You are on page 1of 10

Physical Chemistry Chemical Physics

www.rsc.org/pccp Volume 13 | Number 14 | 14 April 2011 | Pages 63736712

ISSN 1463-9076

COVER ARTICLE Ranocchiari and van Bokhoven Catalysis by metalorganic frameworks: fundamentals and opportunities

HOT ARTICLE Lesarri et al. Structural evidence of anomeric effects in the anesthetic isoflurane

1463-9076(2011)13:14;1-Z

PCCP
Cite this: Phys. Chem. Chem. Phys., 2011, 13, 63886396
Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

Dynamic Article Links

www.rsc.org/pccp

PERSPECTIVE

Catalysis by metalorganic frameworks: fundamentals and opportunities


Marco Ranocchiari*a and Jeroen Anton van Bokhoven*ab
Received 4th November 2010, Accepted 13th December 2010 DOI: 10.1039/c0cp02394a Crystalline porous materials are extremely important for developing catalytic systems with high scientic and industrial impact. Metalorganic frameworks (MOFs) show unique potential that still has to be fully exploited. This perspective summarizes the properties of MOFs with the aim to understand what are possible approaches to catalysis with these materials. We categorize three classes of MOF catalysts: (1) those with active site on the framework, (2) those with encapsulated active species, and (3) those with active sites attached through post-synthetic modication. We identify the tunable porosity, the ability to ne tune the structure of the active site and its environment, the presence of multiple active sites, and the opportunity to synthesize structures in which keylock bonding of substrates occurs as the characteristics that distinguish MOFs from other materials. We experience a unique opportunity to imagine and design heterogeneous catalysts, which might catalyze reactions previously thought impossible.

1. Introduction
Porous crystalline solids are continuously attracting the interest of the scientic community because of their application in
a

Department of Synchrotron Radiation and NanotechnologyLaboratory for Energy and Environment-Paul Scherrer Institute, CH-5232 Villigen PSI, Switzerland. E-mail: marco.ranocchiari@psi.ch b Department of Chemistry and Applied Biosciences-Institute of Chemistry and Bioengineering-ETH Zurich, CH-8093 Zurich, Switzerland. E-mail: jeroen.vanbokhoven@chem.ethz.ch

petrochemical, pharmaceutical, catalytic, and environmental technology. The development of zeolites was a revolution towards the rationalization of surface and solid-state chemistry, and crystalline and porous aluminosilicates are found in numerous industrial applications nowadays.1 More recently, another class of materials drew the attention of scientists from university and industry: metalorganic frameworks (MOFs), also called porous coordination polymers (PCPs). They are constituted by multi-functionalized organic molecules that are bound together by inorganic units to form porous solids with a regular and sometimes predictable geometry.2

Marco Ranocchiari studied at Universita` degli Studi di Milano (Italy), where he obtained his master degree in 2005 with a thesis about cobalt porphyrin metalorganic frameworks and their application in catalysis. In 2009 he completed his doctoral studies on asymmetric aziridination and synthesis of cyclic PNNP complexes under the supervision of Prof. A. Mezzetti at ETH Zurich (Switzerland). Since March 2010 he is working Marco Ranocchiari on catalysis by metalorganic frameworks in the van Bokhoven group at the Paul Scherrer Institute and ETH Zurich (Switzerland).

Jeroen Van Bokhoven studied chemistry at Utrecht University, where he obtained his PhD in inorganic chemistry and catalysis in 2000. After a short junior research position, he was senior researcher at ETH Zurich (20022006). In 2006 he was granted a Swiss National Science Foundation professorship. In 2010 he obtained the chair in heterogeneous catalysis at the Institute for Chemical and Bioengineering at ETH Zurich. Jeroen Anton van Bokhoven He is also head of the Laboratory for Catalysis and Energy and Environment at the Swiss Light Source in the Paul Scherrer Institute, Switzerland. His work focuses on understanding structure/performance relations of catalytic systems based on supported metals and porous materials through the development and application of highly sophisticated X-ray methods of characterization.
This journal is
c

6388

Phys. Chem. Chem. Phys., 2011, 13, 63886396

the Owner Societies 2011

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

The synthesis, characterization, and application of MOFs are strongly growing elds due to their unique properties, and, as a result, an exponential increase in the number of publications has been observed over the last decades. The chemical and structural versatility of such materials makes them potentially great candidates for nding new and unique applications in catalysis. However, although there are nowadays thousands of PCP structures in the Cambridge structural database (CSD), the number of published catalytic applications is still limited, and catalysis by MOF is still in its infancy. Therefore, new strategies to develop unique catalytic systems with such materials that exploit their remarkable structural and chemical exibility must be developed. With this perspective, we aim to give a critical view of the possible catalytic approaches with MOFs. This description starts by explaining their distinctive structural, physical, and chemical properties and aims at trying to nd new and unique catalytic applications by designing specic active sites. The characteristics of MOFs are also compared to those of other materials that have already shown a broad catalytic scope, such as zeolites and mesoporous aluminosilicates, to understand their similarities, dierences, limitations, and challenges. We do not intend to give a comprehensive review of the employment of metalorganic frameworks, which can be found elsewhere,3 but to analyze few but illustrative examples that clarify how properties of porous coordination polymers directly correlate to their potential in nding new heterogeneous transformations. We address it to anyone who is interested in using MOFs in catalysis.

2. MOFs: structure and properties


2.1 Structural characteristics of MOFs MOFs are a class of compounds constituted by multifunctionalized organic molecules that are bound to metal or metal clusters through coordinating moieties, such as amines, pyridines, carboxylates, sulfates, and phosphates, to form crystalline materials that give one-, two-, or three-dimensional structures, which have a cavity sizew that varies from microporous (smaller than 2 nm) to mesoporous (between 2 and 50 nm).z The use of dierent organic building blocks to combine with various inorganic ones generates almost innite possible structures, and more frameworks will be discovered in the years to come. The chemistry of simple coordination polymers is well known since a relatively long timepublications already appeared in the fties4but it took almost fty years to demonstrate their enormous applied potential in areas such as catalysis,3 photocatalysis,5 gas separation and storage, and material chemistry.6 After the pioneer works of the groups of rey,9 the interest in Johnson and Jacobson,7 Bujoli,8 and Fe porous coordination polymers was boosted in the nineties with the publication by Yaghi et al. of the synthesis and characterization of [Zn4O(tpa)3]n (MOF-5, tpa = terephthalic acid),
w The cavity in MOFs is dened as the size of the diameter of the largest sphere that ts the cavity. z At the moment of this publication, the largest cavity found in a MOF is around 4 nm.

which is the most known and studied MOF nowadays.10 This material is a 3D cubic-shaped framework in which the inorganic fragments are tetrahedral Zn4O clusters, which bind six carboxylate moieties of dierent terephthalate organic building blocks in an octahedral fashion (Fig. 1A and 3B). MOF-5 is the framework that more than other ones helped the development of new concepts to facilitate the design and tunability of the chemical and of the pore size of PCPs. Three years after its discovery, Yaghi and co-workers published the synthesis of a set of metalorganic frameworks that were based on the same geometry of MOF-5, but that diered by the presence of functional groups, such as halogens, alkyloxo, hydroxyl, and amino ones (Fig. 1B), or by the size of the organic building-block (Fig. 1C).11 Such isoreticular MOFs (IRMOFs) not only demonstrated the principle of geometrical design and of tunability of the cavity size by using a known material as prototype, but opened also new perspectives in the production of coordination polymers with functional groups, which hold great promise in MOFs catalytic applications (vide infra). The principle of isoreticularity was also proven for other materials like [Zr6O4(OH)4(tpa)6]n (UiO-66)12 and [Fe3O(tpa)6]n (MIL-88),13 and it is still a concept that needs to be further investigated for other kind of PCPs because it is an essential tool to develop new functionalized materials with changeable cavity size. Another unique structural feature of MOFs is the possibility to build frameworks with two or more isoreticular organic linkers, each bearing a dierent functionality that is randomly and homogeneously distributed within the framework. This concept of multivariable or mixed MOFs (MTV-MOFs or MIXMOFs) (Fig. 1D) was pioneered by the groups of Richardson14 and Baiker15 and then recently extended by Yaghi and co-workers,16 who presented the synthesis of several MTV-MOF-5 materials constituted by up to seven dierent functionalized terephthalate derivatives. It is also possible to mix organic building-blocks that would give MOFs with dierent overall geometry if assembled with the same inorganic unit, as shown by Matzger and co-workers for [Zn4O(tpa)(btb)4/3]n (UMCM-1, btb = 1,3,5-tris(4-carboxyphenyl)benzene).17 However, the risk of having heterogeneous or separated solid phases built from dierent organic linkers is relatively high. The homogeneity of the crystalline material must be properly investigated, for example by NMR spectroscopy after digestion of a single crystal, by electron microscopy, and by measuring the elemental analysis of dierent parts of one crystal to overcome any doubt. By modeling the organic starting materials and/or the inorganic units, the number of possible combinations further increases. For example, by exchanging the terephthalic acid building block with a 1,3,5-tris(4-carboxyphenyl)benzene one, framework MOF-177 is produced (Fig. 2A).18 Moreover, if one varies the Zn4O clusters with hexanuclear Zr6O4(OH)4 ones, the structure UiO-66, which has a dierent connectivity through the linker groups, is obtained (Fig. 2B).12 It is even possible to build frameworks that form cages with similar structural features as those in zeolites, such as the zeolitic imidazolate frameworks (ZIFs).19 If rigid organic building blocks are combined with inorganic units with denite coordination geometry, it is also feasible to
Phys. Chem. Chem. Phys., 2011, 13, 63886396 6389

This journal is

the Owner Societies 2011

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

Fig. 1 Schematic representation of IRMOF-3 (B), IRMOF-10 (C), and an MTV-MOF (D), or MIXMOF, based on MOF-5 (A).

and the metalorganic polyhedra22 models, it is still a challenge to start from a drawn model to successfully synthesize the corresponding material. Nevertheless, the possibility of design and the large structural diversity show a unique structural exibility that is dicult to achieve with other solid materials. Interpenetration, which is the catenation in parallel of two polymer chains, is an important negative phenomenon that is often observed especially when big pores are present. The negative consequence of interpenetration is of course a sensible reduction of the surface area and thus of the average pore size. Lillerud and co-workers have shown how two batches of MOF-5, which were synthesized by two dierent procedures, gave dierent adsorption data. A careful singlecrystal XRD analysis revealed that this was caused by the catenation of two cubic chains of the materials.23 Attention must be addressed on the reproducibility of the nitrogen adsorption measurements and on the careful investigations of the diraction patterns.
Fig. 2 Structures of MOF-177 (A) and UiO-66 (B).

2.2 MOFs and porous silica-alumina predict and design the overall framework structure. However, even though the design of MOFs has been rationalized through the node-and-spacer,20 the secondary building unit (SBU),21
6390 Phys. Chem. Chem. Phys., 2011, 13, 63886396

Zeolites are crystalline aluminosilicates with an ordered framework structure built by corner sharing SiO4 and AlO4 units in dierent fractions, but with an Si/Al ratio greater than one.
This journal is
c

the Owner Societies 2011

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

Fig. 3 Space-lla comparison between zeolite-A (A),b MOF-5 (B),b and an MCM-41 (C) with small pores.a,c (aCalculated with van der Waals radii. bStructure obtained from its single crystal X-ray data. cStructure obtained by simulating four silicon atoms per side of the hexagon.)

Table 1

Comparison between the structural, physical, and chemical properties of zeolites, mesoporous silica and alumina, and MOFs Zeolites MOFs Yes Yes Up to 10 400 m2 g1 a Up to 4 nma Low to high Low to mediumb Variable High
b

Mesoporous silicates and aluminosilicates No No o2000 m2 g1 Z 2 nm High Medium High Medium-low

Crystalline? Homogeneous active sites? Surface area Cavity size Diusivity Thermal stability Chemical stability Chemical versatility
a

Yes Yes o600 m2 g1 rca. 1 nm Low High High Low

Maximum value published at the moment of this publication.

Maximum value published at the moment of this publication: 540 1C.

Such units are interconnected in many ways to give rise to numerous structures with cavities and channels of dierent size and connectivity.24 Fig. 3A shows the space-ll model of a zeolite-A with a total of eight Si and Al atoms circumventing one pore. Their synthesis usually requires the presence of a templating agent, which modulates the structure of the framework and thus the pore size and connectivity. Zeolites can be synthesized with dierent atoms in the framework, such as Si, Al, Ge, Fe, and Ti.1 Due to their high thermal and chemical stability, zeolites are a cheap source of materials that can be used in catalysis under harsh conditions. Their pore opening is usually below 1 nm,y which makes them perfect solids for gas phase reactions, but limits their application in reactions where the substrates are relatively large such as in liquid phase transformations. Mesoporous silica alumina materials with dened and regular pores such as MCMs25 (Fig. 3C) and SBAs26 were developed, which showed great promise in applications that are aected by diusion limitation. However, these materials lack the crystallinity and generally do not possess the high catalytic activity and stability of zeolites.27 Catalysis and adsorption are two of the most successful applications of porous materials. Table 1 compares
y The pore size in zeolites is dened as the size of the largest molecule that ts through the pore.

characteristics of MOFs and porous silica alumina structures. Zeolites and MOFs have an inherent geometric regularity that allows us to obtain single crystals suitable for X-ray diraction to determine the exact chemical composition and position of the atoms in space. On the other hand, mesoporous silicates and aluminosilicates are usually not crystalline, and it is not easy to determine their ne structure. As a consequence, it is much easier to have a homogeneous distribution of the active sites in a zeolite and in a MOF than in other mesoporous materials. Moreover, while the synthesis of SBA and MCM type of materials is usually complex, zeolites and PCPs with high crystalline degree are smoothly obtained by solvothermal treatment of the reagents in high boiling solvents. The selfassembly synthesis of metalorganic frameworks is often straightforward and without any templating agent, whereas zeolites usually require the use of it. The crystallinity also permits us to get reproducible surface areas, and the largest are found in MOFs (values above 10 000 m2 g1 are reported).28 Consequently, the pore size can be big in coordination polymers, and there are some, such as .29 MIL-101, with spectacular mesoporous cavities of over 30 A On the contrary, zeolites have a pore size that rarely exceeds 1 nm without signicant loss in stability, whereas mesoporous silicates and aluminosilicates feature channels that overcome . Since PCPs can show a regular mesoporous structure 100 A while maintaining their crystallinity, from a porosity point of
Phys. Chem. Chem. Phys., 2011, 13, 63886396 6391

This journal is

the Owner Societies 2011

view, they can be categorized as overlapping the categories of zeolites and ordered mesoporous solids such as MCMs, SBAs, and other oxides (Fig. 3). Consequently, the diusivity is higher in MOFs than in zeolites. Thermal stability is one of the most important parameters, because it establishes what kind of application the material is suitable for. It is usually measured by thermogravimetric analysis and dierential thermal analysis. For zeolites and mesoporous aluminosilicates, it is well established that there are several factors that inuence it.30 An extensive study on the thermal stability of MOFs has not been performed yet. This is not surprising if one considers that the types of bonds involved and the number of possible frameworks and their geometries make dicult a rationalization of the topic. The bond strengths as well as the nature and topology of the framework are parameters that aect the thermal stability of PCPs. While zeolites show stability that goes from 150 1C to over 800 1C, most MOFs decompose below 300 1C. To increase the thermal stability, it is necessary to produce neutral frameworks with stable and possibly chelating coordinating moieties such as carboxylate units. Materials of this type are the MOF-5 class (Fig. 1) and UiO-66 (Fig. 2B), which are stable above 300 and 500 1C, respectively. At the moment, it is dicult to predict whether more stable coordination polymers will be produced in the near future, but highly stable MOFs are still a challenge that will be aimed at their employment in catalysis under harsh conditions. When dealing with frameworks that contain transition metal complexes or clusters, one has also to take into account possible decomposition with air, moisture, and solvents. MOF-5 derivatives and frameworks, which bear Zn4O cluster nodes, decompose under humid air.31 The decomposition of such materials after removal of the guest molecules into the pores is complete after 24 h and partial decay cannot be excluded also in the solvated framework. Precautions must thus be taken in the storage of and working with sensitive materials, which should be kept in closed vessels and in dry solvents. The biggest advantage of MOFs is the chemical versatility, which is due to the nature of their chemical composition. Besides the possibility of changing linker and inorganic unit, it is also possible to include functional groups and modulate the cavity size by exploiting the concept of isoreticularity. When designing materials with functionalities that bear coordinating groups, one experiences competition between the coordination of the pending group and that of the moieties used to build the framework during the synthesis of the MOF. Therefore, synthesis requires either atom-selective binding to the inorganic unit, such as in IRMOF-3, or the use of protective groups.32 In any case, once synthesized, such functional groups can further react by post-synthetic modication (PSM).33 This changed the way scientists approached the grafting of solid materials. Other crystalline substances, such as microporous zeolites, lack accessible chemical functionality and the size of open space to have a broad scope of post-functionalization, and amorphous mesoporous structures lack the perfect homogeneous surface, although grafting can be performed at hydroxyl groups on the surface of the pores (a) by noncovalent bonds, (b) by covalent interactions through reaction
6392 Phys. Chem. Chem. Phys., 2011, 13, 63886396

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

of the materials with chlorosilanes, organosilanes, or silazanes, and (c) by ion exchange of a cationic transition metal complex with the solids themselves.27 Post-synthetic modication with organic molecules has been performed by several groups on MOF-5,35 UMCM-1,36 and UiO-6634 derivatives; the group of Cohen has done a systematic investigation about the topic with amino-functionalized MOF. The material [Zn4O(atpa)3]n (IRMOF-3, atpa = 2-aminoterephthalic acid) was taken as starting material for nucleophilic substitution reactions at the amino groups with various anhydrides and isocyanates,35 whereas [Zn2(atpa)(DABCO)]n (DMOF-1-NH2, DABCO = 1,4-diazabicyclo(2.2.2)octane) and [Zn4O(atpa)(btb)4/3]n (UMCM-1-NH2) were reacted with anhydrides.36 Such reactions occur also by mixing a single crystal with the reactants to obtain the product as a single crystal. Not only it is possible to react metalorganic frameworks with organic reactants; it is also feasible to use metal precursors to produce metal-functionalized MOFs.37,38 An example was given by the group of Long, which presented the reaction of MOF-5 with Cr(CO)6 to aord [Zn4O[(Z6-tpa)Cr(CO)3]x(tpa)3x]n, a Cr(CO)3-functionalized MOF-5 that was characterized by PXRD and IR spectroscopy.38

3. Catalysis by MOFs
The comparison with other porous materials is helpful to predict what are possible future catalytic applications of MOFs. In fact, whereas it is more or less clear what is the utilization of zeolites and mesoporous aluminosilicates in catalysisoften commercially employed by the industrythe fate of PCPs is not obvious yet. In any case, we can speculate what are possible applications. Due to their high thermal stability and limited pore size, acidic zeolites are widely used in gas phase reaction under harsh conditions and industrial processes such as cracking, isomerization, oligomerization, and alkylation, whereas if an active center such as a transition metal, a metal oxide or sulde, or a metal complex is present, they are used in oxidation and reduction reactions.39 On the other hand, the highly dispersed active sites and large channels make mesoporous materials such as MCMs and SBAs more suitable for the encapsulation of metal particles or molecular catalysts to perform reactions with relatively large substrates.27 Since MOFs feature high crystallinity and do not have theoretical pore size limitations, they provide unique opportunities. This means that it is possible to have a homogeneous distribution of one or more active sites due to the high crystallinity of the material and, at the same time, to overcome diusion and pore size limitations. Moreover, the ne structure and the nature of the active site can be controlled. However, MOFs have still issues concerning stability and cost, due to the nature of the components of the materials. The right combination between cost, stability, and chemical application should be chosen accordingly. As shown in Fig. 4, for how things look at the moment, MOF catalysts tend to be more suitable for ne chemical synthesis than for bulk chemistry and it is not surprising that the examples in the literature reect this. The known strategies for the building of
This journal is
c

the Owner Societies 2011

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

holds the structure together. Therefore, one could experience a rearrangement of the coordination geometry in the inorganic nodes during catalysis, which might result in a collapse of the structure and a quick deactivation of the catalyst with possible negative consequences in activity, reproducibility, and recycling. In any case, the design of an active site, which also has the function of maintaining the framework, is clearly more limited than that of sites that are not responsible for maintaining the structure. Activity at the organic or pseudo-organic linkers. Organic or pseudo-organic building blocks, such as metal complexes with functionalized organic ligands, can also be catalytically active. The earliest instances were reported for Mn(III) and Zn(II) porphyrincarboxylate frameworks, which successfully catalyze the epoxidation of olens45 and acyl transfer to pyridylcarbinols,46 respectively. Such MOFs present the metal at both the inorganic nodes and inside the porphyrin, and it is likely that the active site is the metal that is coordinated by the porphyrin nitrogens. Another example is the amino-functionalized MOF [(Zn4O)(atpa)3] (IRMOF-3, atpa = 2-aminoterephthalate), which is an active basic catalyst for the Knoevenagel condensation of benzaldehyde with ethyl cyanoacetate.47 Asymmetric catalysis has also been performed by using this method. The asymmetric epoxidation of chromene derivatives is catalyzed by an enantiomerically pure Mn(III)salen MOF. The crystalline homochiral MOF, produced by assembling a chiral pyridino-functionalized Mn(III)salen complex with Zn(II) paddle-wheel dimers and biphenyldicarboxylic acid, catalyzes the epoxidation of 2,2-dimethyl-2H-chromene with up to 82% ee, slightly lower than the homogenous analogue (88% ee). Although the system showed remarkable enantioselectivity, such material experienced serious leaching problems that the authors explained by the weak Zn(II)N bond.48 Because of the relatively harsh synthesis conditions, it is not surprising that most crystalline materials with metal complexes are produced with stable metalporphyrins and metalsalen complexes. On the other hand, this approach is ideal for materials with organocatalytic properties such as amino-functionalized PCPs.47 3.2 Encapsulation of active species Encapsulation includes all catalytic reactions in which the active site is positioned within the pores of the MOF structure by a non-covalent interaction. Basically, the framework is used as a support for the catalyst and its main function is to provide a stable pore structure and surface area. Metal particles, complexes, and clusters supported within a PCP fall into this category. Although the examples are not numerous, there is a growing interest in the scientic community due to the large pores that are available in MOFs. BASF is developing an industrial production of MOF-5 to use in gas storage and purication,49 and as support for palladium, platinum, copper, and gold nanoparticles. These materials are then available for applications, such as the hydrogenation of cyclooctene (Pd@MOF-5), the synthesis of methanol from syngas (Cu@MOF-5),50 and H2O2 synthesis (Pt@MOF-5).51 Other examples are the absorption of polyoxometalates [PW11TiO40]5 and [PW11CoO39]5 into MOF-101 and the
Phys. Chem. Chem. Phys., 2011, 13, 63886396 6393

Fig. 4 Successful and potential applications of zeolites, mesoporous silica alumina compounds, and MOFs.

specic catalytic sites in crystalline PCPs can be categorized into the following three classes: (1) framework activity; (2) encapsulation of active species; (3) post-synthetic modication. We will give illustrative examples for each approach and discuss their elements of novelty, their promise, and their limits. Although the categories are analyzed separately, they can be combined and are not mutually exclusive. 3.1 Framework activity

Activity at the inorganic nodes. The earliest reports about catalytic activity of MOFs are those where the active site is an intrinsic part of the framework. Framework activity is shown in the cyanosilylation of aldehydes catalyzed by the cationic two-dimensional cadmium bipyridine (bpy) framework [Cd(bpy)]n(NO3)2n, in which the cadmium center is the active Lewis-acid site.40 Other metalorganic frameworks have shown similar Lewis acid catalytic activity. MOFs with copper paddle wheel dimers as inorganic unit, such as [Cu3(btc)2(H2O)3]n (HKUST-1 or MOF-199, btc = 1,3,5-benzenetricarboxylate), bear H2O molecules that complete the coordination sphere of the Cu(II) atoms.41 By activating these frameworks upon heating in vacuum, the water molecules are removed, and the material catalyzes the Lewis acid-catalyzed cyanosilylation of carbonyl groups. The same organic transformation is catalyzed by the more active [Cr3F(H2O)2O(tpa)3]n (MIL-101) without pre-treatment by heating.42 Examples of intrinsic Brnsted-acid catalytic activity are reported for MOFs such as [M3OF0.85(OH)0.15(H2O)2(btc)2]n (MIL-100, M = Fe, Cr),43 which catalyze the FriedelCrafts benzylation, whereas an outstanding example of catalytic activity of PCPs at the inorganic nodes is the framework [Pd(2-pymo)2]n (2-pymo = 2-hydroxypyrimidolate), which catalyzes the SuzukiMiyaura cross coupling of p-bromoanisole with phenylboronic acid with good conversion and selectivity also after one reuse.44 This approach is interesting, shows promise, and will be further developed. However, the reactive part is the one that
This journal is
c

the Owner Societies 2011

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

subsequent catalytic application in the oxidation of a-pinene to the corresponding alcohol and ketone using hydrogen peroxide and oxygen as oxidants.52 The catalysts that belong to this category could also be supported on other materials that provide porosity, surface area, and stability. MOFs are attractive because of their tunable pore size, high surface area, and crystalline ordering, which might produce shape selective catalysis. The use of enantiopure MOFs as support for particles is potentially intriguing.53 However, no example that demonstrates such concept is reported so far, to the best of our knowledge. An important analysis that should be performed when dealing with encapsulating catalysts is the leaching test to verify whether there is loss of particles or clusters. This test is performed by ltration of the mother liquor under reaction conditions to check whether the catalysis persists even in the absence of the MOF catalyst.54 3.3 Post-synthetic modication

As described above, one of the greatest properties of metal organic frameworks is their chemical versatility. This allows us to produce functionalized materials, to subsequently modify these functional groups, and to include active species, such as metal complexes, through covalent interactions. It is an intriguing aspect of porous coordination polymers as it allows researchers to design the catalytic site according to the need of the reaction. It also renders the process of designing, optimizing, and rationalizing the structure of heterogeneous catalysts more powerful than it ever was. Post-synthesis modication (PSM) of MOFs is generally applied to heterogenize known homogeneous catalysts. Fig. 5 illustrates PSM with both organic and inorganic precursors. A heterogeneous vanadyliminophenol complex was synthesized in two steps from IRMOF-3, an amino-functionalized MOF-5. Such material eectively catalyzes the oxidation of cyclohexene.55 A similar approach was explored by Corma and co-workers, who used the post-synthetic approach to produce a gold iminophenol functional material, which eectively catalyzed the selective hydrogenation of butadiene.56

Enantiomerically pure MOFs have been post-functionalized with metal complexes and then applied in catalysis. The most representative example is probably the one published in 2005 by Lin and co-workers, who structurally characterized an enantiomerically pure 2D cadmium MOF bearing binaphthol moieties. By post-functionalization with Ti(OiPr)4, the resulting complex catalyzed the alkylation of aldehydes with diethylzinc with great enantioselectivity up to 93% ee.57 Supposedly, the active site is a titanium(IV) complex with two isopropanolate ligands and two alcoholate ligands coming from the binaphthol groups on the framework. The fact that enantioselectivity was obtained suggests that the catalysis is heterogeneous. This is an outstanding result, which shows that it is possible to perform asymmetric transformations with crystalline metalorganic frameworks. Catalysts for asymmetric reactions can be produced also within a non-chiral MOF by attaching an enantiomerically pure compound through PSM. Our group is currently investigating this method to perform asymmetric hydrogenation of olens. We have synthesized MOF-5 that contains N(CH3)2 moieties, which serve as anchor points for transition metals, which then serve as catalyst. We have characterized NMe2-MOF-5 by single crystal X-ray diraction and identied the anchoring group by infrared and NMR spectroscopy after digestion. Further modication of the dimethylamino groups with [Rh((S,S)-Me-BPE)(COD)]OTf ((S,S)-Me-BPE = ()-1,2-bis[(2S,5S)-2,5-dimethylphospholano]ethane) produced a catalyst, which was active in the enantioselective hydrogenation of (Z)-2-acetamido-3-phenylacrylic acid (100 : 1 mol vs. initial Rh) in methanol under 3 bars of H2. We overcame the poor activity of the fully functionalized MOF (12% conversion and 80% ee) by using a MIXMOF, where the N(CH3)2 groups were diluted (7%) within the framework: 95% yield and 89% ee were obtained.58 Also for catalysts that belong to this class, a careful investigation of the structure of the active site, of the thermal and chemical stability of the framework, and of the crystallinity of the catalyst sample after the catalysis attempts should be performed. Leaching, surface vs. pores catalysis, and re-use

Fig. 5 Design of a vanadylsalen catalytic site by post-synthetic modication (PSM) of IRMOF-3. Only one pending amino group of the material is shown for clarity.55

6394

Phys. Chem. Chem. Phys., 2011, 13, 63886396

This journal is

the Owner Societies 2011

tests are again essential tools to understand whether the real catalyst is homogeneous or heterogeneous and if it occurs on the surface or within the pores.
Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

4. Towards unique catalytic applications


The properties of MOFs demonstrate that such materials have large applicability from a fundamental and industrial point of view for liquid phase heterogeneous catalysis, enantioselective transformations, and solvent-free reactions, and that can combine the benets of heterogeneous and homogeneous catalysis. Their chemical versatility allows us to create almost innite structures by changing the building blocks, by using the concepts of isoreticularity and that of MIXMOF, and by post-synthetic modication. If we fully exploit all these tools, the strategy for the preparation of the active sites might be revolutionized to design tailor-made catalysts with unique properties and tunable ne-structure environments. To take advantage of the full potential of MOFs, it is important to understand the connection between the design of the active site and its catalytic behavior. This can be done when catalysis is performed with crystalline PCPs, which allow us (a) to determine the exact position of the atoms in the catalytic site and the ne structure of the framework by singlecrystal X-ray diraction, (b) to investigate the crystallinity of the sample and its homogeneity by means of powder XRD, and (c) to correlate diraction patterns obtained with both XRD methods where single crystal XRD data are already available. Whenever single crystals suitable for X-ray diraction cannot be obtained, NMR, UV, IR, X-ray absorption spectroscopy, and ab initio calculations are useful methods to determine the structure of the active site. Since MOFs can be recycled and used several times in catalysis, we advise to check the crystallinity of the sample after any cycle to be sure that the solid maintains its homogenous structure throughout the catalytic application. Up until now, the strategies for the design of the active site have been developed according to the three categories described in the last chapter (framework activity, encapsulation of the active site, and post-synthetic modication). It is important to fully exploit their potential and even go beyond those classes by designing tailor-made pore structures, new (combinations of) active sites, and their direct environment to take advantage of the unique properties of MOFs. Imagining what kind of structures one could synthesize by employing their chemical versatility, one can suggest catalytic applications that are currently very dicult to perform otherwise, and we would like to list a few. The pore structure, which can be tuned by changing the organic/inorganic components, allows MOFs to have active sites bound to cages and to feature shape-selectivity in catalytic transformations. In this way, the reactions performance can be improved through the exclusion of too large reactants, the exclusive formation of certain products, or the selective desorption of products. This phenomenon is also observed in zeolites; MOFs have potential due to their tunable cavity size, which will lead to a broader range of molecules suitable for shape-selective catalysis and separation.
This journal is
c

Novel active sites can give rise to fascinating unique catalysis. For example, two dierent functional groups at opposite ends of a cage could be synthesized by using a templating molecule having two protecting groups at the exact distance bound at the linkers during synthesis. Removal of the template might lead to a key and lock mechanism with intriguing potential application. Additionally, the feasibility of producing materials with multiple functional groups via the MIXMOF or MTV-MOF concepts opens new perspectives in tandem reactions. Multivariable MOFs can also be helpful in designing diluted active sites and anchoring groups. This favors catalysis and PSM with MIXMOFs since they occur mostly within the pores, and not on the crystallite surface. The control of the environment circumventing the active site can produce MOFs that behave similarly to enzymes. This can widen the scope of chemicals and ne chemicals synthesis especially in asymmetric catalysis, asymmetric kinetic resolution, and resolution of enantiomers, which are transformations still dicult to perform with other solid materials. With MOFs, as said above, it should be feasible to introduce one or more functional groups in an oriented position on the framework to produce catalysts featuring multifunctional transformations with substrate and/or functional group selectivity. Such groups could be added either within the inorganic nodes or at the organic linkers, and non-covalent interactions may even be exploited. In this way, one can design catalytic sites and their neighboring environment, and control the polarity of the surroundings of the active site. This may lead to enhanced selectivity through the selective adsorption of a specic reactant or the forcing out of certain products. We are condent that one of the directions of the catalytic application of MOFs will be towards the tailoring of the active site(s) and of the local environment around it to mimic transformations that are still a challenge for classical lab-chemistry. Asymmetric and bio-like applications might come in play. We hope that the scientic community will continue developing new MOFs to solve the stability limitation of these materials to exploit their full capacity and to allow their industrial catalytic applicability. Catalysis by MOFs has just started to intrigue the scientic community and has an enormous scope for development of new materials with structure directing catalysis. Many unique and exciting catalytic applications will be reported in the following years.

References
1 J. B. Nagy, P. Bodart, I. Hannus and I. Kiricsi, Synthesis, Characterization and Use of Zeolitic Microporous Materials, Decagen Ltd., 1998. 2 S. Kitagawa, R. Kitaura and S. Noro, Angew. Chem., Int. Ed., 2004, 43, 23342375; J. Rowsell and O. Yaghi, Microporous Mesoporous Mater., 2004, 73, 314; G. Ferey, Chem. Soc. Rev., 2008, 37, 191214. 3 J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen and J. T. Hupp, Chem. Soc. Rev., 2009, 38, 14501459; L. Ma, C. Abney and W. Lin, Chem. Soc. Rev., 2009, 38, 12481256; D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int. Ed., 2009, 48, 75027513; Z. Wang, G. Chen and K. Ding, Chem. Rev., 2009, 109, 322359; A. Corma, H. Garcia and F. X. Llabres i Xamena, Chem. Rev., 2010, 110, 46064655. 4 F. Knobloch and W. Rauscher, J. Polym. Sci., 1959, 38, 261262.

the Owner Societies 2011

Phys. Chem. Chem. Phys., 2011, 13, 63886396

6395

5 C. G. Silva, I. Luz, F. X. Llabres i Xamena, A. Corma and H. Garcia, Chem.Eur. J., 2010, 16, 1113311138. 6 J.-R. Li, R. J. Kuppler and H.-C. Zhou, Chem. Soc. Rev., 2009, 38, 14771504; S. S. Han, J. L. Mendoza-Cortes and W. A. Goddard, Chem. Soc. Rev., 2009, 38, 14601476; C. Janiak, Dalton Trans., 2003, 27812804; M. D. Allendorf, C. A. Bauer, R. K. Bhakta and R. J. T. Houk, Chem. Soc. Rev., 2009, 38, 13301352; S. Bordiga, F. Bonino, K. P. Lillerud and C. Lamberti, Chem. Soc. Rev., 2010, 39, 48854927. 7 J. Johnson, A. Jacobson, W. Butler, S. Rosenthal, J. Brody and J. Lewandowski, J. Am. Chem. Soc., 1989, 111, 381383. 8 S. Drumel, P. Janvier, P. Barboux, M. Bujolidoeu and B. Bujoli, Inorg. Chem., 1995, 34, 148156. rey, Microporous Mesoporous 9 D. Riou, O. Roubeau and G. Fe Mater., 1998, 23, 2331. 10 H. Li, M. Eddaoudi, M. OKeee and O. M. Yaghi, Nature, 1999, 402, 276279. 11 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. OKeee and O. M. Yaghi, Science, 2002, 295, 469472. 12 J. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga and K. Lillerud, J. Am. Chem. Soc., 2008, 130, 1385013851; S. J. Garibay and S. M. Cohen, Chem. Commun., 2010, 46, 77007702. , C. Serre, C. Mellot-Draznieks, F. Millange and 13 S. Surble rey, Chem. Commun., 2006, 284286. G. Fe 14 A. D. Burrows, C. G. Frost, M. F. Mahon and C. Richardson, Angew. Chem., Int. Ed., 2008, 47, 84828486. 15 W. Kleist, F. Jutz, M. Maciejewski and A. Baiker, Eur. J. Inorg. Chem., 2009, 35523561. 16 H. Deng, C. Doonan, H. Furukawa, R. Ferreira, J. Towne, C. Knobler, B. Wang and O. Yaghi, Science, 2010, 327, 846849. 17 K. Koh, A. G. Wong-Foy and A. J. Matzger, Angew. Chem., Int. Ed., 2008, 47, 677680. 18 H. Chae, D. Siberio-Perez, J. Kim, Y. Go, M. Eddaoudi, A. Matzger, M. OKeee and O. M. Yaghi, Nature, 2004, 427, 523527. 19 B. Wang, A. Cote, H. Furukawa, M. OKeee and O. M. Yaghi, Nature, 2008, 453, 207211. 20 G. Ferey, J. Solid State Chem., 2000, 152, 3748; R. Robson, J. Chem. Soc., Dalton Trans., 2000, 37353744; B. Moulton and M. Zaworotko, Chem. Rev., 2001, 101, 16291658. 21 M. OKeee, M. Eddaoudi, H. L. Li, T. Reineke and O. M. Yaghi, J. Solid State Chem., 2000, 152, 320; M. Eddaoudi, D. Moler, H. LI, B. Chen, T. Reineke, M. OKeee and O. M. Yaghi, Acc. Chem. Res., 2001, 34, 319330. 22 J. J. Perry, J. A. Perman and M. J. Zaworotko, Chem. Soc. Rev., 2009, 38, 14001417. 23 J. Hazovic, M. Bjorgen, U. Olsbye, P. D. C. Dietzel, S. Bordiga, C. Prestipino, C. Lamberti and K. P. Lillerud, J. Am. Chem. Soc., 2007, 129, 36123620. 24 http:// www.iza-structure.org/databases/. 25 J. Beck, J. Vartuli, W. Roth, M. Leonowicz, C. Kresge, K. Schmitt, C. Chu, D. Olson, E. Sheppard, S. McCullen, J. Higgins and J. Schlenker, J. Am. Chem. Soc., 1992, 114, 1083410843. 26 D. Zhao, J. Feng, Q. Huo, N. Melosh, G. Fredrickson, B. Chmelka and G. Stucky, Science, 1998, 279, 548552. 27 A. Wight and M. Davis, Chem. Rev., 2002, 102, 35893614; A. Taguchi and F. Schuth, Microporous Mesoporous Mater., 2005, 77, 145. 28 H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi, A. O. Yazaydin, R. Q. Snurr, M. OKeee, J. Kim and O. M. Yaghi, Science, 2010, 329, 424428. 29 G. Ferey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surble and I. Margiolaki, Science, 2005, 309, 20402042.

30 G. Cruciani, J. Phys. Chem. Solids, 2006, 67, 19731994. 31 S. S. Kaye, A. Dailly, O. M. Yaghi and J. R. Long, J. Am. Chem. Soc., 2007, 129, 1417614178. 32 K. K. Tanabe, C. A. Allen and S. M. Cohen, Angew. Chem., Int. Ed., 2010, 49, 97309733. 33 Z. Wang and S. M. Cohen, Chem. Soc. Rev., 2009, 38, 13151329. 34 M. Kandiah, S. Usseglio, S. Svelle, U. Olsbye, K. P. Lillerud and M. Tilset, J. Mater. Chem., 2010, 20, 9848. 35 K. K. Tanabe, Z. Wang and S. M. Cohen, J. Am. Chem. Soc., 2008, 130, 85088517; E. Dugan, Z. Wang, M. Okamura, A. Medina and S. M. Cohen, Chem. Commun., 2008, 33663368. 36 Z. Wang, K. K. Tanabe and S. M. Cohen, Inorg. Chem., 2009, 48, 296306. 37 S. Chavan, J. G. Vitillo, M. J. Uddin, F. Bonino, C. Lamberti, E. Groppo, K.-P. Lillerud and S. Bordiga, Chem. Mater., 2010, 22, 46024611. 38 M. Dinca and J. R. Long, Angew. Chem., Int. Ed., 2008, 47, 67666779. 39 J. Cejka, A. Corma and S. Zones, Zeolites and Catalysis, Synthesis, Reactions and Applications, Wiley-VCH, Weinheim, 2010. 40 M. Fujita, Y. Kwon, S. Washizu and K. Ogura, J. Am. Chem. Soc., 1994, 116, 11511152. 41 K. Schlichte, T. Kratzke and S. Kaskel, Microporous Mesoporous Mater., 2004, 73, 8188. 42 Y. K. Hwang, D.-Y. Hong, J.-S. Chang, S. H. Jhung, Y.-K. Seo, J. Kim, A. Vimont, M. Daturi, C. Serre and G. Ferey, Angew. Chem., Int. Ed., 2008, 47, 41444148. 43 P. Horcajada, S. Surble, C. Serre, D.-Y. Hong, Y.-K. Seo, J.-S. Chang, J.-M. Greneche, I. Margiolaki and G. Ferey, Chem. Commun., 2007, 28202822. 44 F. X. Llabres i Xamena, A. Abad, A. Corma and H. Garcia, J. Catal., 2007, 250, 294298. 45 K. Suslick, P. Bhyrappa, J. Chou, M. Kosal, S. Nakagaki, D. Smithenry and S. Wilson, Acc. Chem. Res., 2005, 38, 283291. 46 A. M. Shultz, O. K. Farha, J. T. Hupp and S. T. Nguyen, J. Am. Chem. Soc., 2009, 131, 42044205. 47 J. Gascon, U. Aktay, M. D. Hernandez-Alonso, G. P. M. van Klink and F. Kapteijn, J. Catal., 2009, 261, 7587. 48 G. Morris, S. Nguyen and J. Hupp, J. Mol. Catal. A: Chem., 2001, 174, 1520. 49 U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt and J. Pastre, J. Mater. Chem., 2006, 16, 626636; A. U. Czaja, N. Trukhan and U. Mueller, Chem. Soc. Rev., 2009, 38, 12841293. 50 S. Hermes, M. Schroter, R. Schmid, L. Khodeir, M. Muhler, A. Tissler, R. Fischer and R. Fischer, Angew. Chem., Int. Ed., 2005, 44, 62376241. 51 U. Mueller, O. Metelkina, H. Junicke, T. Butz and O. M. Yaghi, US Patent, 2004/081611, 2004. 52 N. V. Maksimchuk, K. A. Kovalenko, S. S. Arzumanov, Y. A. Chesalov, M. S. Melgunov, A. G. Stepanov, V. P. Fedin and O. A. Kholdeeva, Inorg. Chem., 2010, 49, 29202930. 53 B. Kesanli and W. Lin, Coord. Chem. Rev., 2003, 246, 305326. 54 R. Sheldon, M. Wallau, I. Arends and U. Schuchardt, Acc. Chem. Res., 1998, 31, 485493. 55 M. J. Ingleson, J. P. Barrio, J.-B. Guilbaud, Y. Z. Khimyak and M. J. Rosseinsky, Chem. Commun., 2008, 26802682. 56 X. Zhang, F. X. L. i. Xamena and A. Corma, J. Catal., 2009, 265, 155160. 57 C. Wu, A. Hu, L. Zhang and W. Lin, J. Am. Chem. Soc., 2005, 127, 89408941. 58 M. Ranocchiari, J. A. van Bokhoven and F. Mu nch, unpublished results.

Published on 14 January 2011. Downloaded by Universidade Tecnica de Lisboa (UTL) on 26/09/2013 15:44:40.

6396

Phys. Chem. Chem. Phys., 2011, 13, 63886396

This journal is

the Owner Societies 2011

You might also like