CO2 to CO.[31]

Post-synthetic modification edit

Although the three-dimensional structure and internal environment of the pores can be in theory controlled through proper selection of nodes and organic linking groups, the direct synthesis of such materials with the desired functionalities can be difficult due to the high sensitivity of MOF systems. Thermal and chemical sensitivity, as well as high reactivity of reaction materials, can make forming desired products challenging to achieve. The exchange of guest molecules and counter-ions and the removal of solvents allow for some additional functionality but are still limited to the integral parts of the framework.[32] The post-synthetic exchange of organic linkers and metal ions is an expanding area of the field and opens up possibilities for more complex structures, increased functionality, and greater system control.[32][33]

Ligand exchange edit

Post-synthetic modification techniques can be used to exchange an existing organic linking group in a prefabricated MOF with a new linker by ligand exchange or partial ligand exchange.[33][34] This exchange allows for the pores and, in some cases the overall framework of MOFs, to be tailored for specific purposes. Some of these uses include fine-tuning the material for selective adsorption, gas storage, and catalysis.[33][13] To perform ligand exchange prefabricated MOF crystals are washed with solvent and then soaked in a solution of the new linker. The exchange often requires heat and occurs on the time scale of a few days.[34] Post-synthetic ligand exchange also enables the incorporation of functional groups into MOFs that otherwise would not survive MOF synthesis, due to temperature, pH, or other reaction conditions, or hinder the synthesis itself by competition with donor groups on the loaning ligand.[33]

Metal exchange edit

Post-synthetic modification techniques can also be used to exchange an existing metal ion in a prefabricated MOF with a new metal ion by metal ion exchange. The complete metal metathesis from an integral part of the framework has been achieved without altering the framework or pore structure of the MOF. Similarly to post-synthetic ligand exchange, post-synthetic metal exchange is performed by washing prefabricated MOF crystals with solvent and then soaking the crystal in a solution of the new metal.[35] Post-synthetic metal exchange allows for a simple route to the formation of MOFs with the same framework yet different metal ions.[32]

Stratified synthesis edit

In addition to modifying the functionality of the ligands and metals themselves, post-synthetic modification can be used to expand upon the structure of the MOF. Using post-synthetic modification MOFs can be converted from a highly ordered crystalline material toward a heterogeneous porous material.[36] Using post-synthetic techniques, it is possible for the controlled installation of domains within a MOF crystal which exhibit unique structural and functional characteristics. Core-shell MOFs and other layered MOFs have been prepared where layers have unique functionalization but in most cases are crystallographically compatible from layer to layer.[37]

Open coordination sites edit

In some cases MOF metal nodes have an unsaturated environment, and it is possible to modify this environment using different techniques. If the size of the ligand matches the size of the pore aperture, it is possible to install additional ligands to existing MOF structure.[38][39] Sometimes metal nodes have a good binding affinity for inorganic species. For instance, it was shown that metal nodes can perform an extension, and create a bond with the uranyl cation.[40]

Composite materials edit

Another approach to increasing adsorption in MOFs is to alter the system in such a way that chemisorption becomes possible. This functionality has been introduced by making a composite material, which contains a MOF and a complex of platinum with activated carbon. In an effect known as hydrogen spillover, H2 can bind to the platinum surface through a dissociative mechanism which cleaves the hydrogen molecule into two hydrogen atoms and enables them to travel down the activated carbon onto the surface of the MOF. This innovation produced a threefold increase in the room-temperature storage capacity of a MOF; however, desorption can take upwards of 12 hours, and reversible desorption is sometimes observed for only two cycles.[41][42] The relationship between hydrogen spillover and hydrogen storage properties in MOFs is not well understood but may prove relevant to hydrogen storage.

Catalysis edit

 
Electron micrograph and structure of MIL-101. Color codes: red – oxygen, brown – carbon, blue – chromium.

MOFs have potential as heterogeneous catalysts, although applications have not been commercialized.[43] Their high surface area, tunable porosity, diversity in metal and functional groups make them especially attractive for use as catalysts. Zeolites are extraordinarily useful in catalysis.[44] Zeolites are limited by the fixed tetrahedral coordination of the Si/Al connecting points and the two-coordinated oxide linkers. Fewer than 200 zeolites are known. In contrast with this limited scope, MOFs exhibit more diverse coordination geometries, polytopic linkers, and ancillary ligands (F, OH, H2O among others). It is also difficult to obtain zeolites with pore sizes larger than 1 nm, which limits the catalytic applications of zeolites to relatively small organic molecules (typically no larger than xylenes). Furthermore, mild synthetic conditions typically employed for MOF synthesis allow direct incorporation of delicate functionalities into the framework structures. Such a process would not be possible with zeolites or other microporous crystalline oxide-based materials because of the harsh conditions typically used for their synthesis (e.g., calcination at high temperatures to remove organic templates). Metal–organic framework MIL-101 is one of the most used MOFs for catalysis incorporating different transition metals such as Cr.[45] However, the stability of some MOF photocatalysts in aqueous medium and under strongly oxidizing conditions is very low.[46][47]

Zeolites still cannot be obtained in enantiopure form, which precludes their applications in catalytic asymmetric synthesis, e.g., for the pharmaceutical, agrochemical, and fragrance industries. Enantiopure chiral ligands or their metal complexes have been incorporated into MOFs to lead to efficient asymmetric catalysts. Even some MOF materials may bridge the gap between zeolites and enzymes when they combine isolated polynuclear sites, dynamic host–guest responses, and a hydrophobic cavity environment. MOFs might be useful for making semi-conductors. Theoretical calculations show that MOFs are semiconductors or insulators with band gaps between 1.0 and 5.5 eV which can be altered by changing the degree of conjugation in the ligands indicating its possibility for being photocatalysts.

Design edit

 
Example of MOF-5

Like other heterogeneous catalysts, MOFs may allow for easier post-reaction separation and recyclability than homogeneous catalysts. In some cases, they also give a highly enhanced catalyst stability. Additionally, they typically offer substrate-size selectivity. Nevertheless, while clearly important for reactions in living systems, selectivity on the basis of substrate size is of limited value in abiotic catalysis, as reasonably pure feedstocks are generally available.

Metal ions or metal clusters edit

 
Example of zeolite catalyst

Among the earliest reports of MOF-based catalysis was the cyanosilylation of aldehydes by a 2D MOF (layered square grids) of formula Cd(4,4-bpy)2(NO3)2.[48] This investigation centered mainly on size- and shape-selective clathration. A second set of examples was based on a two-dimensional, square-grid MOF containing single Pd(II) ions as nodes and 2-hydroxypyrimidinolates as struts.[49] Despite initial coordinative saturation, the palladium centers in this MOF catalyze alcohol oxidation, olefin hydrogenation, and Suzuki C–C coupling. At a minimum, these reactions necessarily entail redox oscillations of the metal nodes between Pd(II) and Pd(0) intermediates accompanying by drastic changes in coordination number, which would certainly lead to destabilization and potential destruction of the original framework if all the Pd centers are catalytically active. The observation of substrate shape- and size-selectivity implies that the catalytic reactions are heterogeneous and are indeed occurring within the MOF. Nevertheless, at least for hydrogenation, it is difficult to rule out the possibility that catalysis is occurring at the surface of MOF-encapsulated palladium clusters/nanoparticles (i.e., partial decomposition sites) or defect sites, rather than at transiently labile, but otherwise intact, single-atom MOF nodes. "Opportunistic" MOF-based catalysis has been described for the cubic compound, MOF-5.[50] This material comprises coordinatively saturated Zn4O nodes and fully complexed BDC struts (see above for abbreviation); yet it apparently catalyzes the Friedel–Crafts tert-butylation of both toluene and biphenyl. Furthermore, para alkylation is strongly favored over ortho alkylation, a behavior thought to reflect the encapsulation of reactants by the MOF.

Functional struts edit

The porous-framework material [Cu3(btc)2(H2O)3], also known as HKUST-1,[51] contains large cavities having windows of diameter ~6 Å. The coordinated water molecules are easily removed, leaving open Cu(II) sites. Kaskel and co-workers showed that these Lewis acid sites could catalyze the cyanosilylation of benzaldehyde or acetone. The anhydrous version of HKUST-1 is an acid catalyst.[52] Compared to Brønsted vs. Lewis acid-catalyzed pathways, the product selectivity are distinctive for three reactions: isomerization of α-pinene oxide, cyclization of citronellal, and rearrangement of α-bromoacetals, indicating that indeed [Cu3(btc)2] functions primarily as a Lewis acid catalyst. The product selectivity and yield of catalytic reactions (e.g. cyclopropanation) have also been shown to be impacted by defective sites, such as Cu(I) or incompletely deprotonated carboxylic acid moities of the linkers.[20]

MIL-101, a large-cavity MOF having the formula [Cr3F(H2O)2O(BDC)3], is a cyanosilylation catalyst.[53] The coordinated water molecules in MIL-101 are easily removed to expose Cr(III) sites. As one might expect, given the greater Lewis acidity of Cr(III) vs. Cu(II), MIL-101 is much more active than HKUST-1 as a catalyst for the cyanosilylation of aldehydes. Additionally, the Kaskel group observed that the catalytic sites of MIL-101, in contrast to those of HKUST-1, are immune to unwanted reduction by benzaldehyde. The Lewis-acid-catalyzed cyanosilylation of aromatic aldehydes has also been carried out by Long and co-workers using a MOF of the formula Mn3[(Mn4Cl)3BTT8(CH3OH)10].[54] This material contains a three-dimensional pore structure, with the pore diameter equaling 10 Å. In principle, either of the two types of Mn(II) sites could function as a catalyst. Noteworthy features of this catalyst are high conversion yields (for small substrates) and good substrate-size-selectivity, consistent with channellocalized catalysis.

Encapsulated catalysts edit

The MOF encapsulation approach invites comparison to earlier studies of oxidative catalysis by zeolite-encapsulated Fe(porphyrin)[55] as well as Mn(porphyrin)[56] systems. The zeolite studies generally employed iodosylbenzene (PhIO), rather than TPHP as oxidant. The difference is likely mechanistically significant, thus complicating comparisons. Briefly, PhIO is a single oxygen atom donor, while TBHP is capable of more complex behavior. In addition, for the MOF-based system, it is conceivable that oxidation proceeds via both oxygen transfer from a manganese oxo intermediate as well as a manganese-initiated radical chain reaction pathway. Regardless of mechanism, the approach is a promising one for isolating and thereby stabilizing the porphyrins against both oxo-bridged dimer formation and oxidative degradation.[57]

Metal-free organic cavity modifiers edit

Most examples of MOF-based catalysis make use of metal ions or atoms as active sites. Among the few exceptions are two nickel- and two copper-containing MOFs synthesized by Rosseinsky and co-workers.[58] These compounds employ amino acids (L- or D-aspartate) together with dipyridyls as struts. The coordination chemistry is such that the amine group of the aspartate cannot be protonated by added HCl, but one of the aspartate carboxylates can. Thus, the framework-incorporated amino acid can exist in a form that is not accessible for the free amino acid. While the nickel-based compounds are marginally porous, on account of tiny channel dimensions, the copper versions are clearly porous. The Rosseinsky group showed that the carboxylic acids behave as Brønsted acidic catalysts, facilitating (in the copper cases) the ring-opening methanolysis of a small, cavity-accessible epoxide at up to 65% yield. Superior homogeneous catalysts exist however.

Kitagawa and co-workers have reported the synthesis of a catalytic MOF having the formula [Cd(4-BTAPA)2(NO3)2].[59] The MOF is three-dimensional, consisting of an identical catenated pair of networks, yet still featuring pores of molecular dimensions. The nodes consist of single cadmium ions, octahedrally ligated by pyridyl nitrogens. From a catalysis standpoint, however, the most interesting feature of this material is the presence of guest-accessible amide functionalities. The amides are capable of base-catalyzing the Knoevenagel condensation of benzaldehyde with malononitrile. Reactions with larger nitriles, however, are only marginally accelerated, implying that catalysis takes place chiefly within the material's channels rather than on its exterior. A noteworthy finding is the lack of catalysis by the free strut in homogeneous solution, evidently due to intermolecular H-bonding between bptda molecules. Thus, the MOF architecture elicits catalytic activity not otherwise encountered.

In an interesting alternative approach, Férey and coworkers were able to modify the interior of MIL-101 via Cr(III) coordination of one of the two available nitrogen atoms of each of several ethylenediamine molecules.[60] The free non-coordinated ends of the ethylenediamines were then used as Brønsted basic catalysts, again for Knoevenagel condensation of benzaldehyde with nitriles.

A third approach has been described by Kim Kimoon and coworkers.[61] Using a pyridine-functionalized derivative of tartaric acid and a Zn(II) source they were able to synthesize a 2D MOF termed POST-1. POST-1 possesses 1D channels whose cross sections are defined by six trinuclear zinc clusters and six struts. While three of the six pyridines are coordinated by zinc ions, the remaining three are protonated and directed toward the channel interior. When neutralized, the noncoordinated pyridyl groups are found to catalyze transesterification reactions, presumably by facilitating deprotonation of the reactant alcohol. The absence of significant catalysis when large alcohols are employed strongly suggests that the catalysis occurs within the channels of the MOF.

Achiral catalysis edit

 
Schematic Diagram for MOF Catalysis

Metals as catalytic sites edit

The metals in the MOF structure often act as Lewis acids. The metals in MOFs often coordinate to labile solvent molecules or counter ions which can be removed after activation of the framework. The Lewis acidic nature of such unsaturated metal centers can activate the coordinated organic substrates for subsequent organic transformations. The use of unsaturated metal centers was demonstrated in the cyanosilylation of aldehydes and imines by Fujita and coworkers in 2004.[62] They reported MOF of composition {[Cd(4,4-bpy)2(H2O)2] • (NO3)2 • 4H2O} which was obtained by treating linear bridging ligand 4,4-bipyridine (bpy) with Cd(NO3)2. The Cd(II) centers in this MOF possess a distorted octahedral geometry having four pyridines in the equatorial positions, and two water molecules in the axial positions to form a two-dimensional infinite network. On activation, two water molecules were removed leaving the metal centers unsaturated and Lewis acidic. The Lewis acidic character of metal center was tested on cyanosilylation reactions of imine where the imine gets attached to the Lewis-acidic metal centre resulting in higher electrophilicity of imines. For the cyanosilylation of imines, most of the reactions were complete within 1 h affording aminonitriles in quantitative yield. Kaskel and coworkers[63] carried out similar cyanosilylation reactions with coordinatively unsaturated metals in three-dimensional (3D) MOFs as heterogeneous catalysts. The 3D framework [Cu3(btc)2(H2O)3] (btc: benzene-1,3,5-tricarboxylate) (HKUST-1) used in this study was first reported by Williams et al.[64] The open framework of [Cu3(btc)2(H2O)3] is built from dimeric cupric tetracarboxylate units (paddle-wheels) with aqua molecules coordinating to the axial positions and btc bridging ligands. The resulting framework after removal of two water molecules from axial positions possesses porous channel. This activated MOF catalyzes the trimethylcyanosilylation of benzaldehydes with a very low conversion (<5% in 24 h) at 293 K. As the reaction temperature was raised to 313 K, a good conversion of 57% with a selectivity of 89% was obtained after 72 h. In comparison, less than 10% conversion was observed for the background reaction (without MOF) under the same conditions. But this strategy suffers from some problems like 1) the decomposition of the framework with increase of the reaction temperature due to the reduction of Cu(II) to Cu(I) by aldehydes; 2) strong solvent inhibition effect; electron donating solvents such as THF competed with aldehydes for coordination to the Cu(II) sites, and no cyanosilylation product was observed in these solvents; 3) the framework instability in some organic solvents. Several other groups have also reported the use of metal centres in MOFs as catalysts.[54][65] Again, electron-deficient nature of some metals and metal clusters makes the resulting MOFs efficient oxidation catalysts. Mori and coworkers[66] reported MOFs with Cu2 paddle wheel units as heterogeneous catalysts for the oxidation of alcohols. The catalytic activity of the resulting MOF was examined by carrying out alcohol oxidation with H2O2 as the oxidant. It also catalyzed the oxidation of primary alcohol, secondary alcohol and benzyl alcohols with high selectivity. Hill et al.[67] have demonstrated the sulfoxidation of thioethers using a MOF based on vanadium-oxo cluster V6O13 building units.

Functional linkers as catalytic sites edit

Functional linkers can be also utilized as catalytic sites. A 3D MOF {[Cd(4-BTAPA)2(NO3)2] • 6H2O • 2DMF} (4-BTAPA = 1,3,5-benzene tricarboxylic acid tris [N-(4-pyridyl)amide], DMF = N,N-dimethylformamide) constructed by tridentate amide linkers and cadmium salt catalyzes the Knoevenagel condensation reaction.[59] The pyridine groups on the ligand 4-BTAPA act as ligands binding to the octahedral cadmium centers, while the amide groups can provide the functionality for interaction with the incoming substrates. Specifically, the −NH moiety of the amide group can act as electron acceptor whereas the C=O group can act as electron donor to activate organic substrates for subsequent reactions. Ferey et al.[68] reported a robust and highly porous MOF [Cr33-O)F(H2O)2(BDC)3] (BDC: benzene-1,4-dicarboxylate) where instead of directly using the unsaturated Cr(III) centers as catalytic sites, the authors grafted ethylenediamine (ED) onto the Cr(III) sites. The uncoordinated ends of ED can act as base catalytic sites. ED-grafted MOF was investigated for Knoevenagel condensation reactions. A significant increase in conversion was observed for ED-grafted MOF compared to untreated framework (98% vs. 36%). Another example of linker modification to generate catalytic site is iodo-functionalized well-known Al-based MOFs (MIL-53 and DUT-5) and Zr-based MOFs (UiO-66 and UiO-67) for the catalytic oxidation of diols.[69][70]

Entrapment of catalytically active noble metal nanoparticles edit

The entrapment of catalytically active noble metals can be accomplished by grafting on functional groups to the unsaturated metal site on MOFs. Ethylenediamine (ED) has been shown to be grafted on the Cr metal sites and can be further modified to encapsulate noble metals such as Pd.[60] The entrapped Pd has similar catalytic activity as Pd/C in the Heck reaction. Ruthenium nanoparticles have catalytic activity in a number of reactions when entrapped in the MOF-5 framework.[71] This Ru-encapsulated MOF catalyzes oxidation of benzyl alcohol to benzaldehyde, although degradation of the MOF occurs. The same catalyst was used in the hydrogenation of benzene to cyclohexane. In another example, Pd nanoparticles embedded within defective HKUST-1 framework enable the generation of tunable Lewis basic sites.[72] Therefore, this multifunctional Pd/MOF composite is able to perform stepwise benzyl alcohol oxidation and Knoevenagel condensation.

Reaction hosts with size selectivity edit

MOFs might prove useful for both photochemical and polymerization reactions due to the tuneability of the size and shape of their pores. A 3D MOF {[Co(bpdc)3(bpy)] • 4DMF • H2O} (bpdc: biphenyldicarboxylate, bpy: 4,4-bipyridine) was synthesized by Li and coworkers.[73] Using this MOF photochemistry of o-methyl dibenzyl ketone (o-MeDBK) was extensively studied. This molecule was found to have a variety of photochemical reaction properties including the production of cyclopentanol. MOFs have been used to study polymerization in the confined space of MOF channels. Polymerization reactions in confined space might have different properties than polymerization in open space. Styrene, divinylbenzene, substituted acetylenes, methyl methacrylate, and vinyl acetate have all been studied by Kitagawa and coworkers as possible activated monomers for radical polymerization.[74][75] Due to the different linker size the MOF channel size could be tunable on the order of roughly 25 and 100 Å2. The channels were shown to stabilize propagating radicals and suppress termination reactions when used as radical polymerization sites.

Asymmetric catalysis edit

Several strategies exist for constructing homochiral MOFs. Crystallization of homochiral MOFs via self-resolution from achiral linker ligands is one of the way to accomplish such a goal. However, the resulting bulk samples contain both enantiomorphs and are racemic. Aoyama and coworkers[76] successfully obtained homochiral MOFs in the bulk from achiral ligands by carefully controlling nucleation in the crystal growth process. Zheng and coworkers[77] reported the synthesis of homochiral MOFs from achiral ligands by chemically manipulating the statistical fluctuation of the formation of enantiomeric pairs of crystals. Growing MOF crystals under chiral influences is another approach to obtain homochiral MOFs using achiral linker ligands. Rosseinsky and coworkers[78][79] have introduced a chiral coligand to direct the formation of homochiral MOFs by controlling the handedness of the helices during the crystal growth. Morris and coworkers[80] utilized ionic liquid with chiral cations as reaction media for synthesizing MOFs, and obtained homochiral MOFs. The most straightforward and rational strategy for synthesizing homochiral MOFs is, however, to use the readily available chiral linker ligands for their construction.

Homochiral MOFs with interesting functionalities and reagent-accessible channels edit

Homochiral MOFs have been made by Lin and coworkers using 2,2-bis(diphenylphosphino)-1,1-binaphthyl (BINAP) and 1,1-bi-2,2-naphthol (BINOL) as chiral ligands.[81] These ligands can coordinate with catalytically active metal sites to enhance the enantioselectivity. A variety of linking groups such as pyridine, phosphonic acid, and carboxylic acid can be selectively introduced to the 3,3, 4,4, and the 6,6 positions of the 1,1'-binaphthyl moiety. Moreover, by changing the length of the linker ligands the porosity and framework structure of the MOF can be selectively tuned.

Postmodification of homochiral MOFs edit

Lin and coworkers have shown that the postmodification of MOFs can be achieved to produce enantioselective homochiral MOFs for use as catalysts.[82] The resulting 3D homochiral MOF {[Cd3(L)3Cl6] • 4DMF • 6MeOH • 3H2O} (L=(R)-6,6'-dichloro-2,2'-dihydroxyl-1,1'-binaphthyl-bipyridine) synthesized by Lin was shown to have a similar catalytic efficiency for the diethylzinc addition reaction as compared to the homogeneous analogue when was pretreated by Ti(OiPr)4 to generate the grafted Ti- BINOLate species. The catalytic activity of MOFs can vary depending on the framework structure. Lin and others found that MOFs synthesized from the same materials could have drastically different catalytic activities depending on the framework structure present.[83]

Homochiral MOFs with precatalysts as building blocks edit

Another approach to construct catalytically active homochiral MOFs is to incorporate chiral metal complexes which are either active catalysts or precatalysts directly into the framework structures. For example, Hupp and coworkers[84] have combined a chiral ligand and bpdc (bpdc: biphenyldicarboxylate) with Zn(NO3)2 and obtained twofold interpenetrating 3D networks. The orientation of chiral ligand in the frameworks makes all Mn(III) sites accessible through the channels. The resulting open frameworks showed catalytic activity toward asymmetric olefin epoxidation reactions. No significant decrease of catalyst activity was observed during the reaction and the catalyst could be recycled and reused several times. Lin and coworkers[85] have reported zirconium phosphonate-derived Ru-BINAP systems. Zirconium phosphonate-based chiral porous hybrid materials containing the Ru(BINAP)(diamine)Cl2 precatalysts showed excellent enantioselectivity (up to 99.2% ee) in the asymmetric hydrogenation of aromatic ketones.

Biomimetic design and photocatalysis edit

Some MOF materials may resemble enzymes when they combine isolated polynuclear sites, dynamic host–guest responses, and hydrophobic cavity environment which are characteristics of an enzyme.[86] Some well-known examples of cooperative catalysis involving two metal ions in biological systems include: the diiron sites in methane monooxygenase, dicopper in cytochrome c oxidase, and tricopper oxidases which have analogy with polynuclear clusters found in the 0D coordination polymers, such as binuclear Cu2 paddlewheel units found in MOP-1[87][88] and [Cu3(btc)2] (btc=benzene-1,3,5-tricarboxylate) in HKUST-1 or trinuclear units such as {Fe3O(CO2)6} in MIL-88,[89] and IRMOP-51.[90] Thus, 0D MOFs have accessible biomimetic catalytic centers. In enzymatic systems, protein units show "molecular recognition", high affinity for specific substrates. It seems that molecular recognition effects are limited in zeolites by the rigid zeolite structure.[91] In contrast, dynamic features and guest-shape response make MOFs more similar to enzymes. Indeed, many hybrid frameworks contain organic parts that can rotate as a result of stimuli, such as light and heat.[92] The porous channels in MOF structures can be used as photocatalysis sites. In photocatalysis, the use of mononuclear complexes is usually limited either because they only undergo single-electron process or from the need for high-energy irradiation. In this case, binuclear systems have a number of attractive features for the development of photocatalysts.[93] For 0D MOF structures, polycationic nodes can act as semiconductor quantum dots which can be activated upon photostimuli with the linkers serving as photon antennae.[94] Theoretical calculations show that MOFs are semiconductors or insulators with band gaps between 1.0 and 5.5 eV which can be altered by changing the degree of conjugation in the ligands.[95] Experimental results show that the band gap of IRMOF-type samples can be tuned by varying the functionality of the linker.[96] An integrated MOF nanozyme was developed for anti-inflammation therapy.[97]

Mechanical properties edit

Implementing MOFs in industry necessitates a thorough understanding of the mechanical properties since most processing techniques (e.g. extrusion and pelletization) expose the MOFs to substantial mechanical compressive stresses.[98] The mechanical response of porous structures is of interest as these structures can exhibit unusual response to high pressures. While zeolites (microporous, aluminosilicate minerals) can give some insights into the mechanical response of MOFs, the presence of organic linkers as opposed to zeolites, makes for novel mechanical responses.[99] MOFs are very structurally diverse meaning that it is challenging to classify all of their mechanical properties. Additionally, variability in MOFs from batch to batch and extreme experimental conditions (diamond anvil cells) mean that experimental determination of mechanical response to loading is limited, however many computational models have been made to determine structure-property relationships. Main MOF systems that have been explored are zeolitic imidazolate frameworks (ZIFs), Carboxylate MOFs, Zirconium-based MOFs, among others.[99] Generally, the MOFs undergo three processes under compressive loading (which is relevant in a processing context): amorphization, hyperfilling, and/or pressure induced phase transitions. During amorphization linkers buckle and the internal porosity within the MOF collapses. During hyperfilling the MOF which is being hydrostatically compressed in a liquid (typically solvent) will expand rather than contract due to a filling of pores with the loading media. Finally, pressure induced phase transitions where the structure of the crystal is altered during the loading are possible. The response of the MOF is predominantly dependent on the linker species and the inorganic nodes.

Zeolitic imidazolate frameworks (ZIFs) edit

Several different mechanical phenomena have been observed in zeolitic imidazolate frameworks (ZIFs), the most widely studied MOF for mechanical properties due to their many similarities to zeolites.[99] General trends for the ZIF family are the tendency of the Young's modulus and hardness of the ZIFs to decrease as the accessible pore volume increases.[100] The bulk moduli of ZIF-62 series increase with the increasing of benzoimidazolate (bim) concentration. ZIF-62 shows a continuous phase transition from open pore (op) to close pore (cp) phase when bim concentration is over 0.35 per formular unit. The accessible pore size and volume of ZIF-62-bim0.35 can be precisely tuned by applying adequate pressures.[101] Another study has shown that under hydrostatic loading in solvent the ZIF-8 material expands as opposed to contracting. This is a result of hyperfilling of the internal pores with solvent.[102] A computational study demonstrated that ZIF-4 and ZIF-8 materials undergo a shear softening mechanism with amorphizing (at ~ 0.34 GPa) of the material under hydrostatic loading, while still possessing a bulk modulus on the order of 6.5 GPa.[103][104] Additionally, the ZIF-4 and ZIF-8 MOFs are subject to many pressure dependent phase transitions.[100][105]

Carboxylate-based MOFs edit

Carboxylate MOFs come in many forms and have been widely studied. Herein, HKUST-1, MOF-5, and the MIL series are discussed as representative examples of the carboxylate MOF class.

HKUST-1 edit

HKUST-1 consists of a dimeric Cu-paddlewheel that possesses two pore types. Under pelletization MOFs such as HKUST-1 exhibit a pore collapse.[106] Although most carboxylate MOFs have a negative thermal expansion (they densify during heating), it was found that the hardness and Young's moduli unexpectedly decrease with increasing temperature from disordering of linkers.[107] It was also found computationally that a more mesoporous structure has a lower bulk modulus. However, an increased bulk modulus was observed in systems with a few large mesopores versus many small mesopores even though both pore size distributions had the same total pore volume.[108] The HKUST-1 shows a similar, "hyperfilling" phenomenon to the ZIF structures under hydrostatic loading.[109]

MOF-5 edit

MOF-5 has tetranuclear nodes in an octahedral configuration with an overall cubic structure. MOF-5 has a compressibility and Young's modulus (~14.9 GPa) comparable to wood, which was confirmed with density functional theory (DFT) and nanoindentation.[110][111] While it was shown that the MOF-5 can demonstrate the hyperfilling phenomenon within a loading media of solvent, these MOFs are very sensitive to pressure and undergo amorphization/pressure induced pore collapse at a pressure of 3.5 MPa when there is no fluid in the pores.[112]

MIL-53 edit

 
MIL-53 MOF wine rack structure illustrating potential for anisotropy in loading

MIL-53 MOFs possess a "wine rack" structure. These MOFs have been explored for anisotropy in Young's modulus due to the flexibility of loading, and the potential for negative linear compressibility when compressing in one direction, due to the ability of the wine rack opening during loading.[113][114]

Zirconium-based MOFs edit

 
Electron micrograph and structure of UiO-66. Color codes: red – oxygen, brown – carbon, green – zirconium, gray – hydrogen.

Zirconium-based MOFs such as UiO-66 are a very robust class of MOFs (attributed to strong hexanuclear   metallic nodes) with increased resistance to heat, solvents, and other harsh conditions, which makes them of interest in terms of mechanical properties.[115] Determinations of shear modulus and pelletization have shown that the UiO-66 MOFs are very mechanically robust and have high tolerance for pore collapse when compared to ZIFs and carboxylate MOFs.[106][116] Although the UiO-66 MOF shows increased stability under pelletization, the UiO-66 MOFs amorphized fairly rapidly under ball milling conditions due to destruction of linker coordinating inorganic nodes.[117]

Applications edit

Hydrogen storage edit

Molecular hydrogen has the highest specific energy of any fuel. However unless the hydrogen gas is compressed, its volumetric energy density is very low, so the transportation and storage of hydrogen require energy-intensive compression and liquefaction processes.[118][119][120] Therefore, development of new hydrogen storage methods which decrease the concomitant pressure required for practical volumetric energy density is an active area of research.[118] MOFs attract attention as materials for adsorptive hydrogen storage because of their high specific surface areas and surface to volume ratios, as well as their chemically tunable structures.[41]

Compared to an empty gas cylinder, a MOF-filled gas cylinder can store more hydrogen at a given pressure because hydrogen molecules adsorb to the surface of MOFs. Furthermore, MOFs are free of dead-volume, so there is almost no loss of storage capacity as a result of space-blocking by non-accessible volume.[8] Also, because the hydrogen uptake is based primarily on physisorption, many MOFs have a fully reversible uptake-and-release behavior. No large activation barriers are required when liberating the adsorbed hydrogen.[8] The storage capacity of a MOF is limited by the liquid-phase density of hydrogen because the benefits provided by MOFs can be realized only if the hydrogen is in its gaseous state.[8]

The extent to which a gas can adsorb to a MOF's surface depends on the temperature and pressure of the gas. In general, adsorption increases with decreasing temperature and increasing pressure (until a maximum is reached, typically 20–30 bar, after which the adsorption capacity decreases).[8][41][120] However, MOFs to be used for hydrogen storage in automotive fuel cells need to operate efficiently at ambient temperature and pressures between 1 and 100 bar, as these are the values that are deemed safe for automotive applications.[41]

 
MOF-177

The U.S. Department of Energy (DOE) has published a list of yearly technical system targets for on-board hydrogen storage for light-duty fuel cell vehicles which guide researchers in the field (5.5 wt %/40 g L−1 by 2017; 7.5 wt %/70 g L−1 ultimate).[121] Materials with high porosity and high surface area such as MOFs have been designed and synthesized in an effort to meet these targets. These adsorptive materials generally work via physical adsorption rather than chemisorption due to the large HOMO-LUMO gap and low HOMO energy level of molecular hydrogen. A benchmark material to this end is MOF-177 which was found to store hydrogen at 7.5 wt % with a volumetric capacity of 32 g L−1 at 77 K and 70 bar.[122] MOF-177 consists of [Zn4O]6+ clusters interconnected by 1,3,5-benzenetribenzoate organic linkers and has a measured BET surface area of 4630 m2 g−1. Another exemplary material is PCN-61 which exhibits a hydrogen uptake of 6.24 wt % and 42.5 g L−1 at 35 bar and 77 K and 2.25 wt % at atmospheric pressure.[123] PCN-61 consists of [Cu2]4+ paddle-wheel units connected through 5,5,5-benzene-1,3,5-triyltris(1-ethynyl-2-isophthalate) organic linkers and has a measured BET surface area of 3000 m2 g−1. Despite these promising MOF examples, the classes of synthetic porous materials with the highest performance for practical hydrogen storage are activated carbon and covalent organic frameworks (COFs).[124]

Design principles edit

Practical applications of MOFs for hydrogen storage are met with several challenges. For hydrogen adsorption near room temperature, the hydrogen binding energy would need to be increased considerably.[41] Several classes of MOFs have been explored, including carboxylate-based MOFs, heterocyclic azolate-based MOFs, metal-cyanide MOFs, and covalent organic frameworks. Carboxylate-based MOFs have by far received the most attention because

  1. they are either commercially available or easily synthesized,
  2. they have high acidity (pKa ~ 4) allowing for facile in situ deprotonation,
  3. the metal-carboxylate bond formation is reversible, facilitating the formation of well-ordered crystalline MOFs, and
  4. the bridging bidentate coordination ability of carboxylate groups favors the high degree of framework connectivity and strong metal-ligand bonds necessary to maintain MOF architecture under the conditions required to evacuate the solvent from the pores.[41]

The most common transition metals employed in carboxylate-based frameworks are Cu2+ and Zn2+. Lighter main-group metal ions have also been explored. Be12(OH)12(BTB)4, the first successfully synthesized and structurally characterized MOF consisting of a light main group metal ion, shows high hydrogen storage capacity, but it is too toxic to be employed practically.[125] There is considerable effort being put forth in developing MOFs composed of other light main group metal ions, such as magnesium in Mg4(BDC)3.[41]

The following is a list of several MOFs that are considered to have the best properties for hydrogen storage as of May 2012 (in order of decreasing hydrogen storage capacity).[41] While each MOF described has its advantages, none of these MOFs reach all of the standards set by the U.S. DOE. Therefore, it is not yet known whether materials with high surface areas, small pores, or di- or trivalent metal clusters produce the most favorable MOFs for hydrogen storage.[8]


MOFs that are considered to have the best properties for hydrogen storage as of May 2012
Name Formula Structure Hydrogen storage capacity Comments
MOF-210[126] Zn4O(BTE)(BPDC), where BTE3−=4,4,4″-[benzene-1,3,5-triyl-tris(ethyne-2,1-diyl)]tribenzoate and BPDC2−=biphenyl-4,4-dicarboxylate At 77 K: 8.6 excess wt% (17.6 total wt%) at 77 K and 80 bar. 44 total g H2/L at 80 bar and 77 K.[126]
At 298 K: 2.90 delivery wt% (1–100 bar) at 298 K and 100 bar.
MOF-200[126] Zn4O(BBC)2, where BBC3−=4,4,4″-[benzene-1,3,5-triyl-tris(benzene-4,1-diyl)]tribenzoate At 77 K: 7.4 excess wt% (16.3 total wt%) at 77 K and 80 bar. 36 total g H2/L at 80 bar and 77 K.[126]
At 298 K: 3.24 delivery wt% (1–100 bar) at 298 K and 100 bar.
MOF-177[127] Zn4O(BTB)2, where BTB3−=1,3,5-benzenetribenzoate Tetrahedral [Zn4O]6+ units are linked by large, triangular tricarboxylate ligands. Six diamond-shaped channels (upper) with diameter of 10.8 Å surround a pore containing eclipsed BTB3− moieties (lower). 7.1 wt% at 77 K and 40 bar; 11.4 wt% at 78 bar and 77 K. MOF-177 has larger pores, so hydrogen is compressed within holes rather than adsorbed to the surface. This leads to higher total gravimetric uptake but lower volumetric storage density compared to MOF-5.[41]
MOF-5[128] Zn4O(BDC)3, where BDC2−=1,4-benzenedicarboxylate Square openings are either 13.8 or 9.2 Å depending on the orientation of the aromatic rings. 7.1 wt% at 77 K and 40 bar; 10 wt% at 100 bar; volumetric storage density of 66 g/L. MOF-5 has received much attention from theorists because of the partial charges on the MOF surface, which provide a means of strengthening the binding hydrogen through dipole-induced intermolecular interactions; however, MOF-5 has poor performance at room temperature (9.1 g/L at 100 bar).[41]
Mn3[(Mn4Cl)3(BTT)8]2, where H3BTT=benzene-1,3,5-tris(1H-tetrazole)[129] Consists of truncated octahedral cages that share square faces, leading to pores of about 10 Å in diameter. Contains open Mn2+ coordination sites. 60 g/L at 77 K and 90 bar; 12.1 g/L at 90 bar and 298 K. This MOF is the first demonstration of open metal coordination sites increasing strength of hydrogen adsorption, which results in improved performance at 298 K. It has relatively strong metal-hydrogen interactions, attributed to a spin state change upon binding or to a classical Coulombic attraction.[41]
Cu3(BTC)2(H2O)3, where H3BTC=1,3,5-benzenetricarboxylic acid[130] Consists of octahedral cages that share paddlewheel units to define pores of about 9.8 Å in diameter. High hydrogen uptake is attributed to overlapping attractive potentials from multiple copper paddle-wheel units: each Cu(II) center can potentially lose a terminal solvent ligand bound in the axial position, providing an open coordination site for hydrogen binding.[41]

Structural impacts on hydrogen storage capacity edit

To date, hydrogen storage in MOFs at room temperature is a battle between maximizing storage capacity and maintaining reasonable desorption rates, while conserving the integrity of the adsorbent framework (e.g. completely evacuating pores, preserving the MOF structure, etc.) over many cycles. There are two major strategies governing the design of MOFs for hydrogen storage:

1) to increase the theoretical storage capacity of the material, and
2) to bring the operating conditions closer to ambient temperature and pressure. Rowsell and Yaghi have identified several directions to these ends in some of the early papers.[131][132]
Surface area edit

The general trend in MOFs used for hydrogen storage is that the greater the surface area, the more hydrogen the MOF can store. High surface area materials tend to exhibit increased micropore volume and inherently low bulk density, allowing for more hydrogen adsorption to occur.[41]

Hydrogen adsorption enthalpy edit

High hydrogen adsorption enthalpy is also important. Theoretical studies have shown that 22–25 kJ/mol interactions are ideal for hydrogen storage at room temperature, as they are strong enough to adsorb H2, but weak enough to allow for quick desorption.[133] The interaction between hydrogen and uncharged organic linkers is not this strong, and so a considerable amount of work has gone in synthesis of MOFs with exposed metal sites, to which hydrogen adsorbs with an enthalpy of 5–10 kJ/mol. Synthetically, this may be achieved by using ligands whose geometries prevent the metal from being fully coordinated, by removing volatile metal-bound solvent molecules over the course of synthesis, and by post-synthetic impregnation with additional metal cations.[13][129] (C5H5)V(CO)3(H2) and Mo(CO)5(H2) are great examples of increased binding energy due to open metal coordination sites;[134] however, their high metal-hydrogen bond dissociation energies result in a tremendous release of heat upon loading with hydrogen, which is not favorable for fuel cells.[41] MOFs, therefore, should avoid orbital interactions that lead to such strong metal-hydrogen bonds and employ simple charge-induced dipole interactions, as demonstrated in Mn3[(Mn4Cl)3(BTT)8]2.

An association energy of 22–25 kJ/mol is typical of charge-induced dipole interactions, and so there is interest in the use of charged linkers and metals.[41] The metal–hydrogen bond strength is diminished in MOFs, probably due to charge diffusion, so 2+ and 3+ metal ions are being studied to strengthen this interaction even further. A problem with this approach is that MOFs with exposed metal surfaces have lower concentrations of linkers; this makes them difficult to synthesize, as they are prone to framework collapse. This may diminish their useful lifetimes as well.[13][41]

Sensitivity to airborne moisture edit

MOFs are frequently sensitive to moisture in the air. In particular, IRMOF-1 degrades in the presence of small amounts of water at room temperature. Studies on metal analogues have unraveled the ability of metals other than Zn to stand higher water concentrations at high temperatures.[135][136]

To compensate for this, specially constructed storage containers are required, which can be costly. Strong metal-ligand bonds, such as in metal-imidazolate, -triazolate, and -pyrazolate frameworks, are known to decrease a MOF's sensitivity to air, reducing the expense of storage.[137]

Pore size edit

In a microporous material where physisorption and weak van der Waals forces dominate adsorption, the storage density is greatly dependent on the size of the pores. Calculations of idealized homogeneous materials, such as graphitic carbons and carbon nanotubes, predict that a microporous material with 7 Å-wide pores will exhibit maximum hydrogen uptake at room temperature. At this width, exactly two layers of hydrogen molecules adsorb on opposing surfaces with no space left in between.[41][138] 10 Å-wide pores are also of ideal size because at this width, exactly three layers of hydrogen can exist with no space in between.[41] (A hydrogen molecule has a bond length of 0.74 Å with a van der Waals radius of 1.17 Å for each atom; therefore, its effective van der Waals length is 3.08 Å.) [139]

Structural defects edit

Structural defects also play an important role in the performance of MOFs. Room-temperature hydrogen uptake via bridged spillover is mainly governed by structural defects, which can have two effects:

1) a partially collapsed framework can block access to pores; thereby reducing hydrogen uptake, and
2) lattice defects can create an intricate array of new pores and channels causing increased hydrogen uptake.[140]

Structural defects can also leave metal-containing nodes incompletely coordinated. This enhances the performance of MOFs used for hydrogen storage by increasing the number of accessible metal centers.[141] Finally, structural defects can affect the transport of phonons, which affects the thermal conductivity of the MOF.[142]

Hydrogen adsorption edit

Adsorption is the process of trapping atoms or molecules that are incident on a surface; therefore the adsorption capacity of a material increases with its surface area. In three dimensions, the maximum surface area will be obtained by a structure which is highly porous, such that atoms and molecules can access internal surfaces. This simple qualitative argument suggests that the highly porous metal-organic frameworks (MOFs) should be excellent candidates for hydrogen storage devices.

Adsorption can be broadly classified as being one of two types: physisorption or chemisorption. Physisorption is characterized by weak van der Waals interactions, and bond enthalpies typically less than 20 kJ/mol. Chemisorption, alternatively, is defined by stronger covalent and ionic bonds, with bond enthalpies between 250 and 500 kJ/mol. In both cases, the adsorbate atoms or molecules (i.e. the particles which adhere to the surface) are attracted to the adsorbent (solid) surface because of the surface energy that results from unoccupied bonding locations at the surface. The degree of orbital overlap then determines if the interactions will be physisorptive or chemisorptive.[143]

Adsorption of molecular hydrogen in MOFs is physisorptive. Since molecular hydrogen only has two electrons, dispersion forces are weak, typically 4–7 kJ/mol, and are only sufficient for adsorption at temperatures below 298 K.[41]

A complete explanation of the H2 sorption mechanism in MOFs was achieved by statistical averaging in the grand canonical ensemble, exploring a wide range of pressures and temperatures.[144][145]

Determining hydrogen storage capacity edit

Two hydrogen-uptake measurement methods are used for the characterization of MOFs as hydrogen storage materials: gravimetric and volumetric. To obtain the total amount of hydrogen in the MOF, both the amount of hydrogen absorbed on its surface and the amount of hydrogen residing in its pores should be considered. To calculate the absolute absorbed amount (Nabs), the surface excess amount (Nex) is added to the product of the bulk density of hydrogen (ρbulk) and the pore volume of the MOF (Vpore), as shown in the following equation:[146]

 
Gravimetric method edit

The increased mass of the MOF due to the stored hydrogen is directly calculated by a highly sensitive microbalance.[146] Due to buoyancy, the detected mass of adsorbed hydrogen decreases again when a sufficiently high pressure is applied to the system because the density of the surrounding gaseous hydrogen becomes more and more important at higher pressures. Thus, this "weight loss" has to be corrected using the volume of the MOF's frame and the density of hydrogen.[147]

Volumetric method edit

The changing of amount of hydrogen stored in the MOF is measured by detecting the varied pressure of hydrogen at constant volume.[146] The volume of adsorbed hydrogen in the MOF is then calculated by subtracting the volume of hydrogen in free space from the total volume of dosed hydrogen.[148]

Other methods of hydrogen storage edit

There are six possible methods that can be used for the reversible storage of hydrogen with a high volumetric and gravimetric density, which are summarized in the following table, (where ρm is the gravimetric density, ρv is the volumetric density, T is the working temperature, and P is the working pressure):[149]

Storage method ρm (mass%) ρv (kg H2/m3) T (°C) P (bar) Remarks
High-pressure gas cylinders 13 <40 25 800 Compressed H2 gas in lightweight composite cylinder
Liquid hydrogen in cryogenic tanks size-dependent 70.8 −252 1 Liquid H2; continuous loss of a few percent of H2 per day at 25 °C
Adsorbed hydrogen ~2 20 −80 100 Physisorption of H2 on materials
Adsorbed on interstitial sites in a host metal ~2 150 25 1 Atomic hydrogen reversibly adsorbs in host metals
Complex compounds <18 150 >100 1 Complex compounds ([AlH4] or [BH4]); desorption at elevated temperature, adsorption at high pressures
Metal and complexes together with water <40 >150 25 1 Chemical oxidation of metals with water and liberation of H2

Of these, high-pressure gas cylinders and liquid hydrogen in cryogenic tanks are the least practical ways to store hydrogen for the purpose of fuel due to the extremely high pressure required for storing hydrogen gas or the extremely low temperature required for storing hydrogen liquid. The other methods are all being studied and developed extensively.[149]

Electrocatalysis edit

The high surface area and atomic metal sites feature of MOFs make them a suitable candidate for electrocatalysts, especially energy-related ones. Until now, MOFs have been used extensively as electrocatalyst for water splitting (hydrogen evolution reaction and oxygen evolution reaction), carbon dioxide reduction, and oxygen reduction reaction.[150] Currently there are two routes: 1. Using MOFs as precursors to prepare electrocatalysts with carbon support.[151] 2. Using MOFs directly as electrocatalysts.[152][153] However, some results have shown that some MOFs are not stable under electrochemical environment.[154] The electrochemical conversion of MOFs during electrocatalysis may produce the real catalyst materials, and the MOFs are precatalysts under such conditions.[155] Therefore, claiming MOFs as the electrocatalysts requires in situ techniques coupled with electrocatalysis.

Biological imaging and sensing edit

 
MOF-76 crystal, where oxygen, carbon, and lanthanide atoms are represented by maroon, black, and blue spheres, respectively. Includes metal node connectivity (blue polyhedra), infinite-rod SBU, and 3D representation of MOF-76.

A potential application for MOFs is biological imaging and sensing via photoluminescence. A large subset of luminescent MOFs use lanthanides in the metal clusters. Lanthanide photoluminescence has many unique properties that make them ideal for imaging applications, such as characteristically sharp and generally non-overlapping emission bands in the visible and near-infrared (NIR) regions of the spectrum, resistance to photobleaching or "blinking", and long luminescence lifetimes.[156] However, lanthanide emissions are difficult to sensitize directly because they must undergo LaPorte forbidden f-f transitions. Indirect sensitization of lanthanide emission can be accomplished by employing the "antenna effect", where the organic linkers act as antennae and absorb the excitation energy, transfer the energy to the excited state of the lanthanide, and yield lanthanide luminescence upon relaxation.[157] A prime example of the antenna effect is demonstrated by MOF-76, which combines trivalent lanthanide ions and 1,3,5-benzenetricarboxylate (BTC) linkers to form infinite rod SBUs coordinated into a three dimensional lattice.[158] As demonstrated by multiple research groups, the BTC linker can effectively sensitize the lanthanide emission, resulting in a MOF with variable emission wavelengths depending on the lanthanide identity.[159][160] Additionally, the Yan group has shown that Eu3+- and Tb3+- MOF-76 can be used for selective detection of acetophenone from other volatile monoaromatic hydrocarbons. Upon acetophenone uptake, the MOF shows a very sharp decrease, or quenching, of the luminescence intensity.[161]

For use in biological imaging, however, two main obstacles must be overcome:

  • MOFs must be synthesized on the nanoscale so as not to affect the target's normal interactions or behavior
  • The absorbance and emission wavelengths must occur in regions with minimal overlap from sample autofluorescence, other absorbing species, and maximum tissue penetration.[162][163]

Regarding the first point, nanoscale MOF (NMOF) synthesis has been mentioned in an earlier section. The latter obstacle addresses the limitation of the antenna effect. Smaller linkers tend to improve MOF stability, but have higher energy absorptions, predominantly in the ultraviolet (UV) and high-energy visible regions. A design strategy for MOFs with redshifted absorption properties has been accomplished by using large, chromophoric linkers. These linkers are often composed of polyaromatic species, leading to large pore sizes and thus decreased stability. To circumvent the use of large linkers, other methods are required to redshift the absorbance of the MOF so lower energy excitation sources can be used. Post-synthetic modification (PSM) is one promising strategy. Luo et al. introduced a new family of lanthanide MOFs with functionalized organic linkers. The MOFs, deemed MOF-1114, MOF-1115, MOF-1130, and MOF-1131, are composed of octahedral SBUs bridged by amino functionalized dicarboxylate linkers. The amino groups on the linkers served as sites for covalent PSM reactions with either salicylaldehyde or 3-hydroxynaphthalene-2-carboxaldehyde. Both of these reactions extend the π-conjugation of the linker, causing a redshift in the absorbance wavelength from 450 nm to 650 nm. The authors also propose that this technique could be adapted to similar MOF systems and, by increasing pore volumes with increasing linker lengths, larger pi-conjugated reactants can be used to further redshift the absorption wavelengths.[164] Biological imaging using MOFs has been realized by several groups, namely Foucault-Collet and co-workers. In 2013, they synthesized a NIR-emitting Yb3+-NMOF using phenylenevinylene dicarboxylate (PVDC) linkers. They observed cellular uptake in both HeLa cells and NIH-3T3 cells using confocal, visible, and NIR spectroscopy.[165] Although low quantum yields persist in water and Hepes buffer solution, the luminescence intensity is still strong enough to image cellular uptake in both the visible and NIR regimes.

Nuclear wasteform materials edit

 
Schematic representation of different ways to incorporate actinide species inside the MOF.

The development of new pathways for efficient nuclear waste administration is essential in wake of increased public concern about radioactive contamination, due to nuclear plant operation and nuclear weapon decommission. Synthesis of novel materials capable of selective actinide sequestration and separation is one of the current challenges acknowledged in the nuclear waste sector. Metal–organic frameworks (MOFs) are a promising class of materials to address this challenge due to their porosity, modularity, crystallinity, and tunability. Every building block of MOF structures can incorporate actinides. First, a MOF can be synthesized starting from actinide salts. In this case the metal nodes are actinides.[40][166] In addition, metal nodes can be extended, or cation exchange can exchange metals for actinides.[40] Organic linkers can be functionalized with groups capable of actinide uptake.[167][168][169][170][171] Lastly, the porosity of MOFs can be used to incorporate contained guest molecules[172][173][174] and trap them in a structure by installation of additional or capping linkers.[40]

Drug delivery systems edit

The synthesis, characterization, and drug-related studies of low toxicity, biocompatible MOFs has shown that they have potential for medical applications. Many groups have synthesized various low toxicity MOFs and have studied their uses in loading and releasing various therapeutic drugs for potential medical applications. A variety of methods exist for inducing drug release, such as pH-response, magnetic-response, ion-response, temperature-response, and pressure response.[175]

In 2010 Smaldone et al., an international research group, synthesized a biocompatible MOF termed CD-MOF-1 from cheap edible natural products. CD-MOF-1 consists of repeating base units of 6 γ-cyclodextrin rings bound together by potassium ions. γ-cyclodextrin (γ-CD) is a symmetrical cyclic oligosaccharide that is mass-produced enzymatically from starch and consists of eight asymmetric α-1,4-linked D-glucopyranosyl residues.[176] The molecular structure of these glucose derivatives, which approximates a truncated cone, bucket, or torus, generates a hydrophilic exterior surface and a nonpolar interior cavity. Cyclodextrins can interact with appropriately sized drug molecules to yield an inclusion complex.[177] Smaldone's group proposed a cheap and simple synthesis of the CD-MOF-1 from natural products. They dissolved sugar (γ-cyclodextrin) and an alkali salt (KOH, KCl, potassium benzoate) in distilled bottled water and allowed 190 proof grain alcohol (Everclear) to vapor diffuse into the solution for a week. The synthesis resulted in a cubic (γ-CD)6 repeating motif with a pore size of approximately 1 nm. Subsequently, in 2017 Hartlieb et al. at Northwestern did further research with CD-MOF-1 involving the encapsulation of ibuprofen. The group studied different methods of loading the MOF with ibuprofen as well as performing related bioavailability studies on the ibuprofen-loaded MOF. They investigated two different methods of loading CD-MOF-1 with ibuprofen; crystallization using the potassium salt of ibuprofen as the alkali cation source for production of the MOF, and absorption and deprotonation of the free-acid of ibuprofen into the MOF. From there the group performed in vitro and in vivo studies to determine the applicability of CD-MOF-1 as a viable delivery method for ibuprofen and other NSAIDs. In vitro studies showed no toxicity or effect on cell viability up to 100 μM. In vivo studies in mice showed the same rapid uptake of ibuprofen as the ibuprofen potassium salt control sample with a peak plasma concentration observed within 20 minutes, and the cocrystal has the added benefit of double the half-life in blood plasma samples.[178] The increase in half-life is due to CD-MOF-1 increasing the solubility of ibuprofen compared to the pure salt form.

Since these developments many groups have done further research into drug delivery with water-soluble, biocompatible MOFs involving common over-the-counter drugs.[179] In March 2018 Sara Rojas and her team published their research on drug incorporation and delivery with various biocompatible MOFs other than CD-MOF-1 through simulated cutaneous administration. The group studied the loading and release of ibuprofen (hydrophobic) and aspirin (hydrophilic) in three biocompatible MOFs (MIL-100(Fe), UiO-66(Zr), and MIL-127(Fe)). Under simulated cutaneous conditions (aqueous media at 37 °C) the six different combinations of drug-loaded MOFs fulfilled "the requirements to be used as topical drug delivery systems, such as released payload between 1 and 7 days" and delivering a therapeutic concentration of the drug of choice without causing unwanted side effects.[180] The group discovered that the drug uptake is "governed by the hydrophilic/hydrophobic balance between cargo and matrix" and "the accessibility of the drug through the framework". The "controlled release under cutaneous conditions follows different kinetics profiles depending on: (i) the structure of the framework, with either a fast delivery from the very open structure MIL-100 or a slower drug release from the narrow 1D pore system of MIL-127 or (ii) the hydrophobic/hydrophilic nature of the cargo, with a fast (Aspirin) and slow (Ibuprofen) release from the UiO-66 matrix." Moreover, a simple ball milling technique is used to efficiently encapsulate the model drugs 5-fluorouracil, caffeine, para-aminobenzoic acid, and benzocaine. Both computational and experimental studies confirm the suitability of [Zn4O(dmcapz)3] to incorporate high loadings of the studied bioactive molecules.[181]

Recent research involving MOFs as a drug delivery method includes more than just the encapsulation of everyday drugs like ibuprofen and aspirin. In early 2018 Chen et al., published detailing their work on the use of MOF, ZIF-8 (zeolitic imidazolate framework-8) in antitumor research "to control the release of an autophagy inhibitor, 3-methyladenine (3-MA), and prevent it from dissipating in a large quantity before reaching the target."[182] The group performed in vitro studies and determined that "the autophagy-related proteins and autophagy flux in HeLa cells treated with 3-MA@ZIF-8 NPs show that the autophagosome formation is significantly blocked, which reveals that the pH-sensitive dissociation increases the efficiency of autophagy inhibition at the equivalent concentration of 3-MA." This shows promise for future research and applicability with MOFs as drug delivery methods in the fight against cancer.

Semiconductors edit

In 2014 researchers proved that they can create electrically conductive thin films of MOFs (Cu3(BTC)2 (also known as HKUST-1; BTC, benzene-1,3,5-tricarboxylic acid) infiltrated with the molecule 7,7,8,8-tetracyanoquinododimethane) that could be used in applications including photovoltaics, sensors, and electronic materials and a path toward creating semiconductors. The team demonstrated tunable, air-stable electrical conductivity with values as high as 7 siemens per meter, comparable to bronze.[183]

Ni
3
(2,3,6,7,10,11-hexaiminotriphenylene)2 was shown to be a metal-organic graphene analogue that has a natural band gap, making it a semiconductor, and is able to self-assemble. It is an example of conductive metal-organic framework. It represents a family of similar compounds. Because of the symmetry and geometry in 2,3,6,7,10,11-hexaiminotriphenylene (HITP), the overall organometallic complex has an almost fractal nature that allows it to perfectly self-organize. By contrast, graphene must be doped to give it the properties of a semiconductor. Ni3(HITP)2 pellets had a conductivity of 2 S/cm, a record for a metal-organic compound.[184][185]

In 2018 researchers synthesized a two-dimensional semiconducting MOF (Fe3(THT)2(NH4)3, also known as THT, 2,3,6,7,10,11-triphenylenehexathiol) and showed high electric mobility at room temperature.[186] In 2020 the same material was integrated in a photo-detecting device, detecting a broad wavelength range from UV to NIR (400–1575 nm).[187] This was the first time a two-dimensional semiconducting MOF was demonstrated to be used in opto-electronic devices.[188]

  is a 2D MOF structure, and there are limited examples of materials which are intrinsically conductive, porous, and crystalline. Layered 2D MOFs have porous crystalline structure showing electrical conductivity. These materials are constructed from trigonal linker molecules (phenylene or triphenylene) and six functional groups of –OH, - , or –SH. The trigonal linker molecules and square-planarly coordinated metal ions such as  ,  ,  , and   form layers with hexagonal structures which look like graphene in larger scale. Stacking of these layers can build one-dimensional pore systems. Graphene-like 2D MOFs have shown decent conductivities. This makes them a good choice to be tested as electrode material for evolution of hydrogen from water, oxygen reduction reactions, supercapacitors, and sensing of volatile organic compounds (VOCs). Among these MOFs,   has exhibited the lowest conductivity, but also the strongest reaction in sensing of VOCs.[189][190][191]

Bio-mimetic mineralization edit

Biomolecules can be incorporated during the MOF crystallization process. Biomolecules including proteins, DNA, and antibodies could be encapsulated within ZIF-8. Enzymes encapsulated in this way were stable and active even after being exposed to harsh conditions (e.g. aggressive solvents and high temperature). ZIF-8, MIL-88A, HKUST-1, and several luminescent MOFs containing lanthanide metals were used for the biomimetic mineralization process.[192]

Carbon capture edit

Adsorbent edit

MOF's small, tunable pore sizes and high void fractions are promising as an adsorbent to capture CO2.[193] MOFs could provide a more efficient alternative to traditional amine solvent-based methods in CO2 capture from coal-fired power plants.[194]

MOFs could be employed in each of the main three carbon capture configurations for coal-fired power plants: pre-combustion, post-combustion, and oxy-combustion.[195] The post-combustion configuration is the only one that can be retrofitted to existing plants, drawing the most interest and research. The flue gas would be fed through a MOF in a packed-bed reactor setup. Flue gas is generally 40 to 60 °C with a partial pressure of CO2 at 0.13 – 0.16 bar. CO2 can bind to the MOF surface through either physisorption (via Van der Waals interactions) or chemisorption (via covalent bond formation).[196]

Once the MOF is saturated, the CO2 is extracted from the MOF through either a temperature swing or a pressure swing. This process is known as regeneration. In a temperature swing regeneration, the MOF would be heated until CO2 desorbs. To achieve working capacities comparable to the amine process, the MOF must be heated to around 200 °C. In a pressure swing, the pressure would be decreased until CO2 desorbs.[197]

Another relevant MOF property is their low heat capacities. Monoethanolamine (MEA) solutions, the leading capture method, have a heat capacity between 3-4 J/(g⋅K) since they are mostly water. This high heat capacity contributes to the energy penalty in the solvent regeneration step, i.e. when the adsorbed CO2 is removed from the MEA solution. MOF-177, a MOF designed for CO2 capture, has a heat capacity of 0.5 J/(g⋅K) at ambient temperature.[195]

MOFs adsorb 90% of the CO2 using a vacuum pressure swing process. The MOF Mg(dobdc) has a 21.7 wt% CO2 loading capacity. Applied to a large scale power plant, the cost of energy would increase by 65%, while a U.S. NETL baseline amine-based system would cause an increase of 81% (goal is 35%). The capture cost would be $57 / ton, while for the amine system the cost is estimated to be $72 / ton. At that rate the capital required to implement such project in a 580 MW power plant would be $354 million.[198]

Catalyst edit

A MOF loaded with propylene oxide can act as a catalyst, converting CO2 into cyclic carbonates (ring-shaped molecules with many applications). They can also remove carbon from biogas. This MOF is based on lanthanides, which provide chemical stability. This is especially important because the gases the MOF will be exposed to are hot, high in humidity, and acidic.[199] Triaminoguanidinium-based POFs and Zn/POFs are new multifunctional materials for environmental remediation and biomedical applications.[200]

Desalination/ion separation edit

MOF membranes can mimic substantial ion selectivity. This offers the potential for use in desalination and water treatment. As of 2018 reverse osmosis supplied more than half of global desalination capacity, and the last stage of most water treatment processes. Osmosis does not use dehydration of ions, or selective ion transport in biological channels and it is not energy efficient. The mining industry, uses membrane-based processes to reduce water pollution, and to recover metals. MOFs could be used to extract metals such as lithium from seawater and waste streams.[201]

MOF membranes such as ZIF-8 and UiO-66 membranes with uniform subnanometer pores consisting of angstrom-scale windows and nanometer-scale cavities displayed ultrafast selective transport of alkali metal ions. The windows acted as ion selectivity filters for alkali metal ions, while the cavities functioned as pores for transport. The ZIF-8[202] and UiO-66[203] membranes showed a LiCl/RbCl selectivity of ~4.6 and ~1.8, respectively, much higher than the 0.6 to 0.8 selectivity in traditional membranes.[204] A 2020 study suggested that a new MOF called PSP-MIL-53 could be used along with sunlight to purify water in just half an hour.[205]

Gas separation edit

MOFs are also predicted to be very effective media to separate gases with low energy cost using computational high throughput screening from their adsorption[206] or gas breakthrough/diffusion[207] properties. One example is NbOFFIVE-1-Ni, also referred to as KAUST-7 which can separate propane and propylene via diffusion at nearly 100% selectivity.[208] The specific molecule selectivity properties provided by Cu-BDC surface mounted metal organic framework (SURMOF-2) growth on alumina layer on top of back gated Graphene Field Effect Transistor (GFET) can provide a sensor that is only sensitive to ethanol but not to methanol or isopropanol.[209]

Water vapor capture and dehumidification edit

MOFs have been demonstrated that capture water vapor from the air.[210] In 2021 under humid conditions, a polymer-MOF lab prototype yielded 17 liters (4.5 gal) of water per kg per day without added energy.[211][212]

MOFs could also be used to increase energy efficiency in room temperature space cooling applications.[213][214]

 
Schematic diagram for MOF dehumidification, featuring MIL-100(Fe), a MOF with particularly high water adsorption capacity

When cooling outdoor air, a cooling unit must deal with both the air's sensible heat and latent heat. Typical vapor-compression air-conditioning (VCAC) units manage the latent heat in air through cooling fins held below the dew point temperature of the moist air at the intake. These fins condense the water, dehydrating the air and thus substantially reducing the air's heat content. The cooler's energy usage is highly dependent on the cooling coil's temperature and would be improved greatly if the temperature of this coil could be raised above the dew point.[215] This makes it desirable to handle dehumidification through means other than condensation. One such means is by adsorbing the water from the air into a desiccant coated onto the heat exchangers, using the waste heat exhausted from the unit to desorb the water from the sorbent and thus regenerate the desiccant for repeated usage. This is accomplished by having two condenser/evaporator units through which the flow of refrigerant can be reversed once the desiccant on the condenser is saturated, thus making the condenser the evaporator and vice versa.[213]

MOFs' extremely high surface areas and porosities have made them the subject of much research in water adsorption applications.[213][216][217][218] Chemistry can help tune the optimal relative humidity for adsorption/desorption, and the sharpness of the water uptake.[213][219]

Ferroelectrics and multiferroics edit

Some MOFs also exhibit spontaneous electric polarization, which occurs due to the ordering of electric dipoles (polar linkers or guest molecules) below a certain phase transition temperature.[220] If this long-range dipolar order can be controlled by the external electric field, a MOF is called ferroelectric.[221] Some ferroelectric MOFs also exhibit magnetic ordering making them single structural phase multiferroics. This material property is highly interesting for construction of memory devices with high information density. The coupling mechanism of type-I [(CH3)2NH2][Ni(HCOO)3] molecular multiferroic is spontaneous elastic strain mediated indirect coupling.[222]

See also edit

References edit

  1. ^ a b Batten SR, Champness NR, Chen XM, Garcia-Martinez J, Kitagawa S, Öhrström L, O'Keeffe M, Suh MP, Reedijk J (2013). "Terminology of metal–organic frameworks and coordination polymers (IUPAC Recommendations 2013)" (PDF). Pure and Applied Chemistry. 85 (8): 1715–1724. doi:10.1351/PAC-REC-12-11-20. S2CID 96853486.
  2. ^ Bennett, Thomas D.; Cheetham, Anthony K. (2014-05-20). "Amorphous Metal–Organic Frameworks". Accounts of Chemical Research. 47 (5): 1555–1562. doi:10.1021/ar5000314. PMID 24707980.
  3. ^ Bennett, Thomas D.; Coudert, François-Xavier; James, Stuart L.; Cooper, Andrew I. (September 2021). "The changing state of porous materials". Nature Materials. 20 (9): 1179–1187. Bibcode:2021NatMa..20.1179B. doi:10.1038/s41563-021-00957-w. PMID 33859380. S2CID 233239286.
  4. ^ Mon M, Bruno R, Ferrando-Soria J, Armentano D, Pardo E (2018). "Metal–organic framework technologies for water remediation: towards a sustainable ecosystem". Journal of Materials Chemistry A. 6 (12): 4912–4947. doi:10.1039/c8ta00264a.
  5. ^ Cejka J, ed. (2011). Metal-Organic Frameworks Applications from Catalysis to Gas Storage. Wiley-VCH. ISBN 978-3-527-32870-3.
  6. ^ O'Keeffe M, Yaghi OM (2005). "Reticular chemistry—Present and future prospects" (PDF). Journal of Solid State Chemistry. 178 (8): v–vi. Bibcode:2005JSSCh.178D...5.. doi:10.1016/S0022-4596(05)00368-3.
  7. ^ Côté AP, Benin AI, Ockwig NW, O'Keeffe M, Matzger AJ, Yaghi OM (November 2005). "Porous, crystalline, covalent organic frameworks". Science. 310 (5751): 1166–70. Bibcode:2005Sci...310.1166C. doi:10.1126/science.1120411. PMID 16293756. S2CID 35798005.
  8. ^ a b c d e f Czaja AU, Trukhan N, Müller U (May 2009). "Industrial applications of metal-organic frameworks". Chemical Society Reviews. 38 (5): 1284–93. doi:10.1039/b804680h. PMID 19384438.
  9. ^ Cheetham AK, Rao CN, Feller RK (2006). "Structural diversity and chemical trends in hybrid inorganic–organic framework materials". Chemical Communications (46): 4780–4795. doi:10.1039/b610264f. PMID 17345731.
  10. ^ a b Cheetham AK, Férey G, Loiseau T (November 1999). "Open-Framework Inorganic Materials". Angewandte Chemie. 38 (22): 3268–3292. doi:10.1002/(SICI)1521-3773(19991115)38:22<3268::AID-ANIE3268>3.0.CO;2-U. PMID 10602176.
  11. ^ Bucar DK, Papaefstathiou GS, Hamilton TD, Chu QL, Georgiev IG, MacGillivray LR (2007). "Template-controlled reactivity in the organic solid state by principles of coordination-driven self-assembly". European Journal of Inorganic Chemistry. 2007 (29): 4559–4568. doi:10.1002/ejic.200700442.
  12. ^ Parnham ER, Morris RE (October 2007). "Ionothermal synthesis of zeolites, metal-organic frameworks, and inorganic-organic hybrids". Accounts of Chemical Research. 40 (10): 1005–13. doi:10.1021/ar700025k. PMID 17580979.
  13. ^ a b c d Dincă M, Long JR (2008). "Hydrogen storage in microporous metal-organic frameworks with exposed metal sites". Angewandte Chemie. 47 (36): 6766–79. doi:10.1002/anie.200801163. PMID 18688902.
  14. ^ Gitis, Vitaly; Rothenberg, Gadi (2020). Gitis, Vitaly; Rothenberg, Gadi (eds.). Handbook of Porous Materials. Vol. 4. Singapore: WORLD SCIENTIFIC. pp. 110–111. doi:10.1142/11909. ISBN 978-981-12-2328-0.
  15. ^ a b Ni Z, Masel RI (September 2006). "Rapid production of metal-organic frameworks via microwave-assisted solvothermal synthesis". Journal of the American Chemical Society. 128 (38): 12394–5. doi:10.1021/ja0635231. PMID 16984171.
  16. ^ a b Choi JS, Son WJ, Kim J, Ahn WS (2008). "Metal–organic framework MOF-5 prepared by microwave heating: Factors to be considered". Microporous and Mesoporous Materials. 116 (1–3): 727–731. doi:10.1016/j.micromeso.2008.04.033.
  17. ^ Steenhaut, Timothy; Hermans, Sophie; Filinchuk, Yaroslav (2020). "Green synthesis of a large series of bimetallic MIL-100(Fe,M) MOFs". New Journal of Chemistry. 44 (10): 3847–3855. doi:10.1039/D0NJ00257G. S2CID 214492546.
  18. ^ Pichon A, James SL (2008). "An array-based study of reactivity under solvent-free mechanochemical conditions—insights and trends". CrystEngComm. 10 (12): 1839–1847. doi:10.1039/B810857A.
  19. ^ Braga D, Giaffreda SL, Grepioni F, Chierotti MR, Gobetto R, Palladino G, Polito M (2007). "Solvent effect in a "solvent free" reaction". CrystEngComm. 9 (10): 879–881. doi:10.1039/B711983F.
  20. ^ a b Steenhaut, Timothy; Grégoire, Nicolas; Barozzino-Consiglio, Gabriella; Filinchuk, Yaroslav; Hermans, Sophie (2020). "Mechanochemical defect engineering of HKUST-1 and impact of the resulting defects on carbon dioxide sorption and catalytic cyclopropanation". RSC Advances. 10 (34): 19822–19831. Bibcode:2020RSCAd..1019822S. doi:10.1039/C9RA10412G. PMC 9054116. PMID 35520409.
  21. ^ Stassen I, Styles M, Grenci G, Gorp HV, Vanderlinden W, Feyter SD, Falcaro P, Vos DD, Vereecken P, Ameloot R (March 2016). "Chemical vapour deposition of zeolitic imidazolate framework thin films". Nature Materials. 15 (3): 304–10. Bibcode:2016NatMa..15..304S. doi:10.1038/nmat4509. PMID 26657328.
  22. ^ Cruz A, Stassen I, Ameloot, R, et al. (2019). "An integrated cleanroom process for the vapor-phase deposition of large-area zeolitic imidazolate framework thin films". Chemistry of Materials. 31 (22): 9462–9471. doi:10.1021/acs.chemmater.9b03435. hdl:10550/74201. S2CID 208737085.
  23. ^ Virmani, Erika; Rotter, Julian M.; Mähringer, Andre; von Zons, Tobias; Godt, Adelheid; Bein, Thomas; Wuttke, Stefan; Medina, Dana D. (2018-04-11). "On-Surface Synthesis of Highly Oriented Thin Metal–Organic Framework Films through Vapor-Assisted Conversion". Journal of the American Chemical Society. 140 (14): 4812–4819. doi:10.1021/jacs.7b08174. PMID 29542320.
  24. ^ Sebastian Bauer, Norbert Stock (October 2007), "Hochdurchsatz-Methoden in der Festkörperchemie. Schneller zum Ziel", Chemie in unserer Zeit (in German), vol. 41, no. 5, pp. 390–398, doi:10.1002/ciuz.200700404
  25. ^ Gimeno-Fabra, Miquel; Munn, Alexis S.; Stevens, Lee A.; Drage, Trevor C.; Grant, David M.; Kashtiban, Reza J.; Sloan, Jeremy; Lester, Edward; Walton, Richard I. (7 September 2012). "Instant MOFs: continuous synthesis of metal–organic frameworks by rapid solvent mixing". Chemical Communications. 48 (86): 10642–4. doi:10.1039/C2CC34493A. PMID 23000779. Retrieved 22 June 2020.
  26. ^ Rasmussen, Elizabeth G.; Kramlich, John; Novosselov, Igor V. (3 June 2020). "Scalable Continuous Flow Metal–Organic Framework (MOF) Synthesis Using Supercritical CO2". ACS Sustainable Chemistry & Engineering. 8 (26): 9680–9689. doi:10.1021/acssuschemeng.0c01429. S2CID 219915159.
  27. ^ Biemmi, Enrica; Christian, Sandra; Stock, Norbert; Bein, Thomas (2009), "High-throughput screening of synthesis parameters in the formation of the metal-organic frameworks MOF-5 and HKUST-1", Microporous and Mesoporous Materials (in German), vol. 117, no. 1–2, pp. 111–117, doi:10.1016/j.micromeso.2008.06.040
  28. ^ Putnis, Andrew (2009-01-01). "Mineral Replacement Reactions". Reviews in Mineralogy and Geochemistry. 70 (1): 87–124. Bibcode:2009RvMG...70...87P. doi:10.2138/rmg.2009.70.3.
  29. ^ Reboul, Julien; Furukawa, Shuhei; Horike, Nao; Tsotsalas, Manuel; Hirai, Kenji; Uehara, Hiromitsu; Kondo, Mio; Louvain, Nicolas; Sakata, Osami; Kitagawa, Susumu (August 2012). "Mesoscopic architectures of porous coordination polymers fabricated by pseudomorphic replication". Nature Materials. 11 (8): 717–723. Bibcode:2012NatMa..11..717R. doi:10.1038/nmat3359. hdl:2433/158311. PMID 22728321. S2CID 205407412.
  30. ^ Robatjazi, Hossein; Weinberg, Daniel; Swearer, Dayne F.; Jacobson, Christian; Zhang, Ming; Tian, Shu; Zhou, Linan; Nordlander, Peter; Halas, Naomi J. (February 2019). "Metal-organic frameworks tailor the properties of aluminum nanocrystals". Science Advances. 5 (2): eaav5340. Bibcode:2019SciA....5.5340R. doi:10.1126/sciadv.aav5340. PMC 6368424. PMID 30783628.
  31. ^ Kornienko, Nikolay; Zhao, Yingbo; Kley, Christopher S.; Zhu, Chenhui; Kim, Dohyung; Lin, Song; Chang, Christopher J.; Yaghi, Omar M.; Yang, Peidong (2015-11-11). "Metal–Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide". Journal of the American Chemical Society. 137 (44): 14129–14135. doi:10.1021/jacs.5b08212. PMID 26509213. S2CID 14793796.
  32. ^ a b c Das S, Kim H, Kim K (March 2009). "Metathesis in single crystal: complete and reversible exchange of metal ions constituting the frameworks of metal-organic frameworks". Journal of the American Chemical Society. 131 (11): 3814–5. doi:10.1021/ja808995d. PMID 19256486.
  33. ^ a b c d Burrows AD, Frost CG, Mahon MF, Richardson C (2008-10-20). "Post-synthetic modification of tagged metal-organic frameworks". Angewandte Chemie. 47 (44): 8482–6. doi:10.1002/anie.200802908. PMID 18825761.
  34. ^ a b Li T, Kozlowski MT, Doud EA, Blakely MN, Rosi NL (August 2013). "Stepwise ligand exchange for the preparation of a family of mesoporous MOFs". Journal of the American Chemical Society. 135 (32): 11688–91. doi:10.1021/ja403810k. PMID 23688075.
  35. ^ Sun D, Liu W, Qiu M, Zhang Y, Li Z (February 2015). "Introduction of a mediator for enhancing photocatalytic performance via post-synthetic metal exchange in metal-organic frameworks (MOFs)". Chemical Communications. 51 (11): 2056–9. doi:10.1039/c4cc09407g. PMID 25532612.
  36. ^ Liu C, Rosi NL (September 2017). "Ternary gradient metal-organic frameworks". Faraday Discussions. 201: 163–174. Bibcode:2017FaDi..201..163L. doi:10.1039/c7fd00045f. PMID 28621353.
  37. ^ Liu C, Zeng C, Luo TY, Merg AD, Jin R, Rosi NL (September 2016). "Establishing Porosity Gradients within Metal-Organic Frameworks Using Partial Postsynthetic Ligand Exchange". Journal of the American Chemical Society. 138 (37): 12045–8. doi:10.1021/jacs.6b07445. PMID 27593173.
  38. ^ Yuan S, Lu W, Chen YP, Zhang Q, Liu TF, Feng D, Wang X, Qin J, Zhou HC (March 2015). "Sequential linker installation: precise placement of functional groups in multivariate metal-organic frameworks". Journal of the American Chemical Society. 137 (9): 3177–80. doi:10.1021/ja512762r. PMID 25714137.
  39. ^ Yuan S, Chen YP, Qin JS, Lu W, Zou L, Zhang Q, Wang X, Sun X, Zhou HC (July 2016). "Linker Installation: Engineering Pore Environment with Precisely Placed Functionalities in Zirconium MOFs". Journal of the American Chemical Society. 138 (28): 8912–9. doi:10.1021/jacs.6b04501. OSTI 1388673. PMID 27345035.
  40. ^ a b c d Dolgopolova EA, Ejegbavwo OA, Martin CR, Smith MD, Setyawan W, Karakalos SG, Henager CH, Zur Loye HC, Shustova NB (November 2017). "Multifaceted Modularity: A Key for Stepwise Building of Hierarchical Complexity in Actinide Metal-Organic Frameworks". Journal of the American Chemical Society. 139 (46): 16852–16861. doi:10.1021/jacs.7b09496. PMID 29069547.
  41. ^ a b c d e f g h i j k l m n o p q r s Murray LJ, Dincă M, Long JR (May 2009). "Hydrogen storage in metal-organic frameworks". Chemical Society Reviews. 38 (5): 1294–314. CiteSeerX 10.1.1.549.4404. doi:10.1039/b802256a. PMID 19384439. S2CID 10443172.
  42. ^ Li Y, Yang RT (2007). "Hydrogen Storage on Platinum Nanoparticles Doped on Superactivated Carbon". Journal of Physical Chemistry C. 111 (29): 11086–11094. doi:10.1021/jp072867q.
  43. ^ Jacoby, Mitch (2013). "Materials Chemistry: Metal-Organic Frameworks Go Commercial". Chemical & Engineering News. 91 (51).
  44. ^ Cejka J, Corma A, Zones S (27 May 2010). Zeolites and Catalysis: Synthesis, Reactions and Applications. John Wiley & Sons. ISBN 978-3-527-63030-1.
  45. ^ Niknam, Esmaeil; Panahi, Farhad; Daneshgar, Fatemeh; Bahrami, Foroogh; Khalafi-Nezhad, Ali (2018). "Metal–Organic Framework MIL-101(Cr) as an Efficient Heterogeneous Catalyst for Clean Synthesis of Benzoazoles". ACS Omega. 3 (12): 17135–17144. doi:10.1021/acsomega.8b02309. PMC 6643801. PMID 31458334. S2CID 104347751.
  46. ^ López, Jorge; Chávez, Ana M.; Rey, Ana; Álvarez, Pedro M. (2021). "Insights into the Stability and Activity of MIL-53(Fe) in Solar Photocatalytic Oxidation Processes in Water". Catalysts. 11 (4): 448. doi:10.3390/catal11040448.
  47. ^ Chávez, A.M.; Rey, A.; López, J.; Álvarez, P.M.; Beltrán, F.J. (2021). "Critical aspects of the stability and catalytic activity of MIL-100(Fe) in different advanced oxidation processes". Separation and Purification Technology. 255: 117660. doi:10.1016/j.seppur.2020.117660. S2CID 224863042.
  48. ^ Fujita M, Kwon YJ, Washizu S, Ogura K (1994). "Preparation, Clathration Ability, and Catalysis of a Two-Dimensional Square Network Material Composed of Cadmium(II) and 4,4-Bipyridine". Journal of the American Chemical Society. 116 (3): 1151. doi:10.1021/ja00082a055.
  49. ^ Llabresixamena F, Abad A, Corma A, Garcia H (2007). "MOFs as catalysts: Activity, reusability and shape-selectivity of a Pd-containing MOF". Journal of Catalysis. 250 (2): 294–298. doi:10.1016/j.jcat.2007.06.004.
  50. ^ Ravon U, Domine ME, Gaudillère C, Desmartin-Chomel A, Farrusseng D (2008). "MOFs as acid catalysts with shape selectivity properties". New Journal of Chemistry. 32 (6): 937. doi:10.1039/B803953B.
  51. ^ Chui SS, Lo SM, Charmant JP, Orpen AG, Williams ID (1999). "A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n". Science. 283 (5405): 1148–50. Bibcode:1999Sci...283.1148C. doi:10.1126/science.283.5405.1148. PMID 10024237.
  52. ^ Alaerts L, Séguin E, Poelman H, Thibault-Starzyk F, Jacobs PA, De Vos DE (September 2006). "Probing the Lewis acidity and catalytic activity of the metal-organic framework [Cu3(btc)2] (BTC=benzene-1,3,5-tricarboxylate)". Chemistry: A European Journal. 12 (28): 7353–63. doi:10.1002/chem.200600220. hdl:1854/LU-351275. PMID 16881030.
  53. ^ Henschel A, Gedrich K, Kraehnert R, Kaskel S (September 2008). "Catalytic properties of MIL-101". Chemical Communications (35): 4192–4. doi:10.1039/B718371B. PMID 18802526.
  54. ^ a b Horike S, Dinca M, Tamaki K, Long JR (May 2008). "Size-selective Lewis acid catalysis in a microporous metal-organic framework with exposed Mn2+ coordination sites". Journal of the American Chemical Society. 130 (18): 5854–5. doi:10.1021/ja800669j. PMID 18399629.
  55. ^ Chen L, Yang Y, Jiang D (July 2010). "CMPs as scaffolds for constructing porous catalytic frameworks: a built-in heterogeneous catalyst with high activity and selectivity based on nanoporous metalloporphyrin polymers". Journal of the American Chemical Society. 132 (26): 9138–43. doi:10.1021/ja1028556. PMID 20536239.
  56. ^ Rahiman AK, Rajesh K, Bharathi KS, Sreedaran S, Narayanan V (2009). "Catalytic oxidation of alkenes by manganese(III) porphyrin-encapsulated Al, V, Si-mesoporous molecular sieves". Inorganica Chimica Acta. 352 (5): 1491–1500. doi:10.1016/j.ica.2008.07.011.
  57. ^ Mansuy D, Bartoli JF, Momenteau M (1982). "Alkane hydroxylation catalyzed by metalloporhyrins : evidence for different active oxygen species with alkylhydroperoxides and iodosobenzene as oxidants". Tetrahedron Letters. 23 (27): 2781–2784. doi:10.1016/S0040-4039(00)87457-2.
  58. ^ Ingleson MJ, Barrio JP, Bacsa J, Dickinson C, Park H, Rosseinsky MJ (March 2008). "Generation of a solid Brønsted acid site in a chiral framework". Chemical Communications (11): 1287–9. doi:10.1039/B718443C. PMID 18389109.
  59. ^ a b Hasegawa S, Horike S, Matsuda R, Furukawa S, Mochizuki K, Kinoshita Y, Kitagawa S (March 2007). "Three-dimensional porous coordination polymer functionalized with amide groups based on tridentate ligand: selective sorption and catalysis". Journal of the American Chemical Society. 129 (9): 2607–14. doi:10.1021/ja067374y. PMID 17288419.
  60. ^ a b Hwang YK, Hong DY, Chang JS, Jhung SH, Seo YK, Kim J, Vimont A, Daturi M, Serre C, Férey G (2008). "Amine grafting on coordinatively unsaturated metal centers of MOFs: consequences for catalysis and metal encapsulation". Angewandte Chemie. 47 (22): 4144–8. doi:10.1002/anie.200705998. PMID 18435442.
  61. ^ Seo JS, Whang D, Lee H, Jun SI, Oh J, Jeon YJ, Kim K (April 2000). "A homochiral metal-organic porous material for enantioselective separation and catalysis". Nature. 404 (6781): 982–6. Bibcode:2000Natur.404..982S. doi:10.1038/35010088. PMID 10801124. S2CID 1159701.
  62. ^ Ohmori O, Fujita M (July 2004). "Heterogeneous catalysis of a coordination network: cyanosilylation of imines catalyzed by a Cd(II)-(4,4-bipyridine) square grid complex". Chemical Communications (14): 1586–7. doi:10.1039/B406114B. PMID 15263930.
  63. ^ Schlichte K, Kratzke T, Kaskel S (2004). "Improved synthesis, thermal stability and catalytic properties of the metal-organic framework compound Cu3(BTC)2". Microporous and Mesoporous Materials. 73 (1–2): 81–85. doi:10.1016/j.micromeso.2003.12.027. hdl:11858/00-001M-0000-000F-9785-0.
  64. ^ Chui SS, Lo SM, Charmant JP, Orpen AG, Williams ID (1999). "A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n". Science. 283 (5405): 1148–1150. Bibcode:1999Sci...283.1148C. doi:10.1126/science.283.5405.1148. PMID 10024237.
  65. ^ Evans OR, Ngo HL, Lin W (2001). "Chiral Porous Solids Based on Lamellar Lanthanide Phosphonates". Journal of the American Chemical Society. 123 (42): 10395–6. doi:10.1021/ja0163772. PMID 11603994.
  66. ^ Kato C, Hasegawa M, Sato T, Yoshizawa A, Inoue T, Mori W (2005). "Microporous dinuclear copper(II) trans-1,4-cyclohexanedicarboxylate: heterogeneous oxidation catalysis with hydrogen peroxide and X-ray powder structure of peroxo copper(II) intermediate". Journal of Catalysis. 230: 226–236. doi:10.1016/j.jcat.2004.11.032.
  67. ^ Han JW, Hill CL (December 2007). "A coordination network that catalyzes O2-based oxidations". Journal of the American Chemical Society. 129 (49): 15094–5. doi:10.1021/ja069319v. PMID 18020331.
  68. ^ Férey G, Mellot-Draznieks C, Serre C, Millange F, Dutour J, Surblé S, Margiolaki I (September 2005). "A chromium terephthalate-based solid with unusually large pore volumes and surface area" (PDF). Science. 309 (5743): 2040–2. Bibcode:2005Sci...309.2040F. doi:10.1126/science.1116275. PMID 16179475. S2CID 29483796.
  69. ^ Tahmouresilerd B, Larson PJ, Unruh DK, Cozzolino AF (July 2018). "Make room for iodine: systematic pore tuning of multivariate metal–organic frameworks for the catalytic oxidation of hydroquinones using hypervalent iodine". Catalysis Science & Technology. 8 (17): 4349–4357. doi:10.1039/C8CY00794B.
  70. ^ Tahmouresilerd B, Moody M, Agogo L, Cozzolino AF (Apr 2019). "The impact of an isoreticular expansion strategy on the performance of iodine catalysts supported in multivariate zirconium and aluminum metal–organic frameworks". Dalton Transactions. 48 (19): 6445–6454. doi:10.1039/C9DT00368A. PMID 31017171. S2CID 129944197.
  71. ^ Schröder F, Esken D, Cokoja M, van den Berg MW, Lebedev OI, Van Tendeloo G, Walaszek B, Buntkowsky G, Limbach HH, Chaudret B, Fischer RA (May 2008). "Ruthenium nanoparticles inside porous [Zn4O(bdc)3] by hydrogenolysis of adsorbed [Ru(cod)(cot)]: a solid-state reference system for surfactant-stabilized ruthenium colloids". Journal of the American Chemical Society. 130 (19): 6119–30. doi:10.1021/ja078231u. PMID 18402452.
  72. ^ Tan YC, Zeng HC (October 2018). "Lewis basicity generated by localised charge imbalance in noble metal nanoparticle-embedded defective metal-organic frameworks". Nature Communications. 9 (1): 4326. Bibcode:2018NatCo...9.4326T. doi:10.1038/s41467-018-06828-4. PMC 6194069. PMID 30337531.
  73. ^ Pan L, Liu H, Lei X, Huang X, Olson DH, Turro NJ, Li J (February 2003). "RPM-1: a recyclable nanoporous material suitable for ship-in-bottle synthesis and large hydrocarbon sorption". Angewandte Chemie. 42 (5): 542–6. doi:10.1002/anie.200390156. PMID 12569485.
  74. ^ Uemura T, Kitaura R, Ohta Y, Nagaoka M, Kitagawa S (June 2006). "Nanochannel-promoted polymerization of substituted acetylenes in porous coordination polymers". Angewandte Chemie. 45 (25): 4112–6. doi:10.1002/anie.200600333. PMID 16721889.
  75. ^ Uemura T, Hiramatsu D, Kubota Y, Takata M, Kitagawa S (2007). "Topotactic linear radical polymerization of divinylbenzenes in porous coordination polymers". Angewandte Chemie. 46 (26): 4987–90. doi:10.1002/anie.200700242. PMID 17514689.
  76. ^ Ezuhara T, Endo K, Aoyama Y (1999). "Helical Coordination Polymers from Achiral Components in Crystals. Homochiral Crystallization, Homochiral Helix Winding in the Solid State, and Chirality Control by Seeding". Journal of the American Chemical Society. 121 (14): 3279. doi:10.1021/ja9819918.
  77. ^ Wu ST, Wu YR, Kang QQ, Zhang H, Long LS, Zheng Z, Huang RB, Zheng LS (2007). "Chiral symmetry breaking by chemically manipulating statistical fluctuation in crystallization". Angewandte Chemie. 46 (44): 8475–9. doi:10.1002/anie.200703443. PMID 17912730.
  78. ^ Kepert CJ, Prior TJ, Rosseinsky MJ (2000). "A Versatile Family of Interconvertible Microporous Chiral Molecular Frameworks: The First Example of Ligand Control of Network Chirality". Journal of the American Chemical Society. 122 (21): 5158–5168. doi:10.1021/ja993814s.
  79. ^ Bradshaw D, Prior TJ, Cussen EJ, Claridge JB, Rosseinsky MJ (May 2004). "Permanent microporosity and enantioselective sorption in a chiral open framework". Journal of the American Chemical Society. 126 (19): 6106–14. doi:10.1021/ja0316420. PMID 15137776.
  80. ^ Lin Z, Slawin AM, Morris RE (April 2007). "Chiral induction in the ionothermal synthesis of a 3-D coordination polymer". Journal of the American Chemical Society. 129 (16): 4880–1. doi:10.1021/ja070671y. PMID 17394325.
  81. ^ Hu A, Ngo HL, Lin W (May 2004). "Remarkable 4,4-substituent effects on binap: Highly enantioselective Ru catalysts for asymmetric hydrogenation of beta-aryl ketoesters and their immobilization in room-temperature ionic liquids". Angewandte Chemie. 43 (19): 2501–4. doi:10.1002/anie.200353415. PMID 15127435.
  82. ^ Wu CD, Hu A, Zhang L, Lin W (June 2005). "A homochiral porous metal-organic framework for highly enantioselective heterogeneous asymmetric catalysis". Journal of the American Chemical Society. 127 (25): 8940–1. doi:10.1021/ja052431t. PMID 15969557.
  83. ^ Wu CD, Lin W (2007). "Heterogeneous asymmetric catalysis with homochiral metal-organic frameworks: network-structure-dependent catalytic activity". Angewandte Chemie. 46 (7): 1075–8. doi:10.1002/anie.200602099. PMID 17183496.
  84. ^ Cho SH, Ma B, Nguyen ST, Hupp JT, Albrecht-Schmitt TE (June 2006). "A metal-organic framework material that functions as an enantioselective catalyst for olefin epoxidation". Chemical Communications (24): 2563–5. doi:10.1039/b600408c. PMID 16779478.
  85. ^ Hu A, Ngo HL, Lin W (September 2003). "Chiral porous hybrid solids for practical heterogeneous asymmetric hydrogenation of aromatic ketones". Journal of the American Chemical Society. 125 (38): 11490–1. doi:10.1021/ja0348344. PMID 13129339.
  86. ^ Farrusseng D, Aguado S, Pinel C (2009). "Metal-organic frameworks: opportunities for catalysis". Angewandte Chemie. 48 (41): 7502–13. doi:10.1002/anie.200806063. PMID 19691074.
  87. ^ Eddaoudi M, Kim J, Wachter J, Chae HK, O'Keeffe M, Yaghi OM (2001). "Porous Metal-Organic Polyhedra: 25 Å Cuboctahedron Constructed from 12 Cu2(CO2)4 Paddle-Wheel Building Blocks". Journal of the American Chemical Society. 123 (18): 4368–9. doi:10.1021/ja0104352. PMID 11457217.
  88. ^ Furukawa H, Kim J, Ockwig NW, O'Keeffe M, Yaghi OM (September 2008). "Control of vertex geometry, structure dimensionality, functionality, and pore metrics in the reticular synthesis of crystalline metal-organic frameworks and polyhedra". Journal of the American Chemical Society. 130 (35): 11650–61. doi:10.1021/ja803783c. PMID 18693690.
  89. ^ Surblé S, Serre C, Mellot-Draznieks C, Millange F, Férey G (January 2006). "A new isoreticular class of metal-organic-frameworks with the MIL-88 topology". Chemical Communications (3): 284–6. doi:10.1039/B512169H. PMID 16391735.
  90. ^ Sudik AC, Millward AR, Ockwig NW, Côté AP, Kim J, Yaghi OM (May 2005). "Design, synthesis, structure, and gas (N
    2
    , Ar, CO
    2
    , CH
    4
    , and H
    2
    ) sorption properties of porous metal-organic tetrahedral and heterocuboidal polyhedra". Journal of the American Chemical Society. 127 (19): 7110–8. doi:10.1021/ja042802q. PMID 15884953.
  91. ^ Degnan T (2003). "The implications of the fundamentals of shape selectivity for the development of catalysts for the petroleum and petrochemical industries". Journal of Catalysis. 216 (1–2): 32–46. doi:10.1016/S0021-9517(02)00105-7.
  92. ^ Kuc A, Enyashin A, Seifert G (July 2007). "Metal-organic frameworks: structural, energetic, electronic, and mechanical properties". The Journal of Physical Chemistry B. 111 (28): 8179–86. doi:10.1021/jp072085x. PMID 17585800.
  93. ^ Esswein AJ, Nocera DG (October 2007). "Hydrogen production by molecular photocatalysis". Chemical Reviews. 107 (10): 4022–47. doi:10.1021/cr050193e. PMID 17927155.
  94. ^ Yang C, Messerschmidt M, Coppens P, Omary MA (August 2006). "Trinuclear gold(I) triazolates: a new class of wide-band phosphors and sensors". Inorganic Chemistry. 45 (17): 6592–4. doi:10.1021/ic060943i. PMID 16903710.
  95. ^ Fuentes-Cabrera M, Nicholson DM, Sumpter BG, Widom M (2005). "Electronic structure and properties of isoreticular metal-organic frameworks: The case of M-IRMOF1 (M=Zn, Cd, Be, Mg, and Ca)". The Journal of Chemical Physics. 123 (12): 124713. Bibcode:2005JChPh.123l4713F. doi:10.1063/1.2037587. PMID 16392517.
  96. ^ Gascon J, Hernández-Alonso MD, Almeida AR, van Klink GP, Kapteijn F, Mul G (2008). "Isoreticular MOFs as efficient photocatalysts with tunable band gap: an operando FTIR study of the photoinduced oxidation of propylene". ChemSusChem. 1 (12): 981–3. doi:10.1002/cssc.200800203. PMID 19053135.
  97. ^ Liu, Yufeng; Cheng, Yuan; Zhang, He; Zhou, Min; Yu, Yijun; Lin, Shichao; Jiang, Bo; Zhao, Xiaozhi; Miao, Leiying; Wei, Chuan-Wan; Liu, Quanyi; Lin, Ying-Wu; Du, Yan; Butch, Christopher J.; Wei, Hui (July 1, 2020). "Integrated cascade nanozyme catalyzes in vivo ROS scavenging for anti-inflammatory therapy". Science Advances. 6 (29): eabb2695. Bibcode:2020SciA....6.2695L. doi:10.1126/sciadv.abb2695. PMC 7439611. PMID 32832640.
  98. ^ Hindocha, S.; Poulston, S. (2017). "Study of the scale-up, formulation, ageing and ammonia adsorption capacity of MIL-100(Fe), Cu-BTC and CPO-27(Ni) for use in respiratory protection filters". Faraday Discussions. 201: 113–125. Bibcode:2017FaDi..201..113H. doi:10.1039/c7fd00090a. PMID 28612864.
  99. ^ a b c Redfern, Louis R.; Farha, Omar K. (2019). "Mechanical properties of metal–organic frameworks". Chemical Science. 10 (46): 10666–10679. doi:10.1039/C9SC04249K. PMC 7066669. PMID 32190239.
  100. ^ a b Tan, Jin Chong; Bennett, Thomas D.; Cheetham, Anthony K. (2010-05-17). "Chemical structure, network topology, and porosity effects on the mechanical properties of Zeolitic Imidazolate Frameworks". Proceedings of the National Academy of Sciences. 107 (22): 9938–9943. Bibcode:2010PNAS..107.9938T. doi:10.1073/pnas.1003205107. PMC 2890448. PMID 20479264.
  101. ^ Song, Jianbo; Pallach, Roman; Frentzel‐Beyme, Louis; Kolodzeiski, Pascal; Kieslich, Gregor; Vervoorts, Pia; Hobday, Claire L.; Henke, Sebastian (2022-05-16). "Tuning the High‐Pressure Phase Behaviour of Highly Compressible Zeolitic Imidazolate Frameworks: From Discontinuous to Continuous Pore Closure by Linker Substitution". Angewandte Chemie International Edition. 61 (21): e202117565. doi:10.1002/anie.202117565. PMC 9401003. PMID 35119185.
  102. ^ Moggach, Stephen A.; Bennett, Thomas D.; Cheetham, Anthony K. (2009-09-07). "The Effect of Pressure on ZIF-8: Increasing Pore Size with Pressure and the Formation of a High-Pressure Phase at 1.47 GPa" (PDF). Angewandte Chemie International Edition. 48 (38): 7087–7089. doi:10.1002/anie.200902643. hdl:20.500.11820/198bee14-febb-4c8e-b6b4-eb424b7ebac0. PMID 19681088.
  103. ^ Chapman, Karena W.; Halder, Gregory J.; Chupas, Peter J. (2009-11-16). "Pressure-Induced Amorphization and Porosity Modification in a Metal−Organic Framework". Journal of the American Chemical Society. 131 (48): 17546–17547. doi:10.1021/ja908415z. PMID 19916507.
  104. ^ Ortiz, Aurélie U.; Boutin, Anne; Fuchs, Alain H.; Coudert, François-Xavier (2013-05-20). "Investigating the Pressure-Induced Amorphization of Zeolitic Imidazolate Framework ZIF-8: Mechanical Instability Due to Shear Mode Softening" (PDF). The Journal of Physical Chemistry Letters. 4 (11): 1861–1865. doi:10.1021/jz400880p. PMID 26283122.
  105. ^ Widmer, Remo N.; Lampronti, Giulio I.; Chibani, Siwar; Wilson, Craig W.; Anzellini, Simone; Farsang, Stefan; Kleppe, Annette K.; Casati, Nicola P. M.; MacLeod, Simon G.; Redfern, Simon A. T.; Coudert, François-Xavier (2019-05-22). "Rich Polymorphism of a Metal–Organic Framework in Pressure–Temperature Space". Journal of the American Chemical Society. 141 (23): 9330–9337. doi:10.1021/jacs.9b03234. PMC 7007208. PMID 31117654.
  106. ^ a b Peterson, Gregory W.; DeCoste, Jared B.; Glover, T. Grant; Huang, Yougui; Jasuja, Himanshu; Walton, Krista S. (September 2013). "Effects of pelletization pressure on the physical and chemical properties of the metal–organic frameworks Cu3(BTC)2 and UiO-66". Microporous and Mesoporous Materials. 179: 48–53. doi:10.1016/j.micromeso.2013.02.025.
  107. ^ Heinen, Jurn; Ready, Austin D.; Bennett, Thomas D.; Dubbeldam, David; Friddle, Raymond W.; Burtch, Nicholas C. (2018-06-06). "Elucidating the Variable-Temperature Mechanical Properties of a Negative Thermal Expansion Metal–Organic Framework" (PDF). ACS Applied Materials & Interfaces. 10 (25): 21079–21083. doi:10.1021/acsami.8b06604. PMID 29873475. S2CID 46942254.
  108. ^ Dürholt, Johannes P.; Keupp, Julian; Schmid, and Rochus (2016-07-27). "The Impact of Mesopores on the Mechanical Stability of HKUST-1: A Multiscale Investigation". European Journal of Inorganic Chemistry. 2016 (27): 4517–4523. doi:10.1002/ejic.201600566.
  109. ^ Chapman, Karena W.; Halder, Gregory J.; Chupas, Peter J. (2008-07-18). "Guest-Dependent High Pressure Phenomena in a Nanoporous Metal−Organic Framework Material". Journal of the American Chemical Society. 130 (32): 10524–10526. doi:10.1021/ja804079z. PMID 18636710.
  110. ^ Li, Hailian; Eddaoudi, Mohamed; O'Keeffe, M.; Yaghi, O. M. (November 1999). "Design and synthesis of an exceptionally stable and highly porous metal-organic framework". Nature. 402 (6759): 276–279. Bibcode:1999Natur.402..276L. doi:10.1038/46248. hdl:2027.42/62847. S2CID 4310761.
  111. ^ Alexandre, Simone S.; Mattesini, Maurizio; Soler, José M.; Yndurain, Félix (2006-02-22). "Comment on "Magnetism in Atomic-Size Palladium Contacts and Nanowires"". Physical Review Letters. 96 (7): 079701, author reply 079702. Bibcode:2006PhRvL..96g9701A. doi:10.1103/physrevlett.96.079701. PMID 16606151.
  112. ^ Hu, Yun Hang; Zhang, Lei (2010). "Amorphization of metal-organic framework MOF-5 at unusually low applied pressure". Physical Review B. 81 (17): 174103. Bibcode:2010PhRvB..81q4103H. doi:10.1103/PhysRevB.81.174103.
  113. ^ Ortiz, Aurélie U.; Boutin, A.; Fuchs, Alain H.; Coudert, François-Xavier (2013-05-07). "Metal–organic frameworks with wine-rack motif: What determines their flexibility and elastic properties?" (PDF). The Journal of Chemical Physics. 138 (17): 174703. Bibcode:2013JChPh.138q4703O. doi:10.1063/1.4802770. PMID 23656148.
  114. ^ Serra-Crespo, Pablo; Dikhtiarenko, Alla; Stavitski, Eli; Juan-Alcañiz, Jana; Kapteijn, Freek; Coudert, François-Xavier; Gascon, Jorge (2015). "Experimental evidence of negative linear compressibility in the MIL-53 metal–organic framework family". CrystEngComm. 17 (2): 276–280. doi:10.1039/c4ce00436a. PMC 4338503. PMID 25722647.
  115. ^ Chen, Zhijie; Hanna, Sylvia L.; Redfern, Louis R.; Alezi, Dalal; Islamoglu, Timur; Farha, Omar K. (December 2019). "Corrigendum to "Reticular chemistry in the rational synthesis of functional zirconium cluster-based MOFs" [Coord. Chem. Rev. 386 (2019) 32–49]". Coordination Chemistry Reviews. 400: 213050. doi:10.1016/j.ccr.2019.213050. S2CID 203136239.
  116. ^ Wu, Hui; Yildirim, Taner; Zhou, Wei (2013-03-07). "Exceptional Mechanical Stability of Highly Porous Zirconium Metal–Organic Framework UiO-66 and Its Important Implications". The Journal of Physical Chemistry Letters. 4 (6): 925–930. doi:10.1021/jz4002345. PMID 26291357.
  117. ^ Bennett, Thomas D.; Todorova, Tanya K.; Baxter, Emma F.; Reid, David G.; Gervais, Christel; Bueken, Bart; Van de Voorde, B.; De Vos, Dirk; Keen, David A.; Mellot-Draznieks, Caroline (2016). "Connecting defects and amorphization in UiO-66 and MIL-140 metal–organic frameworks: a combined experimental and computational study". Physical Chemistry Chemical Physics. 18 (3): 2192–2201. arXiv:1510.08220. Bibcode:2016PCCP...18.2192B. doi:10.1039/c5cp06798g. PMID 27144237. S2CID 45890138.
  118. ^ a b Committee on Alternatives and Strategies for Future Hydrogen Production and Use, National Research Council, National Academy of Engineering, eds. (2004). The Hydrogen Economy: Opportunities, Costs, Barriers, and R&D Needs. Washington, D.C.: The National Academies Press. pp. 11–24, 37–44. doi:10.2172/882095. ISBN 978-0-309-09163-3.
  119. ^ Ronneau C (2004-11-29). Energie, pollution de l'air et developpement durable. Louvain-la-Neuve: Presses Universitaires de Louvain. ISBN 9782875581716.
  120. ^ a b Praya Y, Mendoza-Cortes JL (October 2009). "Design Principles for High H
    2
    Storage Using Chelation of Abundant Transition Metals in Covalent Organic Frameworks for 0–700 bar at 298 K". Journal of the American Chemical Society. 46 (138): 15204–15213. doi:10.1021/jacs.6b08803. PMID 27792339. S2CID 21366076.
  121. ^ "DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles". Energy.gov.
  122. ^ Thomas KM (March 2009). "Adsorption and desorption of hydrogen on metal-organic framework materials for storage applications: comparison with other nanoporous materials". Dalton Transactions (9): 1487–505. doi:10.1039/B815583F. PMID 19421589.
  123. ^ Yuan D, Zhao D, Sun D, Zhou HC (July 2010). "An isoreticular series of metal-organic frameworks with dendritic hexacarboxylate ligands and exceptionally high gas-uptake capacity". Angewandte Chemie. 49 (31): 5357–61. doi:10.1002/anie.201001009. PMID 20544763.
  124. ^ Liu J (2 November 2016). "Recent developments in porous materials for H
    2
    and CH
    4
    storage". Tetrahedron Letters. 57 (44): 4873–4881. doi:10.1016/j.tetlet.2016.09.085.
  125. ^ Sumida K, Hill MR, Horike S, Dailly A, Long JR (October 2009). "Synthesis and hydrogen storage properties of Be(12)(OH)(12)(1,3,5-benzenetribenzoate)(4)". Journal of the American Chemical Society. 131 (42): 15120–1. doi:10.1021/ja9072707. PMID 19799422.
  126. ^ a b c d Furukawa H, Ko N, Go YB, Aratani N, Choi SB, Choi E, Yazaydin AO, Snurr RQ, O'Keeffe M, Kim J, Yaghi OM (July 2010). "Ultrahigh porosity in metal-organic frameworks". Science. 329 (5990): 424–8. Bibcode:2010Sci...329..424F. doi:10.1126/science.1192160. PMID 20595583. S2CID 25072457.
  127. ^ Rowsell JL, Millward AR, Park KS, Yaghi OM (May 2004). "Hydrogen sorption in functionalized metal-organic frameworks". Journal of the American Chemical Society. 126 (18): 5666–7. doi:10.1021/ja049408c. PMID 15125649.
  128. ^ Rosi NL, Eckert J, Eddaoudi M, Vodak DT, Kim J, O'Keeffe M, Yaghi OM (May 2003). "Hydrogen storage in microporous metal-organic frameworks". Science. 300 (5622): 1127–9. Bibcode:2003Sci...300.1127R. doi:10.1126/science.1083440. PMID 12750515. S2CID 3025509.
  129. ^ a b Dincă M, Dailly A, Liu Y, Brown CM, Neumann DA, Long JR (December 2006). "Hydrogen storage in a microporous metal-organic framework with exposed Mn2+ coordination sites". Journal of the American Chemical Society. 128 (51): 16876–83. doi:10.1021/ja0656853. PMID 17177438.
  130. ^ Lee J, Li J, Jagiello J (2005). "Gas sorption properties of microporous metal organic frameworks". Journal of Solid State Chemistry. 178 (8): 2527–2532. Bibcode:2005JSSCh.178.2527L. doi:10.1016/j.jssc.2005.07.002.
  131. ^ Rowsell JL, Yaghi OM (July 2005). "Strategies for hydrogen storage in metal—organic frameworks". Angewandte Chemie. 44 (30): 4670–9. doi:10.1002/anie.200462786. PMID 16028207.
  132. ^ Rowsell JL, Yaghi OM (February 2006). "Effects of functionalization, catenation, and variation of the metal oxide and organic linking units on the low-pressure hydrogen adsorption properties of metal-organic frameworks". Journal of the American Chemical Society. 128 (4): 1304–15. doi:10.1021/ja056639q. PMID 16433549.
  133. ^ Garrone E, Bonelli B, Arean CO (2008). "Enthalpy-entropy correlation for hydrogen adsorption on zeolites". Chemical Physics Letters. 456 (1–3): 68–70. Bibcode:2008CPL...456...68G. doi:10.1016/j.cplett.2008.03.014.
  134. ^ Kubas, G. J. (2001). Metal Dihydrogen and s-Bond Complexes: Structure, Theory, and Reactivity. New York: Kluwer Academic.
  135. ^ Bellarosa L, Calero S, López N (May 2012). "Early stages in the degradation of metal-organic frameworks in liquid water from first-principles molecular dynamics". Physical Chemistry Chemical Physics. 14 (20): 7240–5. Bibcode:2012PCCP...14.7240B. doi:10.1039/C2CP40339K. PMID 22513503.
  136. ^ Bellarosa L, Castillo JM, Vlugt T, Calero S, López N (September 2012). "On the mechanism behind the instability of isoreticular metal-organic frameworks (IRMOFs) in humid environments". Chemistry: A European Journal. 18 (39): 12260–6. doi:10.1002/chem.201201212. PMID 22907782.
  137. ^ Sengupta, D; Melix, P; Bose, S; Duncan, J; Wang, X; Mian, MR; Kirlikovali, KO; Joodaki, F; Islamoglu, T; Yildirim, T; Snurr, RQ; Farha, OK (20 September 2023). "Air-Stable Cu(I) Metal-Organic Framework for Hydrogen Storage". Journal of the American Chemical Society. 145 (37): 20492–20502. doi:10.1021/jacs.3c06393. PMID 37672758.
  138. ^ Stern AC, Belof JL, Eddaoudi M, Space B (January 2012). "Understanding hydrogen sorption in a polar metal-organic framework with constricted channels". The Journal of Chemical Physics. 136 (3): 034705. Bibcode:2012JChPh.136c4705S. doi:10.1063/1.3668138. PMID 22280775.
  139. ^ Dolgonos G (2005). "How many hydrogen molecules can be inserted into C60? Comments on the paper 'AM1 treatment of endohedrally hydrogen doped fullerene, nH2@C60' by L. Türker and S. Erkoç'". Journal of Molecular Structure: THEOCHEM. 723 (1–3): 239–241. doi:10.1016/j.theochem.2005.02.017.
  140. ^ Tsao CS, Yu MS, Wang CY, Liao PY, Chen HL, Jeng US, Tzeng YR, Chung TY, Wu HC (February 2009). "Nanostructure and hydrogen spillover of bridged metal-organic frameworks". Journal of the American Chemical Society. 131 (4): 1404–6. doi:10.1021/ja802741b. PMID 19140765.
  141. ^ Mulfort KL, Farha OK, Stern CL, Sarjeant AA, Hupp JT (March 2009). "Post-synthesis alkoxide formation within metal-organic framework materials: a strategy for incorporating highly coordinatively unsaturated metal ions". Journal of the American Chemical Society. 131 (11): 3866–8. doi:10.1021/ja809954r. PMID 19292487.
  142. ^ Huang BL, Ni Z, Millward A, McGaughey AJ, Uher C, Kaviany M, Yaghi O (2007). "Thermal conductivity of a metal-organic framework (MOF-5): Part II. Measurement". International Journal of Heat and Mass Transfer. 50 (3–4): 405–411. doi:10.1016/j.ijheatmasstransfer.2006.10.001.
  143. ^ McQuarrie DA, Simon JD (1997). Physical Chemistry: A Molecular Approach. Sausalito, CA: University Science Books.
  144. ^ Belof JL, Stern AC, Eddaoudi M, Space B (December 2007). "On the mechanism of hydrogen storage in a metal-organic framework material". Journal of the American Chemical Society. 129 (49): 15202–10. doi:10.1021/ja0737164. PMID 17999501.
  145. ^ Belof J, Stern A, Space B (2009). "A predictive model of hydrogen sorption for metal-organic materials". Journal of Physical Chemistry C. 113 (21): 9316–9320. doi:10.1021/jp901988e.
  146. ^ a b c Zhao D, Yan D, Zhou HC (2008). "The current status of hydrogen storage in metal–organic frameworks". Energy & Environmental Science. 1 (2): 225–235. doi:10.1039/b808322n. S2CID 44187103.
  147. ^ Furukawa H, Miller MA, Yaghi OM (2007). "Independent verification of the saturation hydrogen uptake in MOF-177 and establishment of a benchmark for hydrogen adsorption in metal–organic frameworks". Journal of Materials Chemistry. 17 (30): 3197–3204. doi:10.1039/b703608f.
  148. ^ Lowell S, Shields JE, Thomas MA, Thommes M (2004). Characterization of Porous Solids and Powders: Surface Area, Pore Size and Density. Springer Netherlands. doi:10.1007/978-1-4020-2303-3. ISBN 978-90-481-6633-6.
  149. ^ a b Züttel A (April 2004). "Hydrogen storage methods". Die Naturwissenschaften. 91 (4): 157–72. Bibcode:2004NW.....91..157Z. doi:10.1007/s00114-004-0516-x. PMID 15085273. S2CID 6985612.
  150. ^ Zheng W, Tsang CS, Lee LY, Wong KY (June 2019). "Two-dimensional metal-organic framework and covalent-organic framework: synthesis and their energy-related applications". Materials Today Chemistry. 12: 34–60. doi:10.1016/j.mtchem.2018.12.002. hdl:10397/101525. S2CID 139305086.
  151. ^ Wang, Wei; Xu, Xiaomin; Zhou, Wei; Shao, Zongping (April 2017). "Recent Progress in Metal-Organic Frameworks for Applications in Electrocatalytic and Photocatalytic Water Splitting". Advanced Science. 4 (4): 1600371. doi:10.1002/advs.201600371. PMC 5396165. PMID 28435777.
  152. ^ Cheng, Weiren; Zhao, Xu; Su, Hui; Tang, Fumin; Che, Wei; Zhang, Hui; Liu, Qinghua (14 January 2019). "Lattice-strained metal–organic-framework arrays for bifunctional oxygen electrocatalysis". Nature Energy. 4 (2): 115–122. Bibcode:2019NatEn...4..115C. doi:10.1038/s41560-018-0308-8. S2CID 139760912.
  153. ^ Liu, Mengjie; Zheng, Weiran; Ran, Sijia; Boles, Steven T.; Lee, Lawrence Yoon Suk (November 2018). "Overall Water-Splitting Electrocatalysts Based on 2D CoNi-Metal-Organic Frameworks and Its Derivative". Advanced Materials Interfaces. 5 (21): 1800849. doi:10.1002/admi.201800849. hdl:10397/101550. S2CID 104572148.
  154. ^ Zheng, Weiran; Liu, Mengjie; Lee, Lawrence Yoon Suk (2020). "Electrochemical Instability of Metal–Organic Frameworks: In Situ Spectroelectrochemical Investigation of the Real Active Sites". ACS Catalysis. 10: 81–92. doi:10.1021/acscatal.9b03790. hdl:10397/100175. S2CID 212979103.
  155. ^ Zheng, Weiran; Lee, Lawrence Yoon Suk (2021-07-22). "Metal–Organic Frameworks for Electrocatalysis: Catalyst or Precatalyst?". ACS Energy Letters. 6 (8): 2838–2843. doi:10.1021/acsenergylett.1c01350. hdl:10397/100058.
  156. ^ Bünzli JC (May 2010). "Lanthanide luminescence for biomedical analyses and imaging". Chemical Reviews. 110 (5): 2729–55. doi:10.1021/cr900362e. PMID 20151630.
  157. ^ Amoroso AJ, Pope SJ (July 2015). "Using lanthanide ions in molecular bioimaging" (PDF). Chemical Society Reviews. 44 (14): 4723–42. doi:10.1039/c4cs00293h. PMID 25588358.
  158. ^ Rosi NL, Kim J, Eddaoudi M, Chen B, O'Keeffe M, Yaghi OM (February 2005). "Rod packings and metal-organic frameworks constructed from rod-shaped secondary building units". Journal of the American Chemical Society. 127 (5): 1504–18. doi:10.1021/ja045123o. PMID 15686384.
  159. ^ Duan TW, Yan B (2014-06-12). "Hybrids based on lanthanide ions activated yttrium metal–organic frameworks: functional assembly, polymer film preparation and luminescence tuning". J. Mater. Chem. C. 2 (26): 5098–5104. doi:10.1039/c4tc00414k.
  160. ^ Xu H, Cao CS, Kang XM, Zhao B (November 2016). "Lanthanide-based metal-organic frameworks as luminescent probes". Dalton Transactions. 45 (45): 18003–18017. doi:10.1039/c6dt02213h. PMID 27470090.
  161. ^ Lian X, Yan B (2016-01-26). "A lanthanide metal–organic framework (MOF-76) for adsorbing dyes and fluorescence detecting aromatic pollutants". RSC Advances. 6 (14): 11570–11576. Bibcode:2016RSCAd...611570L. doi:10.1039/c5ra23681a.
  162. ^ Pansare V, Hejazi S, Faenza W, Prud'homme RK (March 2012). "Review of Long-Wavelength Optical and NIR Imaging Materials: Contrast Agents, Fluorophores and Multifunctional Nano Carriers". Chemistry of Materials. 24 (5): 812–827. doi:10.1021/cm2028367. PMC 3423226. PMID 22919122.
  163. ^ Della Rocca J, Liu D, Lin W (October 2011). "Nanoscale metal-organic frameworks for biomedical imaging and drug delivery". Accounts of Chemical Research. 44 (10): 957–68. doi:10.1021/ar200028a. PMC 3777245. PMID 21648429.
  164. ^ Luo TY, Liu C, Eliseeva SV, Muldoon PF, Petoud S, Rosi NL (July 2017). "2+ Clusters: Rational Design, Directed Synthesis, and Deliberate Tuning of Excitation Wavelengths". Journal of the American Chemical Society. 139 (27): 9333–9340. doi:10.1021/jacs.7b04532. PMID 28618777.
  165. ^ Foucault-Collet A, Gogick KA, White KA, Villette S, Pallier A, Collet G, Kieda C, Li T, Geib SJ, Rosi NL, Petoud S (October 2013). "Lanthanide near infrared imaging in living cells with Yb3+ nano metal organic frameworks". Proceedings of the National Academy of Sciences of the United States of America. 110 (43): 17199–204. Bibcode:2013PNAS..11017199F. doi:10.1073/pnas.1305910110. PMC 3808657. PMID 24108356.
  166. ^ Li Y, Weng Z, Wang Y, Chen L, Sheng D, Diwu J, Chai Z, Albrecht-Schmitt TE, Wang S (January 2016). "Surprising coordination for low-valent actinides resembling uranyl(vi) in thorium(iv) organic hybrid layered and framework structures based on a graphene-like (6,3) sheet topology". Dalton Transactions. 45 (3): 918–21. doi:10.1039/C5DT04183J. PMID 26672441. S2CID 2618108.
  167. ^ Carboni M, Abney CW, Liu S, Lin W (2013). "Highly porous and stable metal–organic frameworks for uranium extraction". Chemical Science. 4 (6): 2396. doi:10.1039/c3sc50230a.
  168. ^ Wang Y, Li Y, Bai Z, Xiao C, Liu Z, Liu W, Chen L, He W, Diwu J, Chai Z, Albrecht-Schmitt TE, Wang S (November 2015). "Design and synthesis of a chiral uranium-based microporous metal organic framework with high SHG efficiency and sequestration potential for low-valent actinides". Dalton Transactions. 44 (43): 18810–4. doi:10.1039/C5DT02337H. PMID 26459775.
  169. ^ Liu W, Dai X, Bai Z, Wang Y, Yang Z, Zhang L, Xu L, Chen L, Li Y, Gui D, Diwu J, Wang J, Zhou R, Chai Z, Wang S (April 2017). "Highly Sensitive and Selective Uranium Detection in Natural Water Systems Using a Luminescent Mesoporous Metal-Organic Framework Equipped with Abundant Lewis Basic Sites: A Combined Batch, X-ray Absorption Spectroscopy, and First Principles Simulation Investigation". Environmental Science & Technology. 51 (7): 3911–3921. Bibcode:2017EnST...51.3911L. doi:10.1021/acs.est.6b06305. PMID 28271891.
  170. ^ Wang LL, Luo F, Dang LL, Li JQ, Wu XL, Liu SJ, Luo MB (2015). "Ultrafast high-performance extraction of uranium from seawater without pretreatment using an acylamide- and carboxyl-functionalized metal–organic framework". Journal of Materials Chemistry A. 3 (26): 13724–13730. doi:10.1039/C5TA01972A.
  171. ^ Demir S, Brune NK, Van Humbeck JF, Mason JA, Plakhova TV, Wang S, Tian G, Minasian SG, Tyliszczak T, Yaita T, Kobayashi T, Kalmykov SN, Shiwaku H, Shuh DK, Long JR (April 2016). "Extraction of Lanthanide and Actinide Ions from Aqueous Mixtures Using a Carboxylic Acid-Functionalized Porous Aromatic Framework". ACS Central Science. 2 (4): 253–65. doi:10.1021/acscentsci.6b00066. PMC 4850516. PMID 27163056.
  172. ^ Banerjee D, Kim D, Schweiger MJ, Kruger AA, Thallapally PK (May 2016). "Removal of TcO4- ions from solution: materials and future outlook". Chemical Society Reviews. 45 (10): 2724–39. doi:10.1039/C5CS00330J. PMID 26947251.
  173. ^ Li B, Dong X, Wang H, Ma D, Tan K, Jensen S, Deibert BJ, Butler J, Cure J, Shi Z, Thonhauser T, Chabal YJ, Han Y, Li J (September 2017). "Capture of organic iodides from nuclear waste by metal-organic framework-based molecular traps". Nature Communications. 8 (1): 485. Bibcode:2017NatCo...8..485L. doi:10.1038/s41467-017-00526-3. PMC 5589857. PMID 28883637.
  174. ^ Abney CW, Mayes RT, Saito T, Dai S (December 2017). "Materials for the Recovery of Uranium from Seawater". Chemical Reviews. 117 (23): 13935–14013. doi:10.1021/acs.chemrev.7b00355. OSTI 1412046. PMID 29165997.
  175. ^ Wu MX, Yang YW (June 2017). "Metal-Organic Framework (MOF)-Based Drug/Cargo Delivery and Cancer Therapy". Advanced Materials. 29 (23): 1606134. Bibcode:2017AdM....2906134W. doi:10.1002/adma.201606134. PMID 28370555. S2CID 30958347.
  176. ^ Smaldone RA, Forgan RS, Furukawa H, Gassensmith JJ, Slawin AM, Yaghi OM, Stoddart JF (November 2010). "Metal-organic frameworks from edible natural products". Angewandte Chemie. 49 (46): 8630–4. doi:10.1002/anie.201002343. PMID 20715239.
  177. ^ Jambhekar SS, Breen P (February 2016). "Cyclodextrins in pharmaceutical formulations I: structure and physicochemical properties, formation of complexes, and types of complex". Drug Discovery Today. 21 (2): 356–62. doi:10.1016/j.drudis.2015.11.017. PMID 26686054.
  178. ^ Hartlieb KJ, Ferris DP, Holcroft JM, Kandela I, Stern CL, Nassar MS, Botros YY, Stoddart JF (May 2017). "Encapsulation of Ibuprofen in CD-MOF and Related Bioavailability Studies". Molecular Pharmaceutics. 14 (5): 1831–1839. doi:10.1021/acs.molpharmaceut.7b00168. PMID 28355489.
  179. ^ Noorian, Seyyed Abbas; Hemmatinejad, Nahid; Navarro, Jorge A. R. (2019-12-01). "BioMOF@cellulose fabric composites for bioactive molecule delivery". Journal of Inorganic Biochemistry. 201: 110818. doi:10.1016/j.jinorgbio.2019.110818. PMID 31518870. S2CID 202571262.
  180. ^ Rojas S, Colinet I, Cunha D, Hidalgo T, Salles F, Serre C, Guillou N, Horcajada P (March 2018). "Toward Understanding Drug Incorporation and Delivery from Biocompatible Metal-Organic Frameworks in View of Cutaneous Administration". ACS Omega. 3 (3): 2994–3003. doi:10.1021/acsomega.8b00185. PMC 5879486. PMID 29623304.
  181. ^ Noorian, Seyyed Abbas; Hemmatinejad, Nahid; Navarro, Jorge A. R. (2020). "Bioactive molecule encapsulation on metal-organic framework via simple mechanochemical method for controlled topical drug delivery systems". Microporous and Mesoporous Materials. 302: 110199. doi:10.1016/j.micromeso.2020.110199. S2CID 216532709.
  182. ^ Chen X, Tong R, Shi Z, Yang B, Liu H, Ding S, Wang X, Lei Q, Wu J, Fang W (January 2018). "MOF Nanoparticles with Encapsulated Autophagy Inhibitor in Controlled Drug Delivery System for Antitumor". ACS Applied Materials & Interfaces. 10 (3): 2328–2337. doi:10.1021/acsami.7b16522. PMID 29286625.
  183. ^ Talin AA, Centrone A, Ford AC, Foster ME, Stavila V, Haney P, Kinney RA, Szalai V, El Gabaly F, Yoon HP, Léonard F, Allendorf MD (January 2014). "Tunable electrical conductivity in metal-organic framework thin-film devices". Science. 343 (6166): 66–9. Bibcode:2014Sci...343...66T. doi:10.1126/science.1246738. OSTI 1254264. PMID 24310609. S2CID 206552714.
  184. ^ "2D self-assembling semiconductor could beat out graphene". www.gizmag.com. 2 May 2014.
  185. ^ Sheberla D, Sun L, Blood-Forsythe MA, Er S, Wade CR, Brozek CK, Aspuru-Guzik A, Dincă M (June 2014). "High electrical conductivity in Ni3(2,3,6,7,10,11-hexaiminotriphenylene)2, a semiconducting metal-organic graphene analogue". Journal of the American Chemical Society. 136 (25): 8859–62. doi:10.1021/ja502765n. PMID 24750124. S2CID 5714037.
  186. ^ Dong, Renhao; Han, Peng; Arora, Himani; Ballabio, Marco; Karakus, Melike; Zhang, Zhe; Shekhar, Chandra; Adler, Peter; Petkov, Petko St; Erbe, Artur; Mannsfeld, Stefan C. B. (November 2018). "High-mobility band-like charge transport in a semiconducting two-dimensional metal–organic framework". Nature Materials. 17 (11): 1027–1032. Bibcode:2018NatMa..17.1027D. doi:10.1038/s41563-018-0189-z. PMID 30323335. S2CID 53027396.
  187. ^ Arora, Himani; Dong, Renhao; Venanzi, Tommaso; Zscharschuch, Jens; Schneider, Harald; Helm, Manfred; Feng, Xinliang; Cánovas, Enrique; Erbe, Artur (2020). "Demonstration of a Broadband Photodetector Based on a Two-Dimensional Metal–Organic Framework". Advanced Materials. 32 (9): 1907063. Bibcode:2020AdM....3207063A. doi:10.1002/adma.201907063. hdl:11573/1555186. PMID 31975468. S2CID 210882482.
  188. ^ Arora, Himani (June 21, 2021). "Charge transport in two-dimensional materials and their electronic applications (Doctoral dissertation)" (PDF).
  189. ^ Liu, Jingjuan; Song, Xiaoyu; Zhang, Ting; Liu, Shiyong; Wen, Herui; Chen, Long (2021-03-08). "2D Conductive Metal–Organic Frameworks: An Emerging Platform for Electrochemical Energy Storage". Angewandte Chemie. 133 (11): 5672–5684. Bibcode:2021AngCh.133.5672L. doi:10.1002/ange.202006102. S2CID 240999292.
  190. ^ Day, Robert W.; Bediako, D. Kwabena; Rezaee, Mehdi; Parent, Lucas R.; Skorupskii, Grigorii; Arguilla, Maxx Q.; Hendon, Christopher H.; Stassen, Ivo; Gianneschi, Nathan C.; Kim, Philip; Dincă, Mircea (2019-12-26). "Single Crystals of Electrically Conductive Two-Dimensional Metal–Organic Frameworks: Structural and Electrical Transport Properties". ACS Central Science. 5 (12): 1959–1964. doi:10.1021/acscentsci.9b01006. PMC 6936098. PMID 31893225.
  191. ^ Hoppe, Bastian; Hindricks, Karen D. J.; Warwas, Dawid P.; Schulze, Hendrik A.; Mohmeyer, Alexander; Pinkvos, Tim J.; Zailskas, Saskia; Krey, Marc R.; Belke, Christopher; König, Sandra; Fröba, Michael; Haug, Rolf J.; Behrens, Peter (2018). "Graphene-like metal–organic frameworks: morphology control, optimization of thin film electrical conductivity and fast sensing applications". CrystEngComm. 20 (41): 6458–6471. doi:10.1039/C8CE01264D.
  192. ^ Liang K, Ricco R, Doherty CM, Styles MJ, Bell S, Kirby N, Mudie S, Haylock D, Hill AJ, Doonan CJ, Falcaro P (June 2015). "Biomimetic mineralization of metal-organic frameworks as protective coatings for biomacromolecules". Nature Communications. 6: 7240. Bibcode:2015NatCo...6.7240L. doi:10.1038/ncomms8240. PMC 4468859. PMID 26041070.
  193. ^ . MOF Technologies. Archived from the original on 2021-02-27. Retrieved 2021-04-07.
  194. ^ Choi S, Drese JH, Jones CW (2009-09-21). "Adsorbent materials for carbon dioxide capture from large anthropogenic point sources". ChemSusChem. 2 (9): 796–854. doi:10.1002/cssc.200900036. PMID 19731282.
  195. ^ a b Sumida K, Rogow DL, Mason JA, McDonald TM, Bloch ED, Herm ZR, Bae TH, Long JR (February 2012). "Carbon dioxide capture in metal-organic frameworks". Chemical Reviews. 112 (2): 724–81. doi:10.1021/cr2003272. PMID 22204561.
  196. ^ Berger AH, Bhown AS (2011-01-01). "Comparing physisorption and chemisorption solid sorbents for use separating CO2 from flue gas using temperature swing adsorption". Energy Procedia. 10th International Conference on Greenhouse Gas Control Technologies. 4: 562–567. doi:10.1016/j.egypro.2011.01.089.
  197. ^ Smit B, Reimer JA, Oldenburg CM, Bourg IC (2014). Introduction to Carbon Capture and Sequestration. London: Imperial College Press. ISBN 978-1-78326-328-8.
  198. ^ Lesch, David A (2010). Carbon Dioxide Removal from Flue Gas Using Microporous Metal Organic Frameworks (Report). doi:10.2172/1003992. OSTI 1003992.
metal, organic, framework, mofs, class, porous, polymers, consisting, metal, clusters, also, known, sbus, coordinated, organic, ligands, form, three, dimensional, structures, organic, ligands, included, sometimes, referred, struts, linkers, example, being, ben. Metal organic frameworks MOFs are a class of porous polymers consisting of metal clusters also known as SBUs 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 Synthesis of the MIL 101 MOF Each green octahedron consists of one Cr atom in the center and six oxygen atoms red balls at the corners Electron micrograph of a MIL 101 crystal showing its supertetrahedraMore formally a metal organic framework is a potentially porous extended structure made from metal ions and organic linkers An extended structure is a structure whose sub units occur in a constant ratio and are arranged in a repeating pattern MOFs are a subclass of coordination networks which is a coordination compound extending through repeating coordination entities in one dimension but with cross links between two or more individual chains loops or spiro links or a coordination compound extending through repeating coordination entities in two or three dimensions Coordination networks including MOFs further belong to coordination polymers which is a coordination compound with repeating coordination entities extending in one two or three dimensions 1 Most of the MOFs reported in the literature are crystalline compounds but there are also amorphous MOFs 2 and other disordered phases 3 In most cases for MOFs the pores are stable during the elimination of the guest molecules often solvents and could be refilled with other compounds Because of this property MOFs are of interest for the storage of gases such as hydrogen and carbon dioxide Other possible applications of MOFs are in gas purification in gas separation in water remediation 4 in catalysis as conducting solids and as supercapacitors 5 The synthesis and properties of MOFs constitute the primary focus of the discipline called reticular chemistry from Latin reticulum small net 6 In contrast to MOFs covalent organic frameworks COFs are made entirely from light elements H B C N and O with extended structures 7 Contents 1 Structure 2 Synthesis 2 1 General synthesis 2 2 High throughput synthesis 2 2 1 High throughput solvothermal synthesis 2 3 Pseudomorphic replication 2 4 Post synthetic modification 2 4 1 Ligand exchange 2 4 2 Metal exchange 2 4 3 Stratified synthesis 2 4 4 Open coordination sites 3 Composite materials 4 Catalysis 4 1 Design 4 1 1 Metal ions or metal clusters 4 1 2 Functional struts 4 1 3 Encapsulated catalysts 4 1 4 Metal free organic cavity modifiers 4 2 Achiral catalysis 4 2 1 Metals as catalytic sites 4 2 2 Functional linkers as catalytic sites 4 2 3 Entrapment of catalytically active noble metal nanoparticles 4 2 4 Reaction hosts with size selectivity 4 3 Asymmetric catalysis 4 3 1 Homochiral MOFs with interesting functionalities and reagent accessible channels 4 3 2 Postmodification of homochiral MOFs 4 3 3 Homochiral MOFs with precatalysts as building blocks 4 4 Biomimetic design and photocatalysis 5 Mechanical properties 5 1 Zeolitic imidazolate frameworks ZIFs 5 2 Carboxylate based MOFs 5 2 1 HKUST 1 5 2 2 MOF 5 5 2 3 MIL 53 5 3 Zirconium based MOFs 6 Applications 6 1 Hydrogen storage 6 1 1 Design principles 6 1 2 Structural impacts on hydrogen storage capacity 6 1 2 1 Surface area 6 1 2 2 Hydrogen adsorption enthalpy 6 1 2 3 Sensitivity to airborne moisture 6 1 2 4 Pore size 6 1 2 5 Structural defects 6 1 3 Hydrogen adsorption 6 1 4 Determining hydrogen storage capacity 6 1 4 1 Gravimetric method 6 1 4 2 Volumetric method 6 1 5 Other methods of hydrogen storage 6 2 Electrocatalysis 6 3 Biological imaging and sensing 6 4 Nuclear wasteform materials 6 5 Drug delivery systems 6 6 Semiconductors 6 7 Bio mimetic mineralization 6 8 Carbon capture 6 8 1 Adsorbent 6 8 2 Catalyst 6 9 Desalination ion separation 6 10 Gas separation 6 11 Water vapor capture and dehumidification 6 12 Ferroelectrics and multiferroics 7 See also 8 References 9 External linksStructure editMOFs are composed of two main components an inorganic metal cluster often referred to as a secondary building unit or SBU and an organic molecule called a linker For this reason the materials are often referred to as hybrid organic inorganic materials 1 The organic units are typically mono di tri or tetravalent ligands 8 The choice of metal and linker dictates the structure and hence properties of the MOF For example the metal s coordination preference influences the size and shape of pores by dictating how many ligands can bind to the metal and in which orientation Classification of hybrid materials based on dimensionality 9 Dimensionality of Inorganic0 1 2 3Dimensionalityof Organic 0 Molecular Complexes Hybrid Inorganic Chains Hybrid Inorganic Layers 3 D Inorganic Hybrids1 Chain Coordination Polymers Mixed Inorganic Organic Layers Mixed Inorganic Organic 3 D Framework2 Layered Coordination Polymer Mixed Inorganic Organic 3 D Framework3 3 D Coordination PolymersTo describe and organize the structures of MOFs a system of nomenclature has been developed Subunits of a MOF called secondary building units SBUs can be described by topologies common to several structures Each topology also called a net is assigned a symbol consisting of three lower case letters in bold MOF 5 for example has a pcu net Attached to the SBUs are bridging ligands For MOFs typical bridging ligands are di and tricarboxylic acids These ligands typically have rigid backbones Examples are benzene 1 4 dicarboxylic acid BDC or terephthalic acid biphenyl 4 4 dicarboxylic acid BPDC and the tricarboxylic acid trimesic acid nbsp SBUs are often derived from basic zinc acetate structure the acetates being replaced by rigid di and tricarboxylates Synthesis editGeneral synthesis edit The study of MOFs has roots in coordination chemistry and solid state inorganic chemistry but it developed into a new field In addition MOFs are constructed from bridging organic ligands that remain intact throughout the synthesis 10 Zeolite synthesis often makes use of a template Templates are ions that influence the structure of the growing inorganic framework Typical templating ions are quaternary ammonium cations which are removed later In MOFs the framework is templated by the SBU secondary building unit and the organic ligands 11 12 A templating approach that is useful for MOFs intended for gas storage is the use of metal binding solvents such as N N diethylformamide and water In these cases metal sites are exposed when the solvent is evacuated allowing hydrogen to bind at these sites 13 Four developments were particularly important in advancing the chemistry of MOFs 14 1 The geometric principle of construction where metal containing units were kept in rigid shapes Early MOFs contained single atoms linked to ditopic coordinating linkers The approach not only led to the identification of a small number of preferred topologies that could be targeted in designed synthesis but was the central point to achieve a permanent porosity 2 The use of the isoreticular principle where the size and the nature of a structure changes without changing its topology led to MOFs with ultrahigh porosity and unusually large pore openings 3 Post synthetic modification of MOFs increased their functionality by reacting organic units and metal organic complexes with linkers 4 Multifunctional MOFs incorporated multiple functionalities in a single framework Since ligands in MOFs typically bind reversibly the slow growth of crystals often allows defects to be redissolved resulting in a material with millimeter scale crystals and a near equilibrium defect density Solvothermal synthesis is useful for growing crystals suitable to structure determination because crystals grow over the course of hours to days However the use of MOFs as storage materials for consumer products demands an immense scale up of their synthesis Scale up of MOFs has not been widely studied though several groups have demonstrated that microwaves can be used to nucleate MOF crystals rapidly from solution 15 16 This technique termed microwave assisted solvothermal synthesis is widely used in the zeolite literature 10 and produces micron scale crystals in a matter of seconds to minutes 15 16 in yields similar to the slow growth methods Some MOFs such as the mesoporous MIL 100 Fe 17 can be obtained under mild conditions at room temperature and in green solvents water ethanol through scalable synthesis methods A solvent free synthesis of a range of crystalline MOFs has been described 18 Usually the metal acetate and the organic proligand are mixed and ground up with a ball mill Cu3 BTC 2 can be quickly synthesised in this way in quantitative yield In the case of Cu3 BTC 2 the morphology of the solvent free synthesised product was the same as the industrially made Basolite C300 It is thought that localised melting of the components due to the high collision energy in the ball mill may assist the reaction The formation of acetic acid as a by product in the reactions in the ball mill may also help in the reaction having a solvent effect 19 in the ball mill It has been shown that the addition of small quantities of ethanol for the mechanochemical synthesis of Cu3 BTC 2 significantly reduces the amounts of structural defects in the obtained material 20 A recent advancement in the solvent free preparation of MOF films and composites is their synthesis by chemical vapor deposition This process MOF CVD 21 was first demonstrated for ZIF 8 and consists of two steps In a first step metal oxide precursor layers are deposited In the second step these precursor layers are exposed to sublimed ligand molecules that induce a phase transformation to the MOF crystal lattice Formation of water during this reaction plays a crucial role in directing the transformation This process was successfully scaled up to an integrated cleanroom process conforming to industrial microfabrication standards 22 Numerous methods have been reported for the growth of MOFs as oriented thin films However these methods are suitable only for the synthesis of a small number of MOF topologies One such example being the vapor assisted conversion VAC which can be used for the thin film synthesis of several UiO type MOFs 23 High throughput synthesis edit High throughput HT methods are a part of combinatorial chemistry and a tool for increasing efficiency There are two synthetic strategies within the HT methods In the combinatorial approach all reactions take place in one vessel which leads to product mixtures In the parallel synthesis the reactions take place in different vessels Furthermore a distinction is made between thin films and solvent based methods 24 Solvothermal synthesis can be carried out conventionally in a teflon reactor in a convection oven or in glass reactors in a microwave oven high throughput microwave synthesis The use of a microwave oven changes in part dramatically the reaction parameters In addition to solvothermal synthesis there have been advances in using supercritical fluid as a solvent in a continuous flow reactor Supercritical water was first used in 2012 to synthesize copper and nickel based MOFs in just seconds 25 In 2020 supercritical carbon dioxide was used in a continuous flow reactor along the same time scale as the supercritical water based method but the lower critical point of carbon dioxide allowed for the synthesis of the zirconium based MOF UiO 66 26 High throughput solvothermal synthesis edit In high throughput solvothermal synthesis a solvothermal reactor with e g 24 cavities for teflon reactors is used Such a reactor is sometimes referred to as a multiclav The reactor block or reactor insert is made of stainless steel and contains 24 reaction chambers which are arranged in four rows With the miniaturized teflon reactors volumes of up to 2 mL can be used The reactor block is sealed in a stainless steel autoclave for this purpose the filled reactors are inserted into the bottom of the reactor the teflon reactors are sealed with two teflon films and the reactor top side is put on The autoclave is then closed in a hydraulic press The sealed solvothermal reactor can then be subjected to a temperature time program The reusable teflon film serves to withstand the mechanical stress while the disposable teflon film seals the reaction vessels After the reaction the products can be isolated and washed in parallel in a vacuum filter device On the filter paper the products are then present separately in a so called sample library and can subsequently be characterized by automated X ray powder diffraction The informations obtained are then used to plan further syntheses 27 Pseudomorphic replication edit Pseudomorphic mineral replacement events occur whenever a mineral phase comes into contact with a fluid with which it is out of equilibrium Re equilibration will tend to take place to reduce the free energy and transform the initial phase into a more thermodynamically stable phase involving dissolution and reprecipitation subprocesses 28 29 Inspired by such geological processes MOF thin films can be grown through the combination of atomic layer deposition ALD of aluminum oxide onto a suitable substrate e g FTO and subsequent solvothermal microwave synthesis The aluminum oxide layer serves both as an architecture directing agent and as a metal source for the backbone of the MOF structure 30 The construction of the porous 3D metal organic framework takes place during the microwave synthesis when the atomic layer deposited substrate is exposed to a solution of the requisite linker in a DMF H2O 3 1 mixture v v at elevated temperature Analogous Kornienko and coworkers described in 2015 the synthesis of a cobalt porphyrin MOF Al2 OH 2TCPP Co TCPP H2 4 4 4 4 porphyrin 5 10 15 20 tetrayl tetrabenzoate the first MOF catalyst constructed for the electrocatalytic conversion of aqueous CO2 to CO 31 Post synthetic modification edit Although the three dimensional structure and internal environment of the pores can be in theory controlled through proper selection of nodes and organic linking groups the direct synthesis of such materials with the desired functionalities can be difficult due to the high sensitivity of MOF systems Thermal and chemical sensitivity as well as high reactivity of reaction materials can make forming desired products challenging to achieve The exchange of guest molecules and counter ions and the removal of solvents allow for some additional functionality but are still limited to the integral parts of the framework 32 The post synthetic exchange of organic linkers and metal ions is an expanding area of the field and opens up possibilities for more complex structures increased functionality and greater system control 32 33 Ligand exchange edit Post synthetic modification techniques can be used to exchange an existing organic linking group in a prefabricated MOF with a new linker by ligand exchange or partial ligand exchange 33 34 This exchange allows for the pores and in some cases the overall framework of MOFs to be tailored for specific purposes Some of these uses include fine tuning the material for selective adsorption gas storage and catalysis 33 13 To perform ligand exchange prefabricated MOF crystals are washed with solvent and then soaked in a solution of the new linker The exchange often requires heat and occurs on the time scale of a few days 34 Post synthetic ligand exchange also enables the incorporation of functional groups into MOFs that otherwise would not survive MOF synthesis due to temperature pH or other reaction conditions or hinder the synthesis itself by competition with donor groups on the loaning ligand 33 Metal exchange edit Post synthetic modification techniques can also be used to exchange an existing metal ion in a prefabricated MOF with a new metal ion by metal ion exchange The complete metal metathesis from an integral part of the framework has been achieved without altering the framework or pore structure of the MOF Similarly to post synthetic ligand exchange post synthetic metal exchange is performed by washing prefabricated MOF crystals with solvent and then soaking the crystal in a solution of the new metal 35 Post synthetic metal exchange allows for a simple route to the formation of MOFs with the same framework yet different metal ions 32 Stratified synthesis edit In addition to modifying the functionality of the ligands and metals themselves post synthetic modification can be used to expand upon the structure of the MOF Using post synthetic modification MOFs can be converted from a highly ordered crystalline material toward a heterogeneous porous material 36 Using post synthetic techniques it is possible for the controlled installation of domains within a MOF crystal which exhibit unique structural and functional characteristics Core shell MOFs and other layered MOFs have been prepared where layers have unique functionalization but in most cases are crystallographically compatible from layer to layer 37 Open coordination sites edit In some cases MOF metal nodes have an unsaturated environment and it is possible to modify this environment using different techniques If the size of the ligand matches the size of the pore aperture it is possible to install additional ligands to existing MOF structure 38 39 Sometimes metal nodes have a good binding affinity for inorganic species For instance it was shown that metal nodes can perform an extension and create a bond with the uranyl cation 40 Composite materials editAnother approach to increasing adsorption in MOFs is to alter the system in such a way that chemisorption becomes possible This functionality has been introduced by making a composite material which contains a MOF and a complex of platinum with activated carbon In an effect known as hydrogen spillover H2 can bind to the platinum surface through a dissociative mechanism which cleaves the hydrogen molecule into two hydrogen atoms and enables them to travel down the activated carbon onto the surface of the MOF This innovation produced a threefold increase in the room temperature storage capacity of a MOF however desorption can take upwards of 12 hours and reversible desorption is sometimes observed for only two cycles 41 42 The relationship between hydrogen spillover and hydrogen storage properties in MOFs is not well understood but may prove relevant to hydrogen storage Catalysis edit nbsp Electron micrograph and structure of MIL 101 Color codes red oxygen brown carbon blue chromium MOFs have potential as heterogeneous catalysts although applications have not been commercialized 43 Their high surface area tunable porosity diversity in metal and functional groups make them especially attractive for use as catalysts Zeolites are extraordinarily useful in catalysis 44 Zeolites are limited by the fixed tetrahedral coordination of the Si Al connecting points and the two coordinated oxide linkers Fewer than 200 zeolites are known In contrast with this limited scope MOFs exhibit more diverse coordination geometries polytopic linkers and ancillary ligands F OH H2O among others It is also difficult to obtain zeolites with pore sizes larger than 1 nm which limits the catalytic applications of zeolites to relatively small organic molecules typically no larger than xylenes Furthermore mild synthetic conditions typically employed for MOF synthesis allow direct incorporation of delicate functionalities into the framework structures Such a process would not be possible with zeolites or other microporous crystalline oxide based materials because of the harsh conditions typically used for their synthesis e g calcination at high temperatures to remove organic templates Metal organic framework MIL 101 is one of the most used MOFs for catalysis incorporating different transition metals such as Cr 45 However the stability of some MOF photocatalysts in aqueous medium and under strongly oxidizing conditions is very low 46 47 Zeolites still cannot be obtained in enantiopure form which precludes their applications in catalytic asymmetric synthesis e g for the pharmaceutical agrochemical and fragrance industries Enantiopure chiral ligands or their metal complexes have been incorporated into MOFs to lead to efficient asymmetric catalysts Even some MOF materials may bridge the gap between zeolites and enzymes when they combine isolated polynuclear sites dynamic host guest responses and a hydrophobic cavity environment MOFs might be useful for making semi conductors Theoretical calculations show that MOFs are semiconductors or insulators with band gaps between 1 0 and 5 5 eV which can be altered by changing the degree of conjugation in the ligands indicating its possibility for being photocatalysts Design edit nbsp Example of MOF 5Like other heterogeneous catalysts MOFs may allow for easier post reaction separation and recyclability than homogeneous catalysts In some cases they also give a highly enhanced catalyst stability Additionally they typically offer substrate size selectivity Nevertheless while clearly important for reactions in living systems selectivity on the basis of substrate size is of limited value in abiotic catalysis as reasonably pure feedstocks are generally available Metal ions or metal clusters edit nbsp Example of zeolite catalystAmong the earliest reports of MOF based catalysis was the cyanosilylation of aldehydes by a 2D MOF layered square grids of formula Cd 4 4 bpy 2 NO3 2 48 This investigation centered mainly on size and shape selective clathration A second set of examples was based on a two dimensional square grid MOF containing single Pd II ions as nodes and 2 hydroxypyrimidinolates as struts 49 Despite initial coordinative saturation the palladium centers in this MOF catalyze alcohol oxidation olefin hydrogenation and Suzuki C C coupling At a minimum these reactions necessarily entail redox oscillations of the metal nodes between Pd II and Pd 0 intermediates accompanying by drastic changes in coordination number which would certainly lead to destabilization and potential destruction of the original framework if all the Pd centers are catalytically active The observation of substrate shape and size selectivity implies that the catalytic reactions are heterogeneous and are indeed occurring within the MOF Nevertheless at least for hydrogenation it is difficult to rule out the possibility that catalysis is occurring at the surface of MOF encapsulated palladium clusters nanoparticles i e partial decomposition sites or defect sites rather than at transiently labile but otherwise intact single atom MOF nodes Opportunistic MOF based catalysis has been described for the cubic compound MOF 5 50 This material comprises coordinatively saturated Zn4O nodes and fully complexed BDC struts see above for abbreviation yet it apparently catalyzes the Friedel Crafts tert butylation of both toluene and biphenyl Furthermore para alkylation is strongly favored over ortho alkylation a behavior thought to reflect the encapsulation of reactants by the MOF Functional struts edit The porous framework material Cu3 btc 2 H2O 3 also known as HKUST 1 51 contains large cavities having windows of diameter 6 A The coordinated water molecules are easily removed leaving open Cu II sites Kaskel and co workers showed that these Lewis acid sites could catalyze the cyanosilylation of benzaldehyde or acetone The anhydrous version of HKUST 1 is an acid catalyst 52 Compared to Bronsted vs Lewis acid catalyzed pathways the product selectivity are distinctive for three reactions isomerization of a pinene oxide cyclization of citronellal and rearrangement of a bromoacetals indicating that indeed Cu3 btc 2 functions primarily as a Lewis acid catalyst The product selectivity and yield of catalytic reactions e g cyclopropanation have also been shown to be impacted by defective sites such as Cu I or incompletely deprotonated carboxylic acid moities of the linkers 20 MIL 101 a large cavity MOF having the formula Cr3F H2O 2O BDC 3 is a cyanosilylation catalyst 53 The coordinated water molecules in MIL 101 are easily removed to expose Cr III sites As one might expect given the greater Lewis acidity of Cr III vs Cu II MIL 101 is much more active than HKUST 1 as a catalyst for the cyanosilylation of aldehydes Additionally the Kaskel group observed that the catalytic sites of MIL 101 in contrast to those of HKUST 1 are immune to unwanted reduction by benzaldehyde The Lewis acid catalyzed cyanosilylation of aromatic aldehydes has also been carried out by Long and co workers using a MOF of the formula Mn3 Mn4Cl 3BTT8 CH3OH 10 54 This material contains a three dimensional pore structure with the pore diameter equaling 10 A In principle either of the two types of Mn II sites could function as a catalyst Noteworthy features of this catalyst are high conversion yields for small substrates and good substrate size selectivity consistent with channellocalized catalysis Encapsulated catalysts edit The MOF encapsulation approach invites comparison to earlier studies of oxidative catalysis by zeolite encapsulated Fe porphyrin 55 as well as Mn porphyrin 56 systems The zeolite studies generally employed iodosylbenzene PhIO rather than TPHP as oxidant The difference is likely mechanistically significant thus complicating comparisons Briefly PhIO is a single oxygen atom donor while TBHP is capable of more complex behavior In addition for the MOF based system it is conceivable that oxidation proceeds via both oxygen transfer from a manganese oxo intermediate as well as a manganese initiated radical chain reaction pathway Regardless of mechanism the approach is a promising one for isolating and thereby stabilizing the porphyrins against both oxo bridged dimer formation and oxidative degradation 57 Metal free organic cavity modifiers edit Most examples of MOF based catalysis make use of metal ions or atoms as active sites Among the few exceptions are two nickel and two copper containing MOFs synthesized by Rosseinsky and co workers 58 These compounds employ amino acids L or D aspartate together with dipyridyls as struts The coordination chemistry is such that the amine group of the aspartate cannot be protonated by added HCl but one of the aspartate carboxylates can Thus the framework incorporated amino acid can exist in a form that is not accessible for the free amino acid While the nickel based compounds are marginally porous on account of tiny channel dimensions the copper versions are clearly porous The Rosseinsky group showed that the carboxylic acids behave as Bronsted acidic catalysts facilitating in the copper cases the ring opening methanolysis of a small cavity accessible epoxide at up to 65 yield Superior homogeneous catalysts exist however Kitagawa and co workers have reported the synthesis of a catalytic MOF having the formula Cd 4 BTAPA 2 NO3 2 59 The MOF is three dimensional consisting of an identical catenated pair of networks yet still featuring pores of molecular dimensions The nodes consist of single cadmium ions octahedrally ligated by pyridyl nitrogens From a catalysis standpoint however the most interesting feature of this material is the presence of guest accessible amide functionalities The amides are capable of base catalyzing the Knoevenagel condensation of benzaldehyde with malononitrile Reactions with larger nitriles however are only marginally accelerated implying that catalysis takes place chiefly within the material s channels rather than on its exterior A noteworthy finding is the lack of catalysis by the free strut in homogeneous solution evidently due to intermolecular H bonding between bptda molecules Thus the MOF architecture elicits catalytic activity not otherwise encountered In an interesting alternative approach Ferey and coworkers were able to modify the interior of MIL 101 via Cr III coordination of one of the two available nitrogen atoms of each of several ethylenediamine molecules 60 The free non coordinated ends of the ethylenediamines were then used as Bronsted basic catalysts again for Knoevenagel condensation of benzaldehyde with nitriles A third approach has been described by Kim Kimoon and coworkers 61 Using a pyridine functionalized derivative of tartaric acid and a Zn II source they were able to synthesize a 2D MOF termed POST 1 POST 1 possesses 1D channels whose cross sections are defined by six trinuclear zinc clusters and six struts While three of the six pyridines are coordinated by zinc ions the remaining three are protonated and directed toward the channel interior When neutralized the noncoordinated pyridyl groups are found to catalyze transesterification reactions presumably by facilitating deprotonation of the reactant alcohol The absence of significant catalysis when large alcohols are employed strongly suggests that the catalysis occurs within the channels of the MOF Achiral catalysis edit nbsp Schematic Diagram for MOF CatalysisMetals as catalytic sites edit The metals in the MOF structure often act as Lewis acids The metals in MOFs often coordinate to labile solvent molecules or counter ions which can be removed after activation of the framework The Lewis acidic nature of such unsaturated metal centers can activate the coordinated organic substrates for subsequent organic transformations The use of unsaturated metal centers was demonstrated in the cyanosilylation of aldehydes and imines by Fujita and coworkers in 2004 62 They reported MOF of composition Cd 4 4 bpy 2 H2O 2 NO3 2 4H2O which was obtained by treating linear bridging ligand 4 4 bipyridine bpy with Cd NO3 2 The Cd II centers in this MOF possess a distorted octahedral geometry having four pyridines in the equatorial positions and two water molecules in the axial positions to form a two dimensional infinite network On activation two water molecules were removed leaving the metal centers unsaturated and Lewis acidic The Lewis acidic character of metal center was tested on cyanosilylation reactions of imine where the imine gets attached to the Lewis acidic metal centre resulting in higher electrophilicity of imines For the cyanosilylation of imines most of the reactions were complete within 1 h affording aminonitriles in quantitative yield Kaskel and coworkers 63 carried out similar cyanosilylation reactions with coordinatively unsaturated metals in three dimensional 3D MOFs as heterogeneous catalysts The 3D framework Cu3 btc 2 H2O 3 btc benzene 1 3 5 tricarboxylate HKUST 1 used in this study was first reported by Williams et al 64 The open framework of Cu3 btc 2 H2O 3 is built from dimeric cupric tetracarboxylate units paddle wheels with aqua molecules coordinating to the axial positions and btc bridging ligands The resulting framework after removal of two water molecules from axial positions possesses porous channel This activated MOF catalyzes the trimethylcyanosilylation of benzaldehydes with a very low conversion lt 5 in 24 h at 293 K As the reaction temperature was raised to 313 K a good conversion of 57 with a selectivity of 89 was obtained after 72 h In comparison less than 10 conversion was observed for the background reaction without MOF under the same conditions But this strategy suffers from some problems like 1 the decomposition of the framework with increase of the reaction temperature due to the reduction of Cu II to Cu I by aldehydes 2 strong solvent inhibition effect electron donating solvents such as THF competed with aldehydes for coordination to the Cu II sites and no cyanosilylation product was observed in these solvents 3 the framework instability in some organic solvents Several other groups have also reported the use of metal centres in MOFs as catalysts 54 65 Again electron deficient nature of some metals and metal clusters makes the resulting MOFs efficient oxidation catalysts Mori and coworkers 66 reported MOFs with Cu2 paddle wheel units as heterogeneous catalysts for the oxidation of alcohols The catalytic activity of the resulting MOF was examined by carrying out alcohol oxidation with H2O2 as the oxidant It also catalyzed the oxidation of primary alcohol secondary alcohol and benzyl alcohols with high selectivity Hill et al 67 have demonstrated the sulfoxidation of thioethers using a MOF based on vanadium oxo cluster V6O13 building units Functional linkers as catalytic sites edit Functional linkers can be also utilized as catalytic sites A 3D MOF Cd 4 BTAPA 2 NO3 2 6H2O 2DMF 4 BTAPA 1 3 5 benzene tricarboxylic acid tris N 4 pyridyl amide DMF N N dimethylformamide constructed by tridentate amide linkers and cadmium salt catalyzes the Knoevenagel condensation reaction 59 The pyridine groups on the ligand 4 BTAPA act as ligands binding to the octahedral cadmium centers while the amide groups can provide the functionality for interaction with the incoming substrates Specifically the NH moiety of the amide group can act as electron acceptor whereas the C O group can act as electron donor to activate organic substrates for subsequent reactions Ferey et al 68 reported a robust and highly porous MOF Cr3 m3 O F H2O 2 BDC 3 BDC benzene 1 4 dicarboxylate where instead of directly using the unsaturated Cr III centers as catalytic sites the authors grafted ethylenediamine ED onto the Cr III sites The uncoordinated ends of ED can act as base catalytic sites ED grafted MOF was investigated for Knoevenagel condensation reactions A significant increase in conversion was observed for ED grafted MOF compared to untreated framework 98 vs 36 Another example of linker modification to generate catalytic site is iodo functionalized well known Al based MOFs MIL 53 and DUT 5 and Zr based MOFs UiO 66 and UiO 67 for the catalytic oxidation of diols 69 70 Entrapment of catalytically active noble metal nanoparticles edit The entrapment of catalytically active noble metals can be accomplished by grafting on functional groups to the unsaturated metal site on MOFs Ethylenediamine ED has been shown to be grafted on the Cr metal sites and can be further modified to encapsulate noble metals such as Pd 60 The entrapped Pd has similar catalytic activity as Pd C in the Heck reaction Ruthenium nanoparticles have catalytic activity in a number of reactions when entrapped in the MOF 5 framework 71 This Ru encapsulated MOF catalyzes oxidation of benzyl alcohol to benzaldehyde although degradation of the MOF occurs The same catalyst was used in the hydrogenation of benzene to cyclohexane In another example Pd nanoparticles embedded within defective HKUST 1 framework enable the generation of tunable Lewis basic sites 72 Therefore this multifunctional Pd MOF composite is able to perform stepwise benzyl alcohol oxidation and Knoevenagel condensation Reaction hosts with size selectivity edit MOFs might prove useful for both photochemical and polymerization reactions due to the tuneability of the size and shape of their pores A 3D MOF Co bpdc 3 bpy 4DMF H2O bpdc biphenyldicarboxylate bpy 4 4 bipyridine was synthesized by Li and coworkers 73 Using this MOF photochemistry of o methyl dibenzyl ketone o MeDBK was extensively studied This molecule was found to have a variety of photochemical reaction properties including the production of cyclopentanol MOFs have been used to study polymerization in the confined space of MOF channels Polymerization reactions in confined space might have different properties than polymerization in open space Styrene divinylbenzene substituted acetylenes methyl methacrylate and vinyl acetate have all been studied by Kitagawa and coworkers as possible activated monomers for radical polymerization 74 75 Due to the different linker size the MOF channel size could be tunable on the order of roughly 25 and 100 A2 The channels were shown to stabilize propagating radicals and suppress termination reactions when used as radical polymerization sites Asymmetric catalysis edit Several strategies exist for constructing homochiral MOFs Crystallization of homochiral MOFs via self resolution from achiral linker ligands is one of the way to accomplish such a goal However the resulting bulk samples contain both enantiomorphs and are racemic Aoyama and coworkers 76 successfully obtained homochiral MOFs in the bulk from achiral ligands by carefully controlling nucleation in the crystal growth process Zheng and coworkers 77 reported the synthesis of homochiral MOFs from achiral ligands by chemically manipulating the statistical fluctuation of the formation of enantiomeric pairs of crystals Growing MOF crystals under chiral influences is another approach to obtain homochiral MOFs using achiral linker ligands Rosseinsky and coworkers 78 79 have introduced a chiral coligand to direct the formation of homochiral MOFs by controlling the handedness of the helices during the crystal growth Morris and coworkers 80 utilized ionic liquid with chiral cations as reaction media for synthesizing MOFs and obtained homochiral MOFs The most straightforward and rational strategy for synthesizing homochiral MOFs is however to use the readily available chiral linker ligands for their construction Homochiral MOFs with interesting functionalities and reagent accessible channels edit Homochiral MOFs have been made by Lin and coworkers using 2 2 bis diphenylphosphino 1 1 binaphthyl BINAP and 1 1 bi 2 2 naphthol BINOL as chiral ligands 81 These ligands can coordinate with catalytically active metal sites to enhance the enantioselectivity A variety of linking groups such as pyridine phosphonic acid and carboxylic acid can be selectively introduced to the 3 3 4 4 and the 6 6 positions of the 1 1 binaphthyl moiety Moreover by changing the length of the linker ligands the porosity and framework structure of the MOF can be selectively tuned Postmodification of homochiral MOFs edit Lin and coworkers have shown that the postmodification of MOFs can be achieved to produce enantioselective homochiral MOFs for use as catalysts 82 The resulting 3D homochiral MOF Cd3 L 3Cl6 4DMF 6MeOH 3H2O L R 6 6 dichloro 2 2 dihydroxyl 1 1 binaphthyl bipyridine synthesized by Lin was shown to have a similar catalytic efficiency for the diethylzinc addition reaction as compared to the homogeneous analogue when was pretreated by Ti OiPr 4 to generate the grafted Ti BINOLate species The catalytic activity of MOFs can vary depending on the framework structure Lin and others found that MOFs synthesized from the same materials could have drastically different catalytic activities depending on the framework structure present 83 Homochiral MOFs with precatalysts as building blocks edit Another approach to construct catalytically active homochiral MOFs is to incorporate chiral metal complexes which are either active catalysts or precatalysts directly into the framework structures For example Hupp and coworkers 84 have combined a chiral ligand and bpdc bpdc biphenyldicarboxylate with Zn NO3 2 and obtained twofold interpenetrating 3D networks The orientation of chiral ligand in the frameworks makes all Mn III sites accessible through the channels The resulting open frameworks showed catalytic activity toward asymmetric olefin epoxidation reactions No significant decrease of catalyst activity was observed during the reaction and the catalyst could be recycled and reused several times Lin and coworkers 85 have reported zirconium phosphonate derived Ru BINAP systems Zirconium phosphonate based chiral porous hybrid materials containing the Ru BINAP diamine Cl2 precatalysts showed excellent enantioselectivity up to 99 2 ee in the asymmetric hydrogenation of aromatic ketones Biomimetic design and photocatalysis edit Some MOF materials may resemble enzymes when they combine isolated polynuclear sites dynamic host guest responses and hydrophobic cavity environment which are characteristics of an enzyme 86 Some well known examples of cooperative catalysis involving two metal ions in biological systems include the diiron sites in methane monooxygenase dicopper in cytochrome c oxidase and tricopper oxidases which have analogy with polynuclear clusters found in the 0D coordination polymers such as binuclear Cu2 paddlewheel units found in MOP 1 87 88 and Cu3 btc 2 btc benzene 1 3 5 tricarboxylate in HKUST 1 or trinuclear units such as Fe3O CO2 6 in MIL 88 89 and IRMOP 51 90 Thus 0D MOFs have accessible biomimetic catalytic centers In enzymatic systems protein units show molecular recognition high affinity for specific substrates It seems that molecular recognition effects are limited in zeolites by the rigid zeolite structure 91 In contrast dynamic features and guest shape response make MOFs more similar to enzymes Indeed many hybrid frameworks contain organic parts that can rotate as a result of stimuli such as light and heat 92 The porous channels in MOF structures can be used as photocatalysis sites In photocatalysis the use of mononuclear complexes is usually limited either because they only undergo single electron process or from the need for high energy irradiation In this case binuclear systems have a number of attractive features for the development of photocatalysts 93 For 0D MOF structures polycationic nodes can act as semiconductor quantum dots which can be activated upon photostimuli with the linkers serving as photon antennae 94 Theoretical calculations show that MOFs are semiconductors or insulators with band gaps between 1 0 and 5 5 eV which can be altered by changing the degree of conjugation in the ligands 95 Experimental results show that the band gap of IRMOF type samples can be tuned by varying the functionality of the linker 96 An integrated MOF nanozyme was developed for anti inflammation therapy 97 Mechanical properties editImplementing MOFs in industry necessitates a thorough understanding of the mechanical properties since most processing techniques e g extrusion and pelletization expose the MOFs to substantial mechanical compressive stresses 98 The mechanical response of porous structures is of interest as these structures can exhibit unusual response to high pressures While zeolites microporous aluminosilicate minerals can give some insights into the mechanical response of MOFs the presence of organic linkers as opposed to zeolites makes for novel mechanical responses 99 MOFs are very structurally diverse meaning that it is challenging to classify all of their mechanical properties Additionally variability in MOFs from batch to batch and extreme experimental conditions diamond anvil cells mean that experimental determination of mechanical response to loading is limited however many computational models have been made to determine structure property relationships Main MOF systems that have been explored are zeolitic imidazolate frameworks ZIFs Carboxylate MOFs Zirconium based MOFs among others 99 Generally the MOFs undergo three processes under compressive loading which is relevant in a processing context amorphization hyperfilling and or pressure induced phase transitions During amorphization linkers buckle and the internal porosity within the MOF collapses During hyperfilling the MOF which is being hydrostatically compressed in a liquid typically solvent will expand rather than contract due to a filling of pores with the loading media Finally pressure induced phase transitions where the structure of the crystal is altered during the loading are possible The response of the MOF is predominantly dependent on the linker species and the inorganic nodes Zeolitic imidazolate frameworks ZIFs edit Several different mechanical phenomena have been observed in zeolitic imidazolate frameworks ZIFs the most widely studied MOF for mechanical properties due to their many similarities to zeolites 99 General trends for the ZIF family are the tendency of the Young s modulus and hardness of the ZIFs to decrease as the accessible pore volume increases 100 The bulk moduli of ZIF 62 series increase with the increasing of benzoimidazolate bim concentration ZIF 62 shows a continuous phase transition from open pore op to close pore cp phase when bim concentration is over 0 35 per formular unit The accessible pore size and volume of ZIF 62 bim0 35 can be precisely tuned by applying adequate pressures 101 Another study has shown that under hydrostatic loading in solvent the ZIF 8 material expands as opposed to contracting This is a result of hyperfilling of the internal pores with solvent 102 A computational study demonstrated that ZIF 4 and ZIF 8 materials undergo a shear softening mechanism with amorphizing at 0 34 GPa of the material under hydrostatic loading while still possessing a bulk modulus on the order of 6 5 GPa 103 104 Additionally the ZIF 4 and ZIF 8 MOFs are subject to many pressure dependent phase transitions 100 105 Carboxylate based MOFs edit Main article Carboxylate based metal organic frameworks Carboxylate MOFs come in many forms and have been widely studied Herein HKUST 1 MOF 5 and the MIL series are discussed as representative examples of the carboxylate MOF class HKUST 1 edit HKUST 1 consists of a dimeric Cu paddlewheel that possesses two pore types Under pelletization MOFs such as HKUST 1 exhibit a pore collapse 106 Although most carboxylate MOFs have a negative thermal expansion they densify during heating it was found that the hardness and Young s moduli unexpectedly decrease with increasing temperature from disordering of linkers 107 It was also found computationally that a more mesoporous structure has a lower bulk modulus However an increased bulk modulus was observed in systems with a few large mesopores versus many small mesopores even though both pore size distributions had the same total pore volume 108 The HKUST 1 shows a similar hyperfilling phenomenon to the ZIF structures under hydrostatic loading 109 MOF 5 edit MOF 5 has tetranuclear nodes in an octahedral configuration with an overall cubic structure MOF 5 has a compressibility and Young s modulus 14 9 GPa comparable to wood which was confirmed with density functional theory DFT and nanoindentation 110 111 While it was shown that the MOF 5 can demonstrate the hyperfilling phenomenon within a loading media of solvent these MOFs are very sensitive to pressure and undergo amorphization pressure induced pore collapse at a pressure of 3 5 MPa when there is no fluid in the pores 112 MIL 53 edit nbsp MIL 53 MOF wine rack structure illustrating potential for anisotropy in loadingMIL 53 MOFs possess a wine rack structure These MOFs have been explored for anisotropy in Young s modulus due to the flexibility of loading and the potential for negative linear compressibility when compressing in one direction due to the ability of the wine rack opening during loading 113 114 Zirconium based MOFs edit nbsp Electron micrograph and structure of UiO 66 Color codes red oxygen brown carbon green zirconium gray hydrogen Zirconium based MOFs such as UiO 66 are a very robust class of MOFs attributed to strong hexanuclear Zr6 displaystyle ce Zr 6 nbsp metallic nodes with increased resistance to heat solvents and other harsh conditions which makes them of interest in terms of mechanical properties 115 Determinations of shear modulus and pelletization have shown that the UiO 66 MOFs are very mechanically robust and have high tolerance for pore collapse when compared to ZIFs and carboxylate MOFs 106 116 Although the UiO 66 MOF shows increased stability under pelletization the UiO 66 MOFs amorphized fairly rapidly under ball milling conditions due to destruction of linker coordinating inorganic nodes 117 Applications editHydrogen storage edit Main article Hydrogen storage Molecular hydrogen has the highest specific energy of any fuel However unless the hydrogen gas is compressed its volumetric energy density is very low so the transportation and storage of hydrogen require energy intensive compression and liquefaction processes 118 119 120 Therefore development of new hydrogen storage methods which decrease the concomitant pressure required for practical volumetric energy density is an active area of research 118 MOFs attract attention as materials for adsorptive hydrogen storage because of their high specific surface areas and surface to volume ratios as well as their chemically tunable structures 41 Compared to an empty gas cylinder a MOF filled gas cylinder can store more hydrogen at a given pressure because hydrogen molecules adsorb to the surface of MOFs Furthermore MOFs are free of dead volume so there is almost no loss of storage capacity as a result of space blocking by non accessible volume 8 Also because the hydrogen uptake is based primarily on physisorption many MOFs have a fully reversible uptake and release behavior No large activation barriers are required when liberating the adsorbed hydrogen 8 The storage capacity of a MOF is limited by the liquid phase density of hydrogen because the benefits provided by MOFs can be realized only if the hydrogen is in its gaseous state 8 The extent to which a gas can adsorb to a MOF s surface depends on the temperature and pressure of the gas In general adsorption increases with decreasing temperature and increasing pressure until a maximum is reached typically 20 30 bar after which the adsorption capacity decreases 8 41 120 However MOFs to be used for hydrogen storage in automotive fuel cells need to operate efficiently at ambient temperature and pressures between 1 and 100 bar as these are the values that are deemed safe for automotive applications 41 nbsp MOF 177The U S Department of Energy DOE has published a list of yearly technical system targets for on board hydrogen storage for light duty fuel cell vehicles which guide researchers in the field 5 5 wt 40 g L 1 by 2017 7 5 wt 70 g L 1 ultimate 121 Materials with high porosity and high surface area such as MOFs have been designed and synthesized in an effort to meet these targets These adsorptive materials generally work via physical adsorption rather than chemisorption due to the large HOMO LUMO gap and low HOMO energy level of molecular hydrogen A benchmark material to this end is MOF 177 which was found to store hydrogen at 7 5 wt with a volumetric capacity of 32 g L 1 at 77 K and 70 bar 122 MOF 177 consists of Zn4O 6 clusters interconnected by 1 3 5 benzenetribenzoate organic linkers and has a measured BET surface area of 4630 m2 g 1 Another exemplary material is PCN 61 which exhibits a hydrogen uptake of 6 24 wt and 42 5 g L 1 at 35 bar and 77 K and 2 25 wt at atmospheric pressure 123 PCN 61 consists of Cu2 4 paddle wheel units connected through 5 5 5 benzene 1 3 5 triyltris 1 ethynyl 2 isophthalate organic linkers and has a measured BET surface area of 3000 m2 g 1 Despite these promising MOF examples the classes of synthetic porous materials with the highest performance for practical hydrogen storage are activated carbon and covalent organic frameworks COFs 124 Design principles edit Practical applications of MOFs for hydrogen storage are met with several challenges For hydrogen adsorption near room temperature the hydrogen binding energy would need to be increased considerably 41 Several classes of MOFs have been explored including carboxylate based MOFs heterocyclic azolate based MOFs metal cyanide MOFs and covalent organic frameworks Carboxylate based MOFs have by far received the most attention because they are either commercially available or easily synthesized they have high acidity pKa 4 allowing for facile in situ deprotonation the metal carboxylate bond formation is reversible facilitating the formation of well ordered crystalline MOFs and the bridging bidentate coordination ability of carboxylate groups favors the high degree of framework connectivity and strong metal ligand bonds necessary to maintain MOF architecture under the conditions required to evacuate the solvent from the pores 41 The most common transition metals employed in carboxylate based frameworks are Cu2 and Zn2 Lighter main group metal ions have also been explored Be12 OH 12 BTB 4 the first successfully synthesized and structurally characterized MOF consisting of a light main group metal ion shows high hydrogen storage capacity but it is too toxic to be employed practically 125 There is considerable effort being put forth in developing MOFs composed of other light main group metal ions such as magnesium in Mg4 BDC 3 41 The following is a list of several MOFs that are considered to have the best properties for hydrogen storage as of May 2012 in order of decreasing hydrogen storage capacity 41 While each MOF described has its advantages none of these MOFs reach all of the standards set by the U S DOE Therefore it is not yet known whether materials with high surface areas small pores or di or trivalent metal clusters produce the most favorable MOFs for hydrogen storage 8 MOFs that are considered to have the best properties for hydrogen storage as of May 2012 Name Formula Structure Hydrogen storage capacity CommentsMOF 210 126 Zn4O BTE BPDC where BTE3 4 4 4 benzene 1 3 5 triyl tris ethyne 2 1 diyl tribenzoate and BPDC2 biphenyl 4 4 dicarboxylate At 77 K 8 6 excess wt 17 6 total wt at 77 K and 80 bar 44 total g H2 L at 80 bar and 77 K 126 At 298 K 2 90 delivery wt 1 100 bar at 298 K and 100 bar MOF 200 126 Zn4O BBC 2 where BBC3 4 4 4 benzene 1 3 5 triyl tris benzene 4 1 diyl tribenzoate At 77 K 7 4 excess wt 16 3 total wt at 77 K and 80 bar 36 total g H2 L at 80 bar and 77 K 126 At 298 K 3 24 delivery wt 1 100 bar at 298 K and 100 bar MOF 177 127 Zn4O BTB 2 where BTB3 1 3 5 benzenetribenzoate Tetrahedral Zn4O 6 units are linked by large triangular tricarboxylate ligands Six diamond shaped channels upper with diameter of 10 8 A surround a pore containing eclipsed BTB3 moieties lower 7 1 wt at 77 K and 40 bar 11 4 wt at 78 bar and 77 K MOF 177 has larger pores so hydrogen is compressed within holes rather than adsorbed to the surface This leads to higher total gravimetric uptake but lower volumetric storage density compared to MOF 5 41 MOF 5 128 Zn4O BDC 3 where BDC2 1 4 benzenedicarboxylate Square openings are either 13 8 or 9 2 A depending on the orientation of the aromatic rings 7 1 wt at 77 K and 40 bar 10 wt at 100 bar volumetric storage density of 66 g L MOF 5 has received much attention from theorists because of the partial charges on the MOF surface which provide a means of strengthening the binding hydrogen through dipole induced intermolecular interactions however MOF 5 has poor performance at room temperature 9 1 g L at 100 bar 41 Mn3 Mn4Cl 3 BTT 8 2 where H3BTT benzene 1 3 5 tris 1H tetrazole 129 Consists of truncated octahedral cages that share square faces leading to pores of about 10 A in diameter Contains open Mn2 coordination sites 60 g L at 77 K and 90 bar 12 1 g L at 90 bar and 298 K This MOF is the first demonstration of open metal coordination sites increasing strength of hydrogen adsorption which results in improved performance at 298 K It has relatively strong metal hydrogen interactions attributed to a spin state change upon binding or to a classical Coulombic attraction 41 Cu3 BTC 2 H2O 3 where H3BTC 1 3 5 benzenetricarboxylic acid 130 Consists of octahedral cages that share paddlewheel units to define pores of about 9 8 A in diameter High hydrogen uptake is attributed to overlapping attractive potentials from multiple copper paddle wheel units each Cu II center can potentially lose a terminal solvent ligand bound in the axial position providing an open coordination site for hydrogen binding 41 Structural impacts on hydrogen storage capacity edit To date hydrogen storage in MOFs at room temperature is a battle between maximizing storage capacity and maintaining reasonable desorption rates while conserving the integrity of the adsorbent framework e g completely evacuating pores preserving the MOF structure etc over many cycles There are two major strategies governing the design of MOFs for hydrogen storage 1 to increase the theoretical storage capacity of the material and 2 to bring the operating conditions closer to ambient temperature and pressure Rowsell and Yaghi have identified several directions to these ends in some of the early papers 131 132 Surface area edit The general trend in MOFs used for hydrogen storage is that the greater the surface area the more hydrogen the MOF can store High surface area materials tend to exhibit increased micropore volume and inherently low bulk density allowing for more hydrogen adsorption to occur 41 Hydrogen adsorption enthalpy edit High hydrogen adsorption enthalpy is also important Theoretical studies have shown that 22 25 kJ mol interactions are ideal for hydrogen storage at room temperature as they are strong enough to adsorb H2 but weak enough to allow for quick desorption 133 The interaction between hydrogen and uncharged organic linkers is not this strong and so a considerable amount of work has gone in synthesis of MOFs with exposed metal sites to which hydrogen adsorbs with an enthalpy of 5 10 kJ mol Synthetically this may be achieved by using ligands whose geometries prevent the metal from being fully coordinated by removing volatile metal bound solvent molecules over the course of synthesis and by post synthetic impregnation with additional metal cations 13 129 C5H5 V CO 3 H2 and Mo CO 5 H2 are great examples of increased binding energy due to open metal coordination sites 134 however their high metal hydrogen bond dissociation energies result in a tremendous release of heat upon loading with hydrogen which is not favorable for fuel cells 41 MOFs therefore should avoid orbital interactions that lead to such strong metal hydrogen bonds and employ simple charge induced dipole interactions as demonstrated in Mn3 Mn4Cl 3 BTT 8 2 An association energy of 22 25 kJ mol is typical of charge induced dipole interactions and so there is interest in the use of charged linkers and metals 41 The metal hydrogen bond strength is diminished in MOFs probably due to charge diffusion so 2 and 3 metal ions are being studied to strengthen this interaction even further A problem with this approach is that MOFs with exposed metal surfaces have lower concentrations of linkers this makes them difficult to synthesize as they are prone to framework collapse This may diminish their useful lifetimes as well 13 41 Sensitivity to airborne moisture edit MOFs are frequently sensitive to moisture in the air In particular IRMOF 1 degrades in the presence of small amounts of water at room temperature Studies on metal analogues have unraveled the ability of metals other than Zn to stand higher water concentrations at high temperatures 135 136 To compensate for this specially constructed storage containers are required which can be costly Strong metal ligand bonds such as in metal imidazolate triazolate and pyrazolate frameworks are known to decrease a MOF s sensitivity to air reducing the expense of storage 137 Pore size edit In a microporous material where physisorption and weak van der Waals forces dominate adsorption the storage density is greatly dependent on the size of the pores Calculations of idealized homogeneous materials such as graphitic carbons and carbon nanotubes predict that a microporous material with 7 A wide pores will exhibit maximum hydrogen uptake at room temperature At this width exactly two layers of hydrogen molecules adsorb on opposing surfaces with no space left in between 41 138 10 A wide pores are also of ideal size because at this width exactly three layers of hydrogen can exist with no space in between 41 A hydrogen molecule has a bond length of 0 74 A with a van der Waals radius of 1 17 A for each atom therefore its effective van der Waals length is 3 08 A 139 Structural defects edit Structural defects also play an important role in the performance of MOFs Room temperature hydrogen uptake via bridged spillover is mainly governed by structural defects which can have two effects 1 a partially collapsed framework can block access to pores thereby reducing hydrogen uptake and 2 lattice defects can create an intricate array of new pores and channels causing increased hydrogen uptake 140 Structural defects can also leave metal containing nodes incompletely coordinated This enhances the performance of MOFs used for hydrogen storage by increasing the number of accessible metal centers 141 Finally structural defects can affect the transport of phonons which affects the thermal conductivity of the MOF 142 Hydrogen adsorption edit Adsorption is the process of trapping atoms or molecules that are incident on a surface therefore the adsorption capacity of a material increases with its surface area In three dimensions the maximum surface area will be obtained by a structure which is highly porous such that atoms and molecules can access internal surfaces This simple qualitative argument suggests that the highly porous metal organic frameworks MOFs should be excellent candidates for hydrogen storage devices Adsorption can be broadly classified as being one of two types physisorption or chemisorption Physisorption is characterized by weak van der Waals interactions and bond enthalpies typically less than 20 kJ mol Chemisorption alternatively is defined by stronger covalent and ionic bonds with bond enthalpies between 250 and 500 kJ mol In both cases the adsorbate atoms or molecules i e the particles which adhere to the surface are attracted to the adsorbent solid surface because of the surface energy that results from unoccupied bonding locations at the surface The degree of orbital overlap then determines if the interactions will be physisorptive or chemisorptive 143 Adsorption of molecular hydrogen in MOFs is physisorptive Since molecular hydrogen only has two electrons dispersion forces are weak typically 4 7 kJ mol and are only sufficient for adsorption at temperatures below 298 K 41 A complete explanation of the H2 sorption mechanism in MOFs was achieved by statistical averaging in the grand canonical ensemble exploring a wide range of pressures and temperatures 144 145 Determining hydrogen storage capacity edit Two hydrogen uptake measurement methods are used for the characterization of MOFs as hydrogen storage materials gravimetric and volumetric To obtain the total amount of hydrogen in the MOF both the amount of hydrogen absorbed on its surface and the amount of hydrogen residing in its pores should be considered To calculate the absolute absorbed amount Nabs the surface excess amount Nex is added to the product of the bulk density of hydrogen rbulk and the pore volume of the MOF Vpore as shown in the following equation 146 Nabs Nex rbulkVpore displaystyle N rm abs N rm ex rho rm bulk V rm pore nbsp Gravimetric method edit The increased mass of the MOF due to the stored hydrogen is directly calculated by a highly sensitive microbalance 146 Due to buoyancy the detected mass of adsorbed hydrogen decreases again when a sufficiently high pressure is applied to the system because the density of the surrounding gaseous hydrogen becomes more and more important at higher pressures Thus this weight loss has to be corrected using the volume of the MOF s frame and the density of hydrogen 147 Volumetric method edit The changing of amount of hydrogen stored in the MOF is measured by detecting the varied pressure of hydrogen at constant volume 146 The volume of adsorbed hydrogen in the MOF is then calculated by subtracting the volume of hydrogen in free space from the total volume of dosed hydrogen 148 Other methods of hydrogen storage edit There are six possible methods that can be used for the reversible storage of hydrogen with a high volumetric and gravimetric density which are summarized in the following table where rm is the gravimetric density rv is the volumetric density T is the working temperature and P is the working pressure 149 Storage method rm mass rv kg H2 m3 T C P bar RemarksHigh pressure gas cylinders 13 lt 40 25 800 Compressed H2 gas in lightweight composite cylinderLiquid hydrogen in cryogenic tanks size dependent 70 8 252 1 Liquid H2 continuous loss of a few percent of H2 per day at 25 CAdsorbed hydrogen 2 20 80 100 Physisorption of H2 on materialsAdsorbed on interstitial sites in a host metal 2 150 25 1 Atomic hydrogen reversibly adsorbs in host metalsComplex compounds lt 18 150 gt 100 1 Complex compounds AlH4 or BH4 desorption at elevated temperature adsorption at high pressuresMetal and complexes together with water lt 40 gt 150 25 1 Chemical oxidation of metals with water and liberation of H2Of these high pressure gas cylinders and liquid hydrogen in cryogenic tanks are the least practical ways to store hydrogen for the purpose of fuel due to the extremely high pressure required for storing hydrogen gas or the extremely low temperature required for storing hydrogen liquid The other methods are all being studied and developed extensively 149 Electrocatalysis edit The high surface area and atomic metal sites feature of MOFs make them a suitable candidate for electrocatalysts especially energy related ones Until now MOFs have been used extensively as electrocatalyst for water splitting hydrogen evolution reaction and oxygen evolution reaction carbon dioxide reduction and oxygen reduction reaction 150 Currently there are two routes 1 Using MOFs as precursors to prepare electrocatalysts with carbon support 151 2 Using MOFs directly as electrocatalysts 152 153 However some results have shown that some MOFs are not stable under electrochemical environment 154 The electrochemical conversion of MOFs during electrocatalysis may produce the real catalyst materials and the MOFs are precatalysts under such conditions 155 Therefore claiming MOFs as the electrocatalysts requires in situ techniques coupled with electrocatalysis Biological imaging and sensing edit nbsp MOF 76 crystal where oxygen carbon and lanthanide atoms are represented by maroon black and blue spheres respectively Includes metal node connectivity blue polyhedra infinite rod SBU and 3D representation of MOF 76 A potential application for MOFs is biological imaging and sensing via photoluminescence A large subset of luminescent MOFs use lanthanides in the metal clusters Lanthanide photoluminescence has many unique properties that make them ideal for imaging applications such as characteristically sharp and generally non overlapping emission bands in the visible and near infrared NIR regions of the spectrum resistance to photobleaching or blinking and long luminescence lifetimes 156 However lanthanide emissions are difficult to sensitize directly because they must undergo LaPorte forbidden f f transitions Indirect sensitization of lanthanide emission can be accomplished by employing the antenna effect where the organic linkers act as antennae and absorb the excitation energy transfer the energy to the excited state of the lanthanide and yield lanthanide luminescence upon relaxation 157 A prime example of the antenna effect is demonstrated by MOF 76 which combines trivalent lanthanide ions and 1 3 5 benzenetricarboxylate BTC linkers to form infinite rod SBUs coordinated into a three dimensional lattice 158 As demonstrated by multiple research groups the BTC linker can effectively sensitize the lanthanide emission resulting in a MOF with variable emission wavelengths depending on the lanthanide identity 159 160 Additionally the Yan group has shown that Eu3 and Tb3 MOF 76 can be used for selective detection of acetophenone from other volatile monoaromatic hydrocarbons Upon acetophenone uptake the MOF shows a very sharp decrease or quenching of the luminescence intensity 161 For use in biological imaging however two main obstacles must be overcome MOFs must be synthesized on the nanoscale so as not to affect the target s normal interactions or behavior The absorbance and emission wavelengths must occur in regions with minimal overlap from sample autofluorescence other absorbing species and maximum tissue penetration 162 163 Regarding the first point nanoscale MOF NMOF synthesis has been mentioned in an earlier section The latter obstacle addresses the limitation of the antenna effect Smaller linkers tend to improve MOF stability but have higher energy absorptions predominantly in the ultraviolet UV and high energy visible regions A design strategy for MOFs with redshifted absorption properties has been accomplished by using large chromophoric linkers These linkers are often composed of polyaromatic species leading to large pore sizes and thus decreased stability To circumvent the use of large linkers other methods are required to redshift the absorbance of the MOF so lower energy excitation sources can be used Post synthetic modification PSM is one promising strategy Luo et al introduced a new family of lanthanide MOFs with functionalized organic linkers The MOFs deemed MOF 1114 MOF 1115 MOF 1130 and MOF 1131 are composed of octahedral SBUs bridged by amino functionalized dicarboxylate linkers The amino groups on the linkers served as sites for covalent PSM reactions with either salicylaldehyde or 3 hydroxynaphthalene 2 carboxaldehyde Both of these reactions extend the p conjugation of the linker causing a redshift in the absorbance wavelength from 450 nm to 650 nm The authors also propose that this technique could be adapted to similar MOF systems and by increasing pore volumes with increasing linker lengths larger pi conjugated reactants can be used to further redshift the absorption wavelengths 164 Biological imaging using MOFs has been realized by several groups namely Foucault Collet and co workers In 2013 they synthesized a NIR emitting Yb3 NMOF using phenylenevinylene dicarboxylate PVDC linkers They observed cellular uptake in both HeLa cells and NIH 3T3 cells using confocal visible and NIR spectroscopy 165 Although low quantum yields persist in water and Hepes buffer solution the luminescence intensity is still strong enough to image cellular uptake in both the visible and NIR regimes Nuclear wasteform materials edit nbsp Schematic representation of different ways to incorporate actinide species inside the MOF The development of new pathways for efficient nuclear waste administration is essential in wake of increased public concern about radioactive contamination due to nuclear plant operation and nuclear weapon decommission Synthesis of novel materials capable of selective actinide sequestration and separation is one of the current challenges acknowledged in the nuclear waste sector Metal organic frameworks MOFs are a promising class of materials to address this challenge due to their porosity modularity crystallinity and tunability Every building block of MOF structures can incorporate actinides First a MOF can be synthesized starting from actinide salts In this case the metal nodes are actinides 40 166 In addition metal nodes can be extended or cation exchange can exchange metals for actinides 40 Organic linkers can be functionalized with groups capable of actinide uptake 167 168 169 170 171 Lastly the porosity of MOFs can be used to incorporate contained guest molecules 172 173 174 and trap them in a structure by installation of additional or capping linkers 40 Drug delivery systems edit The synthesis characterization and drug related studies of low toxicity biocompatible MOFs has shown that they have potential for medical applications Many groups have synthesized various low toxicity MOFs and have studied their uses in loading and releasing various therapeutic drugs for potential medical applications A variety of methods exist for inducing drug release such as pH response magnetic response ion response temperature response and pressure response 175 In 2010 Smaldone et al an international research group synthesized a biocompatible MOF termed CD MOF 1 from cheap edible natural products CD MOF 1 consists of repeating base units of 6 g cyclodextrin rings bound together by potassium ions g cyclodextrin g CD is a symmetrical cyclic oligosaccharide that is mass produced enzymatically from starch and consists of eight asymmetric a 1 4 linked D glucopyranosyl residues 176 The molecular structure of these glucose derivatives which approximates a truncated cone bucket or torus generates a hydrophilic exterior surface and a nonpolar interior cavity Cyclodextrins can interact with appropriately sized drug molecules to yield an inclusion complex 177 Smaldone s group proposed a cheap and simple synthesis of the CD MOF 1 from natural products They dissolved sugar g cyclodextrin and an alkali salt KOH KCl potassium benzoate in distilled bottled water and allowed 190 proof grain alcohol Everclear to vapor diffuse into the solution for a week The synthesis resulted in a cubic g CD 6 repeating motif with a pore size of approximately 1 nm Subsequently in 2017 Hartlieb et al at Northwestern did further research with CD MOF 1 involving the encapsulation of ibuprofen The group studied different methods of loading the MOF with ibuprofen as well as performing related bioavailability studies on the ibuprofen loaded MOF They investigated two different methods of loading CD MOF 1 with ibuprofen crystallization using the potassium salt of ibuprofen as the alkali cation source for production of the MOF and absorption and deprotonation of the free acid of ibuprofen into the MOF From there the group performed in vitro and in vivo studies to determine the applicability of CD MOF 1 as a viable delivery method for ibuprofen and other NSAIDs In vitro studies showed no toxicity or effect on cell viability up to 100 mM In vivo studies in mice showed the same rapid uptake of ibuprofen as the ibuprofen potassium salt control sample with a peak plasma concentration observed within 20 minutes and the cocrystal has the added benefit of double the half life in blood plasma samples 178 The increase in half life is due to CD MOF 1 increasing the solubility of ibuprofen compared to the pure salt form Since these developments many groups have done further research into drug delivery with water soluble biocompatible MOFs involving common over the counter drugs 179 In March 2018 Sara Rojas and her team published their research on drug incorporation and delivery with various biocompatible MOFs other than CD MOF 1 through simulated cutaneous administration The group studied the loading and release of ibuprofen hydrophobic and aspirin hydrophilic in three biocompatible MOFs MIL 100 Fe UiO 66 Zr and MIL 127 Fe Under simulated cutaneous conditions aqueous media at 37 C the six different combinations of drug loaded MOFs fulfilled the requirements to be used as topical drug delivery systems such as released payload between 1 and 7 days and delivering a therapeutic concentration of the drug of choice without causing unwanted side effects 180 The group discovered that the drug uptake is governed by the hydrophilic hydrophobic balance between cargo and matrix and the accessibility of the drug through the framework The controlled release under cutaneous conditions follows different kinetics profiles depending on i the structure of the framework with either a fast delivery from the very open structure MIL 100 or a slower drug release from the narrow 1D pore system of MIL 127 or ii the hydrophobic hydrophilic nature of the cargo with a fast Aspirin and slow Ibuprofen release from the UiO 66 matrix Moreover a simple ball milling technique is used to efficiently encapsulate the model drugs 5 fluorouracil caffeine para aminobenzoic acid and benzocaine Both computational and experimental studies confirm the suitability of Zn4O dmcapz 3 to incorporate high loadings of the studied bioactive molecules 181 Recent research involving MOFs as a drug delivery method includes more than just the encapsulation of everyday drugs like ibuprofen and aspirin In early 2018 Chen et al published detailing their work on the use of MOF ZIF 8 zeolitic imidazolate framework 8 in antitumor research to control the release of an autophagy inhibitor 3 methyladenine 3 MA and prevent it from dissipating in a large quantity before reaching the target 182 The group performed in vitro studies and determined that the autophagy related proteins and autophagy flux in HeLa cells treated with 3 MA ZIF 8 NPs show that the autophagosome formation is significantly blocked which reveals that the pH sensitive dissociation increases the efficiency of autophagy inhibition at the equivalent concentration of 3 MA This shows promise for future research and applicability with MOFs as drug delivery methods in the fight against cancer Semiconductors edit In 2014 researchers proved that they can create electrically conductive thin films of MOFs Cu3 BTC 2 also known as HKUST 1 BTC benzene 1 3 5 tricarboxylic acid infiltrated with the molecule 7 7 8 8 tetracyanoquinododimethane that could be used in applications including photovoltaics sensors and electronic materials and a path toward creating semiconductors The team demonstrated tunable air stable electrical conductivity with values as high as 7 siemens per meter comparable to bronze 183 Ni3 2 3 6 7 10 11 hexaiminotriphenylene 2 was shown to be a metal organic graphene analogue that has a natural band gap making it a semiconductor and is able to self assemble It is an example of conductive metal organic framework It represents a family of similar compounds Because of the symmetry and geometry in 2 3 6 7 10 11 hexaiminotriphenylene HITP the overall organometallic complex has an almost fractal nature that allows it to perfectly self organize By contrast graphene must be doped to give it the properties of a semiconductor Ni3 HITP 2 pellets had a conductivity of 2 S cm a record for a metal organic compound 184 185 In 2018 researchers synthesized a two dimensional semiconducting MOF Fe3 THT 2 NH4 3 also known as THT 2 3 6 7 10 11 triphenylenehexathiol and showed high electric mobility at room temperature 186 In 2020 the same material was integrated in a photo detecting device detecting a broad wavelength range from UV to NIR 400 1575 nm 187 This was the first time a two dimensional semiconducting MOF was demonstrated to be used in opto electronic devices 188 Cu3 HHTP 2 displaystyle ce Cu3 HHTP 2 nbsp is a 2D MOF structure and there are limited examples of materials which are intrinsically conductive porous and crystalline Layered 2D MOFs have porous crystalline structure showing electrical conductivity These materials are constructed from trigonal linker molecules phenylene or triphenylene and six functional groups of OH NH2 displaystyle ce NH2 nbsp or SH The trigonal linker molecules and square planarly coordinated metal ions such as Cu2 displaystyle ce Cu 2 nbsp Ni2 displaystyle ce Ni 2 nbsp Co2 displaystyle ce Co 2 nbsp and Pt2 displaystyle ce Pt 2 nbsp form layers with hexagonal structures which look like graphene in larger scale Stacking of these layers can build one dimensional pore systems Graphene like 2D MOFs have shown decent conductivities This makes them a good choice to be tested as electrode material for evolution of hydrogen from water oxygen reduction reactions supercapacitors and sensing of volatile organic compounds VOCs Among these MOFs Cu3 HHTP 2 displaystyle ce Cu3 HHTP 2 nbsp has exhibited the lowest conductivity but also the strongest reaction in sensing of VOCs 189 190 191 Bio mimetic mineralization edit Biomolecules can be incorporated during the MOF crystallization process Biomolecules including proteins DNA and antibodies could be encapsulated within ZIF 8 Enzymes encapsulated in this way were stable and active even after being exposed to harsh conditions e g aggressive solvents and high temperature ZIF 8 MIL 88A HKUST 1 and several luminescent MOFs containing lanthanide metals were used for the biomimetic mineralization process 192 Carbon capture edit Main article Carbon capture and storage Adsorbent edit MOF s small tunable pore sizes and high void fractions are promising as an adsorbent to capture CO2 193 MOFs could provide a more efficient alternative to traditional amine solvent based methods in CO2 capture from coal fired power plants 194 MOFs could be employed in each of the main three carbon capture configurations for coal fired power plants pre combustion post combustion and oxy combustion 195 The post combustion configuration is the only one that can be retrofitted to existing plants drawing the most interest and research The flue gas would be fed through a MOF in a packed bed reactor setup Flue gas is generally 40 to 60 C with a partial pressure of CO2 at 0 13 0 16 bar CO2 can bind to the MOF surface through either physisorption via Van der Waals interactions or chemisorption via covalent bond formation 196 Once the MOF is saturated the CO2 is extracted from the MOF through either a temperature swing or a pressure swing This process is known as regeneration In a temperature swing regeneration the MOF would be heated until CO2 desorbs To achieve working capacities comparable to the amine process the MOF must be heated to around 200 C In a pressure swing the pressure would be decreased until CO2 desorbs 197 Another relevant MOF property is their low heat capacities Monoethanolamine MEA solutions the leading capture method have a heat capacity between 3 4 J g K since they are mostly water This high heat capacity contributes to the energy penalty in the solvent regeneration step i e when the adsorbed CO2 is removed from the MEA solution MOF 177 a MOF designed for CO2 capture has a heat capacity of 0 5 J g K at ambient temperature 195 MOFs adsorb 90 of the CO2 using a vacuum pressure swing process The MOF Mg dobdc has a 21 7 wt CO2 loading capacity Applied to a large scale power plant the cost of energy would increase by 65 while a U S NETL baseline amine based system would cause an increase of 81 goal is 35 The capture cost would be 57 ton while for the amine system the cost is estimated to be 72 ton At that rate the capital required to implement such project in a 580 MW power plant would be 354 million 198 Catalyst edit A MOF loaded with propylene oxide can act as a catalyst converting CO2 into cyclic carbonates ring shaped molecules with many applications They can also remove carbon from biogas This MOF is based on lanthanides which provide chemical stability This is especially important because the gases the MOF will be exposed to are hot high in humidity and acidic 199 Triaminoguanidinium based POFs and Zn POFs are new multifunctional materials for environmental remediation and biomedical applications 200 Desalination ion separation edit MOF membranes can mimic substantial ion selectivity This offers the potential for use in desalination and water treatment As of 2018 reverse osmosis supplied more than half of global desalination capacity and the last stage of most water treatment processes Osmosis does not use dehydration of ions or selective ion transport in biological channels and it is not energy efficient The mining industry uses membrane based processes to reduce water pollution and to recover metals MOFs could be used to extract metals such as lithium from seawater and waste streams 201 MOF membranes such as ZIF 8 and UiO 66 membranes with uniform subnanometer pores consisting of angstrom scale windows and nanometer scale cavities displayed ultrafast selective transport of alkali metal ions The windows acted as ion selectivity filters for alkali metal ions while the cavities functioned as pores for transport The ZIF 8 202 and UiO 66 203 membranes showed a LiCl RbCl selectivity of 4 6 and 1 8 respectively much higher than the 0 6 to 0 8 selectivity in traditional membranes 204 A 2020 study suggested that a new MOF called PSP MIL 53 could be used along with sunlight to purify water in just half an hour 205 Gas separation edit MOFs are also predicted to be very effective media to separate gases with low energy cost using computational high throughput screening from their adsorption 206 or gas breakthrough diffusion 207 properties One example is NbOFFIVE 1 Ni also referred to as KAUST 7 which can separate propane and propylene via diffusion at nearly 100 selectivity 208 The specific molecule selectivity properties provided by Cu BDC surface mounted metal organic framework SURMOF 2 growth on alumina layer on top of back gated Graphene Field Effect Transistor GFET can provide a sensor that is only sensitive to ethanol but not to methanol or isopropanol 209 Water vapor capture and dehumidification edit MOFs have been demonstrated that capture water vapor from the air 210 In 2021 under humid conditions a polymer MOF lab prototype yielded 17 liters 4 5 gal of water per kg per day without added energy 211 212 MOFs could also be used to increase energy efficiency in room temperature space cooling applications 213 214 nbsp Schematic diagram for MOF dehumidification featuring MIL 100 Fe a MOF with particularly high water adsorption capacityWhen cooling outdoor air a cooling unit must deal with both the air s sensible heat and latent heat Typical vapor compression air conditioning VCAC units manage the latent heat in air through cooling fins held below the dew point temperature of the moist air at the intake These fins condense the water dehydrating the air and thus substantially reducing the air s heat content The cooler s energy usage is highly dependent on the cooling coil s temperature and would be improved greatly if the temperature of this coil could be raised above the dew point 215 This makes it desirable to handle dehumidification through means other than condensation One such means is by adsorbing the water from the air into a desiccant coated onto the heat exchangers using the waste heat exhausted from the unit to desorb the water from the sorbent and thus regenerate the desiccant for repeated usage This is accomplished by having two condenser evaporator units through which the flow of refrigerant can be reversed once the desiccant on the condenser is saturated thus making the condenser the evaporator and vice versa 213 MOFs extremely high surface areas and porosities have made them the subject of much research in water adsorption applications 213 216 217 218 Chemistry can help tune the optimal relative humidity for adsorption desorption and the sharpness of the water uptake 213 219 Ferroelectrics and multiferroics edit Some MOFs also exhibit spontaneous electric polarization which occurs due to the ordering of electric dipoles polar linkers or guest molecules below a certain phase transition temperature 220 If this long range dipolar order can be controlled by the external electric field a MOF is called ferroelectric 221 Some ferroelectric MOFs also exhibit magnetic ordering making them single structural phase multiferroics This material property is highly interesting for construction of memory devices with high information density The coupling mechanism of type I CH3 2NH2 Ni HCOO 3 molecular multiferroic is spontaneous elastic strain mediated indirect coupling 222 See also editBET theory Conjugated microporous polymer Coordination chemistry Coordination polymers Covalent organic framework Cryogenics Electrocatalyst Flexible metal organic framework Gerard Ferey Hydrogen economy Hydrogen Hydrogen bonded organic framework Liquid hydrogen Macromolecular assembly Metal inorganic framework Omar M Yaghi Organometallic chemistry Crystal nets periodic graphs Solid sorbents for carbon capture Susumu Kitagawa United States Department of Energy X ray Crystallography Zeolitic imidazolate frameworksReferences edit a b Batten SR Champness NR Chen XM Garcia Martinez J Kitagawa S Ohrstrom L O Keeffe M Suh MP Reedijk J 2013 Terminology of metal organic frameworks and coordination polymers IUPAC Recommendations 2013 PDF Pure and Applied Chemistry 85 8 1715 1724 doi 10 1351 PAC REC 12 11 20 S2CID 96853486 Bennett Thomas D Cheetham Anthony K 2014 05 20 Amorphous Metal Organic Frameworks Accounts of Chemical Research 47 5 1555 1562 doi 10 1021 ar5000314 PMID 24707980 Bennett Thomas D Coudert Francois Xavier James Stuart L Cooper Andrew I September 2021 The changing state of porous materials Nature Materials 20 9 1179 1187 Bibcode 2021NatMa 20 1179B doi 10 1038 s41563 021 00957 w PMID 33859380 S2CID 233239286 Mon M Bruno R Ferrando Soria J Armentano D Pardo E 2018 Metal organic framework technologies for water remediation towards a sustainable ecosystem Journal of Materials Chemistry A 6 12 4912 4947 doi 10 1039 c8ta00264a Cejka J ed 2011 Metal Organic Frameworks Applications from Catalysis to Gas Storage Wiley VCH ISBN 978 3 527 32870 3 O Keeffe M Yaghi OM 2005 Reticular chemistry Present and future prospects PDF Journal of Solid State Chemistry 178 8 v vi Bibcode 2005JSSCh 178D 5 doi 10 1016 S0022 4596 05 00368 3 Cote AP Benin AI Ockwig NW O Keeffe M Matzger AJ Yaghi OM November 2005 Porous crystalline covalent organic frameworks Science 310 5751 1166 70 Bibcode 2005Sci 310 1166C doi 10 1126 science 1120411 PMID 16293756 S2CID 35798005 a b c d e f Czaja AU Trukhan N Muller U May 2009 Industrial applications of metal organic frameworks Chemical Society Reviews 38 5 1284 93 doi 10 1039 b804680h PMID 19384438 Cheetham AK Rao CN Feller RK 2006 Structural diversity and chemical trends in hybrid inorganic organic framework materials Chemical Communications 46 4780 4795 doi 10 1039 b610264f PMID 17345731 a b Cheetham AK Ferey G Loiseau T November 1999 Open Framework Inorganic Materials Angewandte Chemie 38 22 3268 3292 doi 10 1002 SICI 1521 3773 19991115 38 22 lt 3268 AID ANIE3268 gt 3 0 CO 2 U PMID 10602176 Bucar DK Papaefstathiou GS Hamilton TD Chu QL Georgiev IG MacGillivray LR 2007 Template controlled reactivity in the organic solid state by principles of coordination driven self assembly European Journal of Inorganic Chemistry 2007 29 4559 4568 doi 10 1002 ejic 200700442 Parnham ER Morris RE October 2007 Ionothermal synthesis of zeolites metal organic frameworks and inorganic organic hybrids Accounts of Chemical Research 40 10 1005 13 doi 10 1021 ar700025k PMID 17580979 a b c d Dincă M Long JR 2008 Hydrogen storage in microporous metal organic frameworks with exposed metal sites Angewandte Chemie 47 36 6766 79 doi 10 1002 anie 200801163 PMID 18688902 Gitis Vitaly Rothenberg Gadi 2020 Gitis Vitaly Rothenberg Gadi eds Handbook of Porous Materials Vol 4 Singapore WORLD SCIENTIFIC pp 110 111 doi 10 1142 11909 ISBN 978 981 12 2328 0 a b Ni Z Masel RI September 2006 Rapid production of metal organic frameworks via microwave assisted solvothermal synthesis Journal of the American Chemical Society 128 38 12394 5 doi 10 1021 ja0635231 PMID 16984171 a b Choi JS Son WJ Kim J Ahn WS 2008 Metal organic framework MOF 5 prepared by microwave heating Factors to be considered Microporous and Mesoporous Materials 116 1 3 727 731 doi 10 1016 j micromeso 2008 04 033 Steenhaut Timothy Hermans Sophie Filinchuk Yaroslav 2020 Green synthesis of a large series of bimetallic MIL 100 Fe M MOFs New Journal of Chemistry 44 10 3847 3855 doi 10 1039 D0NJ00257G S2CID 214492546 Pichon A James SL 2008 An array based study of reactivity under solvent free mechanochemical conditions insights and trends CrystEngComm 10 12 1839 1847 doi 10 1039 B810857A Braga D Giaffreda SL Grepioni F Chierotti MR Gobetto R Palladino G Polito M 2007 Solvent effect in a solvent free reaction CrystEngComm 9 10 879 881 doi 10 1039 B711983F a b Steenhaut Timothy Gregoire Nicolas Barozzino Consiglio Gabriella Filinchuk Yaroslav Hermans Sophie 2020 Mechanochemical defect engineering of HKUST 1 and impact of the resulting defects on carbon dioxide sorption and catalytic cyclopropanation RSC Advances 10 34 19822 19831 Bibcode 2020RSCAd 1019822S doi 10 1039 C9RA10412G PMC 9054116 PMID 35520409 Stassen I Styles M Grenci G Gorp HV Vanderlinden W Feyter SD Falcaro P Vos DD Vereecken P Ameloot R March 2016 Chemical vapour deposition of zeolitic imidazolate framework thin films Nature Materials 15 3 304 10 Bibcode 2016NatMa 15 304S doi 10 1038 nmat4509 PMID 26657328 Cruz A Stassen I Ameloot R et al 2019 An integrated cleanroom process for the vapor phase deposition of large area zeolitic imidazolate framework thin films Chemistry of Materials 31 22 9462 9471 doi 10 1021 acs chemmater 9b03435 hdl 10550 74201 S2CID 208737085 Virmani Erika Rotter Julian M Mahringer Andre von Zons Tobias Godt Adelheid Bein Thomas Wuttke Stefan Medina Dana D 2018 04 11 On Surface Synthesis of Highly Oriented Thin Metal Organic Framework Films through Vapor Assisted Conversion Journal of the American Chemical Society 140 14 4812 4819 doi 10 1021 jacs 7b08174 PMID 29542320 Sebastian Bauer Norbert Stock October 2007 Hochdurchsatz Methoden in der Festkorperchemie Schneller zum Ziel Chemie in unserer Zeit in German vol 41 no 5 pp 390 398 doi 10 1002 ciuz 200700404 Gimeno Fabra Miquel Munn Alexis S Stevens Lee A Drage Trevor C Grant David M Kashtiban Reza J Sloan Jeremy Lester Edward Walton Richard I 7 September 2012 Instant MOFs continuous synthesis of metal organic frameworks by rapid solvent mixing Chemical Communications 48 86 10642 4 doi 10 1039 C2CC34493A PMID 23000779 Retrieved 22 June 2020 Rasmussen Elizabeth G Kramlich John Novosselov Igor V 3 June 2020 Scalable Continuous Flow Metal Organic Framework MOF Synthesis Using Supercritical CO2 ACS Sustainable Chemistry amp Engineering 8 26 9680 9689 doi 10 1021 acssuschemeng 0c01429 S2CID 219915159 Biemmi Enrica Christian Sandra Stock Norbert Bein Thomas 2009 High throughput screening of synthesis parameters in the formation of the metal organic frameworks MOF 5 and HKUST 1 Microporous and Mesoporous Materials in German vol 117 no 1 2 pp 111 117 doi 10 1016 j micromeso 2008 06 040 Putnis Andrew 2009 01 01 Mineral Replacement Reactions Reviews in Mineralogy and Geochemistry 70 1 87 124 Bibcode 2009RvMG 70 87P doi 10 2138 rmg 2009 70 3 Reboul Julien Furukawa Shuhei Horike Nao Tsotsalas Manuel Hirai Kenji Uehara Hiromitsu Kondo Mio Louvain Nicolas Sakata Osami Kitagawa Susumu August 2012 Mesoscopic architectures of porous coordination polymers fabricated by pseudomorphic replication Nature Materials 11 8 717 723 Bibcode 2012NatMa 11 717R doi 10 1038 nmat3359 hdl 2433 158311 PMID 22728321 S2CID 205407412 Robatjazi Hossein Weinberg Daniel Swearer Dayne F Jacobson Christian Zhang Ming Tian Shu Zhou Linan Nordlander Peter Halas Naomi J February 2019 Metal organic frameworks tailor the properties of aluminum nanocrystals Science Advances 5 2 eaav5340 Bibcode 2019SciA 5 5340R doi 10 1126 sciadv aav5340 PMC 6368424 PMID 30783628 Kornienko Nikolay Zhao Yingbo Kley Christopher S Zhu Chenhui Kim Dohyung Lin Song Chang Christopher J Yaghi Omar M Yang Peidong 2015 11 11 Metal Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide Journal of the American Chemical Society 137 44 14129 14135 doi 10 1021 jacs 5b08212 PMID 26509213 S2CID 14793796 a b c Das S Kim H Kim K March 2009 Metathesis in single crystal complete and reversible exchange of metal ions constituting the frameworks of metal organic frameworks Journal of the American Chemical Society 131 11 3814 5 doi 10 1021 ja808995d PMID 19256486 a b c d Burrows AD Frost CG Mahon MF Richardson C 2008 10 20 Post synthetic modification of tagged metal organic frameworks Angewandte Chemie 47 44 8482 6 doi 10 1002 anie 200802908 PMID 18825761 a b Li T Kozlowski MT Doud EA Blakely MN Rosi NL August 2013 Stepwise ligand exchange for the preparation of a family of mesoporous MOFs Journal of the American Chemical Society 135 32 11688 91 doi 10 1021 ja403810k PMID 23688075 Sun D Liu W Qiu M Zhang Y Li Z February 2015 Introduction of a mediator for enhancing photocatalytic performance via post synthetic metal exchange in metal organic frameworks MOFs Chemical Communications 51 11 2056 9 doi 10 1039 c4cc09407g PMID 25532612 Liu C Rosi NL September 2017 Ternary gradient metal organic frameworks Faraday Discussions 201 163 174 Bibcode 2017FaDi 201 163L doi 10 1039 c7fd00045f PMID 28621353 Liu C Zeng C Luo TY Merg AD Jin R Rosi NL September 2016 Establishing Porosity Gradients within Metal Organic Frameworks Using Partial Postsynthetic Ligand Exchange Journal of the American Chemical Society 138 37 12045 8 doi 10 1021 jacs 6b07445 PMID 27593173 Yuan S Lu W Chen YP Zhang Q Liu TF Feng D Wang X Qin J Zhou HC March 2015 Sequential linker installation precise placement of functional groups in multivariate metal organic frameworks Journal of the American Chemical Society 137 9 3177 80 doi 10 1021 ja512762r PMID 25714137 Yuan S Chen YP Qin JS Lu W Zou L Zhang Q Wang X Sun X Zhou HC July 2016 Linker Installation Engineering Pore Environment with Precisely Placed Functionalities in Zirconium MOFs Journal of the American Chemical Society 138 28 8912 9 doi 10 1021 jacs 6b04501 OSTI 1388673 PMID 27345035 a b c d Dolgopolova EA Ejegbavwo OA Martin CR Smith MD Setyawan W Karakalos SG Henager CH Zur Loye HC Shustova NB November 2017 Multifaceted Modularity A Key for Stepwise Building of Hierarchical Complexity in Actinide Metal Organic Frameworks Journal of the American Chemical Society 139 46 16852 16861 doi 10 1021 jacs 7b09496 PMID 29069547 a b c d e f g h i j k l m n o p q r s Murray LJ Dincă M Long JR May 2009 Hydrogen storage in metal organic frameworks Chemical Society Reviews 38 5 1294 314 CiteSeerX 10 1 1 549 4404 doi 10 1039 b802256a PMID 19384439 S2CID 10443172 Li Y Yang RT 2007 Hydrogen Storage on Platinum Nanoparticles Doped on Superactivated Carbon Journal of Physical Chemistry C 111 29 11086 11094 doi 10 1021 jp072867q Jacoby Mitch 2013 Materials Chemistry Metal Organic Frameworks Go Commercial Chemical amp Engineering News 91 51 Cejka J Corma A Zones S 27 May 2010 Zeolites and Catalysis Synthesis Reactions and Applications John Wiley amp Sons ISBN 978 3 527 63030 1 Niknam Esmaeil Panahi Farhad Daneshgar Fatemeh Bahrami Foroogh Khalafi Nezhad Ali 2018 Metal Organic Framework MIL 101 Cr as an Efficient Heterogeneous Catalyst for Clean Synthesis of Benzoazoles ACS Omega 3 12 17135 17144 doi 10 1021 acsomega 8b02309 PMC 6643801 PMID 31458334 S2CID 104347751 Lopez Jorge Chavez Ana M Rey Ana Alvarez Pedro M 2021 Insights into the Stability and Activity of MIL 53 Fe in Solar Photocatalytic Oxidation Processes in Water Catalysts 11 4 448 doi 10 3390 catal11040448 Chavez A M Rey A Lopez J Alvarez P M Beltran F J 2021 Critical aspects of the stability and catalytic activity of MIL 100 Fe in different advanced oxidation processes Separation and Purification Technology 255 117660 doi 10 1016 j seppur 2020 117660 S2CID 224863042 Fujita M Kwon YJ Washizu S Ogura K 1994 Preparation Clathration Ability and Catalysis of a Two Dimensional Square Network Material Composed of Cadmium II and 4 4 Bipyridine Journal of the American Chemical Society 116 3 1151 doi 10 1021 ja00082a055 Llabresixamena F Abad A Corma A Garcia H 2007 MOFs as catalysts Activity reusability and shape selectivity of a Pd containing MOF Journal of Catalysis 250 2 294 298 doi 10 1016 j jcat 2007 06 004 Ravon U Domine ME Gaudillere C Desmartin Chomel A Farrusseng D 2008 MOFs as acid catalysts with shape selectivity properties New Journal of Chemistry 32 6 937 doi 10 1039 B803953B Chui SS Lo SM Charmant JP Orpen AG Williams ID 1999 A Chemically Functionalizable Nanoporous Material Cu3 TMA 2 H2O 3 n Science 283 5405 1148 50 Bibcode 1999Sci 283 1148C doi 10 1126 science 283 5405 1148 PMID 10024237 Alaerts L Seguin E Poelman H Thibault Starzyk F Jacobs PA De Vos DE September 2006 Probing the Lewis acidity and catalytic activity of the metal organic framework Cu3 btc 2 BTC benzene 1 3 5 tricarboxylate Chemistry A European Journal 12 28 7353 63 doi 10 1002 chem 200600220 hdl 1854 LU 351275 PMID 16881030 Henschel A Gedrich K Kraehnert R Kaskel S September 2008 Catalytic properties of MIL 101 Chemical Communications 35 4192 4 doi 10 1039 B718371B PMID 18802526 a b Horike S Dinca M Tamaki K Long JR May 2008 Size selective Lewis acid catalysis in a microporous metal organic framework with exposed Mn2 coordination sites Journal of the American Chemical Society 130 18 5854 5 doi 10 1021 ja800669j PMID 18399629 Chen L Yang Y Jiang D July 2010 CMPs as scaffolds for constructing porous catalytic frameworks a built in heterogeneous catalyst with high activity and selectivity based on nanoporous metalloporphyrin polymers Journal of the American Chemical Society 132 26 9138 43 doi 10 1021 ja1028556 PMID 20536239 Rahiman AK Rajesh K Bharathi KS Sreedaran S Narayanan V 2009 Catalytic oxidation of alkenes by manganese III porphyrin encapsulated Al V Si mesoporous molecular sieves Inorganica Chimica Acta 352 5 1491 1500 doi 10 1016 j ica 2008 07 011 Mansuy D Bartoli JF Momenteau M 1982 Alkane hydroxylation catalyzed by metalloporhyrins evidence for different active oxygen species with alkylhydroperoxides and iodosobenzene as oxidants Tetrahedron Letters 23 27 2781 2784 doi 10 1016 S0040 4039 00 87457 2 Ingleson MJ Barrio JP Bacsa J Dickinson C Park H Rosseinsky MJ March 2008 Generation of a solid Bronsted acid site in a chiral framework Chemical Communications 11 1287 9 doi 10 1039 B718443C PMID 18389109 a b Hasegawa S Horike S Matsuda R Furukawa S Mochizuki K Kinoshita Y Kitagawa S March 2007 Three dimensional porous coordination polymer functionalized with amide groups based on tridentate ligand selective sorption and catalysis Journal of the American Chemical Society 129 9 2607 14 doi 10 1021 ja067374y PMID 17288419 a b Hwang YK Hong DY Chang JS Jhung SH Seo YK Kim J Vimont A Daturi M Serre C Ferey G 2008 Amine grafting on coordinatively unsaturated metal centers of MOFs consequences for catalysis and metal encapsulation Angewandte Chemie 47 22 4144 8 doi 10 1002 anie 200705998 PMID 18435442 Seo JS Whang D Lee H Jun SI Oh J Jeon YJ Kim K April 2000 A homochiral metal organic porous material for enantioselective separation and catalysis Nature 404 6781 982 6 Bibcode 2000Natur 404 982S doi 10 1038 35010088 PMID 10801124 S2CID 1159701 Ohmori O Fujita M July 2004 Heterogeneous catalysis of a coordination network cyanosilylation of imines catalyzed by a Cd II 4 4 bipyridine square grid complex Chemical Communications 14 1586 7 doi 10 1039 B406114B PMID 15263930 Schlichte K Kratzke T Kaskel S 2004 Improved synthesis thermal stability and catalytic properties of the metal organic framework compound Cu3 BTC 2 Microporous and Mesoporous Materials 73 1 2 81 85 doi 10 1016 j micromeso 2003 12 027 hdl 11858 00 001M 0000 000F 9785 0 Chui SS Lo SM Charmant JP Orpen AG Williams ID 1999 A Chemically Functionalizable Nanoporous Material Cu3 TMA 2 H2O 3 n Science 283 5405 1148 1150 Bibcode 1999Sci 283 1148C doi 10 1126 science 283 5405 1148 PMID 10024237 Evans OR Ngo HL Lin W 2001 Chiral Porous Solids Based on Lamellar Lanthanide Phosphonates Journal of the American Chemical Society 123 42 10395 6 doi 10 1021 ja0163772 PMID 11603994 Kato C Hasegawa M Sato T Yoshizawa A Inoue T Mori W 2005 Microporous dinuclear copper II trans 1 4 cyclohexanedicarboxylate heterogeneous oxidation catalysis with hydrogen peroxide and X ray powder structure of peroxo copper II intermediate Journal of Catalysis 230 226 236 doi 10 1016 j jcat 2004 11 032 Han JW Hill CL December 2007 A coordination network that catalyzes O2 based oxidations Journal of the American Chemical Society 129 49 15094 5 doi 10 1021 ja069319v PMID 18020331 Ferey G Mellot Draznieks C Serre C Millange F Dutour J Surble S Margiolaki I September 2005 A chromium terephthalate based solid with unusually large pore volumes and surface area PDF Science 309 5743 2040 2 Bibcode 2005Sci 309 2040F doi 10 1126 science 1116275 PMID 16179475 S2CID 29483796 Tahmouresilerd B Larson PJ Unruh DK Cozzolino AF July 2018 Make room for iodine systematic pore tuning of multivariate metal organic frameworks for the catalytic oxidation of hydroquinones using hypervalent iodine Catalysis Science amp Technology 8 17 4349 4357 doi 10 1039 C8CY00794B Tahmouresilerd B Moody M Agogo L Cozzolino AF Apr 2019 The impact of an isoreticular expansion strategy on the performance of iodine catalysts supported in multivariate zirconium and aluminum metal organic frameworks Dalton Transactions 48 19 6445 6454 doi 10 1039 C9DT00368A PMID 31017171 S2CID 129944197 Schroder F Esken D Cokoja M van den Berg MW Lebedev OI Van Tendeloo G Walaszek B Buntkowsky G Limbach HH Chaudret B Fischer RA May 2008 Ruthenium nanoparticles inside porous Zn4O bdc 3 by hydrogenolysis of adsorbed Ru cod cot a solid state reference system for surfactant stabilized ruthenium colloids Journal of the American Chemical Society 130 19 6119 30 doi 10 1021 ja078231u PMID 18402452 Tan YC Zeng HC October 2018 Lewis basicity generated by localised charge imbalance in noble metal nanoparticle embedded defective metal organic frameworks Nature Communications 9 1 4326 Bibcode 2018NatCo 9 4326T doi 10 1038 s41467 018 06828 4 PMC 6194069 PMID 30337531 Pan L Liu H Lei X Huang X Olson DH Turro NJ Li J February 2003 RPM 1 a recyclable nanoporous material suitable for ship in bottle synthesis and large hydrocarbon sorption Angewandte Chemie 42 5 542 6 doi 10 1002 anie 200390156 PMID 12569485 Uemura T Kitaura R Ohta Y Nagaoka M Kitagawa S June 2006 Nanochannel promoted polymerization of substituted acetylenes in porous coordination polymers Angewandte Chemie 45 25 4112 6 doi 10 1002 anie 200600333 PMID 16721889 Uemura T Hiramatsu D Kubota Y Takata M Kitagawa S 2007 Topotactic linear radical polymerization of divinylbenzenes in porous coordination polymers Angewandte Chemie 46 26 4987 90 doi 10 1002 anie 200700242 PMID 17514689 Ezuhara T Endo K Aoyama Y 1999 Helical Coordination Polymers from Achiral Components in Crystals Homochiral Crystallization Homochiral Helix Winding in the Solid State and Chirality Control by Seeding Journal of the American Chemical Society 121 14 3279 doi 10 1021 ja9819918 Wu ST Wu YR Kang QQ Zhang H Long LS Zheng Z Huang RB Zheng LS 2007 Chiral symmetry breaking by chemically manipulating statistical fluctuation in crystallization Angewandte Chemie 46 44 8475 9 doi 10 1002 anie 200703443 PMID 17912730 Kepert CJ Prior TJ Rosseinsky MJ 2000 A Versatile Family of Interconvertible Microporous Chiral Molecular Frameworks The First Example of Ligand Control of Network Chirality Journal of the American Chemical Society 122 21 5158 5168 doi 10 1021 ja993814s Bradshaw D Prior TJ Cussen EJ Claridge JB Rosseinsky MJ May 2004 Permanent microporosity and enantioselective sorption in a chiral open framework Journal of the American Chemical Society 126 19 6106 14 doi 10 1021 ja0316420 PMID 15137776 Lin Z Slawin AM Morris RE April 2007 Chiral induction in the ionothermal synthesis of a 3 D coordination polymer Journal of the American Chemical Society 129 16 4880 1 doi 10 1021 ja070671y PMID 17394325 Hu A Ngo HL Lin W May 2004 Remarkable 4 4 substituent effects on binap Highly enantioselective Ru catalysts for asymmetric hydrogenation of beta aryl ketoesters and their immobilization in room temperature ionic liquids Angewandte Chemie 43 19 2501 4 doi 10 1002 anie 200353415 PMID 15127435 Wu CD Hu A Zhang L Lin W June 2005 A homochiral porous metal organic framework for highly enantioselective heterogeneous asymmetric catalysis Journal of the American Chemical Society 127 25 8940 1 doi 10 1021 ja052431t PMID 15969557 Wu CD Lin W 2007 Heterogeneous asymmetric catalysis with homochiral metal organic frameworks network structure dependent catalytic activity Angewandte Chemie 46 7 1075 8 doi 10 1002 anie 200602099 PMID 17183496 Cho SH Ma B Nguyen ST Hupp JT Albrecht Schmitt TE June 2006 A metal organic framework material that functions as an enantioselective catalyst for olefin epoxidation Chemical Communications 24 2563 5 doi 10 1039 b600408c PMID 16779478 Hu A Ngo HL Lin W September 2003 Chiral porous hybrid solids for practical heterogeneous asymmetric hydrogenation of aromatic ketones Journal of the American Chemical Society 125 38 11490 1 doi 10 1021 ja0348344 PMID 13129339 Farrusseng D Aguado S Pinel C 2009 Metal organic frameworks opportunities for catalysis Angewandte Chemie 48 41 7502 13 doi 10 1002 anie 200806063 PMID 19691074 Eddaoudi M Kim J Wachter J Chae HK O Keeffe M Yaghi OM 2001 Porous Metal Organic Polyhedra 25 A Cuboctahedron Constructed from 12 Cu2 CO2 4 Paddle Wheel Building Blocks Journal of the American Chemical Society 123 18 4368 9 doi 10 1021 ja0104352 PMID 11457217 Furukawa H Kim J Ockwig NW O Keeffe M Yaghi OM September 2008 Control of vertex geometry structure dimensionality functionality and pore metrics in the reticular synthesis of crystalline metal organic frameworks and polyhedra Journal of the American Chemical Society 130 35 11650 61 doi 10 1021 ja803783c PMID 18693690 Surble S Serre C Mellot Draznieks C Millange F Ferey G January 2006 A new isoreticular class of metal organic frameworks with the MIL 88 topology Chemical Communications 3 284 6 doi 10 1039 B512169H PMID 16391735 Sudik AC Millward AR Ockwig NW Cote AP Kim J Yaghi OM May 2005 Design synthesis structure and gas N 2 Ar CO2 CH4 and H2 sorption properties of porous metal organic tetrahedral and heterocuboidal polyhedra Journal of the American Chemical Society 127 19 7110 8 doi 10 1021 ja042802q PMID 15884953 Degnan T 2003 The implications of the fundamentals of shape selectivity for the development of catalysts for the petroleum and petrochemical industries Journal of Catalysis 216 1 2 32 46 doi 10 1016 S0021 9517 02 00105 7 Kuc A Enyashin A Seifert G July 2007 Metal organic frameworks structural energetic electronic and mechanical properties The Journal of Physical Chemistry B 111 28 8179 86 doi 10 1021 jp072085x PMID 17585800 Esswein AJ Nocera DG October 2007 Hydrogen production by molecular photocatalysis Chemical Reviews 107 10 4022 47 doi 10 1021 cr050193e PMID 17927155 Yang C Messerschmidt M Coppens P Omary MA August 2006 Trinuclear gold I triazolates a new class of wide band phosphors and sensors Inorganic Chemistry 45 17 6592 4 doi 10 1021 ic060943i PMID 16903710 Fuentes Cabrera M Nicholson DM Sumpter BG Widom M 2005 Electronic structure and properties of isoreticular metal organic frameworks The case of M IRMOF1 M Zn Cd Be Mg and Ca The Journal of Chemical Physics 123 12 124713 Bibcode 2005JChPh 123l4713F doi 10 1063 1 2037587 PMID 16392517 Gascon J Hernandez Alonso MD Almeida AR van Klink GP Kapteijn F Mul G 2008 Isoreticular MOFs as efficient photocatalysts with tunable band gap an operando FTIR study of the photoinduced oxidation of propylene ChemSusChem 1 12 981 3 doi 10 1002 cssc 200800203 PMID 19053135 Liu Yufeng Cheng Yuan Zhang He Zhou Min Yu Yijun Lin Shichao Jiang Bo Zhao Xiaozhi Miao Leiying Wei Chuan Wan Liu Quanyi Lin Ying Wu Du Yan Butch Christopher J Wei Hui July 1 2020 Integrated cascade nanozyme catalyzes in vivo ROS scavenging for anti inflammatory therapy Science Advances 6 29 eabb2695 Bibcode 2020SciA 6 2695L doi 10 1126 sciadv abb2695 PMC 7439611 PMID 32832640 Hindocha S Poulston S 2017 Study of the scale up formulation ageing and ammonia adsorption capacity of MIL 100 Fe Cu BTC and CPO 27 Ni for use in respiratory protection filters Faraday Discussions 201 113 125 Bibcode 2017FaDi 201 113H doi 10 1039 c7fd00090a PMID 28612864 a b c Redfern Louis R Farha Omar K 2019 Mechanical properties of metal organic frameworks Chemical Science 10 46 10666 10679 doi 10 1039 C9SC04249K PMC 7066669 PMID 32190239 a b Tan Jin Chong Bennett Thomas D Cheetham Anthony K 2010 05 17 Chemical structure network topology and porosity effects on the mechanical properties of Zeolitic Imidazolate Frameworks Proceedings of the National Academy of Sciences 107 22 9938 9943 Bibcode 2010PNAS 107 9938T doi 10 1073 pnas 1003205107 PMC 2890448 PMID 20479264 Song Jianbo Pallach Roman Frentzel Beyme Louis Kolodzeiski Pascal Kieslich Gregor Vervoorts Pia Hobday Claire L Henke Sebastian 2022 05 16 Tuning the High Pressure Phase Behaviour of Highly Compressible Zeolitic Imidazolate Frameworks From Discontinuous to Continuous Pore Closure by Linker Substitution Angewandte Chemie International Edition 61 21 e202117565 doi 10 1002 anie 202117565 PMC 9401003 PMID 35119185 Moggach Stephen A Bennett Thomas D Cheetham Anthony K 2009 09 07 The Effect of Pressure on ZIF 8 Increasing Pore Size with Pressure and the Formation of a High Pressure Phase at 1 47 GPa PDF Angewandte Chemie International Edition 48 38 7087 7089 doi 10 1002 anie 200902643 hdl 20 500 11820 198bee14 febb 4c8e b6b4 eb424b7ebac0 PMID 19681088 Chapman Karena W Halder Gregory J Chupas Peter J 2009 11 16 Pressure Induced Amorphization and Porosity Modification in a Metal Organic Framework Journal of the American Chemical Society 131 48 17546 17547 doi 10 1021 ja908415z PMID 19916507 Ortiz Aurelie U Boutin Anne Fuchs Alain H Coudert Francois Xavier 2013 05 20 Investigating the Pressure Induced Amorphization of Zeolitic Imidazolate Framework ZIF 8 Mechanical Instability Due to Shear Mode Softening PDF The Journal of Physical Chemistry Letters 4 11 1861 1865 doi 10 1021 jz400880p PMID 26283122 Widmer Remo N Lampronti Giulio I Chibani Siwar Wilson Craig W Anzellini Simone Farsang Stefan Kleppe Annette K Casati Nicola P M MacLeod Simon G Redfern Simon A T Coudert Francois Xavier 2019 05 22 Rich Polymorphism of a Metal Organic Framework in Pressure Temperature Space Journal of the American Chemical Society 141 23 9330 9337 doi 10 1021 jacs 9b03234 PMC 7007208 PMID 31117654 a b Peterson Gregory W DeCoste Jared B Glover T Grant Huang Yougui Jasuja Himanshu Walton Krista S September 2013 Effects of pelletization pressure on the physical and chemical properties of the metal organic frameworks Cu3 BTC 2 and UiO 66 Microporous and Mesoporous Materials 179 48 53 doi 10 1016 j micromeso 2013 02 025 Heinen Jurn Ready Austin D Bennett Thomas D Dubbeldam David Friddle Raymond W Burtch Nicholas C 2018 06 06 Elucidating the Variable Temperature Mechanical Properties of a Negative Thermal Expansion Metal Organic Framework PDF ACS Applied Materials amp Interfaces 10 25 21079 21083 doi 10 1021 acsami 8b06604 PMID 29873475 S2CID 46942254 Durholt Johannes P Keupp Julian Schmid and Rochus 2016 07 27 The Impact of Mesopores on the Mechanical Stability of HKUST 1 A Multiscale Investigation European Journal of Inorganic Chemistry 2016 27 4517 4523 doi 10 1002 ejic 201600566 Chapman Karena W Halder Gregory J Chupas Peter J 2008 07 18 Guest Dependent High Pressure Phenomena in a Nanoporous Metal Organic Framework Material Journal of the American Chemical Society 130 32 10524 10526 doi 10 1021 ja804079z PMID 18636710 Li Hailian Eddaoudi Mohamed O Keeffe M Yaghi O M November 1999 Design and synthesis of an exceptionally stable and highly porous metal organic framework Nature 402 6759 276 279 Bibcode 1999Natur 402 276L doi 10 1038 46248 hdl 2027 42 62847 S2CID 4310761 Alexandre Simone S Mattesini Maurizio Soler Jose M Yndurain Felix 2006 02 22 Comment on Magnetism in Atomic Size Palladium Contacts and Nanowires Physical Review Letters 96 7 079701 author reply 079702 Bibcode 2006PhRvL 96g9701A doi 10 1103 physrevlett 96 079701 PMID 16606151 Hu Yun Hang Zhang Lei 2010 Amorphization of metal organic framework MOF 5 at unusually low applied pressure Physical Review B 81 17 174103 Bibcode 2010PhRvB 81q4103H doi 10 1103 PhysRevB 81 174103 Ortiz Aurelie U Boutin A Fuchs Alain H Coudert Francois Xavier 2013 05 07 Metal organic frameworks with wine rack motif What determines their flexibility and elastic properties PDF The Journal of Chemical Physics 138 17 174703 Bibcode 2013JChPh 138q4703O doi 10 1063 1 4802770 PMID 23656148 Serra Crespo Pablo Dikhtiarenko Alla Stavitski Eli Juan Alcaniz Jana Kapteijn Freek Coudert Francois Xavier Gascon Jorge 2015 Experimental evidence of negative linear compressibility in the MIL 53 metal organic framework family CrystEngComm 17 2 276 280 doi 10 1039 c4ce00436a PMC 4338503 PMID 25722647 Chen Zhijie Hanna Sylvia L Redfern Louis R Alezi Dalal Islamoglu Timur Farha Omar K December 2019 Corrigendum to Reticular chemistry in the rational synthesis of functional zirconium cluster based MOFs Coord Chem Rev 386 2019 32 49 Coordination Chemistry Reviews 400 213050 doi 10 1016 j ccr 2019 213050 S2CID 203136239 Wu Hui Yildirim Taner Zhou Wei 2013 03 07 Exceptional Mechanical Stability of Highly Porous Zirconium Metal Organic Framework UiO 66 and Its Important Implications The Journal of Physical Chemistry Letters 4 6 925 930 doi 10 1021 jz4002345 PMID 26291357 Bennett Thomas D Todorova Tanya K Baxter Emma F Reid David G Gervais Christel Bueken Bart Van de Voorde B De Vos Dirk Keen David A Mellot Draznieks Caroline 2016 Connecting defects and amorphization in UiO 66 and MIL 140 metal organic frameworks a combined experimental and computational study Physical Chemistry Chemical Physics 18 3 2192 2201 arXiv 1510 08220 Bibcode 2016PCCP 18 2192B doi 10 1039 c5cp06798g PMID 27144237 S2CID 45890138 a b Committee on Alternatives and Strategies for Future Hydrogen Production and Use National Research Council National Academy of Engineering eds 2004 The Hydrogen Economy Opportunities Costs Barriers and R amp D Needs Washington D C The National Academies Press pp 11 24 37 44 doi 10 2172 882095 ISBN 978 0 309 09163 3 Ronneau C 2004 11 29 Energie pollution de l air et developpement durable Louvain la Neuve Presses Universitaires de Louvain ISBN 9782875581716 a b Praya Y Mendoza Cortes JL October 2009 Design Principles for High H 2 Storage Using Chelation of Abundant Transition Metals in Covalent Organic Frameworks for 0 700 bar at 298 K Journal of the American Chemical Society 46 138 15204 15213 doi 10 1021 jacs 6b08803 PMID 27792339 S2CID 21366076 DOE Technical Targets for Onboard Hydrogen Storage for Light Duty Vehicles Energy gov Thomas KM March 2009 Adsorption and desorption of hydrogen on metal organic framework materials for storage applications comparison with other nanoporous materials Dalton Transactions 9 1487 505 doi 10 1039 B815583F PMID 19421589 Yuan D Zhao D Sun D Zhou HC July 2010 An isoreticular series of metal organic frameworks with dendritic hexacarboxylate ligands and exceptionally high gas uptake capacity Angewandte Chemie 49 31 5357 61 doi 10 1002 anie 201001009 PMID 20544763 Liu J 2 November 2016 Recent developments in porous materials for H 2 and CH4 storage Tetrahedron Letters 57 44 4873 4881 doi 10 1016 j tetlet 2016 09 085 Sumida K Hill MR Horike S Dailly A Long JR October 2009 Synthesis and hydrogen storage properties of Be 12 OH 12 1 3 5 benzenetribenzoate 4 Journal of the American Chemical Society 131 42 15120 1 doi 10 1021 ja9072707 PMID 19799422 a b c d Furukawa H Ko N Go YB Aratani N Choi SB Choi E Yazaydin AO Snurr RQ O Keeffe M Kim J Yaghi OM July 2010 Ultrahigh porosity in metal organic frameworks Science 329 5990 424 8 Bibcode 2010Sci 329 424F doi 10 1126 science 1192160 PMID 20595583 S2CID 25072457 Rowsell JL Millward AR Park KS Yaghi OM May 2004 Hydrogen sorption in functionalized metal organic frameworks Journal of the American Chemical Society 126 18 5666 7 doi 10 1021 ja049408c PMID 15125649 Rosi NL Eckert J Eddaoudi M Vodak DT Kim J O Keeffe M Yaghi OM May 2003 Hydrogen storage in microporous metal organic frameworks Science 300 5622 1127 9 Bibcode 2003Sci 300 1127R doi 10 1126 science 1083440 PMID 12750515 S2CID 3025509 a b Dincă M Dailly A Liu Y Brown CM Neumann DA Long JR December 2006 Hydrogen storage in a microporous metal organic framework with exposed Mn2 coordination sites Journal of the American Chemical Society 128 51 16876 83 doi 10 1021 ja0656853 PMID 17177438 Lee J Li J Jagiello J 2005 Gas sorption properties of microporous metal organic frameworks Journal of Solid State Chemistry 178 8 2527 2532 Bibcode 2005JSSCh 178 2527L doi 10 1016 j jssc 2005 07 002 Rowsell JL Yaghi OM July 2005 Strategies for hydrogen storage in metal organic frameworks Angewandte Chemie 44 30 4670 9 doi 10 1002 anie 200462786 PMID 16028207 Rowsell JL Yaghi OM February 2006 Effects of functionalization catenation and variation of the metal oxide and organic linking units on the low pressure hydrogen adsorption properties of metal organic frameworks Journal of the American Chemical Society 128 4 1304 15 doi 10 1021 ja056639q PMID 16433549 Garrone E Bonelli B Arean CO 2008 Enthalpy entropy correlation for hydrogen adsorption on zeolites Chemical Physics Letters 456 1 3 68 70 Bibcode 2008CPL 456 68G doi 10 1016 j cplett 2008 03 014 Kubas G J 2001 Metal Dihydrogen and s Bond Complexes Structure Theory and Reactivity New York Kluwer Academic Bellarosa L Calero S Lopez N May 2012 Early stages in the degradation of metal organic frameworks in liquid water from first principles molecular dynamics Physical Chemistry Chemical Physics 14 20 7240 5 Bibcode 2012PCCP 14 7240B doi 10 1039 C2CP40339K PMID 22513503 Bellarosa L Castillo JM Vlugt T Calero S Lopez N September 2012 On the mechanism behind the instability of isoreticular metal organic frameworks IRMOFs in humid environments Chemistry A European Journal 18 39 12260 6 doi 10 1002 chem 201201212 PMID 22907782 Sengupta D Melix P Bose S Duncan J Wang X Mian MR Kirlikovali KO Joodaki F Islamoglu T Yildirim T Snurr RQ Farha OK 20 September 2023 Air Stable Cu I Metal Organic Framework for Hydrogen Storage Journal of the American Chemical Society 145 37 20492 20502 doi 10 1021 jacs 3c06393 PMID 37672758 Stern AC Belof JL Eddaoudi M Space B January 2012 Understanding hydrogen sorption in a polar metal organic framework with constricted channels The Journal of Chemical Physics 136 3 034705 Bibcode 2012JChPh 136c4705S doi 10 1063 1 3668138 PMID 22280775 Dolgonos G 2005 How many hydrogen molecules can be inserted into C60 Comments on the paper AM1 treatment of endohedrally hydrogen doped fullerene nH2 C60 by L Turker and S Erkoc Journal of Molecular Structure THEOCHEM 723 1 3 239 241 doi 10 1016 j theochem 2005 02 017 Tsao CS Yu MS Wang CY Liao PY Chen HL Jeng US Tzeng YR Chung TY Wu HC February 2009 Nanostructure and hydrogen spillover of bridged metal organic frameworks Journal of the American Chemical Society 131 4 1404 6 doi 10 1021 ja802741b PMID 19140765 Mulfort KL Farha OK Stern CL Sarjeant AA Hupp JT March 2009 Post synthesis alkoxide formation within metal organic framework materials a strategy for incorporating highly coordinatively unsaturated metal ions Journal of the American Chemical Society 131 11 3866 8 doi 10 1021 ja809954r PMID 19292487 Huang BL Ni Z Millward A McGaughey AJ Uher C Kaviany M Yaghi O 2007 Thermal conductivity of a metal organic framework MOF 5 Part II Measurement International Journal of Heat and Mass Transfer 50 3 4 405 411 doi 10 1016 j ijheatmasstransfer 2006 10 001 McQuarrie DA Simon JD 1997 Physical Chemistry A Molecular Approach Sausalito CA University Science Books Belof JL Stern AC Eddaoudi M Space B December 2007 On the mechanism of hydrogen storage in a metal organic framework material Journal of the American Chemical Society 129 49 15202 10 doi 10 1021 ja0737164 PMID 17999501 Belof J Stern A Space B 2009 A predictive model of hydrogen sorption for metal organic materials Journal of Physical Chemistry C 113 21 9316 9320 doi 10 1021 jp901988e a b c Zhao D Yan D Zhou HC 2008 The current status of hydrogen storage in metal organic frameworks Energy amp Environmental Science 1 2 225 235 doi 10 1039 b808322n S2CID 44187103 Furukawa H Miller MA Yaghi OM 2007 Independent verification of the saturation hydrogen uptake in MOF 177 and establishment of a benchmark for hydrogen adsorption in metal organic frameworks Journal of Materials Chemistry 17 30 3197 3204 doi 10 1039 b703608f Lowell S Shields JE Thomas MA Thommes M 2004 Characterization of Porous Solids and Powders Surface Area Pore Size and Density Springer Netherlands doi 10 1007 978 1 4020 2303 3 ISBN 978 90 481 6633 6 a b Zuttel A April 2004 Hydrogen storage methods Die Naturwissenschaften 91 4 157 72 Bibcode 2004NW 91 157Z doi 10 1007 s00114 004 0516 x PMID 15085273 S2CID 6985612 Zheng W Tsang CS Lee LY Wong KY June 2019 Two dimensional metal organic framework and covalent organic framework synthesis and their energy related applications Materials Today Chemistry 12 34 60 doi 10 1016 j mtchem 2018 12 002 hdl 10397 101525 S2CID 139305086 Wang Wei Xu Xiaomin Zhou Wei Shao Zongping April 2017 Recent Progress in Metal Organic Frameworks for Applications in Electrocatalytic and Photocatalytic Water Splitting Advanced Science 4 4 1600371 doi 10 1002 advs 201600371 PMC 5396165 PMID 28435777 Cheng Weiren Zhao Xu Su Hui Tang Fumin Che Wei Zhang Hui Liu Qinghua 14 January 2019 Lattice strained metal organic framework arrays for bifunctional oxygen electrocatalysis Nature Energy 4 2 115 122 Bibcode 2019NatEn 4 115C doi 10 1038 s41560 018 0308 8 S2CID 139760912 Liu Mengjie Zheng Weiran Ran Sijia Boles Steven T Lee Lawrence Yoon Suk November 2018 Overall Water Splitting Electrocatalysts Based on 2D CoNi Metal Organic Frameworks and Its Derivative Advanced Materials Interfaces 5 21 1800849 doi 10 1002 admi 201800849 hdl 10397 101550 S2CID 104572148 Zheng Weiran Liu Mengjie Lee Lawrence Yoon Suk 2020 Electrochemical Instability of Metal Organic Frameworks In Situ Spectroelectrochemical Investigation of the Real Active Sites ACS Catalysis 10 81 92 doi 10 1021 acscatal 9b03790 hdl 10397 100175 S2CID 212979103 Zheng Weiran Lee Lawrence Yoon Suk 2021 07 22 Metal Organic Frameworks for Electrocatalysis Catalyst or Precatalyst ACS Energy Letters 6 8 2838 2843 doi 10 1021 acsenergylett 1c01350 hdl 10397 100058 Bunzli JC May 2010 Lanthanide luminescence for biomedical analyses and imaging Chemical Reviews 110 5 2729 55 doi 10 1021 cr900362e PMID 20151630 Amoroso AJ Pope SJ July 2015 Using lanthanide ions in molecular bioimaging PDF Chemical Society Reviews 44 14 4723 42 doi 10 1039 c4cs00293h PMID 25588358 Rosi NL Kim J Eddaoudi M Chen B O Keeffe M Yaghi OM February 2005 Rod packings and metal organic frameworks constructed from rod shaped secondary building units Journal of the American Chemical Society 127 5 1504 18 doi 10 1021 ja045123o PMID 15686384 Duan TW Yan B 2014 06 12 Hybrids based on lanthanide ions activated yttrium metal organic frameworks functional assembly polymer film preparation and luminescence tuning J Mater Chem C 2 26 5098 5104 doi 10 1039 c4tc00414k Xu H Cao CS Kang XM Zhao B November 2016 Lanthanide based metal organic frameworks as luminescent probes Dalton Transactions 45 45 18003 18017 doi 10 1039 c6dt02213h PMID 27470090 Lian X Yan B 2016 01 26 A lanthanide metal organic framework MOF 76 for adsorbing dyes and fluorescence detecting aromatic pollutants RSC Advances 6 14 11570 11576 Bibcode 2016RSCAd 611570L doi 10 1039 c5ra23681a Pansare V Hejazi S Faenza W Prud homme RK March 2012 Review of Long Wavelength Optical and NIR Imaging Materials Contrast Agents Fluorophores and Multifunctional Nano Carriers Chemistry of Materials 24 5 812 827 doi 10 1021 cm2028367 PMC 3423226 PMID 22919122 Della Rocca J Liu D Lin W October 2011 Nanoscale metal organic frameworks for biomedical imaging and drug delivery Accounts of Chemical Research 44 10 957 68 doi 10 1021 ar200028a PMC 3777245 PMID 21648429 Luo TY Liu C Eliseeva SV Muldoon PF Petoud S Rosi NL July 2017 2 Clusters Rational Design Directed Synthesis and Deliberate Tuning of Excitation Wavelengths Journal of the American Chemical Society 139 27 9333 9340 doi 10 1021 jacs 7b04532 PMID 28618777 Foucault Collet A Gogick KA White KA Villette S Pallier A Collet G Kieda C Li T Geib SJ Rosi NL Petoud S October 2013 Lanthanide near infrared imaging in living cells with Yb3 nano metal organic frameworks Proceedings of the National Academy of Sciences of the United States of America 110 43 17199 204 Bibcode 2013PNAS 11017199F doi 10 1073 pnas 1305910110 PMC 3808657 PMID 24108356 Li Y Weng Z Wang Y Chen L Sheng D Diwu J Chai Z Albrecht Schmitt TE Wang S January 2016 Surprising coordination for low valent actinides resembling uranyl vi in thorium iv organic hybrid layered and framework structures based on a graphene like 6 3 sheet topology Dalton Transactions 45 3 918 21 doi 10 1039 C5DT04183J PMID 26672441 S2CID 2618108 Carboni M Abney CW Liu S Lin W 2013 Highly porous and stable metal organic frameworks for uranium extraction Chemical Science 4 6 2396 doi 10 1039 c3sc50230a Wang Y Li Y Bai Z Xiao C Liu Z Liu W Chen L He W Diwu J Chai Z Albrecht Schmitt TE Wang S November 2015 Design and synthesis of a chiral uranium based microporous metal organic framework with high SHG efficiency and sequestration potential for low valent actinides Dalton Transactions 44 43 18810 4 doi 10 1039 C5DT02337H PMID 26459775 Liu W Dai X Bai Z Wang Y Yang Z Zhang L Xu L Chen L Li Y Gui D Diwu J Wang J Zhou R Chai Z Wang S April 2017 Highly Sensitive and Selective Uranium Detection in Natural Water Systems Using a Luminescent Mesoporous Metal Organic Framework Equipped with Abundant Lewis Basic Sites A Combined Batch X ray Absorption Spectroscopy and First Principles Simulation Investigation Environmental Science amp Technology 51 7 3911 3921 Bibcode 2017EnST 51 3911L doi 10 1021 acs est 6b06305 PMID 28271891 Wang LL Luo F Dang LL Li JQ Wu XL Liu SJ Luo MB 2015 Ultrafast high performance extraction of uranium from seawater without pretreatment using an acylamide and carboxyl functionalized metal organic framework Journal of Materials Chemistry A 3 26 13724 13730 doi 10 1039 C5TA01972A Demir S Brune NK Van Humbeck JF Mason JA Plakhova TV Wang S Tian G Minasian SG Tyliszczak T Yaita T Kobayashi T Kalmykov SN Shiwaku H Shuh DK Long JR April 2016 Extraction of Lanthanide and Actinide Ions from Aqueous Mixtures Using a Carboxylic Acid Functionalized Porous Aromatic Framework ACS Central Science 2 4 253 65 doi 10 1021 acscentsci 6b00066 PMC 4850516 PMID 27163056 Banerjee D Kim D Schweiger MJ Kruger AA Thallapally PK May 2016 Removal of TcO4 ions from solution materials and future outlook Chemical Society Reviews 45 10 2724 39 doi 10 1039 C5CS00330J PMID 26947251 Li B Dong X Wang H Ma D Tan K Jensen S Deibert BJ Butler J Cure J Shi Z Thonhauser T Chabal YJ Han Y Li J September 2017 Capture of organic iodides from nuclear waste by metal organic framework based molecular traps Nature Communications 8 1 485 Bibcode 2017NatCo 8 485L doi 10 1038 s41467 017 00526 3 PMC 5589857 PMID 28883637 Abney CW Mayes RT Saito T Dai S December 2017 Materials for the Recovery of Uranium from Seawater Chemical Reviews 117 23 13935 14013 doi 10 1021 acs chemrev 7b00355 OSTI 1412046 PMID 29165997 Wu MX Yang YW June 2017 Metal Organic Framework MOF Based Drug Cargo Delivery and Cancer Therapy Advanced Materials 29 23 1606134 Bibcode 2017AdM 2906134W doi 10 1002 adma 201606134 PMID 28370555 S2CID 30958347 Smaldone RA Forgan RS Furukawa H Gassensmith JJ Slawin AM Yaghi OM Stoddart JF November 2010 Metal organic frameworks from edible natural products Angewandte Chemie 49 46 8630 4 doi 10 1002 anie 201002343 PMID 20715239 Jambhekar SS Breen P February 2016 Cyclodextrins in pharmaceutical formulations I structure and physicochemical properties formation of complexes and types of complex Drug Discovery Today 21 2 356 62 doi 10 1016 j drudis 2015 11 017 PMID 26686054 Hartlieb KJ Ferris DP Holcroft JM Kandela I Stern CL Nassar MS Botros YY Stoddart JF May 2017 Encapsulation of Ibuprofen in CD MOF and Related Bioavailability Studies Molecular Pharmaceutics 14 5 1831 1839 doi 10 1021 acs molpharmaceut 7b00168 PMID 28355489 Noorian Seyyed Abbas Hemmatinejad Nahid Navarro Jorge A R 2019 12 01 BioMOF cellulose fabric composites for bioactive molecule delivery Journal of Inorganic Biochemistry 201 110818 doi 10 1016 j jinorgbio 2019 110818 PMID 31518870 S2CID 202571262 Rojas S Colinet I Cunha D Hidalgo T Salles F Serre C Guillou N Horcajada P March 2018 Toward Understanding Drug Incorporation and Delivery from Biocompatible Metal Organic Frameworks in View of Cutaneous Administration ACS Omega 3 3 2994 3003 doi 10 1021 acsomega 8b00185 PMC 5879486 PMID 29623304 Noorian Seyyed Abbas Hemmatinejad Nahid Navarro Jorge A R 2020 Bioactive molecule encapsulation on metal organic framework via simple mechanochemical method for controlled topical drug delivery systems Microporous and Mesoporous Materials 302 110199 doi 10 1016 j micromeso 2020 110199 S2CID 216532709 Chen X Tong R Shi Z Yang B Liu H Ding S Wang X Lei Q Wu J Fang W January 2018 MOF Nanoparticles with Encapsulated Autophagy Inhibitor in Controlled Drug Delivery System for Antitumor ACS Applied Materials amp Interfaces 10 3 2328 2337 doi 10 1021 acsami 7b16522 PMID 29286625 Talin AA Centrone A Ford AC Foster ME Stavila V Haney P Kinney RA Szalai V El Gabaly F Yoon HP Leonard F Allendorf MD January 2014 Tunable electrical conductivity in metal organic framework thin film devices Science 343 6166 66 9 Bibcode 2014Sci 343 66T doi 10 1126 science 1246738 OSTI 1254264 PMID 24310609 S2CID 206552714 2D self assembling semiconductor could beat out graphene www gizmag com 2 May 2014 Sheberla D Sun L Blood Forsythe MA Er S Wade CR Brozek CK Aspuru Guzik A Dincă M June 2014 High electrical conductivity in Ni3 2 3 6 7 10 11 hexaiminotriphenylene 2 a semiconducting metal organic graphene analogue Journal of the American Chemical Society 136 25 8859 62 doi 10 1021 ja502765n PMID 24750124 S2CID 5714037 Dong Renhao Han Peng Arora Himani Ballabio Marco Karakus Melike Zhang Zhe Shekhar Chandra Adler Peter Petkov Petko St Erbe Artur Mannsfeld Stefan C B November 2018 High mobility band like charge transport in a semiconducting two dimensional metal organic framework Nature Materials 17 11 1027 1032 Bibcode 2018NatMa 17 1027D doi 10 1038 s41563 018 0189 z PMID 30323335 S2CID 53027396 Arora Himani Dong Renhao Venanzi Tommaso Zscharschuch Jens Schneider Harald Helm Manfred Feng Xinliang Canovas Enrique Erbe Artur 2020 Demonstration of a Broadband Photodetector Based on a Two Dimensional Metal Organic Framework Advanced Materials 32 9 1907063 Bibcode 2020AdM 3207063A doi 10 1002 adma 201907063 hdl 11573 1555186 PMID 31975468 S2CID 210882482 Arora Himani June 21 2021 Charge transport in two dimensional materials and their electronic applications Doctoral dissertation PDF Liu Jingjuan Song Xiaoyu Zhang Ting Liu Shiyong Wen Herui Chen Long 2021 03 08 2D Conductive Metal Organic Frameworks An Emerging Platform for Electrochemical Energy Storage Angewandte Chemie 133 11 5672 5684 Bibcode 2021AngCh 133 5672L doi 10 1002 ange 202006102 S2CID 240999292 Day Robert W Bediako D Kwabena Rezaee Mehdi Parent Lucas R Skorupskii Grigorii Arguilla Maxx Q Hendon Christopher H Stassen Ivo Gianneschi Nathan C Kim Philip Dincă Mircea 2019 12 26 Single Crystals of Electrically Conductive Two Dimensional Metal Organic Frameworks Structural and Electrical Transport Properties ACS Central Science 5 12 1959 1964 doi 10 1021 acscentsci 9b01006 PMC 6936098 PMID 31893225 Hoppe Bastian Hindricks Karen D J Warwas Dawid P Schulze Hendrik A Mohmeyer Alexander Pinkvos Tim J Zailskas Saskia Krey Marc R Belke Christopher Konig Sandra Froba Michael Haug Rolf J Behrens Peter 2018 Graphene like metal organic frameworks morphology control optimization of thin film electrical conductivity and fast sensing applications CrystEngComm 20 41 6458 6471 doi 10 1039 C8CE01264D Liang K Ricco R Doherty CM Styles MJ Bell S Kirby N Mudie S Haylock D Hill AJ Doonan CJ Falcaro P June 2015 Biomimetic mineralization of metal organic frameworks as protective coatings for biomacromolecules Nature Communications 6 7240 Bibcode 2015NatCo 6 7240L doi 10 1038 ncomms8240 PMC 4468859 PMID 26041070 MOFs for CO2 MOF Technologies Archived from the original on 2021 02 27 Retrieved 2021 04 07 Choi S Drese JH Jones CW 2009 09 21 Adsorbent materials for carbon dioxide capture from large anthropogenic point sources ChemSusChem 2 9 796 854 doi 10 1002 cssc 200900036 PMID 19731282 a b Sumida K Rogow DL Mason JA McDonald TM Bloch ED Herm ZR Bae TH Long JR February 2012 Carbon dioxide capture in metal organic frameworks Chemical Reviews 112 2 724 81 doi 10 1021 cr2003272 PMID 22204561 Berger AH Bhown AS 2011 01 01 Comparing physisorption and chemisorption solid sorbents for use separating CO2 from flue gas using temperature swing adsorption Energy Procedia 10th International Conference on Greenhouse Gas Control Technologies 4 562 567 doi 10 1016 j egypro 2011 01 089 Smit B Reimer JA Oldenburg CM Bourg IC 2014 Introduction to Carbon Capture and Sequestration London Imperial College Press ISBN 978 1 78326 328 8 Lesch David A 2010 Carbon Dioxide Removal from Flue Gas Using Microporous Metal Organic Frameworks Report doi 10 2172 1003992 OSTI 1003992 span, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.