Metal-Organic Framework Composites For Catalysis: Review

Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Review

Metal-Organic Framework
Composites for Catalysis
Liyu Chen1 and Qiang Xu1,2,*

Metal-organic frameworks (MOFs) have emerged as a promising class of mate- Progress and Potential
rials with several unique properties, such as high porosity, diverse composition, Catalysis is of vital importance in
tunable pore structures, and versatile functionality. These characteristics enable many aspects of the modern
MOFs to show potential in the field of heterogeneous catalysis. To satisfy the industrialized society. The
practical applications of MOFs, controllable integration of MOFs and functional efficiency of catalysts is highly
materials (e.g., metal nanoparticles, quantum dots, polyoxometalates, molecu- related to the synergistic work
lar species, enzymes, silica, and polymers) can enhance the characteristics of between active sites, assistant
MOFs through activity improvement and framework stabilization. In MOF com- sites, and microenvironments.
posites/hybrids, functional materials can cooperatively work with MOFs to show Metal-organic frameworks (MOFs)
enhanced catalytic activity, selectivity, and stability in a variety of chemical have been demonstrated to serve
transformations. This review provides an overview of the significant advances as a useful platform for the design
in the development of diverse MOF composites/hybrids with special emphases of active heterogeneous catalysts.
on the preparation and catalytic applications. The controlled integration of
MOFs and functional materials
Introduction can allow each component to
Metal-organic frameworks (MOFs), also known as porous coordination polymers work cooperatively for synergistic
(PCPs), constructed by inorganic nodes (metal ions/clusters) with organic linkers, catalysis, thus greatly enhancing
have emerged as a promising class of materials with several unique properties, the efficiency and stability of
such as high porosity, diverse composition, tunable pore structure, and versatile MOF-based catalysts in a broad
functionality.1,2 These characteristics have attracted significant research interest in range of catalytic processes. Here,
a variety of fields,3–8 especially in catalysis.9–11 However, the employment of pristine we review research progress in the
MOFs as catalysts is restricted by their low activity due to the limited active sites, and preparation and catalytic
low mechanical, thermal, and chemical stability due to the weakness of the metal- applications of functional MOF
linker coordinative bond.12,13 In order to satisfy realistic catalytic applications of composites/hybrids.
MOFs, three main approaches have been proposed to further enhance the catalytic
properties,14–21 through functionalizing the MOF backbone (inorganic nodes or
organic linkers),22–26 integrating MOFs with functional materials to MOF compos-
ites,27–29 and converting MOFs to functional materials.30–34 Here, we focus on the
discussion of integration of MOFs and a variety of functional materials to create
new multifunctional MOF composites/hybrids to combine the merits and mitigate
the shortcomings of both the components.

To date, a variety of functional materials (e.g., metal nanoparticles [NPs], quantum


dots [QDs], polyoxometalates [POMs], molecular species, enzymes, silica, and poly-
mers) have been integrated with MOFs to create MOF composites/hybrids
composed of one MOF and one or more functional materials (Figure 1). In MOF com-
posites/hybrids, the advantages of both MOFs (porous structure, chemical versa-
tility, and structural tailorability) and various kinds of functional materials (unique cat-
alytic, optical, electrical, magnetic properties, and mechanical strength) can be
combined effectively; moreover, new physical and chemical properties can be eli-
cited by the synergistic effect.27,28 Consequently, the remarkable features of MOF
composites make them suitable for catalytic applications with enhanced activity,

Matter 1, 57–89, July 10, 2019 ª 2019 Elsevier Inc. 57


Figure 1. The Integration of MOFs and Various Functional Materials through Encapsulating Functional Materials in the Pores, Matrices, or Layers of
MOFs into Functional Material@MOF Composites or Encapsulating/Coating MOFs in/with Functional Supports/Layers to MOF@Functional Material
Composites

selectivity, and stability. A timeline showing some of the breakthroughs in the study
of MOF composites for catalytic applications is shown in Figure 2.

Integrating MOFs with functional materials can be achieved through (1) encapsu-
lating guest functional materials in the pores, matrices, or layers of MOFs, or (2)
encapsulating/coating MOFs in/with functional supports/layers. In the former
case, MOFs act as porous supports to accommodate functional materials (such as
metal NPs, QDs, POMs), preventing the leaching and aggregation of functional ma-
terials while allowing the free diffusion of substrates and products. In the latter case,
functional materials (such as silica, carbon, and polymers) can act as shelters to
enhance the chemical stability and mechanical strength of MOFs and facilitate cata-
lyst formulation for catalytic applications.35,36 For the successful fabrication of MOF-
functional material composites/hybrids, the nature of the interface is of vital impor- 1AIST-Kyoto University Chemical Energy
tance to control the assembly process (involving nucleation, growth, and orientation) Materials Open Innovation Laboratory
and to trigger synergistic effects for superior catalytic performance.37–39 (ChEM-OIL), National Institute of Advanced
Industrial Science and Technology (AIST),
Yoshida, Sakyo-ku, Kyoto 606-8501, Japan
This paper reviews the recent significant progress in the development of MOF com- 2Schoolof Chemistry and Chemical Engineering,
posites/hybrids, classified on the basis of the kinds of composites/hybrids. The com- Yangzhou University, Yangzhou 225009, China
ponents, structures, and properties of different MOF composites/hybrids are dis- *Correspondence: [email protected]
cussed, with particular emphasis on the preparation and synergistic effects of the https://doi.org/10.1016/j.matt.2019.05.018

58 Matter 1, 57–89, July 10, 2019


Figure 2. Timeline of Important Breakthroughs in the Catalytic Applications of MOF Composites/Hybrids
MOF, metal-organic framework; POM, polyoxometalate; bpy, 2,2 0 -bipyridine; RGO, reduced graphene oxide; ORR, oxygen reduction reaction; IL, ionic
liquid; MOP, metal-organic polyhedral.

composites/hybrids on their catalytic performance. Finally, we present the outlook


and perspective for MOF composites/hybrids in catalytic applications. We sincerely
hope that this review will inspire the interest and enthusiasm of chemists in the inves-
tigation of MOF composites/hybrids for catalysis.

MOF-Metal Nanoparticle Composites


Metal NPs have attracted a variety of attention, due to their high chemical activities
and specificities. However, metal NPs have a high surface area to volume ratio with
high surface energies, and thus tend to aggregate and fuse. Assembling metal NPs
inside porous materials, for example, metal oxides, zeolites, mesoporous silicates,
and carbon, can effectively limit the aggregation of metal NPs in confined cavities.
As a new class of porous materials, MOFs with large surface areas and porosity
are suitable as supports for metal NPs.40–46

Methodologies
For the preparation of MOF-metal composites, there are three well-developed
methodologies, namely, ship-in-bottle, bottle-around-ship, and one-pot synthesis
(Figure 3A). These effective methodologies have enabled the fabrication of mono-
metallic NPs,47 bimetallic metal alloy NPs,48,49 bimetallic core-shell NPs,50,51 and
polyhedral metal nanocrystals52,53 with controlled locations, compositions, and
shapes within MOFs.

The ship-in-bottle strategy involves the introduction of metal precursors to an MOF


using various techniques such as chemical vapor deposition, solution impregnation,
solid grinding, microwave irradiation, followed by reduction of metal precursors to
form metal NPs.47,58 However, due to the inherent microstructure of MOFs, it is diffi-
cult to fully introduce metal precursors into the pores of MOFs, resulting in the depo-
sition of metal NPs on the external surface of MOFs with low stability. To circumvent
the challenge, we developed a ‘‘double solvents’’ method (DSM) to successfully
encapsulate metal NPs inside the pores of an MOF, MIL-101 (Cr3F(H2O)2O[(O2C)

Matter 1, 57–89, July 10, 2019 59


Figure 3. Preparation of MOF-Metal NP Composites
(A) Strategies for fabrication of MOF-metal NP composites including ship-in-bottle, bottle-around-ship, and one-pot synthesis. Cyan, inorganic nodes
of MOFs; gray, organic linkers; orange, metal precursors or NPs; red, stabilizing agents. Adapted with permission from Chen et al.54 Copyright 2014, the
Royal Society of Chemistry.
(B) Schematic representation of immobilization of the AuNi nanoparticles by an MIL-101 matrix using DSM combined with a liquid-phase CCR strategy.
Reprinted with permission from Zhu et al. 55 Copyright 2013, American Chemical Society.
(C) Schematic illustration for the synthesis of the Cu 2 O@ZIF-8 composite. Reprinted with permission from Li et al. 56 Copyright 2018, WILEY-VCH Verlag
GmbH & Co.
(D) Incorporation of Pt NPs in a UiO-66 MOF by means of an in situ one-step protocol with kinetic modulation by H2 /acetic acid (top) and with no
modulation (bottom). Adapted with permission from Liu et al. 57 Copyright 2016, WILEY-VCH Verlag GmbH & Co.

C6H4(CO2)]3,nH2O, where n is 25), without aggregation of metal NPs on the


external surfaces of the framework.59 The DSM is based on a hydrophilic solvent
(water) with a volume set equal to or less than the pore volume of the adsorbent
(MIL-101) to dissolve metal precursors (H2PtCl6), and a hydrophobic solvent (hexane)
in large amount to suspend the adsorbent and facilitate the impregnation process.
This method could facilitate all the aqueous H2PtCl6 to go inside the hydrophilic
pores of MIL-101 by capillary force, greatly minimizing the deposition of H2PtCl6
on the outer surface. After treating in a stream of H2/He at 200 C, ultrafine Pt NPs
with an average size of 1.8 G 0.2 nm were uniformly distributed in the interior cav-
ities of MIL-101 crystals, as demonstrated by electron tomographic reconstruction.
For the fabrication of non-noble metal NPs in MOFs, thermal H2 reduction of non-no-
ble metal precursors needs a higher temperature, which may cause the collapse of
MOFs. To address this challenge, we further developed a liquid-phase concentra-
tion-controlled reduction (CCR) strategy combined with the DSM to encapsulate
AuNi alloy NPs in MIL-101 with size and location control (Figure 3B).55 HAuCl4 and
NiCl2 were first impregnated in MIL-101 via the DSM. In the reduction step, when
an overwhelming reduction (OWR) approach with a high concentration of sodium

60 Matter 1, 57–89, July 10, 2019


borohydride (NaBH4) solution was used, ultrafine AuNi NPs (1.8 nm) were encapsu-
lated into the MIL-101 pores without aggregation on the external surface. However,
when a moderate reduction approach was chosen, serious agglomeration of the
AuNi NPs was observed. The proposed OWR approach could completely reduce
the encapsulated metal precursors in the pores to avoid the redissolution and diffu-
sion of metal precursors out of the pores, and thus avoid the aggregation of metal
NPs on the external surface.

The bottle-around-ship strategy refers to the assembly of MOFs around the metal
NPs. The key to acquiring the metal@MOF core-shell structure is to avoid aggrega-
tion of metal NPs and self-nucleation of MOF shells. The introduction of surfactant
and/or ligand can not only stabilize metal NPs but also counterbalance the interfacial
energy between metal NPs and MOFs thus facilitate the overgrowth of MOF shell on
metal NPs. For example, Huo and co-workers employed polyvinylpyrrolidone (PVP)
as a capping agent to promote the crystallization of ZIF-8 on the surface of a series of
metal NPs (e.g., Au, Pt).60 In addition, by adjusting the time of addition of metal NPs
during the assembly of MOF, they could control the spatial distribution of metal NPs
within ZIF-8 crystals. However, the surfactants are very difficult to remove from the
surface of metal NPs after the fabrication of metal@MOF composites, which would
hinder exposure of metal active sites for catalysis. In this regard, a hard template
method using active metal oxides or inert silica as template is an alternate way for
the synthesis of MOF-encapsulated metal NPs with a capping-agent-free surface.
Active metal oxides (e.g., Cu2O,61 Al2O362) have been used as hard templates to
coat metal NPs for the synthesis of an MOF shell. The active metal oxide shell can
be etched off in the solution of organic linkers, and then the dissolved metal ions
will coordinate with organic linkers to produce MOFs around metal NPs. In addition,
inert SiO2 have been used as hard templates to synthesize yolk-shell metal@MOF
structures.56,63 The typical synthetic process involves a multistep approach: (1) first
is the coating of metal NPs with a layer of SiO2, followed by (2) the growth of MOF
onto the surface of SiO2 to form a sandwich nanostructure, then (3) selective removal
of the SiO2 layer in NaOH solution to afford yolk-shell metal@MOF structures
(Figure 3C).

The one-pot strategy, through direct mixing of the metal precursors and MOF pre-
cursors in one pot, has attracted a great deal of attention recently due to the reduced
production cost, shortened production time, and ease of scaling up.64 It is important
to balance the nucleation and growth of the metal NPs and MOFs together with the
assembly of MOFs around metal NPs.65,66 Recently, Li and co-workers developed a
kinetically modulated one-step strategy to construct Pt@UiO-66 core-shell compos-
ites through delicately controlling the assembly process by modulators and solvent
(Figure 3D).57 They employed H2 coupled with dimethylformamide (DMF) to accel-
erate the rate of Pt ion reduction and acetic acid to slow down the rate of UiO-66 for-
mation. Moreover, the DMF solvent could serve as a ‘‘bridge’’ to stabilize the gener-
ated Pt NPs by its soft C–N group and to anchor Zr4+ by its hard C=O group.
Therefore, the preferential anisotropic growth of the UiO-66 on the Pt surface rather
than self-nucleation could be achieved.

Catalytic Performances of MOF-Metal NP Composites


The early reports of MOF-metal NP composites have shown activity in liquid phase
reduction67 and gas phase oxidation catalytic reactions.68 In general, when applying
MOF-metal NP composites in a catalytic reaction, metal NPs act as active centers for
catalytic transformation, and MOFs act as a stabilizer to protect metal NPs with
enhanced stability54,69 and a regulator to tune the electron environment of metal

Matter 1, 57–89, July 10, 2019 61


NPs with enhanced activity.70 Furthermore, MOFs can participate in the reaction by
stabilizing transition states, organizing molecules, or introducing additional active
sites. In the following section, we focus on the cooperative catalysis between metal
NPs and MOFs.

The Nanostructures of MOFs Cooperating with Metal NPs for Catalysis. The micro-
environment of the MOF pore plays an important role in the adsorption, pre-activa-
tion, and orientation of substrates. The pore/channel wall of MOFs can be systemat-
ically functionalized to tune the surface wettability for accumulation of target
substrates with a higher local concentration around the active metal NPs for
enhanced activity.71 In addition, the confinement of the MOF pores can alter the en-
ergetic situation of the reactants to influence their reactivity. For example, when
confining ammonia borane (NH3BH3, AB) in Pd@MIL-101, we found that the AB
dehydrogenation temperature was significantly reduced and the byproducts were
depressed.59 The dehydrogenation of AB confined in Pd@MIL-101 showed the first
and second H2 evolution peaks at about 88 C and 130 C, respectively, with no NH3
or borazine byproducts produced. Without the Pd catalyst, AB@MIL-101 showed
higher AB decomposition temperature with a broad peak centered at 95 C, and
the evolution of ammonia was not completely suppressed. Pristine AB required
even higher decomposition temperature, with the first and second H2 evolution
peaks at 120 C and 160 C, respectively, accompanied by a large evolution of bor-
azine and NH3. This work clearly demonstrates the synergistic effect of Pt NPs catal-
ysis and nanoconfinement of the MIL-101 framework.

Moreover, the pore/channel of MOFs can geometrically restrict the rotation of the
substrate molecules around the active metal NPs. In this regard, functional groups
with comparable activity but different steric resistance in a molecule may have
different accessibility to interact with the active metal sites in the pore/channel of
MOFs, thus showing different reactivity. For example, Huo and co-workers encapsu-
lated Pt NPs in two MOFs with different pore sizes (3.4 Å for ZIF-8, 6 Å for UiO-66) to
study the site-selective oxidation of diols and reduction of alkadienes (Figure 4A).72
Pt-ZIF-8 selectively transformed the primary groups with the second groups un-
touched. For comparison, Pt-UiO-66 with a larger aperture size only showed
72.9% selectivity for primary groups, and Pt NPs deposited on a carbon nanotube
(Pt-CNT) simultaneously catalyzed two groups without selectivity. The exhibited
site selectivity of ZIF-8 may be due to its limited pore size, which restricts the rotation
of the secondary group to contact with the Pt surface.

The Inorganic Nodes of MOFs Cooperating with Metal NPs for Catalysis. The inor-
ganic nodes of MOFs can serve as Lewis acid catalysts for the activation and trans-
formation of reactants. Some reports have shown cooperative catalysis between
inorganic nodes of MOFs and metal NPs.75,76 However, due to the presence of
organic linkers, the inorganic nodes of MOFs are separated from metal NPs, result-
ing in weak interactions and unsatisfying catalytic performance. Very recently, we
fabricated a ‘‘quasi-MOF’’ through a controlled deligandation process to greatly
expose inorganic nodes to guest metal NPs, resulting in a strong interaction and syn-
ergistic effect between the immobilized metal NPs and the inorganic nodes (Fig-
ure 4B).73 As a proof of concept, Au/MIL-101(Cr) was calcined under an inert atmo-
sphere at different temperatures (i.e., 373, 473, 573, 673, and 1,073 K) to adjust the
interface between Au NPs and the inorganic Cr–O nodes. MIL-101 underwent sol-
vent removal at 373 K, followed by the removal of coordinated H2O molecules
and OH/F groups on the Cr–O nodes at 473 K, consequent partial deligandation
to form quasi-MIL-101 at 573 K, and finally collapse of the framework along with

62 Matter 1, 57–89, July 10, 2019


Figure 4. Catalytic Performances of MOF-Metal NP Composites
(A) Schematic illustration of the selective oxidation of hexane-1,5-diol by Pt/ZIF-8 and Pt/UiO-66. Reprinted with permission from Zhang et al. 72
Copyright 2018, WILEY-VCH Verlag GmbH & Co.
(B) Schematic illustration showing the synthesis of Au/quasi-MIL-101 through controlled deligandation of Au/MIL-101 (top). Schematic illustration of the
calcination-induced structure changes of the Au/MIL-101 composites (bottom). Reprinted with permission from Tsumori et al.73 Copyright 2018,
Elsevier Inc.
(C) Synthetic route for the core-shell Pd@IRMOF-3 hybrids via the mixed solvothermal method (top) and model cascade reactions involving
Knoevenagel condensation of 4-nitrobenzaldehyde and malononitrile via the IRMOF-3 shell and subsequent selective hydrogenation of intermediate
product 2-(4-nitrobenzylidene)malononitrile to 2-(4-aminobenzylidene)-malononitrile via the Pd NP cores (bottom). Reprinted with permission from
Zhao et al. 74 Copyright 2014, American Chemical Society.

carbon formation above 673 K. Partial deligandation of MIL-101 resulted in a strong


interaction between the guest Au NPs and the inorganic Cr–O nodes, while retaining
the porosity of the framework to a certain extent. The resulting Au/MIL-101(573)
composites showed dramatically enhanced catalytic activity for carbon monoxide
(CO) oxidation at temperatures even as low as 193 K, maintaining 100% CO conver-
sion up to 4,000 min. The high catalytic performance obtained is due to the efficient
synergistic effect between the accessible Cr–O sites and Au NPs at the perimeter
interface, in which the accessible Cr–O sites can uptake and activate O2 molecules
and then release active oxygen species to the CO molecules adsorbed on the Au
NPs.

The Organic Linkers of MOFs Cooperating with Metal NPs for Catalysis. The func-
tional groups in MOF linkers can be employed as active sites to cooperate with metal
NPs to promote synergistic catalysis in tandem reactions74 and photocatalytic reac-
tions.77 For example, Tang and co-workers fabricated core-shell Pd@IRMOF-3 nano-
structures as multifunctional catalysts for the tandem Knoevenagel condensation–
hydrogenation reaction (Figure 4C).74 In the core-shell nanocomposites, the amino
groups of the IRMOF-3 shells catalyzed the Knoevenagel condensation of 4-nitro-
benzaldehyde and malononitrile into 2-(4-nitrobenzylidene)malononitrile, and the
Pd NP cores selectively hydrogenated the –NO2 group of the intermediate product
to the target product, 2-(4-aminobenzylidene)-malononitrile. Under optimized

Matter 1, 57–89, July 10, 2019 63


conditions, Pd@IRMOF-3 achieved 100% conversion and 86% selectivity to the
target product. As a comparison, pure IRMOF-3 was only effective for the Knoeven-
gel condensation, but showed no activity for the hydrogenation process. Bare Pd
NPs demonstrated no activity for the Knoevengel condensation but could catalyze
the hydrogenation of the intermediate product.

MOF-Quantum Dot Composites


QDs, with a size range of 2–10 nm, have received considerable attention owing to
their uniquely size-dependent electronic and optical properties. The encapsulation
of QDs within MOFs can enhance their stability and modulate electron-hole recom-
bination rates. Several types of QDs (e.g., oxide-,78 chalcogenide-,79 nitride-,80 and
carbon-based QDs81) have been integrated into MOFs, and the resulting composite
materials exhibit enhanced properties for catalytic applications.82

Methodologies
MOF-QD composites can be synthesized by ‘‘ship-in-bottle’’ and ‘‘bottle-around-
ship’’ strategies. The ‘‘ship-in-bottle’’ strategy involves solution or vapor infiltration
of QD precursors into MOFs, followed by treatment with heating or reduction to
transform the precursors into the corresponding QDs (Figure 5A).80 In this case,
the MOF support needs to be stable under high temperatures and/or redox environ-
ments. Moreover, precise control of the location, size, and shape of the QDs within
the microstructure is very challenging, which could limit their applications as these
parameters of the QDs are closely related to the performance. The ‘‘bottle-
around-ship’’ strategy refers to the assembly of MOFs on preformed QD surfaces.
A capping agent or surfactant is usually required to stabilize the QDs and to facilitate
the heteronucleation of MOFs on the QD surfaces.79 Using this method, the shape
and size of QDs can be optimized for specific applications. However, the capping
agent or surfactant may passivate QDs, thus decreasing their fluorescence. To
address this issue, Banerjee and co-workers developed a capping agent-free strat-
egy for encapsulating CdS within MOFs.83 Uncapped CdS QDs are first trapped in
a low-molecular-weight metallohydrogel, followed by transformation into a xerogel
form and then conversion into a CdS-loaded MOF via a unique process mediated by
sodium chloride (Figure 5B).

Catalytic Performances of MOF-QD Composites


The combination of MOFs and QDs with desirable electronic and optical properties
allows the preparation of composite materials with enhanced properties for applica-
tions in electrocatalysis and photocatalysis.

The application of porous MOFs in electrocatalysis is restricted by their electrical


conductivity. The encapsulation of conductive QDs in the pores of MOFs may in-
crease the electrical conductivity of MOFs. Suh and co-workers developed a sacrifi-
cial-template method for the fabrication of CuS QDs in Cu3(BTC)2 (BTC = 1,3,5-ben-
zenetricarboxylate) by immersing Cu3(BTC)2 in an EtOH solution containing
thioacetamide as a sulfide source (Figure 5C).84 The composites exhibited superior
electrocatalytic activity with respect to the parent Cu-BTC and CuS in the oxygen
reduction reaction (ORR), showing much higher kinetic current density (11.3 mA
cm2) compared with those of Cu-BTC (3.46 mA cm2) and CuS (3.36 mA cm2) at
0.55 V versus reversible hydrogen electrode. Increasing the CuS loading in the com-
posites resulted in an increase in electrical conductivity but a decrease in porosity of
the composites, pointing to an optimized CuS loading amount of 28 WT % with the
best performance. This work demonstrated the synergistic effect between the QDs
(electrical conductivity) and the MOFs (porosity) in the electrocatalytic reaction.

64 Matter 1, 57–89, July 10, 2019


Figure 5. Preparation and Catalytic Performances of MOF-Quantum Dot Composites
(A) Conceptual representation of the confined trimerization of [(CH 3 ) 3 NGaH 3 ] inside ZIF-8 to selectively yield caged (H 2 GaNH 2 ) 3 and finally GaN@ZIF-8.
Reprinted with permission from Esken et al. 80 Copyright 2011, American Chemical Society.
(B) Schematic representation of the production of a CdS-loaded ZAVCl MOF (CdS@ZAVCl-MOF) from CdS-loaded ZAVA gel (CdS@ZAVA gel) and
space-filling representation of the crystal structure of ZAVCl (view through the c axis. Color code: carbon, gray; hydrogen, white; zinc, purple; chlorine,
green; nitrogen, blue; oxygen, red (top)). Real-time digital photographs of the CdS@ZAVA gel to CdS@ZAVCl MOF conversion process (bottom).
Reprinted with permission from Saha et al.83 Copyright 2014, American Chemical Society.
(C) Synthesis of nano-CuS(x WT %)@Cu-BTC and nano-CuS-(99 wt %). Reprinted with permission from Cho et al.84 Copyright 2016, WILEY-VCH Verlag
GmbH & Co.
(D) Schematic illustration of the fabrication process and CO 2 photoreduction process of CsPbBr3 /ZIFs. Reprinted with permission from Kong et al. 85
Copyright 2018, American Chemical Society.
(E) Photocatalytically induced spectral changes to aqueous MB dye (1.55 3 10 5 M) highlighting degradation of the absorption maximum at 664 nm in
the absence of composite (red) and the presence of 1.7 g SnO 2 @monoZIF-8 (green) after 3 h of simulated solar irradiation (left). Degradation of MB in the
presence of 0.4 g (cycles 1–5 and 10) and 1.7 g of SnO 2 @ mono ZIF-8. Error bars are the standard deviation in triplicated readings (right). Reprinted with
permission from Mehta et al. 86 Copyright 2018, WILEY-VCH Verlag GmbH & Co.

The combination of QDs with excellent visible-light responses and MOFs with
gas-capturing ability and active metal centers may show enhanced performance
in photocatalysis. Su and co-workers fabricated zinc/cobalt-based ZIF-coated
halide perovskite QD composites (CsPbBr3@ZIF-8 and CsPbBr3@ZIF-67) as effi-
cient photocatalysts for carbon dioxide (CO2) reduction (Figure 5D).85 The com-
posites exhibited enhanced CO2 reduction activity, showing higher electron con-
sumption rates (15.498 and 29.630 mmol g1 h1 for CsPbBr3@ZIF-8 and
CsPbBr3@ZIF-67, respectively) than that of pure CsPbBr3 (11.14 mmol g1 h1).
An optimized thickness of ZIF shell was of vital importance to the catalytic perfor-
mance of the CsPbBr3@ZIF composite, because the ZIF shell could enhance the
CO2 capture ability and facilitate the charge transfer process but lead to diffusion
resistance for the reactant/product and affect the light harvest. The CsPbBr3 and
ZIF coating played a synergistic action in the photocatalytic reaction, in which
CsPbBr3 was excited to generate electron-hole pairs and ZIFs caught the

Matter 1, 57–89, July 10, 2019 65


photogenerated electrons for the reaction of CO2 molecules adsorbed inside the
ZIF matrix.

The encapsulation of QDs in MOFs have achieved great success in terms of cata-
lytic performance. However, the powder morphology of classic MOFs results in
poor catalyst retention and recyclability. Shaping of powdered MOF supports
into monolithic conformations would enable catalysts with enhanced mechanical
stability for catalytic applications. Wheatley and co-workers immobilized tin
oxide NPs (SnO2-NPs) in situ within a monolithic MOF, monoZIF-8 to form
SnO2@monoZIF-8 for photocatalytic degradation of toxic waste.86 In the degrada-
tion of aqueous methylene blue (MB), 0.4 g of SnO2@monoZIF-8 achieved
41.5% G 3.0% MB degradation in 3 h (Figure 5E). The catalytic activity could
be retained after ten catalytic cycles. A control experiment showed that pure
monoZIF-8 presented no photocatalytic activity under simulated solar irradiation.
In addition, SnO2@monoZIF-8 showed improved mechanical strength than pure
monoZIF-8, attributed to the incorporation of the SnO2 NPs to lead to greater pack-
ing efficiency. The integration of SnO2 and monoZIF-8 to form SnO2@monoZIF-8 com-
posites could not only combine the photocatalytic activity of SnO2 NPs and struc-
tural resilience of the monolithic ZIF-8 but also lead to enhanced stability of the
SnO2 and mechanical strength of the composites. Moreover, the monolithic confor-
mations of the composites will facilitate the handling and recyclability of the cata-
lysts for further catalytic applications.

MOF-Polyoxometalate Composites
POMs are a family of discrete anionic metal-oxygen clusters with rich chemical vari-
ety, tunable shapes and sizes, solubilities, redox potentials, and strong acidities.
These properties provide great opportunities in a variety of catalytic transforma-
tions, especially in acid and oxidation reactions. However, their applications are
limited by their low specific surface area and low stability. Immobilizing POMs into
MOFs is a promising approach for stabilizing and optimizing POMs to enhance their
catalytic properties.87,88 Owing to their compositional diversity and structural versa-
tility, POMs can be employed as versatile building blocks (nodes, or pillars, or tem-
plates within the cages) for the construction of POM-based MOFs. In addition, POMs
can be encapsulated in the pores of MOFs through host-guest interactions to form
POM@MOF composites.

Methodologies
POM-based MOFs are prepared using POM precursors, metal ions, and organic
linkers via hydrothermal,89 solvothermal,90 and ionothermal methods.91 The synthe-
sis of POM-based MOFs is highly dependent on the reaction parameters (e.g., con-
centration, pH, reducing agents, ligands, reaction temperature), resulting in great
difficulty in overall control for successful synthesis. On the other hand, encapsulating
POMs in the pores of MOFs has recently attracted tremendous attention. Several
effective methods for preparing POM@MOF composites have been developed,
including impregnation of POM clusters into MOF pores, synthesis of POMs inside
the pores of presynthesized MOFs, and fabrication of MOFs around POMs.
Although the impregnation method is straightforward, it is limited to MOFs with a
window larger than the POMs,92 and easily results in the inevitable deposition of
POMs on the external surface of MOFs with low stability. The synthesis of POMs in-
side MOFs generally requires strongly acidic conditions, therefore only MOFs with
high stability in acidic solution are suitable as supports. The assembly of MOFs on
POMs has proved to be useful for the construction of hybrid composites.93 The
use of POMs as templates may alter the equilibrium between the metal ions and

66 Matter 1, 57–89, July 10, 2019


organic linkers and thus may affect the nucleation and growth of MOFs to achieve
shape-controlled synthesis of MOF nanocrystals.94

Catalytic Performances of MOF-POM Composites


Due to the redox properties of POMs, MOF-POM composites have been intensively
studied in oxidation reactions.95 In general, confining POMs within MOFs can pre-
vent POMs from leaching to show enhanced stability and recyclability. For example,
Das and co-workers confined a Keggin anion [CoW12O40]6 into the pores of ZIF-8 as
catalysts for electrocatalytic water oxidation (Figure 6A).96 The [H6CoW12O40]@ZIF-8
composite showed a turnover frequency (TOF) of 10.8 mol O2(mol Co)1 s1 at
neutral pH. Moreover, the catalyst was very stable in water oxidation, retaining its
initial activity even after 1,000 catalytic cycles. For comparison, a physical mixture
of POM and ZIF-8 showed similar activity with POM@ZIF-8, but the activity
decreased dramatically within ten cycles due to degradation of POM. These results
demonstrated the importance of the confinement effect of ZIF-8 to improve the sta-
bility of POMs in the catalytic reaction.

Very recently, it was demonstrated that the location of POMs in the MOFs had a
great impact on the reactivity. Farha and co-workers treated POM@MOF composite
(PW12@NU-1000) by using supercritical CO2 drying or heating at 120 C under vac-
uum.97 The POMs were located in the mesopores when drying in supercritical
CO2 and migrated to the micropores when heated to 120 C (Figure 6B). In the oxida-
tion of 2-chloroethyl ethyl sulfide (CEES), the TOF of PW12@NU-1000-scCO2 was
about three times higher than that of PW12@NU-1000-120 C. The different activities
exhibited were related to the location of the POMs in NU-1000. POMs situated in the
mesopores would have no hindered diffusion for the substrate, and thus the sulfide
could readily react with the POM to produce the singly oxidized product. However,
when POMs were located in the micropores, the access to the POM required diffu-
sion through the windows connecting the channels, thus resulting in inferior catalytic
efficiency.

Moreover, the confined POMs within MOF pores can work cooperatively with the
MOF backbone to show synergistic catalysis.

The Inorganic Nodes of MOFs Cooperating with POMs for Catalysis. The interac-
tion between the inorganic nodes of MOFs and the encapsulated POMs may lead to
activity and stability enhancement of POMs for catalysis. Hill and co-workers encap-
sulated a Keggin-type POM ([CuPW11O39]5) in the pores of MOF-199 (HKUST-1) as
highly active and stable catalysts in air-based oxidations.101 The close size matching
of POM diameter and MOF pore allowed strong POM-MOF interactions to show
enhanced activity compared with pure MOF and POM. In the oxidation of H2S,
POM-MOF afforded a TON valve of ca. 4,000 in less than 20 h. For comparison,
neither K5[CuPW11O39] nor MOF-199 showed activity under identical conditions.
Moreover, the POM-MOF was much more stable than the MOF or POM alone.
The authors proposed that the electrostatic interactions between the CuII nodes
of the MOF and the encapsulated POM may in turn stabilize each component and
increase the potential of the Cu centers in the POM with enhanced oxidation ability.

The combination of photoredox POMs and catalytically active inorganic nodes


of MOFs can act as the dual catalysts to individually activate the substrates,
enforcing the highly active intermediates to afford the product and avoiding the un-
wanted side reactions or reverse reactions. Duan and co-workers incorporated a
ruthenium(III) substituted polyoxometalate [SiW11O39Ru(H2O)] 5 within the pores

Matter 1, 57–89, July 10, 2019 67


Figure 6. Catalytic Performances of MOF-POM Composites
(A) Activation of POM for water oxidation by encapsulation in ZIF-8. Reprinted with permission from Mukhopadhyay et al. 96 Copyright 2018, WILEY-VCH
Verlag GmbH & Co.
(B) Structural representations of PW12 @NU-1000-120  C (left) and PW 12 @NU-1000-scCO2 (right). Reprinted with permission from Buru et al.97 Copyright
2018, the Royal Society of Chemistry.
(C) Synthetic procedure for the 3D CR-BPY1 MOF that is composed of wavy-like Cu-BPY sheets, and [SiW 11 O 40 Ru] 7 anions showing the combination of
dual catalytic units and channels for chemical transformations. Reprinted with permission from Shi et al. 98 Copyright 2015, the Royal Society of
Chemistry.
(D) The design concept of achieving a tandem catalyst. Synthetic procedure showing the constitutive/constructional fragments of the MOF; and the
schematic representation of tandem catalysis for asymmetric cyclic carbonate transformation from olefins and carbon dioxide. Reprinted with
permission from Han et al.99 Copyright 2015, Springer Nature.
(E) Schematic representation of the proposed mechanism for the light-driven OER by P2 W 18 Co 4 @MOF-545 (left). Kinetics of visible-light-driven O 2
production measured by gas chromatography analysis over 0.5 mg of P 2 W18 Co 4 @MOF-545 (blue squares), P 2 W18 Co 4 @MOF-545 recycled once (red
triangles), twice (pink stars), and 131 mM TCPP-H2 and 13 mM P 2 W18 Co 4 in solution (green circles). Reaction conditions: 5 mM Na 2 S 2 O 8 in 2 mL of 80 mM
borate buffer solution (pH 8), visible light (l > 420 nm, 280 W) (right). Reprinted with permission from Paille et al.100 Copyright 2018, American Chemical
Society.

68 Matter 1, 57–89, July 10, 2019


of copper(II)-bipyridine MOFs to afford CR-BPY1, merging Cu catalysis/Ru photoca-
talysis within one single MOF for oxidative coupling C–C bond formation (Fig-
ure 6C).98 In the coupling reaction of N-phenyl-tetrahydroisoquinoline and nitro-
methane, CR-BPY1 gave a yield of 90% after 24 h irradiation. Control experiments
using CuII salts or/and K5[SiW11O39Ru(H2O)] as catalysts gave conversions of 39%,
25%, and 42%, respectively. The catalytic results indicated cooperative effects of
the individual parts, that is, ruthenium-containing POM worked as an oxidative pho-
tocatalyst to transform N-phenyl-tetrahydroisoquinolines to iminium ion, and the Cu
in MOF nodes coordinated to nitromethane to form activated nucleophiles for the
C–C coupling reaction.

The Organic Linkers of MOFs Cooperating with POMs for Catalysis. Encapsulation
of POMs in MOFs with chiral organic linkers would offer new opportunities for asym-
metric catalysis. Duan and co-workers reported the integration of a chiral organoca-
talyst (pyrrolidine-2-yl-imidazole, PYI) and a polyoxometalate oxidation catalyst
([ZnW12O40]6) within one single MOF as a tandem catalyst for the efficient conver-
sion of CO2 into chiral cyclic carbonates (Figure 6D).99 The composite proceeded
with asymmetric epoxidation of styrene and the CO2 asymmetric coupling smoothly
and efficiently, affording the target product (R)-phenyl(ethylene carbonate) in 92%
yield with 80% enantiomer excess (ee). In this catalytic system, the good interaction
between organocatalyst PYI, NH2-functionalized bridging links, and the oxidation
catalyst [ZnW12O40]6 allows them to work cooperatively. The hydrogen bonds be-
tween the organocatalyst PYI and the oxidation catalyst [ZnW12O40]6 facilitate the
smooth and stereoselective conversion of olefin into epoxide, which then couples
with the CO2 molecule activated by amine groups on bridging linkers.

In addition, the encapsulation of POMs in MOFs with functional linkers as light-har-


vesting units can drive photocatalytic reactions. Dolbecq and co-workers encapsu-
lated [(PW9O34)2Co4(H2O)2]10 in the channels of ZrIV porphyrinic MOF-545, also
known as PCN-222 and MMPF-6 (Figure 6E).100 The resulting composite showed
good activity in the oxygen evolution reaction (OER), giving a TOF valve of 40 3
103 s1 for the first 15 min. Control experiments with MOF-545 with no encapsu-
lated POM or a solution containing the porphyrin linker and P2W18Co4 did not
show any significant activity. The unique activity of this POM@MOF photosystem
benefits from the cooperation between the porphyrin linker of the MOF and the
encapsulated POMs, in which the porphyrin linker acts as a photosensitizer to cap-
ture visible light and POMs catalyzes the oxidation of water into O2.

MOF-Enzyme Composites
Enzymes are a class of highly efficient biocatalysts, which are very powerful in the
catalysis of numerous reactions, featuring high activity and chemo-, enantio-, and re-
gioselectivity under mild conditions. However, their widespread catalytic applica-
tions are significantly hindered by the fragile nature of enzymes, such as low thermal
stability, narrow optimum pH ranges, and low tolerance to organic solvents and de-
naturants. Moreover, the contamination of enzymes in the desired products requires
cumbersome purification and separation steps. MOFs have been proven to be
promising platforms for the immobilization of enzymes to shield them from deacti-
vating reaction conditions, improve their recyclability, and minimize their contami-
nation in the product.102–105 The precise control over pore size, shape, and structure
of MOFs allows for the confinement of enzyme with matched size, thus minimizing
enzyme self-aggregation and leaching.106 Furthermore, the inorganic nodes and
functional linkers of MOFs may establish some interactions with enzymes through

Matter 1, 57–89, July 10, 2019 69


coordination, covalent bonding, hydrogen bonding, and van der Waals forces to
stabilize enzymes from leaching.107

Methodologies
Several strategies have been reported for the immobilization of enzymes in MOFs,
including physical adsorption, covalent attachment, diffusion, and co-precipita-
tion.104 Physical adsorption relies on weak interactions (e.g., electrostatic interac-
tions, hydrogen bonding, and van der Waals forces) between enzyme and MOF crys-
tals. Such weak interactions improve enzyme stability while preserving enzymatic
structure and activity. To further enhance long-term stability, enzymes can be
anchored to MOF crystals via multiple covalent bonds. For example, the amino
groups on an enzyme surface can covalently couple with the carboxylate groups
on an MOF to form robust peptide bonds, thus preventing the potential leaching
of the enzyme from the support.108 The diffusion of enzymes into MOF pores is
another effective strategy, in which mesoporous MOFs are utilized to achieve high
enzyme loading and efficient substrate diffusion.102 Instead of introducing enzymes
to the presynthesized MOF pores, enzymes can be encapsulated in situ during the
synthetic process of MOFs by a co-precipitation method. In order to avoid denatur-
ation of the enzyme during the encapsulation process, ZIFs with mild synthetic con-
ditions are suitable to be employed as a protective shell for this method.109 More-
over, the small pore size of ZIFs will minimize the leaching of enzyme from the
pores, while being flexible enough to allow the substrates to enter the framework
and reach the encapsulated enzymes.

Catalytic Performances of MOF-Enzyme Composites


In enzyme immobilized systems, the control of enzyme localization, substrate and
product diffusion, and enzyme and coenzyme accessibility, enzyme-support interac-
tions are important parameters to affect the catalytic performance of immobilized
enzyme composites. In this regard, the structure (topology and pore connectivity)
and functionality of MOFs have a great impact on the performance of enzymes.

Influence of Surface Wettability of MOFs on the Performances of Enzymes. It is


known that proteins tend to have a greater affinity for hydrophobic surfaces, and
the hydrophobic interactions often engender conformational changes that denature
the protein and lead to loss of activity. Doonan and co-workers encapsulated cata-
lase and urease within three Zn-based ZIFs with analogous structure metrics but
different hydrophobicity/hydrophilicity to compare the activities of encapsulated
enzymes (e.g., fluorescein-tagged catalase [FCAT]) (Figure 7A).110 In H2O2 decom-
position, FCAT encapsulated in hydrophilic MAF-7 and ZIF-90 retained enzymatic
activity (3.2 3 103 mM s1 for FCAT@ MAF-7, 1.4 3 103 mM s1 for FCAT@ZIF-
90, 4.4 3 103 mM s1 for free FCAT). However, FCAT encapsulated in hydrophobic
ZIF-8 showed no measurable catalytic activity. The loss of activity for FCAT in ZIF-8
may be due to the aggregation of FCAT induced by the hydrophobic interactions
between FCAT and ZIF-8 and may hinder the diffusion of the substrate (H2O2) and
the product (H2O) by the hydrophobic environment of ZIF-8.

Influence of Pore Structure of MOFs on the Performances of Enzymes. In biological


systems, chemical transformations involve the participation of multiple enzymes, co-
enzymes, substrates, and products. The key to achieving a complicated biochemical
transformation in MOF pores is to allow both the substrate and coenzymes to diffuse
and interact with the encapsulated enzymes. Therefore, it is important to design
an MOF structure with suitable topology (pore/channel shape) for enzyme
encapsulation and good connectivity between pores for enzyme accessibility. Farha

70 Matter 1, 57–89, July 10, 2019


Matter 1, 57–89, July 10, 2019 71
Figure 7. Catalytic Performances of MOF-Enzyme Composites
(A) Schematic representations of the different FCAT/ZIF biocomposites formed by encapsulation of enzyme molecules via biomimetic mineralization or
surface adsorption within/on hydrophobic (orange) or hydrophilic (blue) frameworks (left). Catalytic activity of FCAT and different FCAT/ZIF
composites. The assay was performed with an FCAT concentration of 20 nM and H2 O 2 concentration of 0.20 mM (right). Reprinted with permission from
Liang et al. 110 Copyright 2019, American Chemical Society.
(B) Perspective view of one-dimensional channels of PCN-600, CYCU-3, and NU-1000. Purple cylinders represent the channels, and yellow tunnels
indicate the connectivity between neighboring channels (left). Maximum loading capacity (black), enzymes encapsulated in channels (red), and
accessible enzymes (blue) in nano-sized MOFs (right). Reprinted with permission from Li et al. 111 Copyright 2016, Elsevier Inc.
(C) Reaction rate of degradation of methylene blue (MB, positive) and methyl orange (MO, negative) by free MP8 (red circles), immobilized
MP8@nanoMIL-101 (gray squares), and nanoMIL-101 (blue diamonds) (left). Solutions of MB and MO before and after catalytic degradation with
MP8@nanoMIL-101 (right). Reprinted with permission from Gkaniatsou et al. 112 Copyright 2018, WILEY-VCH Verlag GmbH & Co.

and co-workers investigated the immobilization of an enzyme (e.g., cutinase) in three


channel-type MOFs (NU-1000, PCN-600, and CYCU-3) with similar mesopore sizes
(3.0 nm) but different channel systems and connectivity (Figure 7B).111 NU-1000
has a hierarchical pore structure with one-dimensional 3.1 nm hexagonal and
0.8 nm triangular channels connected by microporous windows. The hierarchical
pore structure of NU-1000 could enable the stabilization of cutinase within the mes-
opores and the free diffusion of reactants and products through the micropores with
windows connected to mesopores. PCN-600 features only one-dimensional 3.0 nm
hexagonal channels connected by windows. These mesopores would be blocked af-
ter the encapsulation of cutinase, leaving not enough volume for the diffusion of re-
actants. CYCU-3 has the same hierarchical pore structure as NU-1000 but does not
have interconnecting windows for substrate diffusion through the triangular chan-
nels to access the enzyme in mesopores. The accessibility of cutinase in these
MOFs was demonstrated by a titration experiment using an organophosphorus
ester with a fluorescent resorufin group as an active-site titrant. The titration results
showed that 93% of the encapsulated cutinase in NU-1000 was accessible and cata-
lytically active, while only 6% of cutinase encapsulated in PCN-600 showed activity.
In the hydrolysis of p-nitrophenyl butyrate, NU-1000 with a hierarchical pore struc-
ture and channel-connecting windows exhibited similar overall activity to free cuti-
nase in solution.

In a subsequent work, Farha and co-workers systematically expanded the pores of


NU-1000 to afford an NU-100x (x = 3, 4, 5, 6, 7) series with hierarchical pore networks
consisting of hexagonal (33–67 Å) and triangular channels (13–33 Å) connected by
diamond-shaped windows (8–22 Å) as platforms for the immobilization of enzymes
(e.g., lactate dehydrogenase [LDH]).113 Expansion of the size of the channels and
bridging windows could facilitate the diffusion of substrates and coenzymes with
ample space for enzyme and coenzyme recognition. The activity of encapsulated
LDH increased with increasing pore size of the supports. The initial reaction rate
for LDH immobilized in NU-1005 was nearly equivalent to that of free LDH. They
further built a representative catalytic system consisting of LDH, diaphorase, and
nicotinamide adenine dinucleotide coenzymes (NAD and NADH). The triangular
channels of NU-100x (x = 5, 6, 7) are sufficiently large to accommodate NAD+, while
those of NU-100x (x = 3, 4) are too small. Notably, the three former MOFs attained
reaction rates exceeding that of free LDH by a factor of 1.5–3, while the two latter
MOFs showed slower apparent reaction rates than free LDH. This work demon-
strates that MOF pore size affects the diffusion rates of the substrate and coenzymes
and accessibility of the encapsulated enzymes.

Influence of Electrostatics of MOFs on the Performances of Enzymes. The electro-


statics of MOFs may also play a role in the performances of enzyme@MOF compos-
ites based on the consideration that charged MOFs may show electrostatic
affinity with opposite electrostatic charged substrate. Ricoux and co-workers used

72 Matter 1, 57–89, July 10, 2019


MIL-101(Cr) to encapsulate a peroxidase-type enzyme (microperoxidase-8 [MP8]) as
a stable and reusable biocatalyst for the oxidation reaction of dyes (Figure 7C).112 In
the enzymatic catalytic reaction, MIL-101 could not only act as a stabilizing host to
enhance the stability and reusability of the encapsulated enzyme but also pre-
concentrate the substrate for a higher reaction rate. When selecting dyes with
different electrostatic charges (cationic MB, neutral Sudan II and anionic methyl or-
ange [MO]), MIL-101(Cr) with positive charge in the 5–7 pH range would, therefore,
show selective affinity for the negatively charged MO over the positively charged
MB. Catalytic results showed that MP8@MIL-101(Cr) exhibited much higher reaction
rates compared with the free MP8 in the oxidation of negatively charged dye (MO).
On the contrary, MP8@MIL-101(Cr) showed an almost negligible reaction rate in
the oxidation of positively charged dye (MB) when compared with that observed
with free MP8. Therefore, the electrostatic attractive interactions between MO
and the MOF matrix may strongly favor the diffusion of MO into the pores of the
MOF, resulting in a pre-concentration of MO around the enzyme for enhanced
activity.

MOF-Molecular Species Composites


Homogeneous molecular species (e.g., dye molecules, salen complexes, porphy-
rins, and metalloporphyrins) have been investigated extensively as tools for under-
standing and mimicking the functionalities of biological systems. However, the activ-
ity lifetimes of these molecules are limited owing to self-aggregation as a result of
reactions between the active sites and oxidative self-degradation. The immobiliza-
tion of the molecular species in MOFs, where the active sites can be isolated and
protected, is a promising approach for heterogenization of homogeneous molecular
catalysts to combine the benefits of both classes of catalysts and overcome the
drawbacks.114 Molecular species can be incorporated in an MOF backbone via co-
valent bonding115–118 or encapsulated in MOF pores via non-covalent interactions.
Here, we focus on the latter approach. Direct encapsulation does not perturb the co-
ordination sphere of molecular species and thus maintains the structure and proper-
ties of molecular species. Moreover, direct encapsulation can retain the mobility of
molecular species to a certain extent, allowing them to cooperate with the MOF
backbone.119 Three-dimensional (3D) MOFs with large cavities interconnected by
small windows are particularly appropriate for the stabilization of molecular species,
based on the consideration that the large cavities can accommodate molecular spe-
cies, and the small pore windows prevent molecular species from leaching and
aggregating.

Methodologies
MOF-molecular species composites can be simply prepared by wet infiltration. The
pore windows of the MOFs should be larger than the molecular size of the complex
to allow the diffusion of molecular species into MOFs. Farrusseng and co-workers
systematically investigated the wet infiltration method to encapsulate metal phtha-
locyanine complexes (MPc) of different sizes in MIL-101.120 Small FePcF16 and
RuPcF16 could be successfully incorporated into MIL-101 with loadings of 2.1 and
3.6 WT %, respectively. Large (FePctBu4)2N complex (2.0 3 2.0 nm) could not enter
the pore windows of MIL-101 (1.47 3 1.6 nm) and thus mostly deposited on the
external surface of MIL-101. To address the limitations of the diffusion strategy,
Tsung and co-workers developed a process of dissociative linker exchange for the
encapsulation of large guests with diameters 3–4 times the size of the framework ap-
ertures.121 This strategy takes advantage of the phenomenon of dissociative linker
substitution, resulting in short-lived linker vacancies that can momentarily expand
the pore apertures and thus enable the inclusion of large guests (e.g., Rhodamine

Matter 1, 57–89, July 10, 2019 73


Figure 8. Preparation of MOF-Molecular Species Composites
(A) Incorporation of relatively large, diverse guest molecules into MOFs under linker exchange conditions. Reprinted with permission from Morabito
et al. 121 Copyright 2014, American Chemical Society.
(B) Schematic presentation of de novo assembly of Co-Pc@bio-MOF-1. Reprinted with permission from Li et al. 122 Copyright 2014, American Chemical
Society.
(C) Schematic presentation of porphyrins template the formation of octahemioctahedral cages. Reprinted with permission from Zhang et al. 123
Copyright 2012, American Chemical Society.

6G and triphenylphosphine) into the pores of ZIF-8 (Figure 8A). In addition, an MOF
can serve as a nanoreactor for the assembly of molecular species within its pores. For
example, Ma and co-workers encapsulated a guest molecule (cobalt(II) phthalocya-
nine) into an MOF (bio-MOF-1) via de novo assembly of the component fragments
(CoII and 1,2-dicyanobenzene) of the guest molecule, circumventing the challenge
of encapsulating guest molecules that are larger than the MOF windows
(Figure 8B).122

MOF-molecular species composites can also be prepared by growth of MOFs


around molecular species, known as the bottle-around-ship strategy. To avoid the
aggregation and degradation of molecular species during the assembly process,
MOFs with mild synthesis conditions are preferred. For example, porphyrin mole-
cules could be encapsulated in situ in the pores of MOFs by adding porphyrin mol-
ecules in the synthetic solution of MOFs.124,125 Eddaoudi and co-workers adopted
this strategy to encapsulate a porphyrin, 5,10,15,20-tetrakis-(1-methyl-4-pyridinio)
porphyrin tetra(p-toluenesulfonate) ([H2TMPyP] [p-tosyl]4), in an indium imidazole
dicarboxylate-based rho-ZMOF (topologically analogous to zeolite RHO).124
Furthermore, Zaworotko and co-workers found that porphyrins could act as struc-
ture-directing agents for the synthesis of MOFs that could not be prepared in the
absence of porphyrins (Figure 8C).123

Catalytic Performances of MOF-Molecular Species Composites


The successful encapsulation of molecular species in MOFs can enable improved
performances in many catalytic applications, including organocatalysis,124 photo-
catalysis,126 and asymmetric catalysis.119 Generally, encapsulated molecular spe-
cies exhibit enhanced catalytic stability as a benefit from the confinement of
MOF pores and can be easily separated from the reaction mixture to recover

74 Matter 1, 57–89, July 10, 2019


Figure 9. Catalytic Performances of MOF-Molecular Species Composites
(A) Activity of [Ru]@UiO-66 (TON = mmol HCOO/mmol Ru) upon catalyst recycling and comparison of catalyst activity in the first cycle (dark) with that
upon addition of a second aliquot of 1,8-diazabicyclo(5.4.0)undec-7-ene (DBU) (light) (left). Comparison of the activity of homogeneous (red),
encapsulated (blue), and surface bound (purple) catalysts in the presence of differently sized thiol poisons (right). Reprinted with permission from Li
et al. 127 Copyright 2018, American Chemical Society.
(B) Schematic diagram of redox potentials of MIL-125-NH 2 , BDC-NH2 , the Co II complex and H + /H2 (left). Proposed electron transfer route between MIL-
125-NH 2 and the Co II complex (right top); and between BDC-NH 2 and the Co II complex (right bottom). Reprinted with permission from Li et al. 128
Copyright 2016, American Chemical Society.
(C) Competing pathways for hydrogenation and isomerization of olefinic alcohols with 2-PF 6 (top). Proposed pathway for hydrogenation and
suppression of isomerization with 2@1-SO 3 Na (bottom). Reprinted with permission from Grigoropoulos et al. 129 Copyright 2018, WILEY-VCH Verlag
GmbH & Co.

and reuse for the next run. Tsung and co-workers encapsulated a ruthenium
complex (tBuPNP)Ru(CO)-HCl in Zr-based UiO-66 (tBuPNP = 2,6-bis((di-tert-butyl-
phosphino) methyl)pyridine) to afford [Ru]@UiO-66 composite as a very active
and stable catalyst for the hydrogenation of CO2 to formate.127 Compared
with homogeneous (tBuPNP)Ru(CO)-HCl catalyst, [Ru]@UiO-66 showed similar ac-
tivity but different recyclability. [Ru]@UiO-66 retained its activity through five cy-
cles, while (tBuPNP)Ru(CO)-HCl showed a significant decrease in activity in the
second run due to bimolecular catalyst deactivation (Figure 9A). Moreover, [Ru]
@UiO-66 showed enhanced susceptibility in the presence of thiol poisons
compared with (tBuPNP)Ru(CO)-HCl, benefitting from the shelter effect offered
by MOFs. This work demonstrated that isolation of individual metal complexes
in MOF pores could prevent metal complexes from undergoing bimolecular cata-
lyst deactivation.

Matter 1, 57–89, July 10, 2019 75


Although the pore structure of MOFs can isolate active molecular species within cav-
ities and minimize catalytic degradation, it usually introduces kinetic barriers such as
hindered mass transport.125 Therefore, it is important to design MOF support with a
hierarchical pore structure, in which large cages to encapsulate molecular species
and small cages allow substrates/products to move to and from the encapsulated
active sites freely. Moreover, the pore environment of MOFs should be finely tuned
to show high affinity for substrates and low affinity for products. For example,
Eddaoudi and co-workers applied [H2TMPyP]4+@rho-ZMOF composites in the
oxidation of cyclohexane, which showed a noticeably higher TON valve (23.5)
compared with other systems of metalloporphyrins in zeolites or mesoporous sili-
cates.124 In this catalytic system, the highly polar rho-ZMOF framework showed
low affinity to the polar hydrocarbon products (cyclohexanol and cyclohexanone),
thus facilitating the products to diffuse out of the pores.

Moreover, the backbone of MOFs can work cooperatively with the encapsulated molec-
ular species for synergistic catalysis. Jiang and co-workers encapsulated a CoII molecular
photocatalyst, [CoII(TPA)Cl][Cl] (TPA = tris(2-pyridylmethyl)-amine), in the pores of an
MOF photosensitizer, MIL-125-NH2, for visible-light-driven H2 production.128 The CoII
complex alone could not produce H2 without photosensitizer. Pristine MIL-125-NH2
showed limited activity with a H2 production rate of 17 mmol/g,h. The CoII@MIL-125-
NH2 composites exhibited significantly enhanced photocatalytic activities, possessing
the highest H2 production rate (553 mmol/g,h). For comparison, a physical mixture of
MIL-125-NH2 and the CoII complex with the same Co loading only gave a very low H2
production rate (37 mmol/g,h). These results demonstrated that the encapsulation of
the CoII complex inside the pores of MIL-125-NH2 could trigger a synergistic effect be-
tween the two components for efficient electron transfer from MIL-125-NH2 to the CoII
complex for efficient photocatalysis (Figure 9B). Rosseinsky and co-workers encapsu-
lated the cationic component of Crabtree’s catalyst [Ir(cod) (PCy3) (py)][PF6] (2-PF6)
into a sulfonated MIL-101(Cr) MOF (1-SO3Na) to form 2@1-SO3Na as a catalyst for
the selective hydrogenation of olefinic alcohols by suppressing the competing isomer-
ization reaction.129 In the hydrogenation of trans-crotyl alcohol, free Crabtree’s catalyst
(2-PF6) showed 31% selectivity to alcohol in 3 h and 54% selectivity to aldehyde as the
main product. By contrast, 2@1-SO3Na catalyst significantly suppressed isomerization
to the aldehydes, giving 92% selectivity to the alcohols. The authors proposed that sul-
fonate groups on MOFs could establish H bonding with the hydroxyl group of the
olefinic alcohols, thus disfavoring the coordination of the hydroxyl group to the active
centers in 2-PF6 to give aldehyde byproduct (Figure 9C).

MOF-Silica Composites
Silica NPs and nanostructures provide powerful platforms to accomplish many nano-
scale functions (e.g., porosity, stability, and hydrophilicity), which have attracted
considerable attention in catalytic applications. Integrating silica with MOFs would
combine the unique properties of both materials and lead to novel applications.
There are currently two main types of MOF-silica composites (SiO2@MOFs and
MOFs@SiO2). The former involves the incorporation of dispersed silica NPs within
the pores/channels of MOFs130 or growth of an MOF shell on a preformed silica
sphere,131 while the latter employs silica as a coating shell grown on the MOF sur-
face132 or a support to promote the growth of MOF particles.133

The hydrophilic property of silica greatly depends on the number of silanol moieties
on the silica surface. Decreasing the size of silica may increase the number of surface
silanols, thus enhancing the hydrophilicity of the materials for the adsorption
of hydrophilic molecules. Kitagawa and co-workers employed porous CPL-5

76 Matter 1, 57–89, July 10, 2019


Figure 10. Preparation and Catalytic Performances of MOF-Silica and MOF-Polymer Composites
(A) Formation of nanosized silica dispersed homogeneously inside the channels of porous coordination polymers. Reprinted with permission from
Uemura et al. 134 Copyright 2011, American Chemical Society.
(B) Testosterone esterification with caprylic acid catalyzed by (Zr)UiO-66(NH 2 )/SBA containing varying MOF loadings (a, 6.6 WT %; b, 13.2 WT %; and c,
18.6 WT %) compared with bare SBA-15 (d) and bulk MOF (e, 100 WT %). Yields of testosterone caprylate were obtained by esterification of testosterone
(left), and TOF values were calculated for all the catalysts employed (right). Reaction conditions: 75  C, 5 WT % Zr with respect to testosterone;
testosterone:caprylic acid molar ratio 1:20. Reprinted with permission from Cirujano et al. 135 Copyright 2017, WILEY-VCH Verlag GmbH & Co.
(C) MIL-101-IP with the polymer entrapped within the pores. Reprinted with permission from Aguila et al. 136 Copyright 2018, WILEY-VCH Verlag
GmbH & Co.
(D) Illustration of PDMS coating on the surface of MOFs and the improvement of moisture/water resistance of MOFs. Reprinted with permission from
Zhang et al. 137 Copyright 2014, American Chemical Society.

([Cu2(pzdc)2(dpe)], dpe = 1,2-di(4-pyridyl)-ethylene) as a template to fabricate


sub-nanosized silica NPs in the one-dimensional channels using a sol-gel reaction
of tetramethoxysilane.130 Kitagawa and co-workers further demonstrated that the
encapsulation of silica NPs in the PCP pores allowed remarkable alternations of
the properties of PCPs (Figure 10A).134 Adsorption measurements showed that
CPL-5 pores modified with hydrophilic silica can enhance the adsorption property
for water and 1,4-dioxane, while the diffusion of cyclohexane was highly restricted.
Considering that many chemical reactions involve hydrophilic molecules, this work
will contribute to the development of MOFs in catalytic applications for accelerating
the diffusion rate of hydrophilic molecules into the pores of MOFs.

Confinement of MOF NPs within mesoporous silica can maximize the active uncoor-
dinated sites at the outer surface of MOFs and enhance the chemical and mechanical
stability of MOFs by the shelter of silica. De Vos and co-workers confined a Zr-based

Matter 1, 57–89, July 10, 2019 77


MOF nanocrystal, UiO-66(NH2) in SBA-15 channels and tested the composites as
efficient and reusable heterogeneous catalysts for the synthesis of steroid deriva-
tives (Figure 10B).135 In the esterification of testosterone, (Zr)UiO-66(NH2)/SBA-15
with 6.6 WT % loading of MOF showed the highest activity (TOF, 0.38 h1), while
bulk MOF with the same Zr concentration was less active (0.16 h1). Testosterone
as a bulky substrate could not diffuse into the pores of (Zr)UiO-66(NH2) and thus
could only be catalyzed by active sites located at the outer surface of the MOF.
Therefore, MOF grown within SBA-15 with a smaller size would boost the catalytic
activity. Moreover, (Zr)UiO-66(NH2)/SBA-15 could be reused for at least seven cycles
without significant leaching of Zr from the catalyst.

MOF crystals are fragile and difficult to use as robust catalysts. Moreover, finely
divided MOF crystals are difficult to apply in a fixed-bed reactor, which would cause
very high resistance against the flow. To overcome this issue, immobilizing MOFs in
structured silica monoliths can facilitate processing and formulation into application-
specific configurations. Galarneau and co-workers reported in situ synthesis of Cu-
BTC NPs in the mesopores of silica monolith, then used as catalysts for the Fried-
länder reaction under continuous flow.138 Cu-BTC confined in silica monolith
(CuBTC-MonoSil) showed a steady-state value of 85% conversion with a productivity
of 2.2 mmol min1 gCuBTC1 on flow. The hierarchical silica monolith allowed large
and homogeneous flow to go through the macropore network without a pressure
drop. For comparison, commercial Cu-BTC powder in a batch reactor under similar
conditions gave a smaller productivity of 0.86 mmol min1 gCuBTC1, which may be
due to the larger size of commercial Cu-BTC.

MOF-Polymer Composites
Polymers possess a variety of unique properties, such as softness and thermal and
chemical stability. A combination of MOFs and polymers can produce new and ver-
satile materials that exhibit collective properties for framework stabilization and ac-
tivity enhancement.139

MOFs with regulated and tunable nanochannel structures have been used as nano-
reactors for confined polymerization.140 Confined polymers at nanometer scales
exhibit fascinating properties different from those in the bulk state. Moreover,
the confined polymers in the MOF pore/channel can have abundant interfaces
with the MOF backbone to show synergistic catalysis.136,141–143 Ma and co-workers
reported the insertion of a linear ionic polymer (IP) into MIL-101(Cr) to form MIL-
101(Cr)-IP, which could synergistically catalyze the fixation of carbon dioxide
into epoxides for cyclic carbonates (Figure 10C).136 In the conversion of epichloro-
hydrin to its cyclic carbonate, MIL-101(Cr)-IP composite showed full conversion at
50 C, outperforming the individual components (3% and 32% conversion for IP
and MIL-101(Cr), respectively) and the physical mixture of IP and MIL-101(Cr)
(80% conversion). Moreover, compared with the physical mixture, the composite
showed a much higher rate constant (0.0115 versus 0.00699 s1) at 25 C and lower
activation energy (63.6 versus 91.6 kJ mol1). The enhanced performance of MIL-
101(Cr)-IP composite is attributed to cooperative catalysis between the halide ions
on the linear polymer and CrIII Lewis acid sites of MIL-101(Cr). Specifically, the
Lewis acid sites activate the epoxide and then the nucleophile of the polymer
opens the epoxide ring. The close contact of IP and CrIII Lewis acid sites in the
MIL-101(Cr)-IP composite can work more effectively in a synergistic manner to
reduce the energy barrier, thus showing superior performance than the physical
mixture.

78 Matter 1, 57–89, July 10, 2019


In addition, MOFs can be confined in polymers by coating a polymer shell on the sur-
face of MOFs or dispersing MOF particles within the macroporous polymer mono-
liths/beads. Coating MOFs with a protective hydrophobic polymer layer can
improve the moisture/water stability of MOFs.144 Yu and co-workers modified hy-
drophobic polydimethysiloxane (PDMS) on the surface of MOF materials (MOF-5,
HKUST-1, and ZnBT) to significantly enhance their moisture/water resistance while
retaining their inherent porosity (Figure 10D).137 In the liquid phase cyanosilylation
of benzaldehyde and trimethylsilylcyanide, coated HKUST-1 showed similar activity
(48.2% yield) to that of the pristine one (50.1% yield), suggesting the permeability of
the PDMS coating layer for the substrates. Moreover, the coated HKUST-1 could
retain its activity after saturated water vapor treatment. For comparison, pristine
HKUST-1 showed a sharp decrease of catalytic activity (19.6% yield) after the same
treatment. These results demonstrated that the hydrophobic PDMS coating layer
can effectively prevent the water molecules entering and coordinating to the
exposed CuII sites, while still allowing the diffusion of substrates to reach the inner
MOF.

MOFs are usually obtained as finely dispersed powders, which are unsuitable for
many catalytic applications where shaped bodies are often required for ease of
handling and recyclability. Confinement of MOFs in polymer monoliths/beads
allows processing and formulation into application-specific configurations.145
Bradshaw and co-workers deposited a crystalline MOF, HKUST-1, into millimeter-
sized macroporous polyacrylamide (PAM) beads.36 The tough but flexible polymer
matrix could provide the embedded MOF particles with enhanced resistance
against mechanical attrition. Mechanically stirring the HKUST-1@polymer compos-
ites overnight in ethanol did not lead to bead breakup or MOF degradation, while
significant sample degradation with reduction of crystallite size was observed for
bulk HKUST-1 under the same stirring conditions. Moreover, PAM beads could offer
improved handling over bulk MOF phases, allowing their use in flow-reactor systems
for catalytic applications.

Other MOF-Functional Material Composites


Other functional materials, including carbon, ionic liquids (ILs), and metal-organic
polyhedra (MOPs), have also been integrated with MOFs to generate multifunctional
hybrid materials.

Carbon, including different allotropes (active carbon, nanotubes, fullerene, graphite,


etc.), existing forms (powder, fiber, foam, monolith, etc.), multiple micro-textures with
different dimensionalities from zero-dimensional to 3D and degrees of graphitization,
have attracted increasing interest in a variety of fields. The integration of MOFs and car-
bon-based materials can bring many novel functionalities such as enhanced stability,
improved electrical conductivity, high mechanical strength, and excellent chemical
and thermal robustness to address the weaknesses of MOFs.146,147 Loh and co-workers
assembled graphene oxide (GO) and a Cu-MOF as a trifunctional catalyst for the ORR,
OER, and hydrogen evolution reaction (HER) (Figure 11A).148 The Cu-MOF/GO com-
posites exhibited smaller overpotentials and higher current for all three electrocatalytic
reactions (HER, OER, and ORR) and showed better stability in acid media compared
with pure Cu-MOF. In a single polymer electrolyte membrane fuel cell test, the Cu-
MOF/GO composite achieved a power density of 110.5 mW cm2, which was much
higher than that of pure Cu-MOF (40.4 mW cm2). The higher catalytic activities of
Cu-MOF/GO composites compared with pure Cu-MOF or GO alone should be attrib-
uted to the synergistic effects of framework porosity, rapid charge transfer in GO, and
the catalytically active Cu in the MOF.

Matter 1, 57–89, July 10, 2019 79


Figure 11. Preparation and Catalytic Performances of Other MOF-Functional Material
Composites
(A) Schematic of the chemical structures of: GO (top), Cu-MOF (middle), and the paddle-wheel
secondary building units of pure Cu-MOF (bottom). Reprinted with permission from Jahan et al. 148
Copyright 2013, WILEY-VCH Verlag GmbH & Co.
(B) The polyILs@MIL-101 showed a synergistic effect of good CO 2 enrichment capacity, Lewis acid
sites in the MOF, as well as Lewis base sites in the polyILs. Reprinted with permission from Ding
et al. 149 Copyright 2018, American Chemical Society.
(C) Preparation of M 6 L 4 3MIL-101 (top). Catalytic performance and reusability of the M 6 L 4 and
M 6 L4 3MIL-101 materials in benzyl alcohol oxidation (bottom). Reprinted with permission from Qiu
et al. 150 Copyright 2016, American Chemical Society.

ILs have been regarded as novel and green catalysts due to their negligible vapor pres-
sure, high catalytic activity, good thermal stability, and adjustable physical and chemical
properties. Despite the advantages of ILs, their practical applications are limited to their
weaknesses, such as high cost, difficult catalyst recovery, and low efficiency of utilization.
Incorporating ILs into host materials such as MOFs can tune the properties of ILs through
nanosizing ILs or IL-support interactions and increase utilization to suppress the costs of
ILs.151 IL@MOF composites have shown improved catalytic activity, selectivity, and
reusability compared with their individual counterparts.152 Jiang and co-workers encap-
sulated poly(ionic liquid)s (polyILs) in MIL-101 through in situ polymerization of the
monomers (1-vinyl-3-ethylimidazolium bromide and ortho-divinylbenzene).149 The
encapsulated polyILs and MIL-101 support showed a synergistic effect in the coupling
of CO2 and epoxides (Figure 11B). The porous structure of MIL-101 could enrich CO2
in the pores and facilitate the diffusion of substrates/products. The exposed Lewis
acid sites in the MOF and the Lewis base sites (Br) of polyILs synergistically promoted
the coupling reaction of CO2 and epoxides. The composite catalyst exhibited excellent
catalytic activity (94% yield) toward CO2 cycloaddition with epoxides in the absence of a
co-catalyst. Moreover, polyILs@MIL-101 could be simply separated from the mixture
after reaction and reused, with no reduction of activity during ten runs of the CO2 cyclo-
addition reaction.

80 Matter 1, 57–89, July 10, 2019


MOPs, a new class of discrete inorganic-organic coordination complexes, have at-
tracted considerable attention in catalytic applications owing to their tunable struc-
tures and high symmetry. However, MOPs tend to aggregate under relatively harsh
reaction conditions thus compromising their performance. To address this issue, Li
and co-workers recently encapsulated an MOP, M6L4 (M = (en)Pd(NO3)2, en = ethyl-
enediamine, and L = 1,3,5-tris(4-pyridyl)-2,4,6-triazine) in the pores of MIL-101(Cr)
via a novel hydrophilicity-directed approach (Figure 11C).150 In the selective oxida-
tion of benzyl alcohol to benzaldehyde, M6L43MIL-101 showed significantly
improved activity (95% conversion) and selectivity to benzaldehyde (98%) than those
of M6L4 (60% conversion and 45% selectivity). Moreover, M6L43MIL-101(Cr) re-
tained its activity even after five reuses, while the pristine M6L4 deactivated dramat-
ically. This study indicated that the encapsulation of MOPs in MOF pores can isolate
individual MOPs to protect the highly dispersed MOP active centers from leaching or
aggregation, thus showing enhanced activity and stability.

Combining MOFS with Two or More Functional Materials


Combining MOFs with two or more functional materials can be expected to open
new avenues to multifunctional catalysts.

Metal oxide-supported metal NPs (metal/metal oxides) are one major category of
heterogeneous catalysts. Their interfaces are believed to play critical roles in chem-
ical catalysis. MOFs can serve as functionalized supports to confine ultrafine metal/
metal oxides in the pores to stabilize their interface from phase separation during the
catalytic process. Lin and co-workers used a UiO-bpy MOF, constructed from 2,20 -bi-
pyridine-5,50 -dicarboxylate (bpydc) and Zr6(m3-O)4(m3-OH)4, as support for Cu/ZnOx
catalysts for CO2 hydrogenation to methanol.153 The composites exhibited a space-
time yield to MeOH (STYMeOH) of 2.59 gMeOH kgCu1 h1 at a gas hourly space veloc-
ity of 18,000 h1, far exceeding that of commercial ternary Cu/ZnO/Al2O3 catalyst
(0.83 gMeOH kgCu1 h1). The high efficiency of the Cu/ZnOx@ UiO-bpy could be
attributed to the rich interface sites on well-mixed Cu, ZnOx, and Zr6 nodes for the
adsorption and activation of H2 and CO2 (Figure 12A).

MOF-supported metal NPs have been intensively investigated in various catalytic


transformations. Further coating the MOF-metal NP composites with an MOF/poly-
mer shell can change the chemical environment surrounding the metal NPs to regu-
late the catalytic activity and selectivity. Tang and co-workers coated MIL-101@Pt
NP composites with a thin layer of MIL-101 with inorganic nodes of Fe3+ or Cr3+,
allowing the entire Pt NP surface to have full contact with the support to show selec-
tive catalytic properties in the hydrogenation of cinnamaldehyde to cinnamyl alcohol
(Figure 12B).76 MIL-101@Pt nanostructures showed increased selectivity for cin-
namyl alcohol (86.4% and 44.0% for MIL-101(Fe)@Pt and MIL-101(Cr)@Pt, respec-
tively) compared with that of pure Pt NPs (18.3%), implying that the Lewis acid sites
in MIL-101 could preferentially interact with the C=O bond and activate it. However,
as the Pt NPs were not fully covered with MIL-101, the selectivity exhibited was still
not satisfactory. Further coating MIL-101@Pt with an MIL-101 shell could improve
the selectivity to cinnamyl alcohol; MIL-101@Pt@MIL-101 with MIL-101(Fe) as shells
always gave selectivities higher than 94%. Although coating MIL-101@Pt NP com-
posites with an MIL-101 shell can fully regulate the selective catalytic properties,
the hydrophilic nature of MIL-101 shells is not suitable for the sorption of hydropho-
bic reactants. In a subsequent work, Tang and co-workers coated MOF@metal NP
composites with conjugated micro- and mesoporous polymers (CMPs) for the con-
struction of a sandwich composite MOF@NPs@CMPs as an effective and selective
catalyst for cinnamaldehyde hydrogenation.156 Catalytically active Pt NPs were first

Matter 1, 57–89, July 10, 2019 81


Figure 12. Preparation and Catalytic Performances of Composites Combined with MOFs and Two
or More Functional Materials
(A) Schematic showing the encapsulated active sites in MOFs and the functions of the various
surface sites in catalytic CO 2 hydrogenation. Reprinted with permission from An et al. 153 Copyright
2017, American Chemical Society.
(B) Synthetic route to generating sandwich MIL-101@Pt@MIL-101, comprising Pt nanoparticles
(NPs) sandwiched between a core and a shell of MIL-101. Reprinted with permission from Zhao
et al. 76 Copyright 2016, Springer Nature.
(C) Scheme showing the proposed mechanism for photocatalytic H 2 evolution in the PNPMOF
system (symbols are not to scale). Reprinted with permission from Guo et al. 154 Copyright 2016, the
Royal Society of Chemistry.
(D) H2 evolution from photocatalysts under visible light. Reprinted with permission from Lin et al. 155
Copyright 2014, the Royal Society of Chemistry.

deposited on the surface of MIL-101, followed by coating with hydrophobic CMPs


with iron(III) porphyrin (FeP-CMPs). In the selective hydrogenation of cinnamalde-
hyde to cinnamyl alcohol, MIL-101@Pt achieved 15.0% conversion and 23.3% selec-
tivity to cinnamyl alcohol with a TOF value of 203.4 h1. After FeP-CMP coating, MIL-
101@Pt@FeP-CMP exhibited enhanced activity (97.6%) and better selectivity
(97.3%) for cinnamyl alcohol with a higher TOF value (1,516.1 h1). The FeP-CMP
shell played an important role in the catalytic reaction, serving as a hydrophobic shell
to enrich the reactant around catalytically active Pt NPs and a regulator to active the
C=O bond to improve selectivity to cinnamyl alcohol.

Moreover, metal NP-MOF composites can be integrated with silica to increase the
mechanical strength of the composites with improved stability and reusability in cat-
alytic processes. Li and co-workers prepared a Pt@ZIF-8/SiO2 composite with 3D-or-
dered microporous (3DOM) structure (denoted as 3DOM-Pt@ZIF-8/SiO2) by in situ
self-templated transformation of 3DOM-Pt/ZnO/SiO2 composite.157 Pt NPs were
encapsulated between the ZIF-8 shell and SiO2 skeleton with a 3DOM structure. In
the hydrogenation of olefins (e.g., 1-hexene and cis-cyclooctene) with different mo-
lecular sizes, 3DOM-Pt@ZIF-8/SiO2 showed high activity in the hydrogenation of
1-hexene (89.6% conversion after 24 h, TON = 2,129), but showed no activity using
cis-cyclooctene as substrate. The size selectivity of 3DOM-Pt@ZIF-8/SiO2 could be
attributed to the molecular sieve effect of the ZIF-8 shell. In addition, Pt@ZIF-8/
SiO2 without 3DOM structure showed much lower catalytic activity toward 1-hexene
(42.8% conversion after 24 h, TON = 862) due to inefficient mass transfer in the

82 Matter 1, 57–89, July 10, 2019


absence of macropores. 3DOM-Pt@ZIF-8/SiO2 showed good reusability, retaining
its size-selective catalytic properties even after seven runs, attributed to the good
mechanical properties of the SiO2 support to preserve the 3DOM architecture
from collapsing. This work demonstrated that the integration of the catalytic activity
of Pt NPs, the molecular-sieving effect of ZIF-8 shell, the good mechanical stability of
the SiO2 support, and the high mass transfer efficiency of the 3D-ordered micropo-
rous structure could lead to multifunctional composites with high activity, size selec-
tivity, and stability.

Redox-active POMs, catalytically active metal NPs, and photoactive MOFs can be
assembled into a composite to create more efficient visible-light-driven organic
photocatalysts. Hill and co-workers loaded Keggin POMs (H3PW12O40) and stabi-
lized Pt NPs on the surface of an amine-functionalized MOF, NH2-MIL-53, to afford
POM-Pt NPs@NH2-MIL-53 (PNPMOF) for visible-light-driven catalytic hydrogen
evolution.154 PNPMOF displayed synergistic photocatalytic H2 evolution activity,
achieving a TON of ca. 66 in 6 h with a quantum yield of 1.2 3 104. For comparison,
POM-Pt NPs@NH2-silica showed no production of H2 due to the lack of a light
absorbing unit. Pt black@NH2-MIL-53 only produced a negligible amount of H2,
implying the importance of the POMs to promote electron transfer between NH2-
MIL-53 and the Pt NPs catalyst. In the PNPMOF composites, NH2-MIL-53 formed
electron-hole pairs under visible-light irradiation, then transferred the electrons to
POMs and then to Pt NPs where H2 evolved (Figure 12C). The integration of these
three components in one system could allow them to work cooperatively for
visible-light harvesting and efficient electron transfer to the catalyst surface to
show high activity.

The integration of a photocatalytically active POM, high conductive graphene, and


an MOF semiconductor can also lead to efficient photocatalysts. Wu and co-workers
designed a ternary composite UiO-66/CdS/reduced graphene oxide (RGO) for pho-
tocatalytic hydrogen production.155 The ternary UiO-66/CdS/RGO composites
showed high activity in hydrogen generation under visible-light irradiation, showing
a higher hydrogen production rate (13.8 mmol gCdS 1 h1) than those of binary UiO-
66/CdS composites (11.2 mmol gCdS 1 h1) and CdS (1 mmol gCdS 1 h1) (Fig-
ure 12D). Neither UiO-66 nor RGO gave any activity under a similar reaction condi-
tion. In the ternary system, UiO-66 with a large surface area could well disperse CdS
particles for efficient interface charge separation, and RGO further accelerates the
transfer of photogenerated electrons, thus greatly improving the hydrogen produc-
tion performance of CdS.

Conclusions and Perspectives


The integration of MOFs and a variety of functional materials, such as metal NPs,
QDs, POMs, molecular species, enzymes, silica, and polymers, have been realized
via various well-developed strategies for the preparation of MOF composites/hy-
brids. MOF composites/hybrids can be fabricated by incorporating functional nano-
materials in the MOF framework to greatly enhance the properties of MOFs. On the
other hand, functional materials can act as scaffolds to confine MOF particles, which
would potentially facilitate the processing and formulation of MOF-based catalysts
into application-specific configurations. MOF composites/hybrids show synergistic
effects between the components, exhibiting remarkable features to greatly expand
the catalytic applications for a broad scope of reactions, with high activity, selec-
tivity, and stability. Moreover, MOF composites/hybrids can act as multifunctional
catalysts to realize tandem or cascade reactions. Although tremendous advances
have been achieved, the investigation of MOF composites/hybrids for catalysis is

Matter 1, 57–89, July 10, 2019 83


still in its infancy and many challenges remain to be addressed. In this section, we
discuss the challenges that we believe to be the most important.

(1) The stability issue of MOF composites should be addressed if MOF compos-
ites are to be found in industrial applications. Chemical and thermal stability is
of vital importance for the recyclability of catalysts, and mechanical stability is
required for the shaping of MOF composites for final applications. When
applying MOF composites in catalysis, it is necessary to determine the stabil-
ity of the catalysts and leaching of active sites, and study the mechanical
behavior of the MOF composites under pressure.
(2) The development of scalable and efficient methods for the preparation of MOF
composites is required. Up to now, MOF composites have been fabricated only
on the laboratory scale by low-yield and time-consuming procedures. A one-pot
protocol could simplify the synthetic process and omit energy-consuming steps
such as separation and purification of intermediates, which may open new ave-
nues for the industrial-scale application of MOF composites.
(3) A deeper understanding of interactions between MOFs and functional mate-
rials is vital for successful assembly and improved properties. The nature of
the interface not only affects the assembly process (including nucleation,
growth, and orientation) but also plays an important role in the performance
of the composite. In-depth studies with the aim of characterizing these inter-
facial interactions can be expected to improve fundamental understanding of
the interfaces to establish structure-property relationships and thus guide the
design of the next generation of highly efficient MOF composites for catalysis.
(4) MOF composite catalysts still lack detailed investigations into the nature of
active species and possible deactivation pathways. Common characteriza-
tions of MOF catalysts are based on ex situ methods, which fail to monitor
the reaction dynamics under working conditions. In operando methods are
highly demanded to achieve a deeper understanding of the nature of active
sites and deactivation mechanisms.
(5) MOF composite catalysts are restricted to the transformation of small mole-
cules due to the microstructure of most MOFs. The production of bulky,
high value-added molecules is desirable. The development of effective stra-
tegies to enlarge the pore size of MOFs with hierarchical structures158,159 or
reduce the particle size of MOFs with maximum active sites at the outer sur-
face by confinement in mesoporous scaffolds135,160 will open new avenues
for the applications of MOF composites in transformations involving bulky
molecules (e.g., biomass upgrading toward fine chemicals, pharmacological
applications).

Although many challenges still exist, the rapid development of MOF composites in
recent years has predicted well for the bright future of this new type of functional ma-
terials. Sustained research efforts in this exciting area can be expected to enable
their practical applications.

ACKNOWLEDGMENTS
The authors thank the editor for the kind invitation. We would like to thank AIST for
financial support.

AUTHOR CONTRIBUTIONS
Conceptualization, Q.X.; Investigation, L.C.; Writing – Original Draft, L.C.; Writing –
Review & Editing, L.C. and Q.X.; Supervision, Q.X.

84 Matter 1, 57–89, July 10, 2019


REFERENCES
1. Furukawa, H., Cordova, K.E., O’Keeffe, M., 16. Liu, J., Chen, L., Cui, H., Zhang, J., Zhang, L., template for porous carbon synthesis. J. Am.
and Yaghi, O.M. (2013). The chemistry and and Su, C.-Y. (2014). Applications of metal– Chem. Soc. 130, 5390–5391.
applications of metal–organic frameworks. organic frameworks in heterogeneous
Science 341, 974–986. supramolecular catalysis. Chem. Soc. Rev. 43, 31. Dang, S., Zhu, Q.-L., and Xu, Q. (2017).
6011–6061. Nanomaterials derived from metal–organic
2. Kitagawa, S., Kitaura, R., and Noro, S.-I. frameworks. Nat. Rev. Mater. 3, 17075.
(2004). Functional porous coordination 17. Li, B., Chrzanowski, M., Zhang, Y., and Ma, S.
polymers. Angew. Chem. Int. Ed. 43, 2334– (2016). Applications of metal–organic 32. Pachfule, P., Shinde, D., Majumder, M., and
2375. frameworks featuring multi-functional sites. Xu, Q. (2016). Fabrication of carbon nanorods
Coord. Chem. Rev. 307, 106–129. and graphene nanoribbons from a metal–
3. Simon-Yarza, T., Mielcarek, A., Couvreur, P., organic framework. Nat. Chem. 8, 718–724.
and Serre, C. (2018). Nanoparticles of metal– 18. Huang, Y.-B., Liang, J., Wang, X.-S., and Cao,
organic frameworks: on the road to in vivo R. (2017). Multifunctional metal–organic 33. Shen, K., Chen, X., Chen, J., and Li, Y. (2016).
efficacy in biomedicine. Adv. Mater. 30, framework catalysts: synergistic catalysis and Development of MOF-derived carbon-based
1707365. tandem reactions. Chem. Soc. Rev. 46, nanomaterials for efficient catalysis. ACS
126–157. Catal. 6, 5887–5903.
4. Zhao, X., Wang, Y., Li, D.-S., Bu, X., and Feng,
P. (2018). Metal–organic frameworks for 19. Dhakshinamoorthy, A., Li, Z., and Garcia, H. 34. Chaikittisilp, W., Ariga, K., and Yamauchi, Y.
separation. Adv. Mater. 30, 1705189. (2018). Catalysis and photocatalysis by metal (2013). A new family of carbon materials:
organic frameworks. Chem. Soc. Rev. 47, synthesis of MOF-derived nanoporous
5. Lu, K., Aung, T., Guo, N., Weichselbaum, R., 8134–8172. carbons and their promising applications.
and Lin, W. (2018). Nanoscale metal–organic J. Mater. Chem. A 1, 14–19.
frameworks for therapeutic, imaging, and 20. Tu, W., Xu, Y., Yin, S., and Xu, R. (2018).
sensing applications. Adv. Mater. 30, Rational design of catalytic centers in 35. Ramos-Fernandez, E.V., Garcia-Domingos,
1707634. crystalline frameworks. Adv. Mater. 30, M., Juan-Alcañiz, J., Gascon, J., and Kapteijn,
1707582. F. (2011). MOFs meet monoliths: hierarchical
6. Woellner, M., Hausdorf, S., Klein, N., Mueller, structuring metal organic framework
P., Smith, M.W., and Kaskel, S. (2018). 21. Chughtai, A.H., Ahmad, N., Younus, H.A., catalysts. Appl. Catal. A Gen. 391, 261–267.
Adsorption and detection of hazardous trace Laypkov, A., and Verpoort, F. (2015). Metal–
gases by metal–organic frameworks. Adv. organic frameworks: versatile heterogeneous 36. O’Neill, L.D., Zhang, H., and Bradshaw, D.
Mater. 30, 1704679. catalysts for efficient catalytic organic (2010). Macro-/microporous MOF composite
transformations. Chem. Soc. Rev. 44, 6804– beads. J. Mater. Chem. 20, 5720–5726.
7. Wang, H., Lustig, W.P., and Li, J. (2018). 6849.
Sensing and capture of toxic and hazardous 37. Bradshaw, D., Garai, A., and Huo, J. (2012).
gases and vapors by metal–organic 22. Islamoglu, T., Goswami, S., Li, Z., Howarth, Metal–organic framework growth at
frameworks. Chem. Soc. Rev. 47, 4729–4756. A.J., Farha, O.K., and Hupp, J.T. (2017). functional interfaces: thin films and
Postsynthetic tuning of metal–organic composites for diverse applications. Chem.
8. He, Y., Chen, F., Li, B., Qian, G., Zhou, W., and frameworks for targeted applications. Acc. Soc. Rev. 41, 2344–2381.
Chen, B. (2018). Porous metal–organic Chem. Res. 50, 805–813.
frameworks for fuel storage. Coord. Chem. 38. Doherty, C.M., Buso, D., Hill, A.J., Furukawa,
Rev. 373, 167–198. 23. Wen, Y., Zhang, J., Xu, Q., Wu, X.-T., and Zhu, S., Kitagawa, S., and Falcaro, P. (2014). Using
Q.-L. (2018). Pore surface engineering of functional nano- and microparticles for the
9. Farrusseng, D., Aguado, S., and Pinel, C. metal–organic frameworks for heterogeneous
(2009). Metal–organic frameworks: preparation of metal–organic framework
catalysis. Coord. Chem. Rev. 376, 248–276. composites with novel properties. Acc. Chem.
opportunities for catalysis. Angew. Chem. Int.
Ed. 48, 7502–7513. 24. Kang, Y.-S., Lu, Y., Chen, K., Zhao, Y., Wang, Res. 47, 396–405.
P., and Sun, W.-Y. (2019). Metal–organic 39. Bétard, A., and Fischer, R.A. (2012). Metal–
10. Lee, J., Farha, O.K., Roberts, J., Scheidt, K.A., frameworks with catalytic centers: from
Nguyen, S.T., and Hupp, J.T. (2009). Metal– organic framework thin films: from
synthesis to catalytic application. Coord. fundamentals to applications. Chem. Rev.
organic framework materials as catalysts. Chem. Rev. 378, 262–280.
Chem. Soc. Rev. 38, 1450–1459. 112, 1055–1083.
25. Chen, K., and Wu, C.-D. (2018). Designed
11. Corma, A., Garcia, H., and Llabres i Xamena, 40. Yang, Q., Xu, Q., and Jiang, H.-L. (2017).
fabrication of biomimetic metal–organic
F.X.L.I. (2010). Engineering metal organic Metal–organic frameworks meet metal
frameworks for catalytic applications. Coord.
frameworks for heterogeneous catalysis. nanoparticles: synergistic effect for enhanced
Chem. Rev. 378, 445–465.
Chem. Rev. 110, 4606–4655. catalysis. Chem. Soc. Rev. 46, 4774–4808.
26. Rogge, S.M.J., Bavykina, A., Hajek, J., Garcia,
12. Yuan, S., Feng, L., Wang, K., Pang, J., Bosch, 41. Dhakshinamoorthy, A., and Garcia, H. (2012).
H., Olivos-Suarez, A.I., Sepúlveda-Escribano,
M., Lollar, C., Sun, Y., Qin, J., Yang, X., Zhang, Catalysis by metal nanoparticles embedded
A., Vimont, A., Clet, G., Bazin, P., Kapteijn, F.,
P., et al. (2018). Stable metal–organic on metal–organic frameworks. Chem. Soc.
et al. (2017). Metal–organic and covalent
frameworks: design, synthesis, and Rev. 41, 5262–5284.
organic frameworks as single-site catalysts.
applications. Adv. Mater. 30, 1704303. Chem. Soc. Rev. 46, 3134–3184. 42. Falcaro, P., Ricco, R., Yazdi, A., Imaz, I.,
13. Burtch, N.C., Heinen, J., Bennett, T.D., 27. Zhu, Q.-L., and Xu, Q. (2014). Metal–organic Furukawa, S., Maspoch, D., Ameloot, R.,
Dubbeldam, D., and Allendorf, M.D. (2018). framework composites. Chem. Soc. Rev. 43, Evans, J.D., and Doonan, C.J. (2016).
Mechanical properties in metal–organic 5468–5512. Application of metal and metal oxide
frameworks: emerging opportunities and nanoparticles@MOFs. Coord. Chem. Rev.
challenges for device functionality and 28. Chen, L., Luque, R., and Li, Y. (2017). 307, 237–254.
technological applications. Adv. Mater. 30, Controllable design of tunable
1704124. nanostructures inside metal–organic 43. Meilikhov, M., Yusenko, K., Esken, D., Turner,
frameworks. Chem. Soc. Rev. 46, 4614–4630. S., Van Tendeloo, G., and Fischer, R.A. (2010).
14. Jiao, L., Wang, Y., Jiang, H.-L., and Xu, Q. Metals@MOFs – loading MOFs with metal
(2018). Metal–organic frameworks as 29. Juan-Alcaniz, J., Gascon, J., and Kapteijn, F. nanoparticles for hybrid functions. Eur. J.
platforms for catalytic applications. Adv. (2012). Metal–organic frameworks as scaffolds Inorg. Chem. 2010, 3701–3714.
Mater. 30, 1703663. for the encapsulation of active species: state
of the art and future perspectives. J. Mater. 44. Li, G., Zhao, S., Zhang, Y., and Tang, Z. (2018).
15. Xu, C., Fang, R., Luque, R., Chen, L., and Li, Y. Chem. 22, 10102–10118. Metal–organic frameworks encapsulating
(2019). Functional metal–organic frameworks active nanoparticles as emerging composites
for catalytic applications. Coord. Chem. Rev. 30. Liu, B., Shioyama, H., Akita, T., and Xu, Q. for catalysis: recent progress and
388, 268–292. (2008). Metal–organic framework as a perspectives. Adv. Mater. 30, 1800702.

Matter 1, 57–89, July 10, 2019 85


45. Chen, L., Luque, R., and Li, Y. (2018). heterogeneous nanocatalysts. Angew. Chem. catalytic properties of Pd nanoclusters
Encapsulation of metal nanostructures into Int. Ed. 55, 5019–5023. through their chemical environment at the
metal–organic frameworks. Dalton Trans. 47, atomic level using isoreticular metal–organic
3663–3668. 58. Aijaz, A., and Xu, Q. (2014). Catalysis with frameworks. ACS Catal. 6, 3461–3468.
metal nanoparticles immobilized within the
46. Kim, C.R., Uemura, T., and Kitagawa, S. (2016). pores of metal–organic frameworks. J. Phys. 71. Huang, G., Yang, Q., Xu, Q., Yu, S.-H., and
Inorganic nanoparticles in porous Chem. Lett. 5, 1400–1411. Jiang, H.-L. (2016). Polydimethylsiloxane
coordination polymers. Chem. Soc. Rev. 45, coating for a palladium/MOF composite:
3828–3845. 59. Aijaz, A., Karkamkar, A., Choi, Y.J., Tsumori, highly improved catalytic performance by
N., Rönnebro, E., Autrey, T., Shioyama, H., surface hydrophobization. Angew. Chem. Int.
47. Hermes, S., Schroter, M.K., Schmid, R., and Xu, Q. (2012). Immobilizing highly Ed. 55, 7379–7383.
Khodeir, L., Muhler, M., Tissler, A., Fischer, catalytically active Pt nanoparticles inside the
R.W., and Fischer, R.A. (2005). Metal@MOF: pores of metal–organic framework: a double 72. Zhang, W., Zheng, B., Shi, W., Chen, X., Xu, Z.,
loading of highly porous coordination solvents approach. J. Am. Chem. Soc. 134, Li, S., Chi, Y.R., Yang, Y., Lu, J., Huang, W.,
polymers host lattices by metal organic 13926–13929. et al. (2018). Site-selective catalysis of a
chemical vapor deposition. Angew. Chem. multifunctional linear molecule: the steric
Int. Ed. 44, 6237–6241. 60. Lu, G., Li, S., Guo, Z., Farha, O.K., Hauser, hindrance of metal–organic framework
B.G., Qi, X., Wang, Y., Wang, X., Han, S., Liu, channels. Adv. Mater. 30, 1800643.
48. Chen, F., Shen, K., Chen, J., Yang, X., Cui, J., X., et al. (2012). Imparting functionality to a
and Li, Y. (2019). General immobilization of metal–organic framework material by 73. Tsumori, N., Chen, L., Wang, Q., Zhu, Q.-L.,
ultrafine alloyed nanoparticles within metal– controlled nanoparticle encapsulation. Nat. Kitta, M., and Xu, Q. (2018). Quasi-MOF:
organic frameworks with high loadings for Chem. 4, 310–316. Exposing inorganic nodes to guest metal
advanced synergetic catalysis. ACS Cent. Sci. nanoparticles for drastically enhanced
5, 176–185. 61. Kuo, C.-H., Tang, Y., Chou, L.-Y., Sneed, B.T., catalytic activity. Chem 4, 845–856.
Brodsky, C.N., Zhao, Z., and Tsung, C.-K.
49. Gu, X., Lu, Z.-H., Jiang, H.-L., Akita, T., and Xu, (2012). Yolk–shell nanocrystal@ZIF-8 74. Zhao, M., Deng, K., He, L., Liu, Y., Li, G., Zhao,
Q. (2011). Synergistic catalysis of metal– nanostructures for gas-phase heterogeneous H., and Tang, Z. (2014). Core–shell palladium
organic framework-immobilized Au–Pd catalysis with selectivity control. J. Am. Chem. nanoparticle@metal–organic frameworks as
nanoparticles in dehydrogenation of formic Soc. 134, 14345–14348. multifunctional catalysts for cascade
acid for chemical hydrogen storage. J. Am. reactions. J. Am. Chem. Soc. 136, 1738–1741.
Chem. Soc. 133, 11822–11825. 62. Zhao, Y., Kornienko, N., Liu, Z., Zhu, C.,
Asahina, S., Kuo, T.-R., Bao, W., Xie, C., 75. Pan, Y., Yuan, B., Li, Y., and He, D. (2010).
50. Chen, L., Huang, B., Qiu, X., Wang, X., Luque, Hexemer, A., Terasaki, O., et al. (2015). Multifunctional catalysis by Pd@MIL-101: one-
R., and Li, Y. (2016). Seed-mediated growth of Mesoscopic constructs of ordered and step synthesis of methyl isobutyl ketone over
MOF-encapsulated Pd@Ag core–shell oriented metal–organic frameworks on palladium nanoparticles deposited on a
nanoparticles: toward advanced room plasmonic silver nanocrystals. J. Am. Chem. metal–organic framework. Chem. Commun.
temperature nanocatalysts. Chem. Sci. 7, Soc. 137, 2199–2202. 46, 2280–2282.
228–233.
63. Wang, S., Fan, Y., Teng, J., Fan, Y.-Z., Jiang, 76. Zhao, M., Yuan, K., Wang, Y., Li, G., Guo, J.,
51. Jiang, H.-L., Akita, T., Ishida, T., Haruta, M., J.-J., Wang, H.-P., Grützmacher, H., Wang, D., Gu, L., Hu, W., Zhao, H., and Tang, Z. (2016).
and Xu, Q. (2011). Synergistic catalysis of and Su, C.-Y. (2016). Nanoreactor based on Metal–organic frameworks as selectivity
Au@Ag core–shell nanoparticles stabilized on macroporous single crystals of metal–organic regulators for hydrogenation reactions.
metal–organic framework. J. Am. Chem. Soc. framework. Small 12, 5702–5709. Nature 539, 76–80.
133, 1304–1306.
64. He, L., Liu, Y., Liu, J., Xiong, Y., Zheng, J., Liu, 77. Chen, Y.-Z., Wang, Z.U., Wang, H., Lu, J., Yu,
52. Aijaz, A., Akita, T., Tsumori, N., and Xu, Q. Y., and Tang, Z. (2013). Core–shell noble- S.-H., and Jiang, H.-L. (2017). Singlet oxygen-
(2013). Metal–organic framework- metal@metal–organic-framework engaged selective photo-oxidation over Pt
immobilized polyhedral metal nanocrystals: nanoparticles with highly selective sensing nanocrystals/porphyrinic MOF: the roles of
reduction at solid-gas interface, metal property. Angew. Chem. Int. Ed. 52, 3741– photothermal effect and Pt electronic state.
segregation, core–shell structure, and high 3745. J. Am. Chem. Soc. 139, 2035–2044.
catalytic activity. J. Am. Chem. Soc. 135,
16356–16359. 65. Chen, L., Chen, X., Liu, H., Bai, C., and Li, Y. 78. Feng, P.L., Perry, J.J., Nikodemski, S., Jacobs,
(2015). One-step encapsulation of Pd B.W., Meek, S.T., and Allendorf, M.D. (2010).
53. Li, G., Kobayashi, H., Taylor, J.M., Ikeda, R., nanoparticles in MOFs via a temperature Assessing the purity of metal–organic
Kubota, Y., Kato, K., Takata, M., Yamamoto, control program. J. Mater. Chem. A 3, 15259– frameworks using photoluminescence: MOF-
T., Toh, S., Matsumura, S., et al. (2014). 15264. 5, ZnO quantum dots, and framework
Hydrogen storage in Pd nanocrystals covered decomposition. J. Am. Chem. Soc. 132,
with a metal–organic framework. Nat. Mater. 66. Chen, L., Chen, H., and Li, Y. (2014). One-pot 15487–15489.
13, 802–806. synthesis of Pd@MOF composites without the
addition of stabilizing agents. Chem. 79. Falcaro, P., Hill, A.J., Nairn, K.M., Jasieniak, J.,
54. Chen, L., Chen, H., Luque, R., and Li, Y. (2014). Commun. 50, 14752–14755. Mardel, J.I., Bastow, T.J., Mayo, S.C., Gimona,
Metal–organic framework encapsulated Pd M., Gomez, D., Whitfield, H.J., et al. (2011). A
nanoparticles: towards advanced 67. Sabo, M., Henschel, A., Frode, H., Klemm, E., new method to position and functionalize
heterogeneous catalysts. Chem. Sci. 5, 3708– and Kaskel, S. (2007). Solution infiltration of metal–organic framework crystals. Nat.
3714. palladium into MOF-5: synthesis, Commun. 2, 237.
physisorption and catalytic properties.
55. Zhu, Q.-L., Li, J., and Xu, Q. (2013). J. Mater. Chem. 17, 3827–3832. 80. Esken, D., Turner, S., Wiktor, C., Kalidindi,
Immobilizing metal nanoparticles to metal– S.B., Van Tendeloo, G., and Fischer, R.A.
organic frameworks with size and location 68. Jiang, H.-L., Liu, B., Akita, T., Haruta, M., (2011). GaN@ZIF-8: selective formation of
control for optimizing catalytic performance. Sakurai, H., and Xu, Q. (2009). Au@ZIF-8: CO gallium nitride quantum dots inside a zinc
J. Am. Chem. Soc. 135, 10210–10213. oxidation over gold nanoparticles deposited methylimidazolate framework. J. Am. Chem.
to metal–organic framework. J. Am. Chem. Soc. 133, 16370–16373.
56. Li, B., Ma, J.-G., and Cheng, P. (2018). Silica- Soc. 131, 11302–11303.
protection-assisted encapsulation of Cu2O 81. Biswal, B.P., Shinde, D.B., Pillai, V.K., and
nanocubes into a metal–organic framework 69. Chen, L., Chen, X., Liu, H., and Li, Y. (2015). Banerjee, R. (2013). Stabilization of graphene
(ZIF-8) to provide a composite catalyst. Encapsulation of mono- or bimetal quantum dots (GQDs) by encapsulation
Angew. Chem. Int. Ed. 57, 6834–6837. nanoparticles inside metal–organic inside zeolitic imidazolate framework
frameworks via in situ incorporation of metal nanocrystals for photoluminescence tuning.
57. Liu, H., Chang, L., Bai, C., Chen, L., Luque, R., precursors. Small 11, 2642–2648. Nanoscale 5, 10556–10561.
and Li, Y. (2016). Controllable encapsulation
of ‘‘clean’’ metal clusters within MOFs through 70. Li, X., Goh, T.W., Li, L., Xiao, C., Guo, Z., Zeng, 82. Aguilera-Sigalat, J., and Bradshaw, D. (2016).
kinetic modulation: towards advanced X.C., and Huang, W. (2016). Controlling Synthesis and applications of metal–organic

86 Matter 1, 57–89, July 10, 2019


framework–quantum dot (QD@MOF) polyoxometalates as coordination 106. Deng, H., Grunder, S., Cordova, K.E., Valente,
composites. Coord. Chem. Rev. 307, 267–291. modulators. Sci. China Mater. 58, 370–377. C., Furukawa, H., Hmadeh, M., Gandara, F.,
Whalley, A.C., Liu, Z., Asahina, S., et al. (2012).
83. Saha, S., Das, G., Thote, J., and Banerjee, R. 95. Maksimchuk, N.V., Timofeeva, M.N., Large-pore apertures in a series of
(2014). Photocatalytic metal–organic Melgunov, M.S., Shmakov, A.N., Chesalov, metal–organic frameworks. Science 336,
framework from CdS quantum dot incubated Y.A., Dybtsev, D.N., Fedin, V.P., and 1018–1023.
luminescent metallohydrogel. J. Am. Chem. Kholdeeva, O.A. (2008). Heterogeneous
Soc. 136, 14845–14851. selective oxidation catalysts based on 107. Feng, D., Liu, T.-F., Su, J., Bosch, M., Wei, Z.,
coordination polymer MIL-101 and transition Wan, W., Yuan, D., Chen, Y.-P., Wang, X.,
84. Cho, K., Han, S.-H., and Suh, M.P. (2016). metal-substituted polyoxometalates. J. Catal. Wang, K., et al. (2015). Stable metal–organic
Copper–organic framework fabricated with 257, 315–323. frameworks containing single-molecule traps
CuS nanoparticles: synthesis, electrical for enzyme encapsulation. Nat. Commun. 6,
conductivity, and electrocatalytic activities for 96. Mukhopadhyay, S., Debgupta, J., Singh, C., 5979.
oxygen reduction reaction. Angew. Chem. Kar, A., and Das, S.K. (2018). A keggin
Int. Ed. 55, 15301–15305. polyoxometalate shows water oxidation 108. Jung, S., Kim, Y., Kim, S.-J., Kwon, T.-H., Huh,
activity at neutral PH: POM@ZIF-8, an efficient S., and Park, S. (2011). Bio-functionalization of
85. Kong, Z.-C., Liao, J.-F., Dong, Y.-J., Xu, Y.-F., and robust electrocatalyst. Angew. Chem. Int. metal–organic frameworks by covalent
Chen, H.-Y., Kuang, D.-B., and Su, C.-Y. (2018). Ed. 57, 1918–1923. protein conjugation. Chem. Commun. 47,
Core@shell CsPbBr3@zeolitic imidazolate 2904–2906.
framework nanocomposite for efficient 97. Buru, C.T., Platero-Prats, A.E., Chica, D.G.,
photocatalytic CO2 reduction. ACS Energy 109. Lyu, F., Zhang, Y., Zare, R.N., Ge, J., and Liu, Z.
Kanatzidis, M.G., Chapman, K.W., and Farha,
Lett. 3, 2656–2662. (2014). One-pot synthesis of protein-
O.K. (2018). Thermally induced migration of a
embedded metal–organic frameworks with
polyoxometalate within a metal–organic
86. Mehta, J.P., Tian, T., Zeng, Z., Divitini, G., enhanced biological activities. Nano Lett. 14,
framework and its catalytic effects. J. Mater.
Connolly, B.M., Midgley, P.A., Tan, J.-C., 5761–5765.
Chem. A 6, 7389–7394.
Fairen-Jimenez, D., and Wheatley, A.E.H.
(2018). Sol–gel synthesis of robust metal– 110. Liang, W., Xu, H., Carraro, F., Maddigan, N.K.,
98. Shi, D., He, C., Qi, B., Chen, C., Niu, J., and Li, Q., Bell, S.G., Huang, D.M., Tarzia, A.,
organic frameworks for nanoparticle Duan, C. (2015). Merging of the
encapsulation. Adv. Funct. Mater. 28, Solomon, M.B., Amenitsch, H., et al. (2019).
photocatalysis and copper catalysis in metal– Enhanced activity of enzymes encapsulated in
1705588. organic frameworks for oxidative C–C bond hydrophilic metal–organic frameworks. J. Am.
formation. Chem. Sci. 6, 1035–1042. Chem. Soc. 141, 2348–2355.
87. Ye, J.-J., and Wu, C.-D. (2016). Immobilization
of polyoxometalates in crystalline solids for 99. Han, Q., Qi, B., Ren, W., He, C., Niu, J., and
highly efficient heterogeneous catalysis. 111. Li, P., Modica, J.A., Howarth, A.J., Vargas,
Duan, C. (2015). Polyoxometalate-based L.E., Moghadam, P.Z., Snurr, R.Q., Mrksich,
Dalton Trans. 45, 10101–10112. homochiral metal–organic frameworks for M., Hupp, J.T., and Farha, O.K. (2016). Toward
tandem asymmetric transformation of cyclic design rules for enzyme immobilization in
88. Du, D.-Y., Qin, J.-S., Li, S.-L., Su, Z.-M., and
carbonates from olefins. Nat. Commun. 6, hierarchical mesoporous metal–organic
Lan, Y.-Q. (2014). Recent advances in porous
10007. frameworks. Chem 1, 154–169.
polyoxometalate-based metal–organic
framework materials. Chem. Soc. Rev. 43,
100. Paille, G., Gomez-Mingot, M., Roch-Marchal, 112. Gkaniatsou, E., Sicard, C., Ricoux, R.,
4615–4632.
C., Lassalle-Kaiser, B., Mialane, P., Fontecave, Benahmed, L., Bourdreux, F., Zhang, Q.,
89. An, H.-Y., Wang, E.-B., Xiao, D.-R., Li, Y.-G., M., Mellot-Draznieks, C., and Dolbecq, A. Serre, C., Mahy, J.-P., and Steunou, N. (2018).
Su, Z.-M., and Xu, L. (2006). Chiral 3D (2018). A fully noble metal-free photosystem Enzyme encapsulation in mesoporous
architectures with helical channels based on cobalt-polyoxometalates metal–organic frameworks for
constructed from polyoxometalate clusters immobilized in a porphyrinic metal–organic selective biodegradation of harmful dye
and copper–amino acid complexes. Angew. framework for water oxidation. J. Am. Chem. molecules. Angew. Chem. Int. Ed. 57, 16141–
Chem. Int. Ed. 45, 904–908. Soc. 140, 3613–3618. 16146.

101. Song, J., Luo, Z., Britt, D.K., Furukawa, H., 113. Li, P., Chen, Q., Wang, T.C., Vermeulen, N.A.,
90. Liu, H.-Y., Wu, H., Yang, J., Liu, Y.-Y., Ma, J.-F.,
and Bai, H.-Y. (2011). Solvothermal assembly Yaghi, O.M., Hardcastle, K.I., and Hill, C.L. Mehdi, B.L., Dohnalkova, A., Browning, N.D.,
(2011). A multiunit catalyst with synergistic Shen, D., Anderson, R., Gómez-Gualdrón,
of a series of organic–inorganic hybrid
stability and reactivity: a polyoxometalate– D.A., et al. (2018). Hierarchically engineered
materials constructed from keggin
polyoxometalate clusters and copper(I)– metal organic framework for aerobic mesoporous metal–organic frameworks
decontamination. J. Am. Chem. Soc. 133, toward cell-free immobilized enzyme systems.
organic frameworks. Cryst. Growth Des. 11,
1786–1797. 16839–16846. Chem 4, 1022–1034.

102. Lykourinou, V., Chen, Y., Wang, X.-S., Meng, 114. Wu, C.-D., and Zhao, M. (2017). Incorporation
91. Fu, H., Qin, C., Lu, Y., Zhang, Z.-M., Li, Y.-G.,
L., Hoang, T., Ming, L.-J., Musselman, R.L., of molecular catalysts in metal–organic
Su, Z.-M., Li, W.-L., and Wang, E.-B. (2012). An
and Ma, S. (2011). Immobilization of MP-11 frameworks for highly efficient heterogeneous
ionothermal synthetic approach to porous
into a mesoporous metal–organic framework, catalysis. Adv. Mater. 29, 1605446.
polyoxometalate-based metal–organic
frameworks. Angew. Chem. Int. Ed. 51, 7985– MP-11@mesoMOF: a new platform for
115. Cohen, S.M. (2012). Postsynthetic methods for
7989. enzymatic catalysis. J. Am. Chem. Soc. 133,
the functionalization of metal–organic
10382–10385.
frameworks. Chem. Rev. 112, 970–1000.
92. Férey, G., Mellot-Draznieks, C., Serre, C.,
Millange, F., Dutour, J., Surblé, S., and 103. Drout, R.J., Robison, L., and Farha, O.K. 116. Wang, C., Liu, D., and Lin, W. (2013). Metal–
Margiolaki, I. (2005). A chromium (2019). Catalytic applications of enzymes organic frameworks as a tunable platform for
terephthalate-based solid with unusually encapsulated in metal–organic frameworks. designing functional molecular materials.
large pore volumes and surface area. Science Coord. Chem. Rev. 381, 151–160. J. Am. Chem. Soc. 135, 13222–13234.
309, 2040–2042.
104. Lian, X., Fang, Y., Joseph, E., Wang, Q., Li, J., 117. Kajiwara, T., Fujii, M., Tsujimoto, M.,
93. Sun, C.-Y., Liu, S.-X., Liang, D.-D., Shao, K.-Z., Banerjee, S., Lollar, C., Wang, X., and Zhou, Kobayashi, K., Higuchi, M., Tanaka, K., and
Ren, Y.-H., and Su, Z.-M. (2009). Highly stable H.-C. (2017). Enzyme-MOF (metal–organic Kitagawa, S. (2016). Photochemical reduction
crystalline catalysts based on a microporous framework) composites. Chem. Soc. Rev. 46, of low concentrations of CO2 in a porous
metal–organic framework and 3386–3401. coordination polymer with a ruthenium(II)–CO
polyoxometalates. J. Am. Chem. Soc. 131, complex. Angew. Chem. Int. Ed. 55, 2697–
1883–1888. 105. Doonan, C., Riccò, R., Liang, K., Bradshaw, D., 2700.
and Falcaro, P. (2017). Metal–organic
94. Xu, X., Lu, Y., Yang, Y., Nosheen, F., and frameworks at the biointerface: synthetic 118. Niu, Z., Bhagya Gunatilleke, W.D.C., Sun, Q.,
Wang, X. (2015). Tuning the growth of metal– strategies and applications. Acc. Chem. Res. Lan, P.C., Perman, J., Ma, J.-G., Cheng, Y.,
organic framework nanocrystals by using 50, 1423–1432. Aguila, B., and Ma, S. (2018). Metal–organic

Matter 1, 57–89, July 10, 2019 87


framework anchored with a Lewis pair as a stability and selectivity between competing 143. Dong, X.-W., Yang, Y., Che, J.-X., Zuo, J., Li,
new paradigm for catalysis. Chem 4, 2587– reaction pathways by the MOF chemical X.-H., Gao, L., Hu, Y.-Z., and Liu, X.-Y. (2018).
2599. microenvironment. Angew. Chem. Int. Ed. 57, Heterogenization of homogeneous chiral
4532–4537. polymers in metal–organic frameworks with
119. Bogaerts, T., Van Yperen-De Deyne, A., enhanced catalytic performance for
Liu, Y.-Y., Lynen, F., Van Speybroeck, V., 130. Uemura, T., Hiramatsu, D., Yoshida, K., Isoda, asymmetric catalysis. Green. Chem. 20, 4085–
and Van Der Voort, P. (2013). Mn- S., and Kitagawa, S. (2008). Sol–gel synthesis 4093.
salen@MIL101(al): a heterogeneous, of low-dimensional silica within coordination
enantioselective catalyst synthesized using nanochannels. J. Am. Chem. Soc. 130, 9216– 144. He, S., Wang, H., Zhang, C., Zhang, S., Yu, Y.,
a ‘bottle around the ship’ approach. 9217. Lee, Y., and Li, T. (2019). A generalizable
Chem. Commun. 49, 8021–8023. method for the construction of
131. Jo, C., Lee, H.J., and Oh, M. (2011). One-pot MOF@polymer functional composites
120. Kockrick, E., Lescouet, T., Kudrik, E.V., synthesis of silica@coordination polymer through surface-initiated atom transfer radical
Sorokin, A.B., and Farrusseng, D. (2011). core–shell microspheres with controlled shell polymerization. Chem. Sci. 10, 1816–1822.
Synergistic effects of encapsulated thickness. Adv. Mater. 23, 1716–1719.
phthalocyanine complexes in MIL-101 for the 145. Schwab, M.G., Senkovska, I., Rose, M., Koch,
selective aerobic oxidation of tetralin. Chem. 132. Rieter, W.J., Taylor, K.M.L., and Lin, W. (2007). M., Pahnke, J., Jonschker, G., and Kaskel, S.
Commun. 47, 1562–1564. Surface modification and functionalization of (2008). MOF@polyhipes. Adv. Eng. Mater. 10,
nanoscale metal–organic frameworks for 1151–1155.
121. Morabito, J.V., Chou, L.-Y., Li, Z., Manna, controlled release and luminescence sensing.
C.M., Petroff, C.A., Kyada, R.J., Palomba, J. Am. Chem. Soc. 129, 9852–9853. 146. Zheng, Y., Zheng, S., Xue, H., and Pang, H.
J.M., Byers, J.A., and Tsung, C.-K. (2014). (2018). Metal–organic frameworks/
Molecular encapsulation beyond the aperture 133. Kou, J., and Sun, L.-B. (2018). Fabrication of graphene-based materials: preparations
size limit through dissociative linker exchange metal–organic frameworks inside silica and applications. Adv. Funct. Mater. 28,
in metal–organic framework crystals. J. Am. nanopores with significantly enhanced 1804950.
Chem. Soc. 136, 12540–12543. hydrostability and catalytic activity. ACS Appl.
Mater. Interface 10, 12051–12059. 147. Jahan, M., Bao, Q., and Loh, K.P. (2012).
122. Li, B., Zhang, Y., Ma, D., Ma, T., Shi, Z., and Ma, Electrocatalytically active graphene–
S. (2014). Metal-cation-directed de novo 134. Uemura, T., Kadowaki, Y., Kim, C.R., porphyrin MOF composite for oxygen
assembly of a functionalized guest molecule Fukushima, T., Hiramatsu, D., and Kitagawa, reduction reaction. J. Am. Chem. Soc. 134,
in the nanospace of a metal–organic S. (2011). Incarceration of nanosized silica into 6707–6713.
framework. J. Am. Chem. Soc. 136, 1202– porous coordination polymers: preparation,
characterization, and adsorption property. 148. Jahan, M., Liu, Z., and Loh, K.P. (2013). A
1205.
Chem. Mater. 23, 1736–1741. graphene oxide and copper-centered metal
123. Zhang, Z., Zhang, L., Wojtas, L., Eddaoudi, M., organic framework composite as a tri-
135. Cirujano, F.G., Luz, I., Soukri, M., Van functional catalyst for HER, OER, and ORR.
and Zaworotko, M.J. (2012). Template-
Goethem, C., Vankelecom, I.F.J., Lail, M., and Adv. Funct. Mater. 23, 5363–5372.
directed synthesis of nets based upon
De Vos, D.E. (2017). Boosting the catalytic
octahemioctahedral cages that encapsulate 149. Ding, M., and Jiang, H.-L. (2018).
performance of metal–organic frameworks for
catalytically active metalloporphyrins. J. Am. Incorporation of imidazolium-based
steroid transformations by confinement within
Chem. Soc. 134, 928–933. poly(ionic liquid)s into a metal–organic
a mesoporous scaffold. Angew. Chem. Int.
124. Alkordi, M.H., Liu, Y., Larsen, R.W., Eubank, Ed. 56, 13302–13306. framework for CO2 capture and conversion.
J.F., and Eddaoudi, M. (2008). Zeolite-like ACS Catal. 8, 3194–3201.
136. Aguila, B., Sun, Q., Wang, X., O’Rourke, E., Al-
metal–organic frameworks as platforms for 150. Qiu, X., Zhong, W., Bai, C., and Li, Y. (2016).
Enizi, A.M., Nafady, A., and Ma, S. (2018).
applications: on metalloporphyrin-based Encapsulation of a metal–organic polyhedral
Lower activation energy for catalytic reactions
catalysts. J. Am. Chem. Soc. 130, 12639– in the pores of a metal–organic framework.
through host–guest cooperation within
12641. J. Am. Chem. Soc. 138, 1138–1141.
metal–organic frameworks. Angew. Chem.
125. Larsen, R.W., Wojtas, L., Perman, J., Int. Ed. 57, 10107–10111.
151. Fujie, K., and Kitagawa, H. (2016). Ionic liquid
Musselman, R.L., Zaworotko, M.J., and 137. Zhang, W., Hu, Y., Ge, J., Jiang, H.-L., and Yu, transported into metal–organic frameworks.
Vetromile, C.M. (2011). Mimicking S.-H. (2014). A facile and general coating Coord. Chem. Rev. 307, 382–390.
heme enzymes in the solid state: metal– approach to moisture/water-resistant metal–
organic materials with selectively organic frameworks with intact porosity. 152. Luo, Q.-x., Ji, M., Lu, M.-h., Hao, C., Qiu, J.-s.,
encapsulated heme. J. Am. Chem. Soc. 133, J. Am. Chem. Soc. 136, 16978–16981. and Li, Y.-q. (2013). Organic electron-rich N-
10356–10359. heterocyclic compound as a chemical bridge:
138. Sachse, A., Ameloot, R., Coq, B., Fajula, F., building a Brönsted acidic ionic liquid
126. Kataoka, Y., Sato, K., Miyazaki, Y., Masuda, K., Coasne, B., De Vos, D., and Galarneau, A. confined in MIL-101 nanocages. J. Mater.
Tanaka, H., Naito, S., and Mori, W. (2009). (2012). In situ synthesis of Cu–BTC (HKUST-1) Chem. A 1, 6530–6534.
Photocatalytic hydrogen production from in macro-/mesoporous silica monoliths for
water using porous material [Ru2(p-bdc)2]n. continuous flow catalysis. Chem. Commun. 153. An, B., Zhang, J., Cheng, K., Ji, P., Wang, C.,
Energy Environ. Sci. 2, 397–400. 48, 4749–4751. and Lin, W. (2017). Confinement of ultrasmall
Cu/ZnOx nanoparticles in metal–organic
127. Li, Z., Rayder, T.M., Luo, L., Byers, J.A., and 139. Kitao, T., Zhang, Y., Kitagawa, S., Wang, B., and frameworks for selective methanol synthesis
Tsung, C.-K. (2018). Aperture-opening Uemura, T. (2017). Hybridization of MOFs and from catalytic hydrogenation of CO2. J. Am.
encapsulation of a transition metal catalyst in polymers. Chem. Soc. Rev. 46, 3108–3133. Chem. Soc. 139, 3834–3840.
a metal–organic framework for CO2
hydrogenation. J. Am. Chem. Soc. 140, 8082– 140. Uemura, T., Yanai, N., and Kitagawa, S. (2009). 154. Guo, W., Lv, H., Chen, Z., Sullivan, K.P.,
8085. Polymerization reactions in porous Lauinger, S.M., Chi, Y., Sumliner, J.M., Lian, T.,
coordination polymers. Chem. Soc. Rev. 38, and Hill, C.L. (2016). Self-assembly
128. Li, Z., Xiao, J.-D., and Jiang, H.-L. (2016). 1228–1236. of polyoxometalates, Pt nanoparticles
Encapsulating a Co(II) molecular photocatalyst and metal–organic frameworks
in metal–organic framework for visible-light- 141. Kong, T., Guo, G., Pan, J., Gao, L., and Huo, Y. into a hybrid material for synergistic
driven H2 production: Boosting catalytic (2016). Polystyrene sulfonate threaded in MIL- hydrogen evolution. J. Mater. Chem. A 4,
efficiency via spatial charge separation. ACS 101Cr(III) as stable and efficient acid catalysts. 5952–5957.
Catal. 6, 5359–5365. Dalton Trans. 45, 18084–18088.
155. Lin, R., Shen, L., Ren, Z., Wu, W., Tan, Y., Fu, H.,
129. Grigoropoulos, A., McKay, A.I., Katsoulidis, 142. Bromberg, L., Su, X., and Hatton, T.A. Zhang, J., and Wu, L. (2014). Enhanced
A.P., Davies, R.P., Haynes, A., Brammer, L., (2014). Functional networks of organic photocatalytic hydrogen production activity
Xiao, J., Weller, A.S., and Rosseinsky, M.J. and coordination polymers: catalysis of via dual modification of MOF and reduced
(2018). Encapsulation of Crabtree’s catalyst in fructose conversion. Chem. Mater. 26, 6257– graphene oxide on CdS. Chem. Commun. 50,
sulfonated MIL-101(Cr): enhancement of 6264. 8533–8535.

88 Matter 1, 57–89, July 10, 2019


156. Yuan, K., Song, T., Wang, D., Zhang, X., Gao, of Pt@ZIF-8/Silo2 composite with 3D- 159. Xu, W., Thapa, K.B., Ju, Q., Fang, Z., and
X., Zou, Y., Dong, H., Tang, Z., and Hu, W. ordered macropores and size-selective Huang, W. (2018). Heterogeneous
(2018). Effective and selective catalysts for catalytic properties. Small Methods 2, catalysts based on mesoporous metal–
cinnamaldehyde hydrogenation: 1800219. organic frameworks. Coord. Chem. Rev. 373,
hydrophobic hybrids of metal–organic 199–232.
frameworks, metal nanoparticles, and micro- 158. Shen, K., Zhang, L., Chen, X., Liu, L., Zhang, D.,
and mesoporous polymers. Angew. Chem. Han, Y., Chen, J., Long, J., Luque, R., Li, Y., 160. Fang, R., Tian, P., Yang, X., Luque, R., and Li, Y.
Int. Ed. 57, 5708–5713. et al. (2018). Ordered macro- (2018). Encapsulation of ultrafine metal-oxide
microporous metal–organic nanoparticles within mesopores for biomass-
157. Tian, P., Shen, K., Chen, J., Fan, T., Fang, R., framework single crystals. Science 359, derived catalytic applications. Chem. Sci. 9,
and Li, Y. (2018). Self-templated formation 206–210. 1854–1859.

Matter 1, 57–89, July 10, 2019 89

You might also like