Metal–Organic Frameworks as Biomimetic Catalysts

advertisement
CHEMCATCHEM
MINIREVIEWS
DOI: 10.1002/cctc.201300493
Metal–Organic Frameworks as Biomimetic Catalysts
Zhi-Yuan Gu, Jihye Park, Aaron Raiff, Zhangwen Wei, and Hong-Cai Zhou*[a]
In this Minireview, we have summarized the recent progress of
biomimetic catalysis in the field of metal-organic frameworks
(MOFs) with a focus on the implantation of biomimetic active
sites into a stable MOF. In addition, the potential of creating
highly selective catalytic pockets and diffusion-favored hier-
archical structures in MOFs has also been explored. Furthermore, we have highlighted the achievements of MOF catalysts
in the applications as mimics of peroxidase, cytochromes P450,
hemoglobin, and photosynthetic systems.
Introduction
Biomimicry has been a longstanding scientific and technological endeavor because the imitation of nature not only brings
about new insights into how nature works, but it also provides
a novel perspective on the design of new materials or devices
with a variety of functions.[1] This biomimetic effort is often
based on synthetic materials with extraordinary properties.
With the advent of carbon nanotubes, graphene sheets, and
quantum dots, there arises an unprecedented opportunity to
obtain biomimetic materials that are more efficient than natural systems by employing design principles that are inspired by
nature, and by taking advantage of the characteristics of synthetic systems.[2] Many attempts have been made to incorporate ideas from nature into catalytic systems to improve their
efficiency; however, an artificial enzyme with noticeable catalytic activity, good durability, and substrate selectivity remains
a daunting challenge.
Metal–organic frameworks (MOFs), also known as porous coordination networks (PCNs), are a new class of hybrid inorganic–organic microporous crystalline materials, which consist of
metal ions with bridging organic linkers; these form a 3 D
structure through coordination bonds.[3] The availability of a variety of building blocks, that is, metal ions and organic linkers,
makes it possible to prepare an infinite number of new MOFs
with diverse structures, topologies, and porosity. Owing to
their fascinating structures and unusual properties such as permanent nanoscale porosity, high surface area, good thermostability, and uniformly-structured cavities, MOFs have great potential for hydrogen storage, gas separation, catalysis, sensing,
and bioimaging.[4] Compared with traditional porous materials
such as zeolites and silica, MOFs have much higher surface
areas and better tunability, thus they provide an excellent platform for advancements in heterogeneous catalysis.[5]
MOFs are ideal candidates for building biomimetic systems,
because their uniform cavities can generate a high density of
[a] Dr. Z.-Y. Gu, J. Park, A. Raiff, Z. Wei, Dr. H.-C. Zhou
Department of Chemistry
Texas A&M University
College Station, Texas 77842 (USA)
E-mail: zhou@mail.chem.tamu.edu
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
biomimetic active centers. In addition, thermally and chemically stable MOFs can be built with a variety of metal clusters.
Furthermore, the channels of MOFs provide confined pockets,
which protect the catalytic centers and enhance substrate specificity. Some feature articles previously reviewed the construction of metal–biomolecule frameworks (MBioFs) and the catalytic activity of MOFs, which were usually not stable in water.[6]
To the best of our knowledge, rarely have publications discussed the principles of designing stable MOFs for the potential application of biomimetic catalysis, especially in an aqueous environment. In this Minireview, we summarize the recent
progress of biomimetic catalysis in the MOF field with a focus
on the implantation of biomimetic active sites into a stable
MOF, and we explore the potential of creating highly selective
catalytic pockets and diffusion-favored hierarchical structures
in MOFs. Also, we highlight the achievements of MOF catalysts
in a diverse range of biomimetic applications (Figure 1).
Figure 1. MOFs with active sites of metalloligands as well as stable metal
clusters to create reaction pockets as heterogeneous biomimetic catalysts.
ChemCatChem 2014, 6, 67 – 75
67
CHEMCATCHEM
MINIREVIEWS
Implantation of Biomimetic Active Sites into
a Stable MOF
As a prerequisite of potentially catalytic materials, the incorporation of active sites into MOFs is necessary. The advantages of
MOFs, such as large surface areas and tunability of their pore
sizes and geometries, make them a promising platform for
achieving efficient catalysis. There are several ways to incorporate active sites into MOFs for catalytic purposes based on the
MOF structures. One attractive approach is to construct MOFs
by using metalloligands. The exploration of the metal nodes in
a MOF as catalytically unsaturated metal centers has also been
proven as an alternative methodology. MOFs are also good
containers for active species such as small molecules, metal
complexes, or even nanoparticles.[7] Comprehensive reviews
have summarized the details of these methods elsewhere.[8]
Here we focus on recent construction of water-stable biomimetic MOF catalysts and their potential applications as mimics of
cytochrome P450 enzymes, hemoglobin and photosynthetic
systems.
The stability of a heterogeneous catalytic system is another
crucial factor. The majority of MOFs are composed of coordination bonds between divalent metal ions and carboxylate ligands, which are often prone to hydrolysis. This greatly limits
the applicability of MOFs in an aqueous medium despite the
fact their functionalized analogues show significant catalytic
ability.[9] Zinc and copper in these MOFs show higher affinities
for water molecules than the carboxylate group, which will inevitably cause the framework to decompose. On the other
hand, metals such as aluminum, iron, chromium, titanium, and
zirconium are more reactive to carboxylate ligands, which renders a series of water-stable MOFs, examples include MIL-47,
MIL-53, MIL-100, MIL-101,[10] MIL-125,[11] and UiO structures.[12]
Many efforts to implant catalytically active sites into these
stable MOFs have been successfully demonstrated. MIL-101
has unsaturated metal centers that have catalytic activity for
cyanosilylation and oxidation reactions directly[13] , and through
encapsulation[14] or post-synthetic modification.[15] An amine
functionalized MIL-125 was shown to be a significant catalyst
for the reduction of CO2 under visible light irradiation.[16] The
stable functionalized/doped UiO structures are highly selective
catalysts for the cross-aldol condensation[17] , and they act as
photocatalysts for hydrogen generation,[18] water oxidation and
CO2 reduction.[19]
The other possible route is to incorporate metalloligands
into the aforementioned stable metal clusters to create new
stable catalytic MOFs. The intention of constructing catalytic
MOFs through the metalloligand approach is to incorporate
molecules that have previously been demonstrated as effective
catalysts in homogeneous reactions, such as metalloporphyrins, metallosalens,[20] and Ti-BINOLate complexes (BINOL = 1,1’bi-2-napthol).[21] Here we focus on stable metalloporphyrin
MOFs. In nature, metalloporphyrins are well known for performing diverse biological functions as cofactors for many
enzyme/protein families, which include peroxidases, cytochromes, hemoglobins, and myoglobins.[22] Mimicking the catalytic performance by utilization of metalloporphyrin in MOFs
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
www.chemcatchem.org
has been a sought-after goal for several decades. The pioneering work was reported by Robson and co-workers in 1994. The
large cavities were obtained, but the framework irreversibly decomposed upon the removal of solvent.[23] Years later, Suslick
and coworkers successfully built zeolite structural analogues
(PIZA series) from cobalt and manganese salts with 5,10,15,20tetrakis(4-carboxyphenyl)porphyrin. Remarkable size selectivity
towards small-molecule sorption (water, methanol, ethanol and
1-butylamine) was obtained from PIZA-1, whereas PIZA-3
showed the capability of oxidative catalysis. However, the
small channels in these materials restricted the catalysis of
larger substrates, which occurred only on the exterior surfaces.[24] Choe and coworkers constructed many 2 D metal–organic sheets with a tetra-connected porphyrin ligand, and paddlewheel coordination clusters. Some were further pillared in
the third dimension by dipyridyl ligands.[25] The active sites of
metal ions in the porphyrin centers usually acted as structural
nodes and were therefore blocked for further use as catalysts.
To overcome this disadvantage, Hupp and coworkers employed Zn paddlewheel nodes to build 3 D MOFs with open
channels. Although paddlewheel clusters are known to be less
stable in humid conditions, they succeeded in building a catalytically active MOF with this node by a mixed-ligands
strategy.[26]
To further improve the design of metalloporphyrin MOFs for
real applications in aqueous conditions, our group utilized
stable zirconium clusters as connectors to create an ultrastable
metalloporphyrin-containing MOF, named PCN-222 (See
Figure 2).[27] In an optimized procedure, ZrCl4 was reacted with
Fe-TCPP (TCPP = 4-tetracarboxyphenyl porphyrin) ligands in
the presence of benzoic acid in DMF under solvothermal conditions. Crystalline PCN-222(Fe) was formed, with superior
properties to those of almost all existing MOFs. PCN-222(Fe)
had a large open channel of 3.7 nm in diameter, one of the
largest reported for mesoporous MOFs. N2 adsorption for PCN222 at T = 77 K showed type IV behavior with a sharp increase
at P/P0 of 0.3, which indicates mesoporosity. The Brunauer–
Emmett–Teller (BET) surface area was 2200 m2 g 1, and the
total pore volume was 1.56 cm3 g 1; these are amongst the
highest values reported for all porous materials that contain
porphyrins. Most importantly, it exhibited extraordinary stability compared to other MOFs. From powder XRD patterns, PCN222(Fe) retained its crystallinity upon immersion in boiling
water, and 2, 4, and 8 m, as well as concentrated hydrochloric
acid solutions for 24 h; it showed no phase transition or framework collapse during these treatments. The N2 adsorption isotherms also remained the same after each of these treatments,
which further confirmed the remarkable stability of PCN-222
(Figure 2 D). To the best of our knowledge, PCN-222(Fe) is one
of the most stable MOFs, as it can undergo treatment with
concentrated hydrochloric acid and retain its structure. The zirconium(IV) ion is responsible for the exceptional stability of
PCN-222(Fe). With its high charge density, the zirconium(IV) ion
polarizes the oxygen atoms of the carboxylate groups to form
strong Zr O bonds with significant covalent character.
Similar strategies were also employed by other groups. Rosseinsky et al. recently synthesized an Al-porphyrin MOF, which
ChemCatChem 2014, 6, 67 – 75
68
CHEMCATCHEM
MINIREVIEWS
www.chemcatchem.org
Figure 3. (A) Rietveld refinement of Al-porphyrin MOFs. (B) The porphyrin
ligand. (C–E) The crystal structure of Al-porphyrin MOF viewed from the
[0 0 1], [1 0 0], and [0 1 0] directions, respectively. (Reprinted with permission
from ref. [28], copyright 2012 John Wiley and Sons).
Figure 2. Crystal structure and porosity of PCN-222(Fe). The metalloporphyrin (A) is connected to four Zr6 clusters (B) to generate a 3 D stable MOF with
large 1 D channels that are 3.7 nm in diameter (C). N2 adsorption isotherms
at T = 77 K show the framework stability of PCN-222(Fe) upon treatment
with room-temperature water, boiling water, 2, 4, 8 m, and concentrated HCl
(D). (Reprinted with permission from ref. [27], copyright 2012 John Wiley
and Sons).
utilized stable aluminum clusters and porphyrin ligands. This
unique MOF contains 1 D Al(OH)O4 chains that support highly
porous porphyrin networks and create a water-stable photoactive framework (Figure 3).[28] The pores range from 6 to 11 in
size with the subtraction of van der Waals radii, and are interconnected by rectangular pores, which run along the [0 0 1]
and [11 0] axes to render a 3 D framework. It is thermally stable
up to T = 350 8C in air. N2 adsorption at T = 77 K reveals a surface area of 1400 m2 g 1 and a micropore volume of
0.625 cm3 g 1, which is in good agreement with the calculated
value of 0.63 cm3 g 1. This ultrastable MOF has provided a valuable opportunity for pushing the limits of using MOFs as biomimetic catalysts.
catalytic efficiency have a strong relationship with the molecular diffusion of the substrate towards the reactive center.
Therefore, the implantation of reaction pockets and hierarchical structures into MOFs will be helpful in improving biomimetic catalysis. The creation of reactive pockets has been highly
developed in the homogeneous catalysis field.[29] Also, the synthesis of hierarchical structures is well-established in the creation of mesoporous materials. The requirements for MOF catalyst design are not only the implementation of active sites, but
also the design of the confinement effect and hierarchical
structures, which will allow for accessible catalytic sites, conformational selectivity and efficient diffusion. In this section, we
focus on the MOF catalyst design to fulfill the aforementioned
features by two promising synthetic strategies: 1) the molecular “bottom-up” approach for designing potential reaction
pockets from metal–organic polyhedra (MOPs) to MOFs and
2) the template-directed construction of MOFs that contain hierarchical structures.
Catalyst Design: Reactive Pockets and
Hierarchical Structures
From reaction pockets to MOFs: The possibilities of a molecular “bottom-up” approach
In addition to the use of stable clusters and catalytically active
sites in MOFs, there are two more important factors for the
biomimetic catalyst design. First, the selectivity towards substrate molecules largely relies on the chemical environment
around the active sites. Second, the reaction kinetics and/or
The finite molecular cages, termed in this Minireview as MOPs,
are constructed through the coordination of metal ions and organic ligands, which act as the building blocks. MOPs, therefore, show the same type of chemical bonding as MOFs. This
allows us to build reactive pockets in MOFs through a bottom-
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
ChemCatChem 2014, 6, 67 – 75
69
CHEMCATCHEM
MINIREVIEWS
up approach. Herein we summarize some previously reported
strategies from our group that we can adopt for designing catalytic MOFs.
Zhou et al. synthesized a series of MOPs with different functionalities by a novel strategy; bridging-ligand-substitution
(Figure 4).[30] Direct reactions of Cu(OAc)2 and various linkers
yielded the corresponding MOPs. Further exploration of the
www.chemcatchem.org
stance, it was shown in 2009 that MOFs could be assembled
from MOPs by our group.[32] The results clearly demonstrated
the construction of MOFs in a stepwise manner with molecular
polyhedra as precursors; this paved the way for developing
framework materials from specifically-designed building units
at a molecular level. The other approach is the connection of
these polyhedra into polyhedra-based networks by designing
new ligands. For example, a series of PCN-6X MOFs synthesized by our group[33] are made from a set of dendritic hexacarboxylate ligands to make a series of isoreticular MOFs, which
contain molecular polyhedral subunits. The results indicate
that as the ligands elongated, the cavities increased in size as
well. This is also supported by the BET surface areas that were
measured for PCN-61, -66, and -68, as 3350, 4000, and
5109 m2 g 1, respectively. PCN-610 (also known as NU-100 with
a surface area of 6143 m2 g 1) demonstrated one of the highest
porosities among all existing MOFs.[33b, 34] The aforementioned
molecular bottom-up strategy has proven effective for MOF
design. Further exploration of functionalized molecular cages
in MOFs for catalysis is highly promising and desirable.
Shaping the MOFs: Template-directed synthesis of hierarchical MOFs
Figure 4. The assembly and functionalization of some metal–organic polyhedra based on ligand exchange.
bridging-ligand-substitution method resulted in a variety of
new MOPs with diverse functionalities. The properties of the
MOPs largely relied on the bridging linkers. This opened up
the possibility of targeted synthesis for desirable MOPs with
catalytic activities. Moreover, the structure of the MOPs were
well adjusted, which indicated a high potential for size-selective and shape-refined cavities as well as promising catalytic
activities.[31]
In addition to the construction of MOPs, it is of great importance to demonstrate the potential of the bottom-up building
of MOFs from molecular polyhedral structures. There are two
approaches to achieve this target. One is by using molecular
polyhedra as direct precursors to synthesize MOFs. For in 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
In addition to the molecular bottom-up methodology, the template-directed assembly approach for the creation of hierarchical MOF structures has received extensive attention. Although
some approaches have shown that morphological control can
be achieved through dilution of precursors,[35] the development of template-directed synthesis to build hierarchical MOFs
provides a more tunable path for a range of applications. In
2008, pioneering work was reported by Qiu and coworkers, in
which they applied a surfactant-templated strategy to afford
hierarchical porosity-adjustable micro- and mesoporous MOFs
by using [Cu3(btc)2(H2O)3] (HKUST-1), in which btc is benzenetricarboxylate.[36] A series of hierarchical porous MOF structures
were prepared under solvothermal conditions with copper(II)
solutions and benzene-1,3,5-tricarboxylate precursors in the
presence of cetyltrimethylammonium bromide (CTAB) as
a structure-directing agent. N2 adsorption–desorption isotherms, small- and wide-angle XRD, and transmission electron
microscopy (TEM) images showed that mesopores up to
30 nm in size were generated from microporous MOFs by
tuning the amounts of CTAB. By a similar strategy, MOF nanospheres with long-range ordered mesopores were synthesized
in binary solvent systems of ionic liquids and supercritical
CO2.[37]
Our group demonstrated a cooperative templating effect by
choosing CTAB as a structure-directing agent and citric acid
(CA) as a chelating reagent, which enabled the rational adjustment of the interaction between MOF precursors and surfactant molecules (Figure 5).[38] Structural analysis based on
powder XRD, TEM, and N2 adsorption–desorption isotherms indicates that mesopores are created in MOFs in which the optimal ratio of CTAB/CA is 2.3. This cooperative template-directed
assembly strategy shows a possible synthetic route that one
ChemCatChem 2014, 6, 67 – 75
70
CHEMCATCHEM
MINIREVIEWS
Figure 5. Cooperative template-directed assembly of mesoporous metal–organic frameworks. (Reprinted with permission from ref. [38], copyright 2012
American Chemical Society).
can adopt to build hierarchical frameworks with controllable
mesoporosity.
Liquid-crystal and hard-template approaches were also employed to create hierarchical MOFs. A new route of liquid-crystal templating to combine microporosity and mesoporosity
was also developed by using Prussian Blue.[39] Recently,
Susumu Kitagawa and coworkers demonstrated an alternative
methodology to spatially control the nucleation site, which
leads to the mesoscopic architecture of MOFs.[40] This method
was based on the shaped sacrificial metal oxide as a metal precursor as well as a hard-template structure-directing agent. A
highly ordered analogous MOF hierarchical structure was obtained by the morphological replacement of the metal oxide.
Although the above methodologies are not specifically designed for catalytic applications, molecular bottom-up design
and synthesis of hierarchical structures can facilitate catalytic
processes, especially for the improvement of selectivity and
substrate diffusion. More efforts should be directed towards
the demonstration of such principles in real MOF catalytic
applications.
Solid State Artificial Enzymes
Cytochrome P450-mimic catalysis
Cytochromes P450, which exist as a large family of cysteinatoheme enzymes, are prevalent in various forms of life (plants,
bacteria, and mammals) and play a key role in the oxidative
transformation of endogenous and exogenous molecules. The
prosthetic group of P450 consists of an iron protoporphyrin IX,
which is protected by polypeptides. These enzymes are potent
oxidants capable of catalyzing many useful organic reactions,
such as the hydroxylation of saturated carbon–hydrogen
bonds, the epoxidation of double bonds, and the oxidation of
aromatics.[41] Cytochromes P450 use molecular oxygen, which
inserts one of its oxygen atoms into a substrate (S), and reduces the second oxygen to a water molecule, by utilizing two
electrons provided by nicotinamide adenine dinucleotide
phosphate (NAD(P)H) from a reductase protein. All cytochrome P450 enzymes are involved in a complex multienzymatic
system, which hampers their industrial application for selective
oxygenation of chemicals. The concept of building crystalline
catalytically-active solids to mimic these natural enzymes has
been around for a long time.[42] Many efforts have been made
in this direction.[23, 24, 25f, 26, 43] Greater efforts have been devoted
to the search for water-stable MOFs. To accurately mimic many
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
www.chemcatchem.org
biological systems, water-stable materials will be required.
Water-stable MOFs, which contain the porphyrin group, are
strongly needed to simulate the cytochrome P450 activity. The
newly-developed PCN-222 from our group is an ideal synthetic
platform for mimicking cytochrome P450 enzymes owing to its
stability, mesoporosity and high density of metalloporphyrin
centers.[27, 28]
To show the high potential of PCN-222 as an artificial
enzyme and simplify the catalytic process, we first tested its
peroxidase-like catalytic activity with a variety of metalloporphyrins for the oxidation of several substrates. In natural systems, peroxidases regulate the concentration of hydrogen peroxide in biofluids. PCN-222 with iron centers exhibited excellent peroxidase-like catalytic activity, whereas other MOFs did
not show significant activities under identical conditions,
which proves the stability of PCN-222 as an efficient, accessible
biomimetic catalyst. An enzyme kinetics study for PCN-222(Fe)
was performed and the kinetic parameters kcat and Km were derived (Figure 6). The kcat value gives the maximum number of
Figure 6. A) Peroxidase-like oxidation reaction of pyrogallol catalyzed by
PCN-222(Fe), in which pyrogallol is oxidized to purpurogallin by hydrogen
peroxide. B) The initial pyrogallol oxidation profile catalyzed by PCN-222(Fe)
(2.5 mm active site equivalent) in which the diverse concentrations of pyrogallol are shown. C) The Lineweaver–Burk plot of the pyrogallol oxidation
catalyzed by PCN-222(Fe) (Reprinted with permission from ref. [27], copyright 2012 John Wiley and Sons).
substrate molecules catalyzed per molecule of catalyst per unit
time. Km is the Michaelis constant and indicates the affinity of
the catalyst molecules for the substrate. For the pyrogallol oxidation reaction, the kcat value of the PCN-222(Fe) catalyst is
16.1 min 1, which is 7 times higher than the value for free
hemin (2.4 min 1). The Km value ( 0.33 mm) is lower than that
of the natural enzyme HRP (horseradish peroxidase, 0.81 mm),
which indicates a better affinity of the substrate to PCN222(Fe); this results from the high porosity of PCN-222(Fe).
Other substrates such as 3,3’,5,5’-tetramethylbenzidine and
o-phenylenediamine have also been tested for peroxidase-like
oxidation to demonstrate the general applicability of PCN222(Fe) as an enzyme mimic. PCN-222(Fe) showed superior catalytic activity over free hemin as its kcat value was nearly
10 times higher than that of free hemin. Other porphyrin-containing systems, such as porphyrin-encapsulated MOFs also
ChemCatChem 2014, 6, 67 – 75
71
CHEMCATCHEM
MINIREVIEWS
show significant catalytic activity for oxidation reactions in organic solvents.[26, 43a, c–e, g, 44]
If one considers the peroxide shunt in the P450 catalytic
cycle and the similar high-valent iron-oxo intermediates shared
by oxygenases and peroxidases, this test not only reflects the
ability of PCN-222 to oxidize the substrates but also shows its
similarity to cytochromes P450 (Scheme 1).[41] However, there is
www.chemcatchem.org
separating CO2 from N2 in postcombustion capture, which is
beneficial to our environment.
Long and coworkers employed unsaturated copper(II) and
iron(II)-based MOFs to systematically investigate their O2-binding activities, which is the main role of hemoglobin in human
blood. An activated Cr3(btc)2 MOF was obtained from heating
and vacuum treatment after the reaction of Cr(CO)6 with trimesic acid in DMF. The BET surface area from N2 gas adsorption measurements was 1810 m2 g 1. At T = 298 K, the O2
uptake for the Cr3(btc)2 MOF rises sharply to 11 wt %, whereas
the corresponding N2 uptake shows a capacity of 0.58 wt % at
P = 1 bar. An excellent O2/N2 selectivity factor of 22 was obtained. Desorption of O2 was achieved by heating at T = 50 8C
under vacuum. Infrared and X-ray absorption spectra as well as
neutron powder diffraction indicated the location of O2 molecules that were near to the chromium(II) unsaturated centers
(Figure 7).[45]
Scheme 1. Representation of the consensus P450 catalytic cycle and peroxide shunt pathway (Reprinted with permission from ref. [41], copyright 2012
American Chemical Society).
still a large discrepancy between the synthetic materials and
natural P450 enzymes, which capture oxygen as oxidants and
protect the iron center to finish the catalytic cycle, not to mention the construction of the complex multienzymatic system.
However, the benefits that MOF catalysts will provide, such as
ultrastability in water and organic solvents, a high density of
reaction centers, and good recyclability, continue to inspire researchers to pursue these goals to finish the catalytic cycle of
cytochromes P450 in MOF systems, which will definitely be
a groundbreaking result.
Hemoglobin-like MOFs
The design of MOFs could also benefit from consideration of
other natural proteins, which perform intriguing activities, such
as selective binding of O2 to hemoglobin, and light-harvesting
pigments in photosystems I and II. Clever usage of active sites
in MOFs shows significant similarity towards such applications,
but in a stable solid state. This solid-state material is beneficial
to the industrial separation of O2 from air, which implies the
potential for using purified O2 combustion as a means of eliminating CO2 emissions from fossil fuel-fired power plants. O2 is
separated from the N2 in air prior to combustion instead of
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Figure 7. Structure of Cr3(btc)2 MOF, in which green, red, and gray spheres
represent Cr, O, and C atoms, respectively, whereas large red spheres represent either bound DMF molecules in the solvated structure or bound O2
molecules upon desolvation and exposure to O2. The atomic positions
shown correspond to the results obtained from Rietveld refinement of
powder neutron diffraction data for activated (B) and O2-loaded (A and C)
samples. (Reprinted with permission from ref. [45], copyright 2010 American
Chemical Society).
Similar unsaturated iron(II) sites that are effective at binding
O2 were also observed in a Fe2(dobdc) MOF (dobdc4 = 2,5-dioxido-1,4-benzenedicarboxylate), which had a BET surface area
of 1360 m2 g 1. It was synthesized from the air-free reaction between FeCl2 and H4dobdc in a mixture of DMF and methanol.[46] At T = 298 K, the Fe2(dobdc) MOF irreversibly binds O2
with a capacity of 9.3 wt % (molar ratio of O2 to iron is 1:2). Interestingly, at a lower temperature of 211 K, reversible adsorption is observed and O2 uptake is 18.2 wt % (molar ratio of O2
to iron is 1:1). This temperature-dependent O2-binding phenomenon was investigated by Mçssbauer and infrared spectroscopy, as well as powder neutron diffraction. The results revealed complete charge transfer to form iron(III) and O22 ions
at room temperature, which rendered a strong affinity and
lower uptake, whereas partial charge transfer was observed
from iron(II) ions to O2 molecules at low temperature, which afChemCatChem 2014, 6, 67 – 75
72
CHEMCATCHEM
MINIREVIEWS
forded reversible adsorption and higher uptake. This interpretation was consistent with the results of Rietveld analyses of
powder neutron diffraction data, which revealed the geometry
of O2 molecules that were bound towards iron centers. These
isolated metal centers together with large open volumes in
the MOFs not only gave the high uptake of desired O2 molecules but also new insights into a solid-state artificial enzyme
mimic.
Solar energy conversion and light-driven catalytic
applications
Natural photosystems I and II gave insights into how nature
harvests solar energy and inspired researchers to solve the
energy crisis with recent developments in photovoltaic cells,
dye-sensitized solar cells, and solar-driven water splitting. Pioneering contributions were made from the MOF field to mimic
natural systems for light harvesting, electron transfer, energy
transfer, and photocatalytic hydrogen evolution. Kent and Lin
investigated efficient MOF luminescence quenching in the
presence of either oxidative or reductive quenchers to demonstrate the energy migration and electron transfer,[47] whereas
other groups highlighted the advantages of the spatial
energy/charge transfer effect in specific MOF structures.[48]
Nanoscale coordination polymers with guest-induced lightharvesting properties were reported from Zhang and coworkers.[48a–c] The combination of linear conjugated dicarboxylate
ligands and lanthanide metal ions in organic solvent resulted
in a series of self-assembled nanoscale coordination polymers
that exhibit long-range ordered structures. Efficient light harvesting with energy transfer either from the framework to the
guest molecule or from the framework to framework was observed. Other than the nanoscale well-dispersed system, studies of MOF solids were also performed. The highly-ordered porphyrin-like pigments (chlorophylls) in natural photosynthesis
inspired us to investigate the analogues of highly-ordered porphyrin-based MOFs, from which we might observe a similar
energy migration behavior. Hupp and coworkers designed and
synthesized antenna-like light-harvesting MOFs for solar
energy conversion. A bodipy- and porphyrin-based MOF was
capable of harvesting light across the entire visible spectrum
by the efficient “strut-to-strut” energy transfer through the ordered MOF structure.[48d] The same group also explored the
first example of long-distance energy migration within a MOF.
The photogenerated exciton in a MOF with enhanced p-conjugation could migrate over a net distance of up to approximately 45 porphyrin struts within its lifetime, whereas in the
other porphyrin MOF it could only migrate over approximately
3 struts.[48e] Not only pure MOFs, but also MOF-quantum dot
hybrid materials were tested. Owing to the spectral coverage
limit of porphyrin MOFs, quantum dots with broad absorption
bands were introduced into MOF pores to enhance light harvesting by energy transfer from quantum dots to MOFs. Eighty
percent efficiency was observed for energy transfer from quantum dots to the MOFs, which greatly enhanced the light harvesting in solar energy conversion.[48f]
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
www.chemcatchem.org
In addition to light harvesting and energy transfer, waterstable MOFs have also contributed to light-driven water splitting. Garca and coworkers[18] applied the water-resistant Zr
MOFs UiO-66 and UiO-66-NH2 to this process. Both of these
MOFs had photocatalytic activities for hydrogen generation
upon irradiation at wavelengths longer than 300 nm. The
amino group generated a bathochromic shift, which is promising for more efficient water splitting. The apparent quantum
yield of 3.5 % was obtained for UiO-66-NH2 at l = 370 nm UV irradiation. Recently, visible-light-driven hydrogen generation
was reported by Rosseinsky and coworkers in a water-stable
aluminum porphyrin MOF (Figure 8).[28] The porous MOF Al-
Figure 8. The photocatalytic reaction of Al-porphyrin(Zn) MOFs under two
different conditions. A) The other reagents are methyl viologen, colloidal
platinum, and sacrificial EDTA. B) The other reagents are colloidal platinum,
and sacrificial EDTA. (Reprinted with permission from ref. [28], copyright
2012 John Wiley and Sons).
PMOF with a BET surface area of 1400 m2 g 1 performed visible-light-driven hydrogen generation from water with the help
of Pt as a co-catalyst and ethylenediaminetetracetic acid
(EDTA) as a sacrificial electron donor. The MOF/EDTA/Pt system
produced H2 at the rate of up to 200 mmol g 1 h 1. A control
experiment in the MOF-filtered solution confirmed that H2 was
produced by heterogeneous photocatalytic activity of the
MOF. It is highly promising to push the envelope of MOF research in mimicking photosynthetic systems and to develop
highly efficient energy and electron transfer solid-state lightharvesting materials.
Merits compared to other biomimetic systems
In the category of solid materials, MOFs offer the highest surface areas, and chemical tunability. Compared with traditional
heterogeneous porous catalysts such as zeolites and silica,
MOFs have greater opportunity for tuning their functions, thus
providing an excellent platform for biomimetic catalysis. The
merits of MOFs as biomimetic catalysts over other solid-state
catalysts and natural enzyme systems[49] are provided by
1) spatially and accurately isolating catalytic centers that typically degrade in solution through undesirable reactions such
as m-oxo heme formation; 2) the provision of protection to
enzyme active sites under extremely harsh conditions, such as
ChemCatChem 2014, 6, 67 – 75
73
CHEMCATCHEM
MINIREVIEWS
in hydrochloric acid, through confinement in mesopores within
ultrastable aluminum(III), iron(III), chromium(III), hafnium(IV),
and zirconium(IV)-constructed MOFs, instead of polypeptide
environments; 3) enhancing the density of porphyrin centers
by incorporating them into the struts of the framework to
ensure sufficient interaction between open metal sites and
substrate molecules; 4) creating MOFs with large pores and
channels, which will enable fast diffusion of guest molecules,
which therefore increases the overall reaction rate; 5) introducing channels with multiple sizes, which allow for tuning of the
reactivity of the catalytic active center through selective axial
ligand binding and for tweaking of substrate selectivity
through size/shape control; and 6) adjusting catalytic reactivity
of metalloporphyrin moieties by systematically varying the
metal ions coordinated in the porphyrin centers.
www.chemcatchem.org
[5]
[6]
[7]
[8]
Outlook
With regards to MOF catalyst design, extremely stable MOFs
with a high loading of active sites, predesigned reaction pockets, as well as hierarchical structures for the heterogeneous
catalysis are highly desirable. The recent progress in the design
of stable metal clusters as connection nodes for MOFs will be
beneficial for stabilizing the MOF catalysts in future developments. More efforts should be put into the design of highly selective catalytic pockets in stable MOFs, as well as into the exploration of hierarchical MOF structures to reduce the diffusion
resistance. For real applications, highly efficient biomimetic
MOF catalysts, especially with regards to solid-state artificial
enzymes, are still underdeveloped despite the significant progresses made in the synthesis of stable MOFs. More fundamental innovations in MOF research will bring biomimetic catalysis
a bright future.
Acknowledgements
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
This work was supported as part of the Center for Gas Separations Relevant to Clean Energy Technologies, an Energy Frontier
Research Center funded by the U.S. Department of Energy (DOE),
Office of Science, Office of Basic Energy Sciences under Award
Number DE-SC0001015. We acknowledge financial support by
the DOE EERE Hydrogen Program under contract number DEFC36-07GO17033.
Keywords: artificial enzyme · biomimetic synthesis ·
heterogeneous catalysis · mesoporous materials · metal–
organic frameworks · microporous materials
[1] S. Mann in Biomimetic Materials Chemistry (Ed.: S. Mann), Wiley-VCH,
Weinheim, 1997.
[2] a) S. Iijima, Nature 1991, 354, 56 – 58; b) D. Zhao, J. Feng, Q. Huo, N.
Melosh, G. H. Fredrickson, B. F. Chmelka, G. D. Stucky, Science 1998, 279,
548 – 552; c) S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen, I. D.
Williams, Science 1999, 283, 1148 – 1150; d) C. Sanchez, H. Arribart, M. M.
Giraud Guille, Nat. Mater. 2005, 4, 277 – 288.
[3] H.-C. Zhou, J. R. Long, O. M. Yaghi, Chem. Rev. 2012, 112, 673 – 674.
[4] a) J.-R. Li, J. Sculley, H.-C. Zhou, Chem. Rev. 2012, 112, 869 – 932; b) Z.-Y.
Gu, C.-X. Yang, N. Chang, X.-P. Yan, Acc. Chem. Res. 2012, 45, 734 – 745;
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
c) Y. Cui, Y. Yue, G. Qian, B. Chen, Chem. Rev. 2012, 112, 1126 – 1162;
d) L. E. Kreno, K. Leong, O. K. Farha, M. Allendorf, R. P. Van Duyne, J. T.
Hupp, Chem. Rev. 2012, 112, 1105 – 1125; e) C. Wang, T. Zhang, W. Lin,
Chem. Rev. 2012, 112, 1084 – 1104; f) H.-L. Jiang, Q. Xu, Chem. Commun.
2011, 47, 3351 – 3370; g) C.-Y. Huang, M. Song, Z.-Y. Gu, H.-F. Wang, X.-P.
Yan, Environ. Sci. Technol. 2011, 45, 4490 – 4496.
a) M. Yoon, R. Srirambalaji, K. Kim, Chem. Rev. 2012, 112, 1196 – 1231;
b) A. Dhakshinamoorthy, H. Garcia, Chem. Soc. Rev. 2012, 41, 5262 –
5284; c) J.-L. Wang, C. Wang, W. Lin, ACS Catal. 2012, 2, 2630 – 2640.
a) K. Lillerud, U. Olsbye, M. Tilset, Top. Catal. 2010, 53, 859 – 868; b) I.
Imaz, M. Rubio-Martinez, J. An, I. Sole-Font, N. L. Rosi, D. Maspoch,
Chem. Commun. 2011, 47, 7287 – 7302.
J. Juan-AlcaÇiz, J. Gascon, F. Kapteijn, J. Mater. Chem. 2012, 22, 10102 –
10118.
a) D. Farrusseng, S. Aguado, C. Pinel, Angew. Chem. 2009, 121, 7638 –
7649; Angew. Chem. Int. Ed. 2009, 48, 7502 – 7513; b) J. Lee, O. K. Farha,
J. Roberts, K. A. Scheidt, S. T. Nguyen, J. T. Hupp, Chem. Soc. Rev. 2009,
38, 1450 – 1459; c) L. Ma, C. Abney, W. Lin, Chem. Soc. Rev. 2009, 38,
1248 – 1256; d) A. Corma, H. Garca, F. X. Llabrs i Xamena, Chem. Rev.
2010, 110, 4606 – 4655; e) M. Ranocchiari, J. A. van Bokhoven, Phys.
Chem. Chem. Phys. 2011, 13, 6388 – 6396; f) C. Wang, M. Zheng, W. Lin,
J. Phys. Chem. Lett. 2011, 2, 1701 – 1709.
J. Gascon, U. Aktay, M. D. Hernandez-Alonso, G. P. M. van Klink, F. Kapteijn, J. Catal. 2009, 261, 75 – 87.
G. Frey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surbl, I.
Margiolaki, Science 2005, 309, 2040 – 2042.
M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin, C. Sanchez, G. Frey,
J. Am. Chem. Soc. 2009, 131, 10857 – 10859.
J. H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga,
K. P. Lillerud, J. Am. Chem. Soc. 2008, 130, 13850 – 13851.
a) A. Henschel, K. Gedrich, R. Kraehnert, S. Kaskel, Chem. Commun.
2008, 4192 – 4194; b) J. Kim, S. Bhattacharjee, K.-E. Jeong, S.-Y. Jeong,
W.-S. Ahn, Chem. Commun. 2009, 3904 – 3906.
a) Y. K. Hwang, D.-Y. Hong, J.-S. Chang, S. H. Jhung, Y.-K. Seo, J. Kim, A.
Vimont, M. Daturi, C. Serre, G. Frey, Angew. Chem. 2008, 120, 4212 –
4216; Angew. Chem. Int. Ed. 2008, 47, 4144 – 4148; b) Y. Huang, Z. Lin, R.
Cao, Chem. Eur. J. 2011, 17, 12706 – 12712; c) E. Kockrick, T. Lescouet,
E. V. Kudrik, A. B. Sorokin, D. Farrusseng, Chem. Commun. 2011, 47,
1562 – 1564; d) C. M. Granadeiro, A. D. S. Barbosa, P. Silva, F. A. A. Paz,
V. K. Saini, J. Pires, B. de Castro, S. S. Balula, L. Cunha-Silva, Appl. Catal. A
2013, 453, 316 – 326; e) O. V. Zalomaeva, K. A. Kovalenko, Y. A. Chesalov,
M. S. Mel’gunov, V. I. Zaikovskii, V. V. Kaichev, A. B. Sorokin, O. A. Kholdeeva, V. P. Fedin, Dalton Trans. 2011, 40, 1441 – 1444.
M. Banerjee, S. Das, M. Yoon, H. J. Choi, M. H. Hyun, S. M. Park, G. Seo,
K. Kim, J. Am. Chem. Soc. 2009, 131, 7524 – 7525.
Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu, Z. Li, Angew. Chem.
2012, 124, 3420 – 3423; Angew. Chem. Int. Ed. 2012, 51, 3364 – 3367.
F. Vermoortele, R. Ameloot, A. Vimont, C. Serre, D. De Vos, Chem.
Commun. 2011, 47, 1521 – 1523.
C. Gomes Silva, I. Luz, F. X. Llabrs i Xamena, A. Corma, H. Garca, Chem.
Eur. J. 2010, 16, 11133 – 11138.
C. Wang, Z. Xie, K. E. deKrafft, W. Lin, J. Am. Chem. Soc. 2011, 133,
13445 – 13454.
a) S.-H. Cho, B. Ma, S. T. Nguyen, J. T. Hupp, T. E. Albrecht-Schmitt, Chem.
Commun. 2006, 2563 – 2565; b) F. Song, C. Wang, J. M. Falkowski, L. Ma,
W. Lin, J. Am. Chem. Soc. 2010, 132, 15390 – 15398.
a) C.-D. Wu, A. Hu, L. Zhang, W. Lin, J. Am. Chem. Soc. 2005, 127, 8940 –
8941; b) L. Ma, J. M. Falkowski, C. Abney, W. Lin, Nat. Chem. 2010, 2,
838 – 846.
a) B. Meunier, S. P. de Visser, S. Shaik, Chem. Rev. 2004, 104, 3947 – 3980;
b) S. Shaik, S. Cohen, Y. Wang, H. Chen, D. Kumar, W. Thiel, Chem. Rev.
2010, 110, 949 – 1017.
B. F. Abrahams, B. F. Hoskins, D. M. Michail, R. Robson, Nature 1994, 369,
727 – 729.
a) M. E. Kosal, J.-H. Chou, S. R. Wilson, K. S. Suslick, Nat. Mater. 2002, 1,
118 – 121; b) K. S. Suslick, P. Bhyrappa, J. H. Chou, M. E. Kosal, S. Nakagaki, D. W. Smithenry, S. R. Wilson, Acc. Chem. Res. 2005, 38, 283 – 291.
a) E.-Y. Choi, P. M. Barron, R. W. Novotny, H.-T. Son, C. Hu, W. Choe, Inorg.
Chem. 2009, 48, 426 – 428; b) E.-Y. Choi, P. M. Barron, R. W. Novotney, C.
Hu, Y.-U. K. Kwon, W. Choe, CrystEngComm 2008, 10, 824 – 826; c) E.-Y.
Choi, L. D. DeVries, R. W. Novotny, C. Hu, W. Choe, Cryst. Growth Des.
ChemCatChem 2014, 6, 67 – 75
74
CHEMCATCHEM
MINIREVIEWS
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
2010, 10, 171 – 176; d) E.-Y. Choi, C. A. Wray, C. Hu, W. Choe, CrystEngComm 2009, 11, 553 – 555; e) J. M. Verduzco, H. Chung, C. Hu, W. Choe,
Inorg. Chem. 2009, 48, 9060 – 9062; f) P. M. Barron, C. A. Wray, C. H. Hu,
Z. Y. Guo, W. Choe, Inorg. Chem. 2010, 49, 10217 – 10219.
a) A. M. Shultz, O. K. Farha, J. T. Hupp, S. T. Nguyen, J. Am. Chem. Soc.
2009, 131, 4204 – 4205; b) O. K. Farha, A. M. Shultz, A. A. Sarjeant, S. T.
Nguyen, J. T. Hupp, J. Am. Chem. Soc. 2011, 133, 5652 – 5655.
D. Feng, Z.-Y. Gu, J.-R. Li, H.-L. Jiang, Z. Wei, H.-C. Zhou, Angew. Chem.
2012, 124, 10453 – 10456; Angew. Chem. Int. Ed. 2012, 51, 10307 – 10310.
A. Fateeva, P. A. Chater, C. P. Ireland, A. A. Tahir, Y. Z. Khimyak, P. V.
Wiper, J. R. Darwent, M. J. Rosseinsky, Angew. Chem. 2012, 124, 7558 –
7562; Angew. Chem. Int. Ed. 2012, 51, 7440 – 7444.
a) M. J. Wiester, P. A. Ulmann, C. A. Mirkin, Angew. Chem. 2011, 123,
118 – 142; Angew. Chem. Int. Ed. 2011, 50, 114 – 137; b) M. Hatano, K. Ishihara, Chem. Commun. 2012, 48, 4273 – 4283.
J.-R. Li, H.-C. Zhou, Nat. Chem. 2010, 2, 893 – 898.
W. Lu, D. Yuan, A. Yakovenko, H.-C. Zhou, Chem. Commun. 2011, 47,
4968 – 4970.
J.-R. Li, D. J. Timmons, H.-C. Zhou, J. Am. Chem. Soc. 2009, 131, 6368 –
6369.
a) D. Zhao, D. Yuan, D. Sun, H.-C. Zhou, J. Am. Chem. Soc. 2009, 131,
9186 – 9188; b) D. Yuan, D. Zhao, D. Sun, H.-C. Zhou, Angew. Chem.
2010, 122, 5485 – 5489; Angew. Chem. Int. Ed. 2010, 49, 5357 – 5361.
O. K. Farha, A. zgr Yazaydın, I. Eryazici, C. D. Malliakas, B. G. Hauser,
M. G. Kanatzidis, S. T. Nguyen, R. Q. Snurr, J. T. Hupp, Nat. Chem. 2010, 2,
944 – 948.
Z. Xin, J. Bai, Y. Pan, M. J. Zaworotko, Chem. Eur. J. 2010, 16, 13049 –
13052.
L.-G. Qiu, T. Xu, Z.-Q. Li, W. Wang, Y. Wu, X. Jiang, X.-Y. Tian, L.-D. Zhang,
Angew. Chem. 2008, 120, 9629 – 9633; Angew. Chem. Int. Ed. 2008, 47,
9487 – 9491.
Y. Zhao, J. Zhang, B. Han, J. Song, J. Li, Q. Wang, Angew. Chem. 2011,
123, 662 – 665; Angew. Chem. Int. Ed. 2011, 50, 636 – 639.
L.-B. Sun, J.-R. Li, J. Park, H.-C. Zhou, J. Am. Chem. Soc. 2012, 134, 126 –
129.
X. Roy, M. J. MacLachlan, Chem. Eur. J. 2009, 15, 6552 – 6559.
J. Reboul, S. Furukawa, N. Horike, M. Tsotsalas, K. Hirai, H. Uehara, M.
Kondo, N. Louvain, O. Sakata, S. Kitagawa, Nat. Mater. 2012, 11, 717 –
723.
R. Fasan, ACS Catal. 2012, 2, 647 – 666.
J. T. Hupp, Nat. Chem. 2010, 2, 432 – 433.
2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
www.chemcatchem.org
[43] a) S. Roy, C. B. George, M. A. Ratner, J. Phys. Chem. C 2012, 116, 23494 –
23502; b) W. Morris, B. Volosskiy, S. Demir, F. Gndara, P. L. McGrier, H.
Furukawa, D. Cascio, J. F. Stoddart, O. M. Yaghi, Inorg. Chem. 2012, 51,
6443 – 6445; c) Z. Zhang, L. Zhang, L. Wojtas, P. Nugent, M. Eddaoudi,
M. J. Zaworotko, J. Am. Chem. Soc. 2012, 134, 924 – 927; d) Z. Zhang, L.
Zhang, L. Wojtas, M. Eddaoudi, M. J. Zaworotko, J. Am. Chem. Soc. 2012,
134, 928 – 933; e) R. W. Larsen, L. Wojtas, J. Perman, R. L. Musselman,
M. J. Zaworotko, C. M. Vetromile, J. Am. Chem. Soc. 2011, 133, 10356 –
10359; f) R. W. Larsen, J. Miksovska, R. L. Musselman, L. Wojtas, J. Phys.
Chem. A 2011, 115, 11519 – 11524; g) M. H. Alkordi, Y. Liu, R. W. Larsen,
J. F. Eubank, M. Eddaoudi, J. Am. Chem. Soc. 2008, 130, 12639 – 12641.
[44] C. Zou, Z. Zhang, X. Xu, Q. Gong, J. Li, C.-D. Wu, J. Am. Chem. Soc. 2012,
134, 87 – 90.
[45] L. J. Murray, M. Dinca, J. Yano, S. Chavan, S. Bordiga, C. M. Brown, J. R.
Long, J. Am. Chem. Soc. 2010, 132, 7856 – 7857.
[46] E. D. Bloch, L. J. Murray, W. L. Queen, S. Chavan, S. N. Maximoff, J. P. Bigi,
R. Krishna, V. K. Peterson, F. Grandjean, G. J. Long, B. Smit, S. Bordiga,
C. M. Brown, J. R. Long, J. Am. Chem. Soc. 2011, 133, 14814 – 14822.
[47] a) C. A. Kent, B. P. Mehl, L. Ma, J. M. Papanikolas, T. J. Meyer, W. Lin, J.
Am. Chem. Soc. 2010, 132, 12767 – 12769; b) C. A. Kent, D. Liu, L. Ma,
J. M. Papanikolas, T. J. Meyer, W. Lin, J. Am. Chem. Soc. 2011, 133,
12940 – 12943; c) C. A. Kent, D. Liu, T. J. Meyer, W. Lin, J. Am. Chem. Soc.
2012, 134, 3991 – 3994.
[48] a) X. Zhang, Z.-K. Chen, K. P. Loh, J. Am. Chem. Soc. 2009, 131, 7210 –
7211; b) X. Zhang, M. A. Ballem, M. Ahrn, A. Suska, P. Bergman, K.
Uvdal, J. Am. Chem. Soc. 2010, 132, 10391 – 10397; c) X. Zhang, M. A.
Ballem, Z.-J. Hu, P. Bergman, K. Uvdal, Angew. Chem. 2011, 123, 5847 –
5851; Angew. Chem. Int. Ed. 2011, 50, 5729 – 5733; d) C. Y. Lee, O. K.
Farha, B. J. Hong, A. A. Sarjeant, S. T. Nguyen, J. T. Hupp, J. Am. Chem.
Soc. 2011, 133, 15858 – 15861; e) H.-J. Son, S. Jin, S. Patwardhan, S. J.
Wezenberg, N. C. Jeong, M. So, C. E. Wilmer, A. A. Sarjeant, G. C. Schatz,
R. Q. Snurr, O. K. Farha, G. P. Wiederrecht, J. T. Hupp, J. Am. Chem. Soc.
2013, 135, 862 – 869; f) S. Jin, H.-J. Son, O. K. Farha, G. P. Wiederrecht,
J. T. Hupp, J. Am. Chem. Soc. 2013, 135, 955 – 958.
[49] a) R. A. Sheldon, Curr. Opin. Solid State Mater. Sci. 1996, 1, 101 – 106;
b) V. Artero, M. Fontecave, Chem. Soc. Rev. 2013, 42, 2338 – 2356.
Received: June 25, 2013
Revised: August 13, 2013
Published online on October 18, 2013
ChemCatChem 2014, 6, 67 – 75
75
Download