Вы находитесь на странице: 1из 11

View Article Online / Journal Homepage / Table of Contents for this issue

APPLICATION

www.rsc.org/materials | Journal of Materials Chemistry

Metalorganic frameworksprospective industrial applications{


U. Mueller,* M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt and J. Pastre
Received 22nd August 2005, Accepted 26th October 2005 First published as an Advance Article on the web 23rd November 2005 DOI: 10.1039/b511962f The generation of metalorganic framework (MOF) coordination polymers enables the tailoring of novel solids with regular porosity from the micro to nanopore scale. Since the discovery of this new family of nanoporous materials and the concept of so called reticular design, nowadays several hundred different types of MOF are known. The self assembly of metal ions, which act as coordination centres, linked together with a variety of polyatomic organic bridging ligands, results in tailorable nanoporous host materials as robust solids with high thermal and mechanical stability. Describing examples of different zinc-containing structures, e.g. MOF-2, MOF-5 and IRMOF8 verified synthesis methods will be given, as well as a totally novel electrochemical approach for transition metal based MOFs will be presented for the first time. With sufficient amounts of sample now being available, the testing of metalorganic frameworks in fields of catalysis and gas processing is exemplified. Report is given on the catalytic activation of alkynes (formation of methoxypropene from propyne, vinylester synthesis from acetylene). Removal of impurities in natural gas (traces of tetrahydrothiophene in methane), pressure swing separation of rare gases (krypton and xenon) and storage of hydrogen (3.3 wt% at 2.0 MPa/77 K on Cu-BTC-MOF) will underline the prospective future industrial use of metal organic frameworks in gas processing. Whenever possible, comparison is made to state-of-art applications in order to outline possibilities which might be superior by using MOFs.

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

1. Introduction
As early as 1965 a first publication by Tomic1 on novel solids was introduced which, nowadays, would be categorized and addressed as metalorganic frameworks, coordination polymers or supramolecular structures. Already in the aforementioned contribution simple syntheses of coordination polymers
BASF Aktiengesellschaft, Chemicals Research & Engineering, D-67056, Ludwigshafen, Germany. E-mail: ulrich.mueller@basf-ag.de { Presented at Symposium T: Porous materials for emerging applications, International Conference on Materials for Advanced Technologies (ICMAT 2005), Singapore, 38 July 2005.

based on metals like zinc, nickel, iron, aluminium (but also on thorium and uranium) employing bi- to tetravalent aromatic carboxylic acids are described . Interesting features of these compounds such as high thermal stability and high metal content were already investigated. However, decades later interest in the field was stimulated by the group of O. M. Yaghi, which published the structure of MOF-5 in late 1999,2 and the concept of reticular design, with totally different carboxylate linkers, in 2002.35 Meanwhile, numerous reviews have addressed this fast growing research efforts, the most comprehensive ones given by Kitagawa6 and Yaghi.7 Structures, properties and possible applications as

Ulrich Mueller, born 1957 in Katzenelnbogen, Germany. 1977: studied chemistry in Mainz (thesis on the synthesis of large zeolite crystals and sorption properties) and recieved his PhD in the group of Prof. K.K. Unger; research activities at CNRS Tian&Calvet, Marseille, ILL Grenoble, and with G.T. Kokotailo, Univ. Pennsylvania. 1989: Ammonia Laboratory BASF: zeolite synthesis and application in catalysis and adsorption. 1999: Senior Scientist, zeolite catalysis: CFC-free polyurethane foams, catalysts for crop protection agents, chemical intermediates, sorptive olefin feedstream
626 | J. Mater. Chem., 2006, 16, 626636

purification, piloting of propylene epoxidation catalysts. 1999: Synthesis, scale-up, modification and testing of various metalorganic framework compositions. 2005: BASF Research Director. Markus M. Schubert, born 1971 in Munich, Germany. 1991: study of chemistry in Ulm and PhD in group of Prof. Behm on catalysis and surface chemistry. 2000: Postdoc at ETH Zurich with Prof. Baiker. 2001: Ammonia Laboratory BASF: catalysts carriers, acidbase catalysis. 2004: Scale-up and piloting of metalorganic frameworks for gas processing.
This journal is The Royal Society of Chemistry 2006

View Article Online

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

storage media were again studied by Rowsell, from Yaghis group.7,8 Comparisons with oxides, molecular sieves, porous carbon and heteropolyanion salts has been filed by Barton and coauthors.9 Nowadays several hundred different MOFs are known. The self assembly of metal ions, which act as coordination centres, linked together with a variety of polyatomic organic bridging ligands, results in tailored nanoporous host materials as robust solids with high thermal and mechanical stability (Fig. 1). Interestingly, unlike other solid matter, e.g. zeolites, carbons and oxides, a number of coordination compounds are known to exhibit high framework flexibility and shrinkage/expansion due to interaction with guest molecules.6 The most striking difference to state-ofart materials is probably the total lack of non-accessible bulk volume in metalorganic framework structures. Although high surface areas are already known from activated carbons and zeolites as well, it is the absence of any dead volume in MOFs which principally gives them (on a weight-specific basis) the highest porosities and world record surface areas (Fig. 2), especially with MOF-177, for which values of 4500 m2 g21 are reported.5 Of course, properties like the drastically increased

velocity of molecular traffic through these open structures are closely related to the regularity of pores in nanometer size as well. Thus the combination of so far unreached porosity, surface area, pore size and wide chemical inorganicorganic composition recently brought these materials to the attention of many researchers in both academia and industry, with about 1000 publications on coordination polymers per annum.6 This paper, however, aims to describe how MOF-materials can be synthesized using verified synthetic methods as well as by a totally novel electrochemical approach.10 With a large range of samples now available, the testing of metalorganic frameworks in fields of catalysis and gas processing is enabled. A report is given on the catalytic activation of alkynes (the formation of methoxypropene from propyne, vinylester synthesis from acetylene).11 Further examples like olefin polymerization, DielsAlder reaction, transesterification6 or cyanosilylation12 are referenced in the literature. The removal of impurities in natural gas (i.e. traces of tetrahydrothiophene in methane), pressure swing separation of

Friedhelm Teich, born 1955 in Trier, Germany. 1973: study of Chemical Engineering at Karlsruhe (PhD on gas processing). 1986: Dyestuff & Pigments laboratory BASF. 2004: BASF Chemicals Research & Engineering, 20 years experience in designing and developing processes for the production of pigments and fine chemicals. 2004: process simulation for application of metalorganic framework materials. Hermann Putter, born 1944 in Duesseldorf, Germany. 1951 1964: school in Duesseldorf. 19641972: study of chemistry in Wuerzburg (thesis on the preparation and elucidation of the electrochemical properties of squaric acid derivatives). 1969: summer/autumn: electroanalytical studies at the Heyrovsky Institute in Prague. 19731985: Main Laboratory BASF,

Ludwigshafen, development of direct and indirect organic electrosyntheses, commercialisation of the first electrodialysis processes in our company. 19851992: plant manager of a chloralkali plant in Ludwigshafen. 1990: establishment of the first chlorine based chemistree of Germany for VCI. During this time: papers, public discussions and lectures on chlorine chemistry. Since 1993: Manager of R&D activities on all electrochemical processes of our company. Since 1994: lectures on environmental and sustainability aspects of chemistry. 1999: BASF innovation award for the first technical paired electrosynthesis, a process with high atom efficiency that avoids emissions and halves the energy demand of an electrosynthesis. 2001: BASF Research Fellow. 2003, Synthesis of metalorganic frameworks using electrochemistry.

Kerstin Schierle-Arndt, born 1971 in Koln, Germany. 1990: study of chemistry in Bonn; PhD in organic electrochemistry. 1998: Ammonia Laboratory BASF: electrochemical research. 2003: Chemicals Research & Engineering, Head of Controlling. 2005: New Business development at BASFs Inorganic Specialties.

Joerg Pastre, born 1970, in Gro-Gerau, Germany. 1977: study of chemistry in Darmstadt. 2000: PhD on Chemical Engineering at ETH Zu rich. 2000: Ammonia laboratory BASF: process development department. 2005: New business development at BASF Inorganic Specialties.

This journal is The Royal Society of Chemistry 2006

J. Mater. Chem., 2006, 16, 626636 | 627

View Article Online

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

Fig. 3 Principle flowsheet scheme of industrial MOF synthesis procedure. Cost efficiency and sustainability requires solvent recycling and zinc oxide rather than zinc nitrate. Fig. 1 Illustration of metalcarboxylate building units of MOFs (upper left: MOF-5 with a Zn4O-cluster linked to the terephthalic acid molecules, upper right: IRMOF-8 with a Zn4O-cluster attached to the 2,6-naphthalene dicarboxylic acid, lower left: Cu-BTC with a dimeric Cu-cluster terminated by 1,3,5-benzenetricarboxylic acid, lower right: MOF-2 showing the paddle-wheel of a dimeric Zn-cluster linked to the terephthalic acid units).

repeatedly checked and scaled-up to the order of kg. A typical scheme of a semi-technical process is given in Fig. 3, indicating the different steps of preparation, as well as recycling of the solvent and further processing of the dried powders into shaped particles. 2.1 Verified lab-scale synthesis recipes of MOF samples MOF-2. A glass reactor equipped with a reflux condenser and a teflon-lined stirrer was filled with 24.9 g of terephthalic acid (BDC) and 52.2 g of zinc nitrate tetrahydrate (Merck) were dissolved in a mixture of 43.6 g of N-methyl-2pyrrolidone, 8.6 g of chlorobenzene and 24.9 g of dimethylformamide (Merck) and heated up to 70 uC for a total of 3 h. After about 60 min, 30 g triethylamine was added. The white precipitate formed was filtered off, dried at room temperature and finally heated at 200 uC for 8 h. The molar yield based on zinc amounted to 87%. MOF-5. Uniformly large crystals of MOF-5 were synthesized by using an optimized procedure starting from terephthalic acid, zinc nitrate and diethylformamide as organic solvent. In a glass reactor equipped with a reflux condenser and a teflon-lined stirrer, 41 g of terephthalic acid (BDC) and 193 g of zinc nitrate tetrahydrate (Merck) were dissolved in 5650 g of diethylformamide (BASF AG; ,100 ppm water) and heated up to 130 uC for 4 h. After about 45 min, crystallization started and the formerly clear solution turned slightly opaque. After a total of 4 h, the reaction product was cooled down to room temperature. The solid was filtered off, washed three times with 1 L of dry acetone and dried under a stream of flowing nitrogen. Finally the product was activated at 60 uC for at least 3 h under a reduced pressure of ,0.2 mbar. The wet chemical analysis of the thus obtained solid yielded 33 wt% Zn, equivalent to 91% molar yield of MOF-5 calculated as Zn4O(BDC)3. The concentration from residual nitrate amounted to 0.05 wt% N. Cubic shaped crystals in between 50150 mm size are to be observed by scanning micrographs (Fig. 4). The PXRD pattern is shown in Fig. 5. Specific surface area measurements with nitrogen at 77 K were determined as 3400 m2 g21.
This journal is The Royal Society of Chemistry 2006

Fig. 2 Framework of MOF-5 displaying free access to nanosized voids and the absence of non-accessible bulk-volume. The picture shows a MOF-5 particle of about 500 nanocells with a cube edge length of about 100 nm.

rare gases (krypton and xenon) and storage of hydrogen (3.3 wt% at 2.5 MPa/77 K on Cu-BTC-MOF) will underline the prospective industrial use of metalorganic frameworks. Whenever possible, comparison is made to state-of-art applications in order to outline possibilities of processes which might be beneficially run by using MOFs.

2. Experimentals
In order to enable readers to look into the properties of MOFs, some easy-to-follow recipes are listed hereafter. They were all
628 | J. Mater. Chem., 2006, 16, 626636

View Article Online

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

Fig. 4 SEM-picture of large MOF-5 crystals from laboratory preparation (scale bar: 1 mm).

Using argon adsorption at liquid argon temperature (87 K) the adsorption properties were compared to state-of-art materials like zeolite X and activated carbon (Ceca, carbon AC40). As depicted in Fig. 6 the outstanding uptake behaviour of argon on MOF-5 is obvious and clearly exceeds zeolites and carbon components. IRMOF-8. Large crystals of IRMOF-8 were grown from a starting mixture containing 2,6-naphthalenedicarboxylic acid,

zinc nitrate tetrahydrate, diethylformamide and N-methyl-2pyrollidone as organic solvent. In a glass reactor equipped with a reflux condenser and a teflon-lined stirrer 7.5 g of naphthalenedicarboxylic acid (NDC) and 67.4 g of zinc nitrate tetrahydrate (Merck) were dissolved in 883 g of diethylformamide (BASF AG) and heated up to 130 uC for 4 h. After about 75 min crystallization started and the formerly clear solution became turbid. After the reaction the product was cooled down to room temperature and the precipitate was filtered off, washed three times with 1 L of dry chloroform and dried under a stream of flowing nitrogen. Finally the product was activated at 60 uC for 3 h under reduced pressure of ,0.2 mbar giving 9.5 g of the final product. Wet chemical analysis of the solid yielded 27.3 wt% Zn equivalent to 15% molar yield of IRMOF-8 calculated on zinc and 92% on NDC (calculated as Zn4O(NDC)3). Concentration from residual nitrate amounted to 0.78 wt% nitrogen. The Langmuir specific surface area reached 1750 m2 g21. All crystals of IRMOF-8 had a cubic shape of about 100 mm size, however, the scaly morphology indicated a high degree of intergrowth. Cu-MOF. For the first time, to our knowledge, MOFs are synthesized using an electrochemical route. Bulk copper plates, thickness 5 mm, are arranged as the anodes in an electrochemical cell with the carboxylate linker, viz. 1,3,5-benzenetricarboxylic acid, dissolved in methanol as solvent and a copper cathode. Details are to be found in.10 During a period of 150 min at a voltage of 1219 V and a

Fig. 5 PXRD of large MOF-5 crystals from laboratory preparation indicating superior crystallinity.

This journal is The Royal Society of Chemistry 2006

J. Mater. Chem., 2006, 16, 626636 | 629

View Article Online

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

Fig. 6 Adsorption isotherms of argon at 87 K on MOF-5 compared to activated carbon and zeolite X.

Fig. 8 Synthesis cell for electrochemical preparation of MOFs with Cu-plates as electrode material (laboratory scale).

Fig. 7 SEM-picture of Cu-MOF crystals from novel electrochemical preparation (scale bar: 1 mm).

currency of 1.3 A, a greenish blue precipitate was formed. After separation by filtration and drying at 120 uC overnight pure Cu-MOF was obtained. The surface area determination yielded 1649 m2 g21 and after activation at 250 uC finally 1820 m2 g21 giving a dark blue coloured solid of octahedral crystals from 0.5 to 5 mm size (Fig. 7) The experimental setup of the electrochemical cell in a laboratory glass reactor being under operation is depicted in Fig. 8. 2.2 Analysis and characterization Adsorption measurements were performed with commercially available equipment (Autosorb 6, Quantachrome Corp.) using nitrogen sorption at 77 K. Prior to measurement, samples were activated down to 1024 mbar, first at 60 uC and finally at 120 uC, until a constant vacuum was achieved for 14 h. Surface area values were calculated according to the Langmuir equation. X-Ray powder diffraction was monitored on sealed samples stored under dry nitrogen using Cu Ka radiation (D 5000, Siemens). Scanning electron micrographs were taken at a 10 kV electron beam using field emission cathode arrangement (JSM 6400F; Jeol).
630 | J. Mater. Chem., 2006, 16, 626636

Selected samples of electrochemically prepared Cu-MOF were compared to conventionally synthesized materials by EXAFS and XANES. Spectra were collected on beamline E4 at HASYLAB in the DESY synchrotron, Hamburg. Single crystal data of monoclinic tenorite mineral (Cc-symmetry) served as model basis. Although the X-ray powder diffraction pattern of both Cu-BTC-MOF conventionally synthesized and electrochemically prepared are very close to each other, the samples offer clear differences with regard to the local finestructure of the copper. The electrochemically prepared Cu-MOF gives CuK edge spectra which nicely agree with the spectra of the mineral tenorite having as model a fourfold coordination of Cu, whereas the Cu-BTC-MOF following literature recipes1214 is indicative of having a disturbed Cu-environment and an additional peak at 8.98 keV photon energy. The latter one is presumably due to the existence of occluded nitrate moieties close to the open metal copper ligand site, which could also explain the lower surface areas of only 917 m2 g21.14 Beneficially, this does not occur on the electrochemically prepared material.10 The latter is a very useful adsorbent with a strong possibility to attract electronrich molecules on the open copper-sites, as is illustrated in our gas purification example in section 2.5, below. In order to measure intracrystalline self-diffusion, pulsed field gradient NMR measurements were performed on the NMR spectrometer FEGRIS 400 NT using the 13-interval stimulated spin echo pulse sequence with two pairs of alternating pulsed magnetic field gradients at the group of Karger.15 The results were compared to literature data obtained on NaX zeolite crystals and will be discussed in detail elsewhere.15 Nevertheless, for ethane and benzene, in summary from Fig. 9, it can be seen that diffusion in MOF-5 is clearly orders
This journal is The Royal Society of Chemistry 2006

View Article Online

Table 1 Catalyst MOF-2 MOF-5

Vinyl group esterification on MOFs Surface Area/m2 g21 330 3400 Conversion (mol% acid) 94 91 Selectivity (mol% ester) 83 56

Fig. 9 Comparison of diffusivity of ethane and benzene molecules in MOF-5 over zeolite X experimentally determined from PFG-NMR measurements.15

of magnitude faster than in zeolite NaX. This is considered to be a consequence of the difference between the diameters of the two large nanoporous cavities in the MOF-5 (1.11.3 nm2) over the smaller NaX supercage (1.2 nm) with an even lower entrance window size (0.7 nm). The effective self-diffusion coefficient of benzene in the MOF-5 structure is only slightly smaller than the self-diffusion coefficient in the neat liquids at the same temperature (C6H6: 2.5 61029 m2 s21). For both catalysis and gas processing, this is an important observation and promises fast molecular transport and low mass transfer resistance as additional benefit when using MOF materials rather than zeolites. In industrial applications of MOFs this might contribute to permanent benefits in variable energy cost over state-of-art solids. 2.3 Catalysis For testing the catalytic activity in methoxypropene formation, a differentially-running reactor (volume 200 ml) filled with 55 g of MOF-2 tablet material, obtained according to the preparation, using pyrrolidone as solvent, were packed in a fixed bed basket-type arrangement. The reactor was heated to 250 uC and fed with a mixture of methanolcyclohexane at a 10 : 1 ratio and at a rate of 1.5 g h21. Propyne was added at a flux of 6 g h21. Product analysis was done on line using gas chromatography. When the system had reached steady state conditions the conversion of propyne was calculated to be 30% with a selectivity of 80% into 2-methoxypropene.11 Using zinc silicate as catalyst in a comparative example the selectivity towards 2-methoxypropene was found to be 77%, however, full conversion at temperatures of only 180200 uC was obtained.21 In a different reaction, the esterification of 4-tert-butylbenzoic acid was performed in a batch type test configuration. An autoclave was filled with 2.5 g of MOF-2 material suspended in 100 g of N-methyl-2-pyrrolidone and 40 g of 4-tertbutylbenzoic acid. After inflating to 0.5 MPa of nitrogen, the reactor vessel was heated to 180 uC and 2 MPa of acetylene were introduced. The acetylene pressure was kept constant for 24 h. The final liquid product after cooling down was analyzed by means of gas chromatography.11 In a repeated trial MOF-5 was used instead of MOF-2. Table 1 collects the results.
This journal is The Royal Society of Chemistry 2006

Interestingly, although both MOF catalyst materials reach at almost the same conversion of tert-butylbenzoic acid, the selectivity towards the vinyl ester is expressed far more by the zinc paddle-wheel containing MOF-2, with about a tenfold lower surface area.16 MOF-5 with fully saturated zinc-coordination obviously is quite poor in performing in the selective esterification. Although a clear understanding of the underlying mechanism is not yet available, there is without any doubt proof that MOFs can act as catalysts. Successful catalysis on zinc-containing MOFs in the activation of alkoxides and carbon dioxide into polypropylene carbonate has already been reported.17,18 Even enantioselective conversions with an enatiomeric excess of 8% on homochiral metalorganic POST-1 have been addressed by Kim and for heterogeneous asymmetric catalysis by Lin19,20 towards chiral secondary alcohols. Further catalytic reactions from other research groups on MOFs were recently reviewed and collected by Kitagawa.6 ZieglerNatta-type polymerization, DielsAlder-reaction, transesterification, cyanosilylation of aldehydes, hydrogenation and isomerization reactions were reported. Future research is now dedicated to find out if the metal centers, the ligands or functionalized ligands, or even metal ligand interactions or differences in particle size, may cause unusual catalytic properties and if MOFs can compete with well-known heterogeneous industrial catalysts. Beyond activity and selectivity, also the questions of time-on-stream behaviour and leaching stability will be crucial. 2.4 Gas purification Cu-BTC-MOF, prepared as described above, was used in demonstrating the removal of sulfur odorant components from natural gas. In a fixed bed reactor vessel (inner diameter of 10 mm) about 10 g of granular Cu-MOF of a particle size fraction of 12 mm were thoroughly packed. At a temperature of 25 uC a gas stream of methane odorized with 13 ppm of tetrahydrothiophene (THT) was fed over the packing and analyzed in the reactor effluent by means of gas chromatography until breakthrough occurred (Fig. 10). The used catalyst was removed from the reactor and analyzed for residual sulfur content using bulk wet chemical analysis. For comparison, the trials were repeated using two commercially available activated carbon materials as adsorbents, viz. Norit (type RB4) and CarboTech (type C38/4). Taking the breakthrough curve of tetrahydrothiophene as depicted in Fig. 10, it can be clearly seen that the sulfur odorant is completely omitted from the natural gas in the effluent and even detection by smelling is not any longer indicative. The analysis of the used Cu-MOF after the test cycle reveals an volume specific uptake of 70 g THT L21 MOF
J. Mater. Chem., 2006, 16, 626636 | 631

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

View Article Online

Fig. 10 Breakthrough-curve of continuous tetrahydrothiopheneremoval from natural gas using Cu-BTC-MOF out of electrochemical synthesis.

which is considerably higher than 0.5 g THT L21 for Noritcarbon or 6.5 g THT L21 for the CarboTech sample. Interestingly, the colour of the Cu-MOF had changed from a deep blue into a light greenish one. Obviously, Cu-MOF is a powerful material for the separation of polar components from nonpolar gases. Looking at the structure, the special arrangement of channels in Cu-BTC-MOF together with open metal ligand sites offers a dual type sorption behaviour.1214 Bulk adsorption, on one hand, exemplified e.g. during nitrogen loading at 77 K, aiming at pore filling to deduce pore volumes, and secondly, the predominant existence of open ligands at the copper paddle wheel-cluster (cf. Fig. 1, lower left structure unit), give rise to possible chemisorptive sites, the latter being nicely kept free and unaffected during BASFs electrochemical synthesis starting from bulk metal. The anion-free preparation omits the unwanted occlusion of nitrates into the metal organic framework that is usually applied in state-of-the-art preparations.13,14,22,23 Further electron-rich molecules that are successfully removed by Cu-MOF in a similar manner are amines and ammonia, water traces, alcohols and oxygenates. In addition, in all cases a dominant color change readily allows visible detection of breakthrough and contaminant saturation on the MOF. During removal of the contaminant by vacuum treatment or heating, the original color reappears, indicating a possible regeneration of the adsorbent. Depending on time and temperature applied, chemical analysis of the sulfur content as well as the final weight of the used Cu-MOF versus the fresh sample helps to monitor the degree of regeneration. 2.5 Gas storage Gas storage both at room temperature and 77 K up to 10 MPa was measured on a homebuilt gravimetric prototype equipment comprising two highly sensitive balances A and B (Mettler, PG 5002-S, 0.01 g displayed). Balance A carried a stainless steel (material type 1.4541) fabricated container A of about 460 ml content which was filled up to the neck with about 150 g of in-situ activated MOF sample sitting in
632 | J. Mater. Chem., 2006, 16, 626636

a thermostatted Dewar vessel. This experimental setup guaranteed a sufficiently high resolution of storage capacity at a sensitivity of about 104. The reproducibility usually was better than 2%. Balance B monitored the weight of a high pressure hydrogen feed flask B (e.g. 20 MPa). By controlled stepwise opening of valves in the connecting piping steel system between containers A and B, there was a redundant checking of weight and pressure increase in the MOF-containing vessel, whereas weight and pressure drop was indicated in the hydrogen feeding container B. Of course uptake on one side and release data on the other side had to fit to each other in terms of proving a leak-free mass balance. Blank runs without MOFsample were repeated prior to each measurement. In order to obtain weight-specific uptake the data was corrected for bulk volume using the helium density of the MOF-sample. As for the MOF-5 sample adsorption isotherms with either argon, krypton or xenon at room temperature were registered and compared to their conventional compression curve up to about 5 MPa. Obviously, as shown in Fig. 11, for all rare gases under investigation (Ar, Kr, Xe) the volume specific uptake is higher in case of a gas cylinder being filled with MOF-5. This effect is becoming even more expressed from argon over krypton to xenon. In the latter case, at about 10 bar pressure, the amount of xenon stored in a MOF-filled container is about fourfold higher (if one compares the curves vertically). Of course such an enhanced uptake of one gas over another can be exploited to separate gas mixtures of e.g. krypton and xenon. Details on the performance of such a process are given in section 2.6. Looking at a storage level in the xenon curve of about 100 g Xe L21 in Fig. 11, it becomes clear from a horizontal comparison of the isotherms at a constant uptake value that with a MOF-filled container the pressure which is needed for holding a gas in a given volume is reduced from about 17 to 2 bar. Thus, storage of a gas in MOF-filled canisters can be used either way to enhance capacity in a given volume or to transport an equivalent amount of gas at a far lower pressure.

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

Fig. 11 Compression curves of rare gases Ar, Kr, Xe and comparison over inflation curves into MOF-5-filled gas containers (room temperature, up to 60 bar; lecture bottles).

This journal is The Royal Society of Chemistry 2006

View Article Online

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

Fig. 12 Compression of propane into gas container with and without MOF-filling (MOF-5 tablets in lecture bottles, room temperature).

Similarly, other gases like methane and hydrocarbons can be held at a denser state in the same manner.22,24 The same finding holds for propane as well. In Fig. 12 are depicted the curves for shaped MOF-5 pellets in gas containers over non-filled containers. Again the inflation curves are markedly different. The non-MOF-filled container represents the almost linear uptake behaviour, whereas the MOF-filled can is far higher and with a steeper increase at the beginning. Taking the pressure at about 10 bar it becomes obvious that one MOF-filled container substitutes the amount of three state-of-art pressurized cylinders. Already today much interest is attributed to the storage of hydrogen for mobile and portable fuel-cell application by using hydrogen as a carrier for electric power supply.25 Due to the need for alternative fuel sources and energy carriers, the target values of US Department of Energy were recently reviewed,8 compiling storage data of some 0.23.8 wt% of hydrogen on MOFs, the maximum on a weight-specific calculation being found by Ferey on MIL-537 derived from aluminium salts and BDC. Pillaring with secondary amine (triethylenediamine) linkers as strategy was employed by Kim27 and Seki22 and MOF-505 again by Yaghi-group28 with Cu paddle-wheels connected by 3,39,5,59-biphenyltetracarboxylic acid ranged up to more than 2 wt%. Doubly interpenetrated nets of zinc frameworks built by NTB-linkers (4,49,40-nitrilotrisbenzoic acid) by Suh29 were reported to reach 1.9 wt%. of hydrogen uptake at 77 K. However, it is not yet possible to foresee if large surface area materials like MOF-177,29 MIL-10032 or MOF-5 and isoreticular members,24 or materials with an average surface of between 1000 to 1500 m2 g2126,29 or even small pore MOFs like30,33 will be the most promising storage media. Neither can it be concluded if divalent or trivalent metal-clusters34 are the most favourable ones. Simulations obviously support metalorganic frameworks as being favorable over zeolites.31 From results of our prototype equipment (77 K temperature, up to 40 bar) it can be seen, how different MOF-materials contribute differently to volume-specific hydrogen storage (cf. Fig. 13). Comparing to the pressurizing of an empty container with hydrogen MOF-5, IRMOF-8 and Cu-BTC-MOF increasingly take up higher amounts of hydrogen, all of them exceeding the standard pressurevolumetemperature (PVT) uptake curve of
This journal is The Royal Society of Chemistry 2006

Fig. 13 Volume-specific storage curves of hydrogen on different MOF-materials in comparison to compression curve of hydrogen into empty gas containers (77 K, 40 bar; measurement from BASF prototype setup as described in section 2.5).

the empty container and with steepest incline below 10 bar. At 40 bar the PVT-relation of hydrogen in an empty canister is registered as 12.8 g H2 L21, whereas Cu-BTC-MOF filled into containers reach a plus 44% capacity of up to 18.5 g H2 L21. For comparison, the volume-specific density of liquid hydrogen at its boiling point (20 K) is 70 g H2 L21. Above 10 bar, the curves run mostly parallel to the conventional H2-pressurevolume relation. On a per weightcalculation it becomes clear that the saturation of MOFs with hydrogen is already achieved at pressures of less than 15 bar (Fig. 14). For electrochemically-prepared Cu-BTC-MOF this attributes to about 3.3 wt% H2-uptake. This data from our large scale prototype compares reasonably well with the literature.8 It should be mentioned here, that for many volume-limited fuel-cell applications, i.e. the mobile and portable cases, it will be industrially much more relevant to compare storage data on a volume-specific rather than on a weight-specific storage capacity basis. Typically the packing densities of MOF powders are around 0.2 to 0.4 g cm23 increasing to 0.50.8 g cm23 when shaped into tablets or extrudates. So far, this material density is so low that weight limitation in an application is non-existent, in contrast of course to the alternative use of metal hydride as storage media. Much more importantly, however, there is only a strictly confined volume available in both mobile automotive as well as in

Fig. 14 Volume-specific versus weight-specific storage curve of hydrogen on Cu-BTC-MOF from electrochemical preparation.

J. Mater. Chem., 2006, 16, 626636 | 633

View Article Online

portable electronic device applications. Hence, storage data communicated as weight-specific rather than volume-specific numbers can be misleading by merely focusing on only one of them. Depending on the intended application and the given boundary conditions, it might be necessary in future to indicate both types of values. Anyhow, the most important issue is the amount of hydrogen which, in a reasonable time-scale, can be discharged from storage media. Here, MOFs really do have a fully reversible uptake and release behaviour. As the storage mechanism is predominately physisorption, there are no huge activation energy barriers to be overcome as compared to e.g. metal hydrides when trying to liberate the stored hydrogen. Simple pressure reduction by controlled valve opening is sufficient to draw off hydrogen from MOFs within a few seconds. Energy density values of 1.1 kW h L21, as requested in the European Hydrogen and Fuel Cell Strategic Research Agenda and Deployment Strategy,25 are equivalent to a volumetric hydrogen storage capacity of about 33 g H2 L21. As demonstrated, almost two thirds of this value can be reached by storing hydrogen in MOFs at 77 K and a moderate pressure of 4050 bar. Increasing this volumetric capacity by shaping MOF-powders into tablets and extrudates is proved to be feasible. Presently, a high amount of storage at room temperature has not been attained. Further research concepts are dedicated to increase the storage capacity and to shift it to higher temperatures. Once it is clarified where the hydrogen molecules are favourably adsorbed and attached in the different MOFs, it will be easy to browse through the huge variety of several hundred (and constantly still-increasing) numbers of structures and to identify the most promising examples. In this respect, molecular modelling tools might become as important as elaborate experimental synthesis efforts.31 It should be kept in mind that depending on the temperature required for a possible application, highly porous MOFs might be favourable for low temperatures, whereas rather small pore materials,30,33 or highly attractive and flexible ones,26,35 could be favourites for room temperature storage. New mechanisms bridging chemisorption (as in hydrides) and physisorption (as in metal organic frameworks) might be also requested to meet future challenges. 2.6 Gas separation Already during the monitoring of the diverse uptake behaviour of rare gases on MOF-5 (Fig. 11) it became obvious that this property could possibly be used to separate mixtures of rare gases by adsorption on MOFs. Such mixtures are usually to be found technically in cryogenic air separation units and once the gases are separated, xenon and krypton can be marketed separately, e.g. xenon as narcotic medical gas and krypton as filler for lamp industry. Unlike cryogenic distillation, the following experiments indicate a far simpler process of pressureswing adsorption to separate rare gases mixtures. A mixture of Kr (ca. 94% mol) and Xe (ca. 6% mol) was fed continuously to an isothermal tubular reactor (2 m length,
634 | J. Mater. Chem., 2006, 16, 626636

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

Fig. 15 Gas separation of KrXe mixture by continuous adsorption on electrochemically produced Cu-BTC-MOF.

1.6 cm diameter) filled with 193 g Cu-MOF at a given temperature of 55 uC and 40 bar (Fig. 15). The flow rate (60 L h21) and reactor dimensions were chosen appropriately in order to prevent complete saturation of the MOF material during the time-span of the measurement. The pressure was adjusted manually by a needle valve at the reactor outlet. The gas composition leaving the adsorber was detected by an on-line mass spectrometer (Pfeiffer Vacuum OmniStar2 QMS-200). In the gas stream leaving the adsorber, Xe was reduced to a level of ca. 50 ppm. After more than 100 min on-stream the MOF became saturated with Xe and a rapid breakthrough was observed. The calculated capacity of the Cu-MOF for Xe was more than 60 wt%. This is almost twice as much Xe as a high surface active carbon (Ceca, AC 40, ca. 2000 m2 g21) could take up under identical conditions. Due to the fact that MOFs exhibit a gas molecule mobility of about two to three orders higher than state-of-art molecular sieves or active carbons, a faster swing operation period between adsorption and desorption cycle seems possible. In terms of economic consideration, this contributes to fairly reduced purge time, energy consumption and overall variable costs thus rendering MOFs beneficial over established technology. Again novel material properties of metalorganic frameworks, viz. their porosity with nanometer size pores in regular arrays and the absence of blocked bulk volume contribute to principal differences in performance. Looking into the literature where some zeolite-like (MTN topology) metalorganic frameworks have already been observed, it will be interesting to watch if, with the small pore materials, separations based on molecular sieving might become possible.30,33,36 Given this possibility, MOFs would grow to be competitors to zeolitic molecular sieves, which currently have a market volume of some hundred thousand tons capacity per year.

3. Conclusion and prospects


In this article, we have shown that metalorganic frameworks or coordination polymers are not merely a new class of porous materials for combining inorganic and organic chemistry classifications. From an industrial point of view, they offer,
This journal is The Royal Society of Chemistry 2006

View Article Online

in principle, many interesting and promising features over prior art, viz. N world records in surface area N ultimate porosity with absence of blocked volume in solid matter N combined flexible and robust frameworks N full exposure of metal sites N high mobility of guest species in regular framework nanopores N fast growing number of novel inorganicorganic chemical compositions. Obviously, many applications might (and surely will) be tested once verified synthesis recipes of MOFs are available. The recipes given in the Experimental section will allow the reader to prepare these new compounds in laboratory-scale amounts. However, industrial synthesis at BASF is understood to be far more advanced, already into barrel-size pilot scale, and additional issues need to be taken into account during the manufacturing procedures, which of course are beyond the scope of this paper. Unlike many other novel materials, e.g. carbon polymorphs, fullerenes, bucky-balls, CNT, the metalorganic framework materials preparation and fabrication does not necessarily need additional capital investment into a totally new synthesis technology. Simply adaptation of conventionally available precipitation and crystallisation manufacturing methods needs to be done. Shaping of metalorganic framework powders into industrially widespread geometries of mm-sized tablets, extrudates, honeycombs, etc. can be performed on MOFs as well without any major obstacle. The examples which we gave for catalysis as well as for gas processing and storage already indicate that there is much room left for many future research efforts (viz. storage of alternative energy carriers like small hydrocarbons, odour removal in both stationarye.g. householdas well as mobile (bus, train, subway, ship) environments or the adverse odorization in carrying perfumes, etc. Pick-up of liquids without swelling of solids could be of interest as well, e.g. for food packaging or removal of hazard liquids like organic solvents, oils, brake fluids and the like). It is worthwhile to mention that, unlike state-of-art heterogeneous catalysts, the metal sites in MOFs are usually fully exposed, therefore giving an ultimately high degree of metal-dispersion. From supplementary work we already know that these metal sites usually behave differently from bulk metals. Compared to zeolites the amount of metals in MOFs are by almost a factor of ten higher and many of the metal species belong for chemists to the interesting class of transition metals. In summary and perspective, all this might lead to a fast growing, prosperous and widespread innovation in materials science, both in academia and industry. However, one certainly has to keep attention to find superior performance by applying MOFs over state-of-art technologies. It will never be sufficient just to find a me-too solution instead of looking for considerable improvement of the best existing one. Only the latter approach will finally contribute to true innovation and value-added growth of industrial companies and society.
This journal is The Royal Society of Chemistry 2006

Acknowledgements
Technical assistance of Dr O. Metelkina-Schubert, Dr Cox, S. Lutter, W. Kippenberger, U. Diehlmann, R. Hess, R. Ruetz, I. Schwabauer, R. Senk, and H. Sichler is gratefully acknowledged. U.M. thanks O.M. Yaghi and S. Kitagawa for many stimulating discussions on metalorganic frameworks (MOFs) and coordination polymers (CPLs).

References
1 E. A. Tomic, J. Appl. Polym. Sci., 1965, 9, 37453752. 2 H. Li, M. Eddaoudi, M. OKeeffe and O. M. Yaghi, Nature, 1999, 402, 276279. 3 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. OKeefe and O. M. Yaghi, Science, 2002, 295, 469472. 4 O. M. Yaghi, M. Eddaoudi, H. Li, J. Kim and N. Rosi, WO 2002/ 088148, 2002, University of Michigan. 5 H. K. Chae, D. Y. Siberio-Perez, J. Kim, Y. B. Go, M. Eddaoudi, A. J. Matzger, M. OKeeffe and O. M. Yaghi, Nature, 2004, 427, 523527. 6 S. Kitagawa, R. Kitaura and S. Noro, Angew. Chem., Int. Ed., 2004, 43, 23342375. 7 J. L. C. Rowsell and O.M. Yaghi, Microporous Mesoporous Mater., 2004, 73, 314. 8 J. L. C. Rowsell and O. M. Yaghi, Angew. Chem., 2005, 117, 47484758. 9 T. J. Barton, L. M. Bull, W. G. Klemperer, D. A. Loy, B. McEnaney, M. Misono, P. A. Monson, G. Pez, G. W. Scherer, J. C. Vartuli and O. M. Yaghi, Chem. Mater., 1999, 11, 26332656. 10 U. Mueller, H. Puetter, M. Hesse and H. Wessel, WO 2005/049892, 2005, BASF Aktiengesellschaft. 11 U. Mueller, M. Hesse, L. Lobree, M. Hoelzle, J. D. Arndt and P. Rudolf, WO 2002/070526, 2002, BASF Aktiengesellschaft. 12 K. Schlichte, T. Kratzke and S. Kaskel, Microporous Mesoporous Mater., 2004, 73, 8188. 13 Q. M. Wang, D. Shen, M. Buelow, M. L. Lau, F. R. Fitch and S. Deng, US Pat. 6 491 740, 2002, The BOC Group, Inc. 14 S. S.-Y. Chui, S. M.-F. Lo, J. P. H. Charmant, A. G. Orpen and I. D. Williams, Science, 1999, 283, 11481150. 15 F. Stallmach, S. Groeger, V. Kuenzel, J. Kaerger, O. M. Yaghi, M. Hesse and U. Mueller, Angew. Chem., Int. Ed. (submitted). 16 M. Eddaoudi, H. Li and O.M. Yaghi, J. Am. Chem. Soc., 2000, 122, 13911397. 17 U. Mueller, G. Luinstra and O. M. Yaghi, US Pat. 6 617 467, 2004, BASF Aktiengesellschaft. 18 R. Eberhardt, M. Allmendiger, M. Zintl, C. Troll, G. A. Luinstra and B. Rieger, Macromol. Chem. Phys., 2004, 205, 4247. 19 Chuan-De Wu, A. Hu, L. Zhang and W. Lin, J. Am. Chem. Soc., 2005, 127, 8940. 20 J. S. Seo, D. Whang, H. Lee, S. I. Jun, J. Oh, Y. J. Jeon and K. Kim, Nature, 2000, 404, 982986. 21 Ch. Miller, P. Rudolf and H. J. Teles, WO 2004/009523, 2004, BASF Aktiengesellschaft. 22 K. Seki and W. Mori, J. Phys. Chem. B, 2002, 106, 13801385. 23 O. M. Yaghi, US Pat. 5 648 508, 1997, Nalco Chemical Company. 24 U. Mueller, K. Harth, M. Hoelzle, M. Hesse, L. Lobree, W. Harder and O. M. Yaghi, WO 2003/064030, 07.08.2003, BASF Aktiengesellschaft. 25 S. Barrett, Fuel Cells Bull., 2005, 1219. 26 G. Ferey, M. Latroche, C. Serre, F. Millange, T. Loiseau and A. Percheron-Guegan, Chem. Commun., 2003, 29762977. 27 D. N. Dybtsev, H. Chun and K. Kim, Angew. Chem., 2004, 116, 51435146. 28 B. Chen, N. W. Ockwig, A. R. Millward, D. S. Contreras and O. M. Yaghi, Angew. Chem., 2005, 117, 48234827. 29 E. Y. Lee, S. Y. Jang and M. P. Suh, J. Am. Chem. Soc., 2005, 127, 63746381. 30 D. N. Dybtsev, H. Chun, S. H. Yoon, D. Kim and K. Kim, J. Am. Chem. Soc., 2004, 126, 3233. 31 Q. Yang and Ch. Zhong, J. Phys. Chem. B, 2005, 109, 24, 118624. 32 G. Ferey, C. Serre, C. Mellot-Draznieks, F. Millange, S. Surble, J. Dutour and I. Margiolaki, Angew. Chem., 2004, 116, 64566461.

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

J. Mater. Chem., 2006, 16, 626636 | 635

View Article Online


33 R. Matsuda, R. Kitaura, S. Kitagawa, Y. Kubota, R. V. Belosludov, T. C. Kobayashi, H. Sakamoto, T. Chiba, M. Takata, Y. Kawazoe and Y. Mita, Nature, 2005, 436, 238241. 34 C. Serre, F. Millange, S. Surble and G. Ferey, Angew. Chem., 2004, 116, 64466449. 35 E. Choi, K. Park, C. Yang, H. Kim, J. Son, S. W. Lee, Y. H. Lee, D. Min and Y. Kwon, Chem.Eur. J., 2004, 10, 55355540. 36 Q. Fang, G. Zhu, M. Xue, J. Sun, Y. Wei, S. Qiu and R. Xu, Angew. Chem., 2005, 117, 39133916.

Downloaded by Imperial College London Library on 24 January 2013 Published on 23 November 2005 on http://pubs.rsc.org | doi:10.1039/B511962F

636 | J. Mater. Chem., 2006, 16, 626636

This journal is The Royal Society of Chemistry 2006

Вам также может понравиться