Industrial applications of metal–organic frameworks.

11
This article was published as part of the 2009 Metal–organic frameworks issue Reviewing the latest developments across the interdisciplinary area of metal–organic frameworks from an academic and industrial perspective Guest Editors Jeffrey Long and Omar Yaghi Please take a look at the issue 5 table of contents to access the other reviews. Downloaded by NATIONAL TECHNICAL UNIVERSITY OF ATHENS on 22 February 2013 Published on 16 March 2009 on http://pubs.rsc.org | doi:10.1039/B804680H View Article Online / Journal Homepage / Table of Contents for this issue

description

Metal–organic frameworks (MOFs) constitute anemerging class of materials useful in gas storage, gas purification and separation applications aswell as heterogeneous catalysis.

Transcript of Industrial applications of metal–organic frameworks.

Page 1: Industrial applications of metal–organic frameworks.

This article was published as part of the

2009 Metal–organic frameworks issueReviewing the latest developments across the interdisciplinary area of

metal–organic frameworks from an academic and industrial perspective Guest Editors Jeffrey Long and Omar Yaghi

Please take a look at the issue 5 table of contents to access the other reviews.

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

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

Page 2: Industrial applications of metal–organic frameworks.

Industrial applications of metal–organic frameworksw

Alexander U. Czaja,*aNatalia Trukhan

band Ulrich Muller

b

Received 10th February 2009

First published as an Advance Article on the web 16th March 2009

DOI: 10.1039/b804680h

New materials are prerequisite for major breakthrough applications influencing our daily life, and

therefore are pivotal for the chemical industry. Metal–organic frameworks (MOFs) constitute an

emerging class of materials useful in gas storage, gas purification and separation applications as

well as heterogeneous catalysis. They not only offer higher surface areas and the potential for

enhanced activity than currently used materials like base metal oxides, but also provide shape/size

selectivity which is important both for separations and catalysis. In this critical review an

overview of the potential applications of MOFs in the chemical industry is presented.

Furthermore, the synthesis and characterization of the materials are briefly discussed from the

industrial perspective (88 references).

Introduction

Already in 1965, roughly three decades before the commonly

assumed birth year of metal–organic frameworks (MOFs),

Tomic mentioned materials which would nowadays be called

MOFs, metal–organic polymers or supramolecular structures.1

Bi- and trivalent aromatic carboxylic acids were used to form

frameworks with zinc, nickel, iron, aluminium, thorium

and uranium. Some interesting features of MOFs, like high

thermal stability and high metal content were already

reported. Also in 1965, Biondi et al. reported on Cu(II)

tricyanomethanide being a crystalline, polymeric compound.2

In 1990, Hoskins and Robson reported on the design of

scaffold-like materials using Cu(I) centres and tetracyano-

tetraphenylmethane.3

Interest in the field was again kindled by the group of O. M.

Yaghi, which published the structure of MOF-5 in late 1999,4

and the concept of reticular design, with totally different

carboxylate linkers, in 2002.5–7

Numerous reviews summarize the fast growing research

efforts in the field. The most comprehensive ones are by

S. Kitagawa,8 O. M. Yaghi,9 and G. Ferey.10 The review by

Gerard Ferey is also highly recommended as an introduction

to the concepts applied in MOF research.

Structures, properties and possible applications of MOFs as

storage media were studied.9,10 Comparisons with oxides,

molecular sieves, porous carbon and heteropolyanion salts

have been made by Barton and co-authors.11 Nowadays

several hundred different MOFs have been identified. The

self-assembly of metal ions, which act as coordination centres,

linked together with a variety of polyatomic organic bridging

ligands, has resulted in tailored nanoporous host materials

as robust solids with high thermal and mechanical stability.

a Chemicals Research and Engineering, GCC/PZ, BASF SE,California NanoSystems Institute, University of California,Los Angeles, 570 Westwood Plaza, Los Angeles, CA 90095, USA.E-mail: [email protected]

bChemicals Research and Engineering, GCC/PZ, BASF SE,67056 Ludwigshafen, Germany

w Part of the metal–organic frameworks themed issue.

Alexander U. Czaja

Alexander Czaja, born 1977 inNurnberg, Germany. 1997:studied chemistry at theFriedrich-Alexander Univer-sity Erlangen-Nurnberg. 2005:Dr. rer. nat. in the group ofProf. Dieter Sellmann/Prof.Horst Kisch (thesis on compe-titive catalysts for nitrogenfixation); subsequently joinedBASF SE (Ammonia Labora-tory) as a research manager,working on heterogeneousoxidation catalysis and highthroughput catalyst screening.2008: BASF scholar at the

University of California in Los Angeles, currently workingon catalysis with metal–organic frameworks (MOFs) incollaboration with Prof. Omar Yaghi.

Natalia Trukhan

Natalia Trukhan, born 1978 inPetropavlovsk-Kamchatskij,Russia. 1995: studied chemis-try at Novosibirsk StateUniversity, Department ofNatural Sciences. 2003: PhDin the group of Dr Kholdeevaon speciality ‘‘Kinetics andCatalysis’’ (thesis on theselective oxidation of organicsubstrates over mesoporoustitanium- and vanadium-silicates). 2004: Postdocat Stuttgart University,Germany, with Prof. Rodunersupported by an Alexander

von Humboldt Fellowship. 2005: research manager atBASF SE. Research activity: zeolite and MOF synthesis andapplications.

1284 | Chem. Soc. Rev., 2009, 38, 1284–1293 This journal is �c The Royal Society of Chemistry 2009

CRITICAL REVIEW www.rsc.org/csr | Chemical Society Reviews

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 3: Industrial applications of metal–organic frameworks.

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 interactions with guest molecules.8 The most striking

difference to state-of-the-art materials is probably the total

lack of non-accessible bulk volume in MOF 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. It was reported, for example, that in the case of

MOF-177 the surface area reaches 5640 m2 g�1 12 and for

MIL-101 up to 5900 m2 g�1.13 Of course, properties like the

drastically increased velocity of molecular traffic through these

open structures are closely related to the regularity of pores in

nanometre size as well.

Thus, the combination of such unbelievable levels of

porosity, surface area, pore size and wide chemical inorganic–

organic composition has recently brought these materials to

the attention of many researchers both in academia and

industry, with over 1000 publications on ‘‘metal–organic

frameworks’’ and ‘‘coordination polymers’’ per annum.8

In this review, we focus on the topics of synthesis,

characterisation and application of MOF materials from an

industrial point of view. Whenever possible, comparison is

made to state-of-the-art applications in order to outline

possibilities of processes which might be beneficially run by

using MOFs. Recent findings, like the electrochemical reactivity

of MIL-53(Fe) with a promise for application in Li-based

batteries,14 the first completely new class of scintillation

materials using Zn-MOFs with stilbene dicarboxylic acid as

linker15 and the storage and delivery of pharmaceuticals by

MOFs,16 are acknowledged and may find application in

industry some time. However, they would go beyond the scope

of this review which focuses on the application in chemical

industry.

1. Synthesis

Usually, the synthesis of MOFs is straightforward, using

well-soluble salts as the source for the metal component, e.g.

metal nitrates, sulfates or acetates. The organic ingredients,

which mostly are mono-, di-, tri- and tetracarboxylic acids,

are supplied in a polar organic solvent, typically an amine

(triethylamine) or amide (diethylformamide, dimethylformamide).

After combination of these inorganic and organic components

under stirring, the metal–organic structures are formed by

self-assembly at temperatures starting at room temperature

and up to solvothermal conditions at 200 1C within a few

hours. A typical scheme of a semi-technical process is given in

Fig. 1, where the scale-up synthesis of lightweight Mg-MOF

(Basolite M050) with formic acid as linker is shown as an

example. The scheme indicates not only the different steps of

preparation but also recycling of the solvent and further

processing of the dried powder into shaped material. Some

details about preparing shaped bodies from MOF materials

(Fig. 2) are published.17,18

It should be pointed out, however, that the filtering and

drying of metal–organic compounds in the wet processing

steps need to be carried out very carefully, as—due to

their high porosity and surface area—MOFs may easily carry

Fig. 1 Simplified flow diagram of industrial metal–organic frame-

work synthesis procedures via solvothermal (Mg-MOF) and electro-

chemical (Cu-EMOF) routes.

Fig. 2 Shaped bodies made from MOF materials.

Ulrich Muller

Ulrich Muller, born 1957 inKatzenelnbogen, Germany.1977: studied chemistry inMainz (thesis on the synthesisof large zeolite crystals andsorption properties) andreceived his PhD in thegroup of Prof. K. K. Unger;research activities at CNRS‘Tian&Calvet’, Marseille,ILL Grenoble, and with G. T.Kokotailo, Univ. Pennsylvania.1989: BASF SE: zeolitesynthesis and application incatalysis and adsorption.1999: Senior Scientist, zeolite

catalysis: CFC-free polyurethane foams, catalysts for cropprotection agents, chemical intermediates, sorptive olefin feed-stream purification, piloting of propylene epoxidation catalysts.1999: Synthesis, scale-up, modification and testing of variousmetal–organic framework compositions. 2005: BASF ResearchDirector.

This journal is �c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 1284–1293 | 1285

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 4: Industrial applications of metal–organic frameworks.

50 to 150 wt% of occluded solvent. This is an order of

magnitude higher than in a zeolite or base metal oxide

preparation. Therefore, it is advisable to first remove the large

amount of adsorbed solvent under gentle conditions regarding

pressure and temperature, prior to high thermal activation.

Furthermore, the crystallization of MOFs is an equilibrium

reaction between an inorganic base metal salt and an

organic acid. The preparation of, for example, MOF-5 is

stoichiometrically expressed as:

4[Zn(NO3)2�4H2O] þ 3[H2BDC] þ 8[OH�]

# [Zn4O(BDC)3] þ 8[NO3�] þ 23[H2O]

where H2BDC is terephthalic acid and Zn4O(BDC)3 is the

MOF-5 unit composition. It is clearly seen, that the

equilibrium can be shifted to the MOF-product side by

working on the concentration profiles of, for example, the

solvent, liberated water and nitrate, respectively. By taking

these equilibrium conditions into account, it became possible

to improve the MOF quality; i.e. to increase the surface area

from 2600 m2 g�1 up to 3400 m2 g�1.19

As the reaction can be easily driven in both directions, it is

obvious that the stability and integrity of MOFs in their use

depend on the polarity of the environment and the pH-value.

The conditions under which a material is stable have to be

considered on a case-by-case basis.

On a larger production scale, attention has to be paid

to safety issues whenever high nitrate concentrations are

involved, especially in addition to the build-up of large surface

area volumes in adiabatic reactor vessels. Thus, an alternative

salt-free procedure has been developed by BASF SE, using an

electrochemical route. Bulk metal sacrificial anodes are

oxidized in the presence of dissolved carboxylates in an

electrochemical cell (Fig. 1).20,21 Simple recovery of the formed

solid by filtration directly yields the final MOF powder

after drying. This procedure is especially beneficial for

MOF-structures containing open metal sites, as no anions

from added salts can block the access to the sites. By the same

method the successful synthesis of Zn-EZIF (electrochemically

made zeolite imidazolate framework) with 2-methylimidazole

as a linker was reported.22

It is necessary to note that an issue of prime importance for

scale-up production is space–time–yield (STY) (kg of MOF

per m3 of reaction mixture per day), which should be as high as

possible. In Table 1, the STY for some industrial MOFs in

comparison to zeolite production are presented.

A summary of some verified synthesis procedures for

laboratory and larger scale has been presented.21

2. Characterization

As MOFs are both crystalline and highly porous materials,

most frequently X-ray diffraction (XRD) is used to characterize

the crystallinity and phase purity, adsorption measurements

are performed to check for the porosity, and more elaborate

studies make use of neutron scattering to determine sorption

sites.23 Commercially available equipment using nitrogen

sorption at 77 K or argon uptake at 87 K are applied and

equivalent surface areas are calculated according to, e.g., the

Langmuir equation. However, it should be considered that the

underlying model of independent, equivalent and non-infringing

sorption sites might be different on a molecular level. Many

reports have described localized rather than bulk volume

adsorption phenomena on metal–organic materials, with some

even clearly differentiating between various crystallographic

sites and adsorption strengths.23–26

For industrial applications, sorption studies are simply a

convenient and fast way to compare the higher sorption

capacities of MOFs over state-of-the-art sorbents, although

care has to be taken whenever applications might rely upon

gravimetric or volumetric uptake later on.

While mass transfer can heavily influence results in catalysis,

it is clearly necessary to have information at hand on the

crystal sizes and size distribution of the MOF samples, which

are routinely collected by scanning electron microscopy

(Fig. 3).

As the metal content reaches values between 20–40 wt%, it

is also desirable to check the local metal cluster arrangements

and environments using more elaborate methodology such as

extended X-ray adsorption fine structure (EXAFS), X-ray

adsorption near-edge structure (XANES) or X-ray photon

spectroscopy (XPS).

The adsorbates in the pores of MOFs can be investigated by

UV-VIS, IR- and Raman-spectroscopy.7

3. Applications of metal–organic frameworks

Only a limited number of possible applications have been

discussed for MOFs so far and, to the best of our knowledge,

none have been industrially realized as yet. However, taking

a closer look at specific properties, it becomes clear that

especially the high porosity and the absence of hidden volumes

in these new frameworks in principal render them quite useful

for volume specific applications like adsorption, separations,

purification purposes, and catalysis. Typical drop-in technologies

might use MOFs first, for example, substituting zeolitic

molecular sieves, activated carbons and base metal oxides in

Table 1 Space–time–yield for MOF synthesis

Materials Langmuir surface area/m2 g�1 Space–time–yield/kg m�3 day�1

Basolite A100 (Al-MOF) Al-terephthalate 1100–1500 160Basolite C300 (Cu-BTC-MOF) Cu-benzene-1,3,5-tricarboxylate 1500–2100 225Basolite F300 (Fe-EMOF) Fe-benzene-1,3,5-tricarboxylate 1300–1600 20Basolite Z1200 (Zn-EZIF) Zn-2-methylimidazole 1300–1800 100Basolite M050 (Mg-MOF) Mg-formate 400–600 4300Zeolite — 300–800 50–150

1286 | Chem. Soc. Rev., 2009, 38, 1284–1293 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 5: Industrial applications of metal–organic frameworks.

existing plants to gain better performance in variable costs

without any fixed capital expenditure.

3.1 Gas purification

One possible field where MOFs can be beneficially used might

be the removal of ppm-traces of sulfur components from

various gases. Especially MOF structures with accessible

open metal sites are well suited to strongly (430 kJ mol�1)

chemisorb electron-rich, odour-generating molecules, like

amines, phosphines, oxygenates, alcohols, water, or sulfur-

containing molecules.

One example, which was experimentally evaluated in

continuous breakthrough trials, is the removal of tetra-

hydrothiophene (THT, odorant) from natural gas. At room

temperature, traces of 10–15 ppm sulfur were fully captured

down to less than 1 ppm using an electrochemically prepared

shaped Cu-EMOF (Fig. 4) in a fixed bed reactor.21 The

overall capacity of the MOF material (70 g THT LMOF�1)

outperformed, by about an order of magnitude, commercially

available activated carbon materials as adsorbents, namely

Norit (type RB4) and CarboTech (type C38/4).

3.2 Gas separation

In gas separation processes, the gas mixtures usually consist of

components having concentrations in the same order of

magnitude. This is in contrast to the gas purification processes.

Usually, either distillation or pressure and/or thermal swing

adsorption–desorption are used to separate the mixtures.

Examples of existing technologies are nitrogen–oxygen (air),

nitrogen–methane, and noble gas (e.g. Kr–Xe) separations,

some of which use zeolitic adsorbents.

Recently, the separation of Kr–Xe by pressure swing

adsorption, as well as the purification of methane in natural

gas were piloted on MOF-adsorbents.27,28

Fig. 3 Morphology of crystals from different MOF samples determined by SEM analysis.

Fig. 4 Feed stream purification and removal of tetrahydrothiophene

from natural gas using Cu-BTC-EMOF.

This journal is �c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 1284–1293 | 1287

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 6: Industrial applications of metal–organic frameworks.

Interestingly, even small pore MOFs have been described

which are able to separate molecules by size or kinetic

diameter,29–33 e.g., while water uptake occurs, slightly larger

molecules such as N2, O2, CO2, and methane are occluded

from adsorption.30 To demonstrate this concept, the adsorption

isotherms of nitrogen and hydrogen for the small-pore

Mg-MOF (Basolite M050) are shown in Fig. 5. Preferential

adsorption of hydrogen over nitrogen is observed.

CO2 separation is of great interest due to the awareness of

global warming over the past years. MOFs were considered as

CO2 adsorbents.10,34,35 Currently, elimination of CO2 from

low pressure flue gases as emitted, for example, by fossil fuel

burning power plants is done very effectively by gas scrubbing.

The partial pressure of CO2 in these streams lies between

60 and 130 mbar. The separation is done using the reversible

stoichiometric reaction (chemisorption) of CO2 with an amine

in aqueous solution (amine wash) at room temperature. CO2

in MOFs is physisorbed in the same temperature region. Due

to the lower energy of adsorption for the physisorption

process, MOF adsorbents, e.g. MIL-53(Cr),36 need a higher

CO2 partial pressure to reach sufficient adsorption capacities

(Fig. 6). This shows that MOFs currently can only be

effectively used in CO2 separation at high pressure. However,

even under these conditions they do not perform as well as an

amine scrubber.

3.3 Gas storage

Due to the unique structure of MOFs and especially due to the

absence of dead volume, it becomes possible to increase the

volumetric specific gas storage above previously known levels.

This effect can be very pronounced and depends on the type,

temperature and pressure of the gas, as well as on the specific

MOF material being used.

The mechanism of increased storage in MOF-filled gas

cylinders over empty gas bottles is easy to understand, once

the underlying principles are considered. The filling of a

conventional gas cylinder is simply applying physical effects

depending on the pVT-characteristic of the gas under

investigation. Considering MOF-filled gas cylinders, the above

mentioned effect is overlaid by an additional adsorption effect

inside the MOF. As these frameworks are free of dead-volume,

there is almost no notable loss of storage capacity due to space

blocking by non-accessible volume. Summarizing, pVT-filling

plus adsorption contribute to an enhanced volumetric storage

capacity.

The uptake curves resulting from this effect are shown in

Fig. 7 for the storage of CH4 at room temperature and

pressures up to 200 bar in gas cylinders filled with different

MOF materials. A non-linear uptake behaviour can be

monitored as the pressure is increased. Release of the gas

follows the same curve when the valves are opened and the

pressure is reduced. At a pressure of 150 bar, it is clearly seen

that about 35% more capacity is reached over the state-of-

the-art filling curve. It should be noted that the benefit of

enhanced capacity with MOF-filled vessels can only occur if

the gas is in a true gaseous state and not in the liquid-phase.

Furthermore, the density of the liquid phase is the limiting case

for the storage capacity of the MOF. Graphically speaking, a

MOF cannot compress gas to a higher density than its liquid

state density.

Recently, exceptionally high uptakes of even acetylene were

reported by Kitagawa, in this case clearly exceeding the

packing density usually obtained in today’s storage devices.37

The greatest challenge is, of course, the storage of hydrogen.

The underlying possibility to use hydrogen as a fuel for mobile

or portable fuel-cell applications raises a very high interest in

Fig. 5 N2- and H2-adsorption isotherms (Mg-MOF, Basolite M050)

indicating a molecular sieving effect.

Fig. 6 Adsorption isotherms of carbon dioxide on MIL-53(Cr) in

comparison to carbon dioxide removal from flue gas by amine wash. Fig. 7 CH4-storage in MOFs (prototype trials).

1288 | Chem. Soc. Rev., 2009, 38, 1284–1293 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 7: Industrial applications of metal–organic frameworks.

hydrogen storage possibilities. MOF-storage for hydrogen

works fully reversibly, avoids complicated heat treatments,

and recharging proceeds within seconds or minutes. This

clearly is an advantage over, e.g., metal hydrides as storage

materials. In MOF-5 even the distinct locations of adsorbed

hydrogen molecules have been elucidated using inelastic

neutron scattering,23 while density functional theory (DFT)

studies have proposed a storage capacity of 16–20 molecules of

H2 per Zn4O-cluster.26

The need for alternative fuel sources and energy carriers

together with target values by the US Department of Energy

was recently reviewed.38 Storage data are reported of up to

7.5 wt% of hydrogen on MOFs, (e.g.MOF-177) at 9 MPa and

77 K.12,27,37 Many other research groups look into this

challenge as well. 3.8 wt% hydrogen capacity has been

reported by Ferey’s group onMIL-5339 derived from aluminium

salts and terephthalic acid. Pillaring with secondary amine

linkers (triethylenediamine) as a strategy was employed by

Kim40 and Seki,41 which resulted in hydrogen uptakes up to

2.0 wt% at 0.1 MPa and 87 K. Similar values were achieved

with MOF-505 by Yaghi’s group,42 where Cu-paddle-wheels

are connected by 3,30,5,50-biphenyltetracarboxylic acid.

Doubly interpenetrated nets of zinc frameworks built by

NTB-linkers (4,40,400-nitrilotrisbenzoic acid) by Suh43 were

reported to reach 1.9 wt% of hydrogen uptake at 77 K.

However, it is not yet known whether large surface area

materials like MOF-177,7 MIL-10044 or MOF-5 and isoreticular

compounds,4–6 or materials with an average surface area of

between 1000 to 1500 m2 g�1 39,43 or even small pore

MOFs24,33 will be the most promising storage media. Neither

can it be concluded whether divalent or trivalent metal

clusters45 are the most favourable ones. Nevertheless, a

comparison to NaX-zeolite very clearly indicates the superior

behaviour of MOFs over conventional microporous inorganic

media.46

From our results on prototype equipment (77 K and up to

40 bar) it can be seen how different MOF materials contribute

differently to volume-specific hydrogen storage (Fig. 8).

MOF-177 demonstrates the most promising results for the

selected MOFs. This data from a large scale prototype

compares reasonably well with the literature.38

For many volume-limited mobile and portable fuel-cell

applications it will be industrially much more relevant to

compare storage data on a volume-specific rather than a

weight-specific storage capacity. Typically, the packing

densities of MOF powders are around 0.2 to 0.4 g cm�3

increasing to 0.5–0.8 g cm�3 when shaped into tablets or

extrudates. The density of this material is so low that weight

limitation for an application, generally, is not relevant. This is

in contrast, of course, to the alternative use of metal hydrides

as storage media.

The most important issue here is the amount of hydrogen

which, on a reasonable time-scale, can be discharged from the

storage media. Within this context, MOFs really have a fully

reversible uptake- and release-behaviour. As the storage

mechanism is based predominantly on physisorption, there

are no huge activation energy barriers to be overcome when

liberating the stored hydrogen. A simple pressure reduction by

controlling valve opening is sufficient to draw off hydrogen

from MOFs within a few seconds.

Energy density values of 1.1 kW h L�1, as requested in the

European hydrogen and fuel cell Strategic Research Agenda

and Deployment Strategy,47 are equivalent to a volumetric

hydrogen stored capacity of about 33 g H2 L�1. As has been

demonstrated, this value can be reached by storing hydrogen

in MOF-177 at 77 K at a moderate pressure of 40–50 bar.

However, due to the low degree of interaction between

hydrogen molecules and its low heat of adsorption, until

now significant storage capacity at room temperature has

not been reached.

Getting better information on the preferred adsorption sites

for hydrogen in MOF structures should principally enable the

prediction of storage capacities. In this respect, molecular

modelling tools might become as important as elaborate

experimental synthesis efforts.46 Depending on the temperature

range of a possible application, highly porous MOFs might be

favourable for low temperatures, whereas rather small pore

materials,24,33 or highly attractive and flexible ones,39,48 could

be favourites for room temperature hydrogen storage.

More recent reports have addressed biomimetic approaches,

like Co4(m4-O)(carboxylate)4 units49 being related to the structure

of active centres like Fe in haemoglobin. Such materials have

quite high sorption enthalpies of 10 kJ mol�1 for hydrogen

and similarly very high enthalpies for O2, CO or CH4.

3.4 Heterogeneous catalysis

3.4.1 Relevance. Due to its crucial role in many chemical

processes, heterogeneous catalysis is one of the key elements of

our industrialized society, and thus has direct impact on the

global economy. It is reported that approximately nine-out-

of-ten chemical processes utilize heterogeneous catalysts.50

The global catalyst market is estimated to be between

15–20 billion USD annually. Approximately half of this

market is geared directly towards the chemical industry

and the rest is divided between environmental and refinery

applications. The annual combined worth of the products

obtained from industrial catalytic processes (including refinery

operations) is estimated to be at a multi-trillion USD level.

Therefore, it can be stated that the value created by utilizingFig. 8 H2-storage capacities for different MOFs (prototype trials).

This journal is �c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 1284–1293 | 1289

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 8: Industrial applications of metal–organic frameworks.

Table 2 Summary of catalytic reactions employing MOFsa

MOF type Reaction type Reactants Ref.

MOF as classical catalyst support

Ag@[Zn4O(BDC)3] (MOF-5) Epoxidation Propylene þ O2 53Pt@[Zn4O(BDC)3] (MOF-5) H2O2 synthesis H2 þ O2 54Cu@[Zn4O(BDC)3] (MOF-5) Methanol synthesis Syngas 55Pd@[Zn4O(BDC)3] (MOF-5) Hydrogenation Cyclooctene þ H2 55

Link as active centre

Ti(OiPr)4[Cd3Cl6(L1)3]�4DMF�6MeOH�3H2O Addition tocarbonyls

ZnEt2 þ aromatic aldehydes 56, 57Ti(OiPr)4[Cd3(NO3)6(L1)4]�7MeOH�5H2OTi(OiPr)4[Cd(L1)2(H2O)2][ClO4]2�DMF�4MeOH�3H2O[Rh2(M

21TCPP)2] (M21 ¼ Cu, Ni, Pd) Hydrogenation Propene þ H2 58

1-Butene þ H2 59[Zn2(BPDC)2(L2)]�10DMF�8H2O Epoxidation 2,2-Dimethyl-2H-chromene þ

2-(tert-butylsulfonyl)iodosylbenzene60

SBU (secondary building unit) as active centre

[Zn3O(L3-H)]�(H3O)2(H2O)12 (D-POST-1) Transesterification Acetic acid 2,4-dinitrophenyl ester þ EtOH 61

[Sm(L4-H2)(L4-H3)(H2O)4]�(H2O)x Esterification Meso-2,3-dimethylsuccinic anhydride þ MeOH 62[Cu(bpy)(H2O)2(BF4)(bpy)] Alcoholysis of

epoxidesEpoxide þ alcohol 63

[Zn4O(BDC)3] (MOF-5) Alkoxylation Dipropylene glycol þ PO, methyl dipropyleneglycol þ EO, acrylic acid þ EO

64

Zn-DHBDC-MOF, Zn-BTC-MOF Polyalkylenecarbonateformation

PO þ CO2 65

[Ln(OH)(H2O)(naphthalenedisulfonate)],Ln ¼ Nd, Pr, La

Oxidation Linalool þ H2O2 66

[In4(OH)6(BDC)3] Oxidation Methylphenylsulfide,(2-ethylbutyl)phenylsulfide þ H2O2

67

[Cu3(BTC)2] Oxidation Polyphenols þ H2O2 68

[Sc2(BDC)3] Oxidation Methylphenylsulfide þ H2O2 69

[Sc2(BDC)2.5(OH)],[Y2(BDC)3(H2O)2]�H2O, [La2(BDC)3(H2O)2]�H2O

70

[Pd(2-pymo)2] Oxidation Cinnamyl alcohol þ air 71

[Rh2(H2TCPP)2]BF4 Oxidation Alcohol þ air/O2 72

[Cu2(trans-1,4-cyclohexanedicarboxylate)2]�H2O Oxidation Alcohol þ H2O2 72

[Cu(2-pymo)2] Oxidation Tetralin þ air 73[Co(PhIM)2][In2(BDC)3(bpy)2], [In2(BDC)2(OH)2(phen)2],[In(BTC)(H2O)(bpy)], [In(BTC)(H2O)(phen)]

Acetalization Benzaldehyde þ trimethylorthoformate 74

[Sc2(BDC)2.5(OH)], [Y2(BDC)3(H2O)2]�H2O,[La2(BDC)3(H2O)2]�H2O

70

[Cu3(BTC)2] Cyanosilylation Benzaldehyde þ cyanotrimethylsilane 75

[Cd(4,40-bpy)2(H2O)2](NO3)2�(H2O)4 Cyanosilylation Benzaldehyde, aromatic imines þcyanotrimethylsilane

76, 77

[Sm(L4-H2)(L4-H3)(H2O)4]�(H2O)x Cyanosilylation Benzaldehyde þ cyanotrimethylsilane 62Mn3[(Mn4Cl)(BTT)8(MeOH)10]2 Cyanosilylation Aromatic aldehyde þ cyanotrimethylsilane 78

[Zn4O(BDC)3] (MOF-5) Addition to alkynes 4-tert-butylbenzoic acid þ C2H2, MeOH þpropyne

79

Ti-(2,7-dihydroxynaphthalene)-MOFs Ziegler–Nattapolymerization

Ethylene, propylene 80

[Pd(2-pymo)2] C–C coupling 4-Bromoanisole þ phenylboronic acid 71

[Cu3(BTC)2] Isomerization a-Pinene oxide 81

[Cu3(BTC)2] Cyclization Citronellal 81

[Cu3(BTC)2] Rearrangement 2-Bromopropiophenone 81

[Rh2(L5)] Hydrogenation Ethylene, propylene þ H2 82

[Rh(BDC)], [Rh(fumarate)] H–D exchange,hydrogenation

Propylene þ H2 83

[Ru(1,4-diisocyanobenzene)2]Cl2 Hydrogenation 1-Hexene þ H2 84, 85[In4(OH)6(BDC)3] Hydrogenation 1-Nitro-2-methylnaphthalene,

nitrobenzene þH2

67

[Ru2(BDC)2], [Ru2(BPDC)2], [Ru2(BDC)2(dabco)],[Ru2(BPDC)2(dabco)]

Hydrogenation Ethylene þ H2 86

[Rh2(fumarate)2], [Rh2(BDC)2], [Rh2(H2TCPP)2] Hydrogenation Ethylene þ H2 58H–D exchange Propylene þ H2

[Pd(2-pymo)2] Hydrogenation Octane/cyclododecene þ H2 71

1290 | Chem. Soc. Rev., 2009, 38, 1284–1293 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 9: Industrial applications of metal–organic frameworks.

catalysts is approximately three orders of magnitude higher

than the amount invested in them.

Shape and size selectivity is a vital consideration for many

industrial catalytic processes, and it is typically attained by

utilizing catalysts of nanoporous nature. The selectivity can be

based on the shape/size of the reactant, product or intermediate.

In order to provide such shape/size selective behaviour, the

catalyst must have uniform pores with molecular dimensions.

Uniform porosity is the outcome of a highly ordered structure,

such as MOFs, where pores/channels are part of the building

blocks (i.e., repetitive units). Each MOF has the potential to

offer unique structural and chemical features that can be

beneficial for an industrial application.

The most remarkable characteristic of MOFs relevant

for catalysis, which makes them unique, is the lack of

non-accessible bulk volume (‘‘dead volume’’). Furthermore,

due to the very open architecture, the self diffusion coefficients

of molecules in the pore system are only slightly lower than in

the bulk solvent.51 This means that mass transport in the pore

system is not hindered. In addition, the ordered structure

offers the opportunity to spatially separate active centres. As

a result of their high surface areas, MOF-based catalysts

contain a very high density of fully exposed active sites per

volume.52 This characteristic results in enhanced activity, and

thus a more effective catalytic system.

3.4.2 Different approaches towards MOF catalysts.

Approaches towards heterogeneous catalysis with MOFs fall

into three categories. The first, and most straightforward

approach is the use of the material as the carrier for an active

mass. This concept is well known for classical catalysts,

e.g. platinum on carbon. Patents by BASF SE53,54 show that

silver and platinum on MOF-5 are catalytically active. With

Ag@MOF-5 the epoxidation of propylene with molecular

oxygen was achieved. Pt@MOF-5 is active in the synthesis

of hydrogen peroxide from the elements. The group of Fischer

in Bochum, Germany used MOF-5 as a carrier for copper,

palladium and gold.55 Cu@MOF-5 was active in the synthesis

of methanol from synthesis gas. Pd@MOF-5 was used for the

reduction of cyclooctene with hydrogen. Au@MOF-5 proved

to be too labile to do any catalysis with. Remarkable is the

method of preparation employed. The metal precursors were

introduced by chemical vapour deposition (CVD).

The second approach is fixation of catalytically active

species in the framework. This can be done by modifying

known catalysts in a way so that they are capable of forming

MOF frameworks. Knowledge about the catalytic activity

existed before (a priori) the formation of the MOF. Therefore,

this approach can be termed the a priori approach to MOF

catalysis. For that type of catalyst, it is important to set the

results from the catalytic test into the context of the results for

catalysis experiments with the free linker as catalyst. The work

by Hupp et al.60 on manganese salen complexes for epoxidations

illustrates that exemplary. The reaction rate for the epoxidation

was only slightly lower for the MOF catalyst compared to the

homogeneous one. Additionally, there was no deactivation of

the MOF-based catalyst during the reaction, because the

active centres were spatially separated from each other at fixed

positions, limiting the amount of side reactions that could

result in catalyst deactivation. This can be considered the first

unequivocal description of a ‘‘MOF effect’’ in catalysis.

Third, coordination species can be generated in and

stabilised by the framework which are either completely

unknown in the homogeneous complex chemistry of the metal

Table 2 (continued)

MOF type Reaction type Reactants Ref.

Active centre introduced by post-synthesis modification

[Ni(L-aspartate)bpy0.5]�HCl0.9�MeOH0.5 Methanolysis of epoxides Cis-2,3-epoxybutane 87[Cu(L-aspartate)bpy0.5]�HCl[Cu(D-aspartate)bpy0.5]�HCl[Cu(L-aspartate)bpe0.5]�HCl[Cu(D-aspartate)bpe0.5]�HClCr3(F,OH)(en)2O(BDC)3 (ED-MIL-101) Knoevenagel condensation Benzaldehyde þ ethyl cyanoacetate 88Cr3(F,OH)(en)2O(BDC)3 (ED-MIL-101) Heck coupling Iodobenzene þ acrylic acid 88

a Abbr.: BDC ¼ 1,4-benzenedicarboxylate, BTC ¼ 1,3,5-benzenetricarboxylate, bpy ¼ 4,40-bipyridine, phen ¼ phenanthroline, BPDC ¼biphenyldicarboxylate, DHBDC ¼ 2,5-dihydroxy-1,4-benzenedicarboxylate, TCPP ¼ 5,10,15,20-tetrakis-(4-benzoate)-porphyrin, 2-pymo ¼2-hydroxy-pyrimidinolate, PhIM ¼ phenylimidazolate, H3BTT ¼ 1,3,5-benzene-tristetrazol-5-yl, bpe ¼ 1,2-bis(4-pyridyl)ethane, en ¼ ethylene-

diamine.

This journal is �c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 1284–1293 | 1291

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 10: Industrial applications of metal–organic frameworks.

or which are extremely sensitive. This will mean the active

centre is created from scratch or de novo. An example of this

approach is the free metal-site in fully activated Cu3(BTC)2units in Cu-BTC-MOF. If the apical aquo-ligand is removed

from the molecular parent compound Cu2(CH3COO)4�2H2O,

the unsaturated Cu-centre will coordinate an oxygen donor

from a second complex fragment to form a m2-O bridge. In the

MOF, this way of saturation is prohibited by the framework

and the open metal sites are stable and highly reactive.

Another example for the de novo approach is the protonation

of chiral MOF frameworks and their usage as catalysts in the

enantioselective methanolysis of cis-2,3-epoxybutane.87

The classification of MOF catalysts to the a priori and

de novo categories may not always be unequivocal. However,

it shows that for MOF catalysts one also should think outside

the realm of conventional complex chemistry. For example, in

a MOF structure the stabilization of highly reactive species,

like open metal-sites, is much easier than in the homogeneous

phase—see the example of Cu-BTC-MOF above. Due to the

self-organisation of simple (cheap) ligands and metal salts,

structural motifs can be realized that would require a huge

amount of sterically demanding decoration on the ligand

sphere of the transition metal in the homogenous phase. This

concept may also be applicable for other catalytically active

species.

In Table 2, a list of publications on MOFs as catalysts is

given. The classification in Table 2 is according to the nature

of the active centre or the way the catalyst was gained. Recent

publications on catalytic properties of MOFs make also use of

post-synthesis modification. Noteworthy is the modification of

MIL-101 by ethylenediamine to yield a basic catalyst that is

able to catalyze the Knoevenagel condensation of benzaldehyde

and ethyl cyanoacetate with a selectivity of 99.3% towards the

condensation product.88 Since the reaction takes place inside

the pore system, it can be hypothesized that a follow up

Michael addition of benzaldehyde and the large condensation

product is unfavourable. Therefore, the high selectivity is

directly linked to the structural features of the MOF.

Conclusions

It can be stated that MOF materials are of great interest to the

chemical industry. Promising fields of applications are gas

storage, gas purification, separations and catalysis. Gas

storage, gas purification and separation are the most mature

fields of research. Therefore, it is most likely that the first

application will come from one of these fields. However,

research on the catalytic properties of MOFs is gaining

momentum. Due to their unique properties, MOFs are likely

to give new impulses to catalysis research as a whole and may

also be beneficial for existing processes. All in all, as an

emerging class of porous materials, MOFs are being investigated

more and more. Consequently, an increasing number of new

materials are being discovered and novel applications are

being identified. Since there is virtually an infinite number of

possible combinations of linker molecules and metal ions, it

can be expected that academic and industrial research activities

in this field will continue to be vigorous.

References

1 E. A. Tomic, J. Appl. Polym. Sci., 1965, 9, 3745.2 C. Biondi, M. Bonamico, L. Torelli and A. Vaciago, Chem.Commun., 1965, 191.

3 B. F. Hoskins and R. Robson, J. Am. Chem. Soc., 1990, 112,1546.

4 H. Li, M. Eddaoudi, M. O’Keeffe and O. M. Yaghi, Nature, 1999,402, 276.

5 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O’Keefeand O. M. Yaghi, Science, 2002, 295, 469.

6 O. M. Yaghi, M. Eddaoudi, H. Li, J. Kim and N. Rosi, Patent,WO 2002/088148, 2002.

7 H. K. Chae, D. Y. Siberio-Perez, J. Kim, Y. B. Go, M. Eddaoudi,A. J. Matzger, M. O’Keeffe and O. M. Yaghi, Nature, 2004, 427,523.

8 S. Kitagawa, R. Kitaura and S. Noro, Angew. Chem., Int. Ed.,2004, 43, 2334.

9 J. L. C. Rowsell and O. M. Yaghi, Microporous MesoporousMater., 2004, 73, 3.

10 G. Ferey, Chem. Soc. Rev., 2008, 37, 191.11 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 andO. M. Yaghi, Chem. Mater., 1999, 11, 2633.

12 A. G. Wong-Foy, A. J. Matzger and O. M. Yaghi, J. Am. Chem.Soc., 2006, 128, 3494.

13 G. Ferey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour,S. Surble and I. Margiolaki, Science, 2005, 309, 2040.

14 G. Ferey, F. Millange, M. Morcrette, C. Serre, M.-L. Doublet,J.-M. Greneche and J.-M. Tarascon, Angew. Chem., 2007, 119,3323.

15 F. P. Doty, C. A. Bauer, A. J. Skulan, P. G. Grant andM. D. Allendorf, Adv. Mater., 2009, 21, 95.

16 P. Horcajada, C. Serre, M. Vallet-Regı, M. Sebban, F. Taulelle andG. Ferey, Angew. Chem., 2006, 118, 6120.

17 U. Mueller, L. Lobree, M. Hesse, O. Yaghi and M. Eddaoudi,Patent, US 2003222023 A1, 2003.

18 M. Hesse, U. Muller, U. Mueller and O. M. Yaghi, Patent, US2006099398 A1, 2006.

19 M. Schubert, U. Muller and H. Mattenheimer, Patent, WO2007090864 A1, 2007.

20 U. Mueller, H. Puetter, M. Hesse and H. Wessel, Patent, WO2005/049892, 2005.

21 U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndtand J. Pastre, J. Mater. Chem., 2006, 16, 626.

22 U. Mueller, I. Richter and M. Schubert, Patent, WO 2007131955A1, 2007.

23 N. L. Rosi, J. Eckert, M. Eddaoudi, D. T. Vodak, J. Kim,M. O’Keefe and O. M. Yaghi, Science, 2003, 300, 1127.

24 Y. Kubota, M. Takata, R. Matsuda, R. Kitaura, S. Kitagawa,K. Kato, M. Sakata and T. C. Kobayashi, Angew. Chem., 2005,117, 942.

25 T. Yildirim and M. R. Hartman, Phys. Rev. Lett., 2005, 95,215504.

26 T. Mueller and G. Ceder, J. Phys. Chem. B, 2005, 109, 17974.27 U. Mueller, M. Hesse, H. Puetter, M. Schubert and D. Mirsch,

Patent, EP 1,674,555, 2005.28 U. Mueller, M. Hesse, H. Puetter, M. Kamieth, K. Schierle-Arndt

and Kerstin, Patent, WO 2006072573 A2, 2006.29 M. Dinca and J. R. Long, J. Am. Chem. Soc., 2005, 127,

9376.30 T. K. Maji, G. Mostafa, H. Ch. Chang and S. Kitagawa, Chem.

Commun., 2005, 2436.31 L. Pan, B. Parker, X. Huang, D. H. Olson, J. Y. Lee and J. Li,

J. Am. Chem. Soc., 2006, 128, 4180.32 L. Pan, D. H. Olson, L. R. Ciemnolonski, R. Heddy and J. Li,

Angew. Chem., Int. Ed., 2006, 45, 616.33 D. N. Dybtsev, H. Chun, S. H. Yoon, D. Kim and K. Kim, J. Am.

Chem. Soc., 2004, 126, 32.34 B. Arstad, H. Fjellvag, K. O. Kongshaug, O. Swang and R. Blom,

Adsorption, 2008, 14, 755.35 M. Schubert, U. Mueller and C. Kiener, Patent, WO 2008061958

A1, 2008.36 P. L. Llewellyn, S. Bourrelly, C. Serre, Y. Filinchuk and G. Ferey,

Angew. Chem., 2006, 118, 7915.

1292 | Chem. Soc. Rev., 2009, 38, 1284–1293 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online

Page 11: Industrial applications of metal–organic frameworks.

37 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, 238.

38 J. L. C. Rowsell and O. M. Yaghi, Angew. Chem., 2005, 117, 4748.39 G. Ferey, M. Latroche, C. Serre, F. Millange, T. Loiseau and

A. Percheron-Guegan, Chem. Commun., 2003, 2976.40 D. N. Dybtsev, H. Chun and K. Kim, Angew. Chem., 2004, 116,

5143.41 K. Seki and W. Mori, J. Phys. Chem. B, 2002, 106, 1380.42 B. Chen, N. W. Ockwig, A. R. Millward, D. S. Contreras and

O. M. Yaghi, Angew. Chem., 2005, 117, 4823.43 E. Y. Lee, S. Y. Jang and M. P. Suh, J. Am. Chem. Soc., 2005, 127,

6374.44 G. Ferey, C. Serre, C. Mellot-Draznieks, F. Millange, S. Surble,

J. Dutour and I. Margiolaki, Angew. Chem., 2004, 116, 6456.45 C. Serre, F. Millange, S. Surble and G. Ferey, Angew. Chem., 2004,

116, 6446.46 Q. Yang and Ch. Zhong, J. Phys. Chem. B, 2005, 109, 11862.47 S. Barrett, Fuel Cells Bulletin, 2005, 12.48 K. Uemura, R. Matsuda and S. Kitagawa, J. Solid State Chem.,

2005, 178, 2420.49 S. Ma and H.-C. Zhou, J. Am. Chem. Soc., 2006, 128, 11734.50 National Research Council, in Catalysis Looks to the Future,

(Panel on New Directions in Catalytic Science and Technology),National Academy Press, Washington DC, 1992, pp. 1–18.

51 F. Stallmach, S. Groeger, V. Kuenzel, J. Kaerger, O. M. Yaghi,M. Hesse and U. Mueller, Angew. Chem., 2006, 118, 2177.

52 T. Mueller and G. Ceder, J. Phys. Chem. B, 2005, 109, 17974.53 U. Mueller, L. Lobree, M. Hesse, O. M. Yaghi and M. Eddaoudi,

Patent, WO 03/101975, 2003.54 U. Mueller, O. Metelkina, H. Junicke, T. Butz and O. M. Yaghi,

Patent, US 2004/081611, 2004.55 S. Hermes, M.-K. Schroter, R. Schmid, L. Khodeir, M. Muhler,

A. Tissler, R. W. Fischer and R. A. Fischer, Angew. Chem., 2005,117, 6394.

56 C.-D. Wu, A. Hu, L. Zhang and W. Lin, J. Am. Chem. Soc., 2005,127, 8940.

57 C.-D. Wu and W. Lin, Angew. Chem., 2007, 119, 1093.58 W. Mori, T. Sato, T. Ohmura, C. N. Kato and T. Takei, J. Solid

State Chem., 2005, 178, 2555.59 T. Sato, W. Mori, C. N. Kato, E. Yanaoka, T. Kuribayashi,

R. Ohtera and Y. Shiraishi, J. Catal., 2005, 232, 186.60 S.-H. Cho, B. Ma, S. T. Nguyen, J. T. Hupp and T. E. Albrecht-

Schmitt, Chem. Commun., 2006, 2563.61 J. S. Seo, D. Whang, H. Lee, S. I. Jun, J. Oh, Y. J. Jeon and

K. Kim, Nature, 2000, 404, 982.62 O. R. Evans, H. L. Ngo and W. Lin, J. Am. Chem. Soc., 2001, 123,

10395.63 D. Jiang, T. Mallat, F. Krumeich and A. Baiker, J. Catal., 2008,

257, 290.

64 U. Mueller, M. Stoesser, R. Ruppel, E. Baum, E. Bohres, M. Sigl,L. Lobree, O. M. Yaghi and M. Eddaoudi, Patent, WO 03/035717,2003.

65 U. Mueller, G. Luinstra, O. M. Yaghi, O. Metelkina andM. Stoesser, Patent, WO 2004/037895, 2004.

66 E. Guiterrez-Puebla, C. Cascales-Sedano, B. Gomez-Lor,M. M. Iglesias-Hernandez, M. A. Monge-Bravo, C. Ruiz-Valeroand N. Snejko, Patent, ES 2,200,681, 2004.

67 B. Gomez-Lor, E. Guiterrez-Puebla, M. Iglesias, M. A. Monge,C. Ruiz-Valero and N. Snejko, Inorg. Chem., 2002, 41, 2429.

68 S. De Rosa, G. Giordano, T. Granato, A. Katovic, A. Sicilianoand F. Tripicchio, J. Agric. Food Chem., 2005, 53, 8306.

69 J. Perles, M. Iglesias, M.-A. Martın-Luengo, M. A. Monge,C. Ruiz-Valero and N. Snejko, Chem. Mater., 2005, 17, 5837.

70 J. Perles, M. Iglesias, C. Ruiz-Valero and N. Snejko, J. Mater.Chem., 2004, 14, 2683.

71 F. X. Llabres i Xamena, A. Abad, A. Corma and H. Garcia,J. Catal., 2007, 250, 294.

72 C. Nozaki and W. Mori, C. R. Chim., 2007, 10, 284.73 F. X. Llabres i Xamena, O. Casanova, R. Galiasso Tailleur,

H. Garcia and A. Corma, J. Catal., 2008, 255, 220.74 B. Gomez-Lor, E. Guiterrez-Puebla, M. Iglesias, M. A. Monge,

C. Ruiz-Valero and N. Snejko, Chem. Mater., 2005, 17,2568.

75 K. Schlichte, T. Kratzke and S. Kaskel, Microporous MesoporousMater., 2004, 73, 81.

76 M. Fujita, Y. J. Kwon, S. Washizu and K. Ogura, J. Am. Chem.Soc., 1994, 116, 1151.

77 O. Ohmori and M. Fujita, Chem. Commun., 2004, 1586.78 S. Horike, M. Dinca, K. Tamaki and J. Long, J. Am. Chem. Soc.,

2008, 130, 5854.79 U. Mueller, M. Hesse, L. Lobree, M. Hoelzle, J. D. Arndt and

P. Rudolf, Patent, WO 2002/070526, 2002.80 J. M. Tanski and P. T. Wolczanski, Inorg. Chem., 2001, 40,

2026.81 L. Alaerts, E. Seguin, H. Poelman, F. Thibault-Starzyk,

P. A. Jacobs and D. E. De Vos, Chem.–Eur. J., 2006, 12, 7353.82 T. Sato, W. Mori, C. N. Kato, T. Ohmura, T. Sato, K. Yokoyama,

S. Takamizawa and S. Naito, Chem. Lett., 2003, 32, 854.83 S. Naito, T. Tanibe, E. Saito, T. Miyao and W. Mori, Chem. Lett.,

2001, 30, 1178.84 R. Tannenbaum, Chem. Mater., 1994, 6, 550.85 R. Tannenbaum, J. Mol. Catal. A: Chem., 1996, 107, 207.86 T. Ohmura, W. Mori, H. Hiraga, M. Ono and Y. Nishimota,

Chem. Lett., 2003, 32, 468.87 M. Ingleson, J. Perez Barrio, J. Bacsa, C. Dickinson, H. Park and

M. Rosseinsky, Chem. Commun., 2008, 1287.88 Y. Hwang, D.-Y. Hong, J.-S. Chang, S. Jhung, Y.-K. Seo, J. Kim,

A. Vimont, M. Dautri, C. Serre and G. Ferey, Angew. Chem., 2008,120, 4212.

This journal is �c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 1284–1293 | 1293

Dow

nloa

ded

by N

AT

ION

AL

TE

CH

NIC

AL

UN

IVE

RSI

TY

OF

AT

HE

NS

on 2

2 Fe

brua

ry 2

013

Publ

ishe

d on

16

Mar

ch 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B80

4680

H

View Article Online