Metal-organic framework

Metal-organic framework

Metal-Organic Frameworks are crystalline compounds consisting of metal ions or clusters coordinated to often rigid organic molecules to form one-, two-, or three-dimensional structures that can be porous. In some cases, the pores are stable to elimination of the guest molecules (often solvents) and can be used for the storage of gases such as hydrogen and carbon dioxide. Other possible applications of MOFs are in gas purification, in gas separation, in catalysis and as sensors.

Contents

Metal-organic framework structure

A metal-organic framework (MOF) is composed of two major components: a metal ion or cluster of metal ions and an organic molecule called a linker. The organic units are typically mono-, di-, tri-, or tetravalent ligands.[1] The choice of metal and linker has significant effects on the structure and 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.

Coordination Polymers and MOFs

There is no consensus in the scientific literature about the definitions of the terms coordination polymer and metal-organic framework. Some authors suggest definitions based on chemical bonding[2] others propose that the terms coordination polymer and metal-organic framework are synonyms.[3] An IUPAC project was initiated in 2009 to address the terminology issues in this area and will deliver its final report in 2011.[4][5]

Classification of hybrid materials based on dimensionality [6]
Dimensionality of Inorganic
Dimensionality of Organic 0 1 2 3
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


Describing and organizing the complex structures of MOFs could be a difficult and confusing task without a logical, unambiguous set of classifications. Recently, a system of nomenclature has been developed to fill this need. Inorganic sections of a MOF, called secondary building units (SBU), 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. The database of net structures can be found at the Reticular Chemistry Structure Resource.[2]

Common ligands in MOFs

Common name IUPAC name Chemical formula Structural formula
Bidentate Carboxylics
Oxalic acid ethanedioic acid HOOC-COOH Oxalic acid.png
Malonic acid propanedioic acid HOOC-(CH2)-COOH Malonic acid structure.png
Succinic acid butanedioic acid HOOC-(CH2)2-COOH Succinic acid.png
Glutaric acid pentanedioic acid HOOC-(CH2)3-COOH Glutaric acid.png
Phthalic acid benzene-1,2-dicarboxylic acid
o-phthalic acid
C6H4(COOH)2 Phthalic-acid-2D-skeletal.png
Isophthalic acid benzene-1,3-dicarboxylic acid
m-phthalic acid
C6H4(COOH)2 Isophthalic-acid-2D-skeletal.png
Terephthalic acid benzene-1,4-dicarboxylic acid
p-phthalic acid
C6H4(COOH)2 Terephthalic-acid-2D-skeletal.png
Tridentate Carboxylates
Citric Acid 2-Hydroxy-1,2,3-propanetricarboxylic acid (HOOC)CH2C(OH)(COOH)CH2(COOH) Zitronensäure - Citric acid.svg
Trimesic acid benzene-1,3,5-tricarboxylic acid C9H6O6 Trimesic acid.svg
Azoles
1,2,3-Triazole 1H-1,2,3-triazole C2H3N3 1,2,3-triazole numbering.png
pyrrodiazole 1H-1,2,4-triazole C2H3N3 1,2,4-triazole numbering.png
Other
Squaric acid 3,4-Dihydroxy-3-cyclobutene-1,2-dione C4H2O4 Squaric acid.png

Synthesis of MOFs

The study of MOFs developed from the study of zeolites, with very little change in synthetic technique. MOFs and zeolites alike are produced almost exclusively by hydrothermal or solvothermal techniques, where crystals are slowly grown from a hot solution of metal precursor, such as metal nitrates, and bridging ligands.[7] Zeolite synthesis often makes use of a variety of templates, or structure-directing compounds, and a few examples of templating, particularly by organic anions, are seen in the MOF literature as well.[8][9] A particular 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 fully evacuated, allowing hydrogen to bind at these sites.[10]

Post-synthetic modification of MOFs opens up another dimension of structural possibilities that might not be achieved by conventional synthesis. A great deal of recent work explores covalent modification of the bridging ligands.[11] Of particular interest to MOFs for hydrogen storage are modifications which expose metal sites. This has been demonstrated with post-synthetic coordination of additional metal ions to sites on the bridging ligands,[10][11] and addition and removal of metal atoms to the metal site.[10][12]

Since ligands in MOFs typically bind reversibly, the slow growth of crystals 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.[13][14] This technique, termed “microwave-assisted solvothermal synthesis”, is widely used in the zeolite literature,[7] and produces micron-scale crystals in a matter of seconds to minutes,[13][14] in yields similar to the slow growth methods.

A solvent-free synthesis of a range of crystalline MOFs have been described by Stuart James and his group[15]. Usually the metal acetate is reacted with the organic ligand of choice, placed in a stainless steel vessel with a ball bearing, mixed and grinded in a ball mill mixer. Cu3(BTC)2 has been synthesised in this way with a quantitative yield. The reaction is quick and can be monitored with XRPD. In the ball mill it is found that the morphology of the MOF may differ from the expected. In the case of Cu3(BTC)2 the morphology of the solvent free synthesised product was the same as the industrial made, Basolite C300. It is thought that localised melting when the molecules collide 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[16] in the ball mill.

Composite MOF materials

Another approach to increasing adsorption in MOFs is to alter the system in such a way that chemisorption becomes possible. This has been achieved 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 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.[17][18] The relationship between hydrogen spillover and hydrogen storage properties in MOFs is not well understood, but further research in this direction may provide inexpensive boosts in hydrogen storage capacity.

MOFs for hydrogen storage

Given the imminent depletion of petroleum reserves, there is considerable interest in the development of non-petroleum energy carriers for use in transportation. Hydrogen has the potential to be an attractive option because it has a high energy content (120 MJ/kg compared to 44 MJ/kg for gasoline), produces clean exhaust product (water vapor without CO2 or NOx), and can be derived from a variety of primary energy sources. However, the specific energy of uncompressed hydrogen gas is very low, and considerable attention must be given to denser storage methods if hydrogen is to emerge as a serious option for energy storage.[19]

Proposed forms of reversible hydrogen storage include: compressed gas, cryogenic liquid, adsorption to high surface-area materials, chemical storage as metal hydrides, and various reactions of liquid fuels high in hydrogen content (whose products must be collected and recycled after use).[20][21] Of these, compressed and liquid hydrogen are the most mature technologies and are the most suitable for immediate deployment.[21] The United States Department of Energy (USDOE) projects that with further technological development, adsorptive or chemical storage may prove most effective for storage.[20]

Metal Organic Frameworks (MOFs) attract attention as materials for adsorptive hydrogen storage because of their exceptionally high specific surface areas and chemically-tunable structures.[17] MOFs can be thought of as a three-dimensional grid in which the vertices are metal ions or clusters of metal ions that are connected to each other by organic molecules called linkers. Hydrogen molecules are stored in a MOF by adsorbing to its surface. Compared to an empty gas cylinder, a MOF-filled gas cylinder can store more gas because of adsorption that takes place on the surface of MOFs. (Note that wikt:molecular hydrogen adsorbs to the surface, not wikt:atomic hydrogen.) 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.[1] Also, MOFs have a fully reversible uptake-and-release behavior: since the storage mechanism is based primarily on physisorption, there are no large activation barriers to be overcome when liberating the adsorbed hydrogen.[1] 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.[1]

In order to realize the benefits provided, such as adsorption, by MOFs hydrogen cannot be stored in them at densities greater than its liquid-phase density.[1] 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[1] (until a maximum is reached, typically 20-30 bar, after which the adsorption capacity decreases).[17],[1] 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.[17]

US Department of Energy hydrogen storage guidelines

Despite the fact that the US DOE Secretary has questioned the viability of existing hydrogen storage methods, the search for high-capacity hydrogen storage materials remains a highly competitive area of research: the race is on to develop MOFs that can meet all of the targets set by the DOE. The DOE 2015 targets for a hydrogen storage system are [22] : 1) a capacity of 40 g H2 per L, 2) a refueling time of 10 min or less, 3) a lifetime of 1000 refueling cycles, and 4) an ability to operate within the temperature range 30 to 50 °C. Note that these targets are for the entire storage system; therefore, the performance of a storage material must be even higher in order to account for the storage container and, if necessary, a temperature regulating apparatus.[17] MOF-177 currently boasts the hydrogen absorption record, with a surface area of 4526 m2/g and an excess hydrogen uptake of 1.23 wt% and 32.1 g/L at 1 bar and 77 K.[23]

Examples of MOFs for hydrogen storage

The most important challenge for creating hydrogen adsorbents that operate at room temperature is increasing the hydrogen binding energy.[17] 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 in the literature 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.[17]

The most common transition metals employed in carboxylate-based frameworks are Cu2+ or 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.[24] 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.[17]

The following is a list several MOFs that are considered to have the best properties for hydrogen storage as of November 2009 (in order of decreasing hydrogen storage capacity).[17] While each MOF described has its advantages, none of these MOFs reach all of the standards set by the USDOE. 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.[1]

Zn4O(BTB)2, where BTB3- = 1,3,5-benzenetribenzoate (MOF-177) [25]

Structure: 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).

Hydrogen storage capacity: 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.[17]

Zn4O(BDC)3, where BDC2- = 1,4-benzenedicarboxylate (MOF-5) [26]

Structure: Square openings are either 13.8 or 9.2 Å depending on the orientation of the aromatic rings.

Hydrogen storage capacity: 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).[17]

Mn3[(Mn4Cl)3(BTT)8]2, where H3BTT = benzene-1,3,5-tris(1H-tetrazole) [27]

Structure: Consists of truncated octahedral cages that share square faces, leading to pores of about 10 Å in diameter. Contains open Mn2+ coordination sites.

Hydrogen storage capacity: 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.[17]

Cu3(BTC)2(H2O)3, where H3BTC = 1,3,5-benzenetricarboxylic acid [28]

Structure: 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.[17]

Structural impacts on hydrogen storage capacity

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.[29][30]

Surface area

The general trend in MOFs used for hydrogen storage is that the greater the surface area, the more hydrogen the MOF can store. This is because high surface area materials tend to exhibit increased micropore volume and inherently low bulk density, allowing for more hydrogen adsorption to occur.[17]

Hydrogen adsorption enthalpy

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.[31] The interaction between hydrogen and uncharged organic linkers is not this strong, and so a considerable amount of work has gone into 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.[10][27] (C5H5)V(CO)3(H2) and Mo(CO)5(H2) are great examples of increased binding energy due to open metal coordination sites;[32] 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.[17] 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.[17] 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.[10][17]

Sensitivity to air

MOFs are frequently air-sensitive. 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.[17]

Pore size

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.[17] 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.[17] (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.) [33]

Structural defects

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.[34]

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.[35] Finally, structural defects can affect the transport of phonons, which affects the thermal conductivity of the MOF.[36]

Hydrogen adsorption

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.[37]

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.[17]

Determining hydrogen storage capacity

For the characterization of MOFs as hydrogen storage materials, there are two hydrogen-uptake measurement methods: 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:[38]

Nabs=Nex+(ρbulk)(Vpore)

Gravimetric method

The increased mass of the MOF due to the stored hydrogen is directly calculated by a highly sensitive microbalance.[38] The mass of the adsorbed hydrogen decreases when high pressure is applied to the system due to its buoyancy. This weight loss is calculated by the volume of the MOF’s frame and the density of hydrogen.[39]

Volumetric method

The changing of amount of hydrogen stored in the MOF is measured by detecting the varied pressure of hydrogen at constant volume.[38] 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.[40]

Other methods of hydrogen storage

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):[41]

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 pressure required for storing hydrogen liquid. The other methods are all being studied and developed extensively.[41]

Other applications of MOFs

There are many uses of MOFs other than hydrogen storage, such as gas purification, gas separation, gas storage (other than hydrogen), and heterogeneous catalysis. MOFs are used for gas purification because of strong chemisorption that takes place between electron-rich, odor-generating molecules (such as amines, phosphines, oxygenates, alcohols, water, or sulfur-containing molecules) and the framework, allowing the desired gas to pass through the MOF. Gas separation can be performed with MOFs because they can allow certain molecules to pass through their pores based on size and kinetic diameter. This is particularly important for separating out carbon dioxide. Regarding gas storage, MOFs can store molecules such as carbon dioxide, carbon monoxide, methane, and oxygen due to their high adsorption enthalpies (similar to hydrogen). Finally, MOFs are used for catalysis because of their shape and size selectivity and their accessible bulk volume. Also, because of their very porous architecture, mass transport in the pores is not hindered.[1]

See also


References

  1. ^ a b c d e f g h i Czaja, Alexander U.; Trukhan, Natalia; Müller, Ulrich (2009). "Industrial applications of metal-organic frameworks". Chemical Society Reviews 38 (5): 1284–1293. doi:10.1039/b804680h. PMID 19384438. 
  2. ^ a b Tranchemontagne, David J.; Mendoz-Cortés, José L.; O'Keefe, Michael; Yaghi, Omar M. (2009). "Secondary building units, nets and bonding in the chemistry of metal-organic frameworks". Chemical Society Reviews 38 (5): 1257–1283. doi:10.1039/b906666g. PMID 19384437. 
  3. ^ Janiak, Christoph (2003). "Engineering coordination polymers towards applications". Dalton Transactions: 2781–2804. doi:10.1039/b305705bg. 
  4. ^ IUPAC Current Projects. May 09, 2010. Retrieved on May 09, 2010.
  5. ^ CP and MOF Project. May 09, 2010. Retrieved on May 09, 2010.
  6. ^ Cheetham, Rao, and Feller. Structural diversity and chemical trends in hybrid inorganic-organic framework materials. Chem. Comm. 46 (2006) 4780. doi:10.1039/b610264f
  7. ^ a b Cheetham, Anthony K.; Ferey, Gerard; Loiseau, Thierry (1999). "Open-framework inorganic materials". Angewandte Chemie, International Edition 38 (22): 3268–3292. doi:10.1002/(SICI)1521-3773(19991115)38:22<3268::AID-ANIE3268>3.0.CO;2-U. PMID 10602176. 
  8. ^ Bucar, Dejan-Kresimir; Papaefstathiou, Giannis S.; Hamilton, Tamara D.; Chu, Quanli L.; Georgiev, Ivan G.; MacGillivray, Leonard R. (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. 
  9. ^ Parnham, Emily R.; Morris, Russell E. (2007). "Ionothermal Synthesis of Zeolites, Metal-Organic Frameworks, and Inorganic-Organic Hybrids.". Accounts of Chemical Research 40 (10): 1005–1013. doi:10.1021/ar700025k. PMID 17580979. 
  10. ^ a b c d e Dincǎ, Mircea; Long, Jeffrey R. (2008). "Hydrogen storage in microporous metal-organic frameworks with exposed metal sites". Angewandte Chemie International Edition 47 (36): 6766–6779. doi:10.1002/anie.200801163. PMID 18688902. 
  11. ^ a b Wang, Zhenqiang; Cohen, Seth M. (2009). "Postsynthetic modification of metal–organic frameworks". Chemical Society Reviews 38 (5): 1315–1329. doi:10.1039/b802258p. PMID 19384440. 
  12. ^ Das, Sunirban; Kim, Hyunuk; Kim, Kimoon (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): 3814–3815. doi:10.1021/ja808995d. PMID 19256486. 
  13. ^ a b Ni, Zheng; Masel, Richard I. (2006). "Rapid production of metal−organic frameworks via microwave-assisted solvothermal synthesis". Journal of the American Chemical Society 128 (38): 12394–12395. doi:10.1021/ja0635231. PMID 16984171. 
  14. ^ a b Choi, Jung-Sik; Son, Won-Jin; Kim, Jahoen; Ahn, Wha-Seung (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. 
  15. ^ S.James and A. Pichon CrystEngComm, 2008,10,p1839
  16. ^ D. Braga CrystEngComm 2007, 9, p879
  17. ^ a b c d e f g h i j k l m n o p q r s t u Murray, Leslie J.; Dincǎ, Mircea; Long, Jeffrey R. (2009). "Hydrogen storage in metal-organic frameworks". Chemical Society Reviews 38 (5): 1294–1314. doi:10.1039/b802256a. PMID 19384439. 
  18. ^ Li, Yingwei; Yang, Ralph T. (2007). "Hydrogen Storage on Platinum Nanoparticles Doped on Superactivated Carbon". Journal of Physical Chemistry C 111 (29): 11086–11094. doi:10.1021/jp072867q. 
  19. ^ Committee on Alternatives and Strategies for Future Hydrogen Production and Use, National Research Council, National Academy of Engineering. (2004). The Hydrogen Economy: Opportunities, Costs, Barriers, and R&D Needs. Washington, D.C.: National Academies. pp. 11–24, 37–44. ISBN 0-309-09163-2. 
  20. ^ a b United States Department of Energy (2004). "Basic Research Needs for the Hydrogen Economy." 
  21. ^ a b Eberle, Ulrich; Felderhoff, Michael; Schüth, Ferdi (2009). "Chemical and physical solutions for hydrogen storage". Angewandte Chemie International Edition 48 (36): 6608–6630. doi:10.1002/anie.200806293. PMID 19598190. 
  22. ^ Yang, J.; A. Sudik, C. Wolverton, D. J. Siegel (2010). "High capacity hydrogen storage materials: attributes for automotive applications and techniques for materials discovery". Chem. Soc. Rev. 39 (2): 656–675. doi:10.1039/b802882f. PMID 20111786. http://pubs.rsc.org/en/Content/ArticleLanding/2010/CS/b802882f. 
  23. ^ Yan, Yong; Lin, Xiang; Yang, Sihai; Blake, Alexander J.; Dailly, Anne; Champness, Neil R.; Hubberstey, Peter; Schröder, Martin (2009). "Exceptionally high H2 storage by a metal-organic polyhedral framework". Chemical Communications (9): 1025–1027. doi:10.1039/b900013e. 
  24. ^ Sumida, Kenji; Hill, Matthew R.; Horike, Satoshi; Dailly, Anne; Long, Jeffrey R. (2009). "Synthesis and hydrogen storage properties of Be12(OH)12(1,3,5-benzenetribenzoate)4". Journal of the American Chemical Society 131 (42): 15120–15121. doi:10.1021/ja9072707. PMID 19799422. 
  25. ^ Rowsell, Jesse L. C.; Millward, Andrew R.; Park, Kyo Sung; Yaghi, Omar M. (2004). "Hydrogen sorption in functionalized metal−organic frameworks". Journal of the American Chemical Society 126 (18): 5666–5667. doi:10.1021/ja049408c. PMID 15125649. 
  26. ^ Rosi, Nathaniel L.; Eckert, Juergen; Eddaoudi, Mohamed; Vodak, David T.; Kim, Jaheon; O'Keefe, Michael; Yaghi, Omar M. (2003). "Hydrogen storage in microporous metal-organic frameworks". Science 300 (5622): 1127–1129. doi:10.1126/science.1083440. PMID 12750515. 
  27. ^ a b Dincǎ, Mircea; Dailly, Anne; Liu, Yun; Brown, Craig M.; Neumann, Dan A.; Long, Jeffrey R. (2006). "Hydrogen storage in a microporous metal−organic framework with exposed Mn2+ coordination sites". Journal of the American Chemical Society 128 (51): 16876–16883. doi:10.1021/ja0656853. PMID 17177438. 
  28. ^ Lee, JeongYong; Li, Jing; Jagiello, Jacek (2005). "Gas sorption properties of microporous metal organic frameworks". Journal of Solid State Chemistry 178 (8): 2527–2532. doi:10.1016/j.jssc.2005.07.002. 
  29. ^ Rowsell, Jesse L. C.; Yaghi, Omar M. (2005). "Strategies for hydrogen storage in metal-organic frameworks". Angewandte Chemie International Edition 44 (30): 4670–4679. doi:10.1002/anie.200462786. PMID 16028207. 
  30. ^ Rowsell, Jesse L. C.; Yaghi, Omar M. (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): 1304–1315. doi:10.1021/ja056639q. PMID 16433549. 
  31. ^ Garrone, E; Bonelli, B; Arean, C. O. (2008). "Enthalpy-entropy correlation for hydrogen adsorption on zeolites". Chemical Physics Letters 456 (1–3): 68–70. doi:10.1016/j.cplett.2008.03.014. 
  32. ^ Kubas, G. J. (2001). Metal Dihydrogen and σ-Bond Complexes: Structure, Theory, and Reactivity. New York: Kluwer Academic. 
  33. ^ Dolgonos, Grygoriy (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 Ş. Erkoç'". Journal of Molecular Structure: THEOCHEM 723 (1–3): 239–241. doi:10.1016/j.theochem.2005.02.017. 
  34. ^ Tsao, Cheng-Si; Yu, MingSheng; Wang, Cheng-Yu; Liao, Pin-Yen; Chen, Hsin-Lung; Jeng, U-Ser; Tzeng, Yi-Ren; Chung, Tsui-Yun et al. (2009). "Nanostructure and hydrogen spillover of bridged metal-organic frameworks". Journal of the American Chemical Society 131 (4): 1404–1406. doi:10.1021/ja802741b. PMID 19140765. 
  35. ^ Mulfort, Karen L.; Farha, Omar K.; Stern, Charlotte L.; Sarjeant, Amy A.; Hupp, Joseph T. (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): 3866–3868. doi:10.1021/ja809954r. PMID 19292487. 
  36. ^ Huang, B. L.; Ni, Z.; Millward, A.; McGaughey, A. J. H.; 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. 
  37. ^ McQuarrie, D. A.; Simon, J. D. (1997). Physical Chemistry: A Molecular Approach. Sausalito, CA: University Science Books. 
  38. ^ a b c Zhao, Dan; Yan, Daqiang; Zhou, Hong-Cai (2008). "The current status of hydrogen storage in metal–organic frameworks". Energy & Environmental Science 1 (1): 225–235. doi:10.1039/b808322n. 
  39. ^ Furukawa, Hiroyasu; Miller, Michael A.; Yaghi, Omar M. (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. 
  40. ^ Lowell, S.; Shields, Joan E.; Thomas, Martin A.; Thommes, Matthias (2004). Characterization of Porous Solids and Powders: Surface Area, Pore Size and Density. Kluwer Academic Publishers. 
  41. ^ a b Züttel, Andreas (2004). "Hydrogen storage methods". Naturwissenschaften 91 (4): 157–172. doi:10.1007/s00114-004-0516-x. PMID 15085273. 

Wikimedia Foundation. 2010.

Игры ⚽ Нужен реферат?

Look at other dictionaries:

  • Metal Organic Framework — Metall organische Gerüste (engl. metal organic frameworks, MOF) sind mikroporöse kristalline Materialien, die aus metallischen Knotenpunkten, den sogenannten SBUs (Structural Building Units) und organischen Molekülen (Linkern) als… …   Deutsch Wikipedia

  • Metal-organic Framework — Metall organische Gerüste (engl. metal organic frameworks, MOF) sind poröse Materialien mit wohlgeordneter kristalliner Struktur. Sie bestehen aus Komplexen mit Übergangsmetallen (meist Cu , Zn , Ni oder Co) als Knoten und organischen Molekülen… …   Deutsch Wikipedia

  • metal-organic framework — noun Any of several highly porous, crystalline substances having a cage structure of metal ions coordinated to organic compounds; they have an application in the bulk storage of gases such as hydrogen …   Wiktionary

  • Covalent organic framework — The design and synthesis of crystalline extended organic structures in which the building blocks are linked by strong covalent bonds are core concepts of covalent organic frameworks (COFs). COFs are porous, and crystalline, and made entirely from …   Wikipedia

  • Metal nitrosyl complex — Metal nitrosyl complexes are complexes that contain nitric oxide, NO, bonded to a transition metal.[1] Many kinds of nitrosyl complexes are known, which vary both in structure and coligand. Contents 1 Bonding and structure 1.1 Linear vs bent… …   Wikipedia

  • alkali metal — Chem. any of the group of univalent metals including potassium, sodium, lithium, rubidium, cesium, and francium, whose hydroxides are alkalis. [1880 85] * * * Any of the six chemical elements in the leftmost group of the periodic table (lithium,… …   Universalium

  • Wasserstoffspeicher — Die Wasserstoffspeicherung ist Teil einer Wasserstoffwirtschaft und der Wasserstoffherstellung nachgelagert. Die Eigenschaften des Wasserstoffs stellen bei seiner Speicherung ein besonderes wirtschaftliches und sicherheitstechnisches Problem dar …   Deutsch Wikipedia

  • Wasserstofftank — Die Wasserstoffspeicherung ist Teil einer Wasserstoffwirtschaft und der Wasserstoffherstellung nachgelagert. Die Eigenschaften des Wasserstoffs stellen bei seiner Speicherung ein besonderes wirtschaftliches und sicherheitstechnisches Problem dar …   Deutsch Wikipedia

  • Wasserstoffspeicherung — Die Wasserstoffspeicherung ist Teil der Wasserstoffwirtschaft. Konventionelle Methoden der Speicherung und Lagerung von Wasserstoff sind: Druckgasspeicherung (Speicherung in Druckbehältern durch Verdichten mit Kompressoren) Flüssiggasspeicherung… …   Deutsch Wikipedia

  • Omar M. Yaghi — (* 1965 in Amman, Jordanien) ist Professor für Chemie und Biochemie an der University of California, Los Angeles. Omar M. Yaghi studierte an der University at Albany, The State University of New York, wo er 1984 seinen Bachelor erhielt. 1990… …   Deutsch Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”