You are on page 1of 29

Review

Metal-Organic Frameworks
for Energy Applications
Hailong Wang,1 Qi-Long Zhu,1 Ruqiang Zou,2,* and Qiang Xu1,*

Recently, the development of clean sustainable energy storage and conversion The Bigger Picture
technologies to deal with environmental pollution and the forthcoming energy Metal-organic frameworks
crisis has attracted much attention in the energy research community. It is (MOFs), also known as porous
critical to develop carriers to store energy or to facilitate mass and electron coordination polymers (PCPs),
transportation in energy storage and conversion. The emerging metal-organic have attracted great interest
frameworks (MOFs) are well suited for this purpose because of their inherent because of their unique porous
advantages, including structural diversity, functionality, tailorability, and versa- structures, synthetic advantages,
tile applications. Moreover, when utilized as supports and sacrificial precursors, organic-inorganic hybrid nature,
MOFs can immobilize active functional materials and create highly controllable and versatile applications.
nanostructures, respectively, gaining new momentum for energy applications. Recently, the applications of
In this review, we present the recent progress in the field of energy storage MOFs in energy fields such as fuel
and conversion using MOFs and their composites and derivatives by focusing storage, photo-induced hydrogen
on the correlation of structure, composition, and function. evolution, fuel cells, batteries, and
supercapacitors have
INTRODUCTION experienced a new surge of
After the economic boom, the global demand for energy resources has increased interest in both the chemistry and
continuously and dramatically, giving rise to urgent global concerns on the sustain- materials science communities.
ability of fuel reserves and environmental problems, including climate change, Research on the various
melting glaciers, and rise in sea levels. The development of renewable, safe, clean, applications of MOFs has shown
and sustainable energy storage and conversion technologies has thus become a hot that they are promising porous
research topic. Considerable efforts have been devoted to mitigating the strong reli- materials for energy storage and
ance on fossil fuel and solving environmental issues.1–5 Hydrogen is considered an conversion technologies.
ideal energy carrier, discharging no harmful chemicals while supplying energy.1,2 Furthermore, MOFs have been
In addition, methane has been suggested as a transition fuel for clean and renewable used as support substrates to
energy for the future,1 because it is abundant in nature and yields relatively environ- accommodate metals, metal
mentally friendly combustion products. In the proposed alternative energy cycle, the oxides, semiconductors, and
energy sources are generated by photo- and electro-driven management of water complexes and have been used as
and CO2.4 Meanwhile, a variety of technologies such as solar cells, fuel cells, super- sacrificial materials for the
capacitors, and batteries have been proposed to address the foreseeable severe generation of various
energy issues.5 Therefore, confidence in overcoming the energy crisis and the envi- nanostructures for energy
ronment pollution has been strengthened considerably. applications. Here, we present the
latest research to highlight energy
During the development of innovative energy technologies, porous materials, which applications based on MOFs and
are able to store energy carriers or to facilitate fast mass and electron transportation their composites and derivatives.
for energy storage and conversion, have been explored extensively to identify
the best materials for photo- and electrochemical energy applications. Suitable sur-
face area, rational pore-size distribution, desired morphology, and superior func-
tionality are the most important factors for mass and electron transportation and
substance reaction kinetics, thus determining the performance of energy devices.
In comparison with traditional inorganic porous materials, metal-organic frame-
works (MOFs) have a highly porous structure, uniform spatial dispersion of compo-
nents, tunable pore sizes and topologies, a hybrid organic-inorganic nature, and

52 Chem 2, 52–80, January 12, 2017 ª 2017 Elsevier Inc.


multifunctionality,6,7 which enable them to outperform other porous materials in
energy applications. Furthermore, the well-defined structure is helpful for the direct
and exact rationalization of the related structure-property relationships, which is an
important aspect for optimizing materials.

In recent years, the spectacular development of MOFs has drawn wide attention
from many fields and is stimulating the emergence of an innovation wave of conven-
tional energy applications. The excellent gas storage capacity of MOFs at ambient
temperature has driven their rapid growth, providing a large number of new adsor-
bents with potential uses in vehicle gas tanks, fuel cells, and stationary power facil-
ities.1,8–12 MOFs with photo- and electroactive ligands and metal ions can be used as
energy acceptors or catalytic sites for boosting energy-related chemical reactions in
alternative energy cycles.4,13–17 The advent of nanotechnology and modern charac-
terization facilities has led to new and efficient approaches to rationally design
and modify these materials.18–20 Nanoscale MOFs with controllable size and
morphology have opened up new avenues for traditional MOF applications. More-
over, a variety of functional materials have been integrated with MOFs for the pur-
pose of affording multifunctional composites and even new advanced materials.19,20
These MOF composites serve as new platforms for the exploration of energy appli-
cations. In the case of thermal treatment, the MOF-template approach guarantees
efficient pyrolysis and uniform distribution of porosity in MOF derivatives, including
porous carbons, metals, metal oxides, metal sulfides, and multicomponent compos-
ites, extending the application potential of MOFs.21 After more than 20 years of
exploration of the syntheses and fundamental properties of MOFs, their energy ap-
plications have become a hot topic.

The remarkable functionalities of MOFs, MOF composites, and MOF derivatives


have attracted a surge of interest and investment, stimulating the emergence of
innovative materials for energy applications. This review provides an overview of
the recent progress in a wide range of energy-related applications of MOFs and their
composites and derivatives. We first summarize the components, structures, and
unique advantages of these materials for energy storage and conversion, and sub-
sequently provide an up-to-date appraisal of the various energy applications of
these materials, including fuel storage, photo-induced hydrogen evolution and car-
bon dioxide reduction, solar cells, fuel cells, supercapacitors, lithium-based batte-
ries, and electrolysis of water. Representative syntheses and energy applications
of MOFs, MOF composites, and MOF derivatives are discussed. We hope that this
review and the cited articles will inspire chemists and material scientists to explore
MOFs, MOF composites, and MOF derivatives for energy applications.

MOFS, MOF COMPOSITES, AND MOF DERIVATIVES


MOFs
MOFs, also known as porous coordination polymers (PCPs), came on the scene more
than 20 years ago, and work is ongoing in terms of the continuously increasing
1Research Institute of Electrochemical Energy,
numbers and applications.6,7 This new kind of crystalline material is made up of inor-
National Institute of Advanced Industrial Science
ganic secondary building units (SBUs, metal ions or clusters) and single or multiple and Technology (AIST), Ikeda, Osaka 563-8577,
organic linkers connected by coordination interactions.6 The structures and proper- Japan
2Department of Materials Science and
ties of MOFs are easily tailored by judiciously changing the species, geometry, size,
Engineering, College of Engineering, Peking
and functionality of the modules in pre-design or post-synthetic modification (PSM). University, Beijing 100871, China
Thus far, advances in syntheses and PSMs have led to more than 20,000 MOFs6 and *Correspondence: rzou@pku.edu.cn (R.Z.),
some fundamental breakthroughs in the fields of sensing22 and molecule separa- q.xu@aist.go.jp (Q.X.)
tion.23 The high surface area (the highest value, 10,000 m2 g1 [Langmuir]), http://dx.doi.org/10.1016/j.chempr.2016.12.002

Chem 2, 52–80, January 12, 2017 53


Figure 1. Schematic Illustration of MOFs, MOF Composites, and MOF Derivatives as well as Their
Conversions

controllable pore size (from a few Angstroms to 98 Å), and low density (the lowest
value, 0.13 g cm3) allow MOFs to act in inherent applications as adsorbents for
gas storage or as reaction containers to selectively catalyze some organic reac-
tions.6,7 Moreover, nano-MOFs with controllable size and morphology represent a
new direction for traditional pristine MOFs.18 In recent years, applications of
MOFs in energy storage and conversion, such as photocatalytic hydrogen evolution,
fuel cells, batteries, and supercapacitors, have attracted much interest in both the
chemistry and materials science communities.24 The research on MOF applications
demonstrates that MOFs are excellent porous materials for energy storage and con-
version. In addition to direct utilization, MOFs have also been used either as support
substrates to integrate functional metals, metal oxides, semiconductors, and com-
plexes or as sacrificial materials to generate various nanostructures (Figure 1).20,21

MOF Composites
Integration of MOFs with a variety of functional materials is a very effective and
feasible strategy to further improve MOF performance or introduce new function-
ality for practical use.20 Thus far, numerous MOF composites have been successfully
prepared by assembling MOFs and functional species, including graphene, carbon
nanotubes (CNTs), metal nanoparticles and nanorods, metal oxides, complexes, and
even enzymes, and have shown striking performance in catalysis,25 photo-induced
H2 generation,14 proton conduction,26 and so on. In these MOF composites, the in-
dividual functions of the MOFs and functional materials synergistically merge
together not only to provide multifunctionality as a whole but also to generate
new physical and chemical properties that are not present in the individual

54 Chem 2, 52–80, January 12, 2017


components. The combination of MOFs and other functional materials has extended
their applications. Moreover, research on high-performance MOF hybrids with
sophisticated architectures, in combination with enrichment of the MOF database,
has led to new design tactics for MOF composites.

MOF Derivatives
It is well known that porous carbons, metals, metal oxides, metal sulfides, and their
multicomponent hybrids are important inorganic materials in the field of nano-
science, with potential for energy and environmental applications. Nanoporous car-
bon (NPC) materials have a high surface area and large pore volume as well as
outstanding chemical and mechanical stability; their applications are determined
by the pore structure and micromorphology.21 Because MOFs have a highly ordered
pore structure and abundant organic components, their carbonization successfully
leads to NPC structures with high surface area and narrow pore-size distribution.21
In the MOF-to-carbon (MTC) process with27–30 or without29–32 a secondary carbon
source, the heating rate, calcination temperature, and MOF crystal-size and compo-
nents are vital to afford carbon materials with tunable structures and functionality. In
a recent report, a KOH-assisted sonochemical treatment of as-synthesized MOF-
derived carbon nanorods followed by thermal activation produced two- to six-
layered graphene nanoribbons with excellent supercapacitor performance.33 This
opens up a new perspective for developing PSM methods for MOF-derived carbon
materials.

In addition, numerous synthetic methodologies have been developed to prepare


inorganic materials of metals, metal oxides, metal sulfides, etc. with controllable
size and morphology. Fabrication of these nanostructured materials derived from
the pyrolysis of MOFs has been demonstrated as an effective route.21,24 In contrast
to other methods reported previously, MOF precursors are obtained from simple
and low-cost processing procedures, and the size and morphology of the resulting
inorganic nanomaterials are controllable because of the unique roles of MOFs as
templates and precursors.21 MOF pyrolysis is also facile to prepare metals, metal
oxides, metal sulfides, and their composites effectively wrapped by heteroatom-
doped carbon or a highly graphitic carbon layer, with significantly enhanced photo-
and electrochemical activity and durability.34,35 MOF composites have also been
used as pyrolysis precursors to effectively control the morphology and enhance
the conductivity of MOF derivatives.21 Briefly, MOF thermolysis guarantees efficient
carbonization of MOF precursors and uniform porosity of MOF derivatives, including
porous carbons, metals, metal oxides, metal sulfides, and their hybrids. This origi-
nates from the unique crystalline structures and orderly dispersed organic-inorganic
ingredients of MOF precursors and the porosity of MOF composites.

HYDROGEN AND METHANE STORAGE


Physical Adsorption Storage
Hydrogen Storage
Hydrogen is an ideal clean-energy carrier, and hydrogen storage materials with high
gravimetric and volumetric densities are prerequisites for practical on-board appli-
cation. Successional US Department of Energy (DOE) hydrogen programs have
significantly accelerated hydrogen storage research, which is also true for the devel-
opment of MOFs.1,8,9 The first discovery of MOF-5 (Figure 2A) with hydrogen stor-
age capability was reported in 2003.8 Since then, porous MOFs have attracted
increasing attention as hydrogen adsorbents, and the number of MOFs for high
hydrogen storage has increased significantly. Thus far, the highest hydrogen storage
capacity with total H2 gravimetric uptake of 17.6 wt % at 80 bar and 77 K is still

Chem 2, 52–80, January 12, 2017 55


Figure 2. Crystal Structures of Four Representative MOFs
(A) MOF-5.
(B) HKUST-1.
(C) Mg-MOF-74.
(D) ZIF-8.
C, gray; O, red; and N, blue. The yellow ball represents the porosities, and the yellow line shows the
topology of ZIF-8.

exhibited by MOF-210 with the highest surface area.9 Although the overall gravi-
metric H2 storage capacity (77 K and high pressure) has been estimated to be pro-
portional to MOF surface areas,1 this cannot be enhanced infinitely because of the
synthetic difficulty in achieving MOFs with much higher surface area. In addition,
the hydrogen storage performance of MOFs can be boosted by introducing open
metal sites (OMSs), alkaline or alkaline-earth metal ions (such as Li+ and Mg2+), metal
nanoparticles, and small pores.1 Although these MOF-based adsorbents are able to
achieve the desired hydrogen capacity at low temperature, it is still a challenge to
reach the DOE hydrogen storage target at ambient temperature.

Methane Storage
Methane, the main component of natural gas, has the highest hydrogen to carbon
ratio among all hydrocarbons together with a high octane number of 107. It is crucial
to develop safe, cheap, and convenient technologies for high-capacity methane
storage for implementation in natural-gas-fueled vehicles. The emerging porous
MOFs are promising candidates for adsorbed natural gas (ANG) technology. It is
well known that the pore volume and Brunauer-Emmett-Teller (BET) surface area,
in combination with pore size, play vital roles in the physical adsorption of methane.1
The total gravimetric uptake of MOFs is proportional to the pore volume and BET
surface area under high pressure, and the proper pore size will be significantly help-
ful for enhancing total methane storage capacity.10 The introduction of OMSs and

56 Chem 2, 52–80, January 12, 2017


functional groups is able to improve methane binding affinity and thus increases the
volumetric methane uptake.1,7 With a few exceptions, the former strategy is usually
much more effective than the latter. However, the introduction of OMSs often results
in an undesired increase in adsorbent density and a quick increase in gas uptake at
low pressure (0–5 bar), causing limited and even negative enhancement of working
capacity (the working capacity represents the total uptake difference between the
pressure at 5 bar and a higher pressure).

There is an unexpected increase in the volumetric uptake when the benzene ring of
NOTT-101 is replaced with the pyrimidine ring of UTSA-76.11 This MOF exhibits an
extraordinarily high volumetric uptake of 257 cm3 (STP) cm3 (298 K and 65 bar) and
a high working capacity of 197 cm3 (STP) cm3. The noticeably maximized uptake
capacities are associated with the occasional optimal combination of Lewis-basic
N sites and the dynamic freedom of pyrimidine groups. Long’s group described a
new strategy to use flexible MOFs, which undergo single or multiple phase transi-
tions in response to the exoteric stimulation, as adsorbents to maximize the working
capacity of CH4.12 In this regard, {Co(bdp)}n (bdp2 = 1,4-benzenedipyrazolate)
makes use of gate-opening behavior to realize a high working capacity of 197 cm3
(STP) cm3 for methane storage.12 Although the methane storage capacity of
MOF materials is favorable, in terms of volumetric and gravimetric storage or work-
ing capacities, relative to other porous materials, it is still challenging to provide
desired adsorbents for commercially available ANG technology. In the next step,
cooperative efforts from both engineering science and materials science are
required to exploit the available methane storage techniques and harvest high-per-
formance adsorbents.

Chemical Hydrogen Storage


Chemical hydrogen storage is defined as the storage of hydrogen in chemical
bonds, which is promising for fuel-cell applications. It is also a safe and efficient alter-
native to physical hydrogen storage. The chemical storage of hydrogen in solid and
liquid states has been extensively investigated in the past few decades. MOFs are
able to confine chemical hydrides in their nanopores, releasing hydrogen under
mild conditions with less undesirable volatile byproducts. Moreover, MOF-sup-
ported metal nanoparticle catalysts enhance the hydrogen liberation kinetics of
liquid-phase chemical hydrides.3,36–41 A broad overview of these two aspects is
given below.

Nanoconfinement of Chemical Hydrides


Nanoconfinement of chemical hydrides in the pores of MOFs is of great interest for
chemical hydrogen storage, changing the thermodynamics of chemical hydrides
and significantly enhancing dehydrogenation kinetics with suppressed evolution
of undesirable volatile byproducts.36–38 The nanoconfinement of chemical hydrides
in MOFs was first attempted with HKUST-1 (Figure 2B) to confine solid NaAlH4.36
A comparative study on the H2 desorption behavior of NaAlH4@HKUST-1 and
NaAlH4 revealed that the former species desorbs H2 at a much lower temperature
(70 C) than the latter (250 C). With the aim of increasing the hydrogen release
rate and suppressing volatile byproduct formation in the dehydrogenation process,
a metal-nanoparticle-embedded MOF scaffold, Pt@MIL-101, was used to confine
ammonia borane (AB).38 The comparative pyrolysis of neat AB, AB@MIL-101,
and AB@Pt@MIL-101 revealed that only AB@Pt@MIL-101 generates hydrogen
without any volatile byproduct. In addition, the dehydrogenation temperature of
AB@Pt@MIL-101 is much lower than that for neat AB. In this context, the catalytically
active metal nanoparticles (NPs) coupled with the nanoconfinement effect play a

Chem 2, 52–80, January 12, 2017 57


Figure 3. Schematic Illustration of Immobilization of the Au/Ni Nanoparticles by the MIL-101
Matrix by a Double-Solvent Method Combined with a Liquid-Phase Concentration-Controlled
Reduction Strategy
Reprinted with permission from Zhu et al. 41 Copyright 2013 American Chemical Society.

significant role in preventing generation of volatile byproducts. In these chemical hy-


dride-encapsulated MOFs, the outstanding nanoconfinement effect is closely asso-
ciated with the interaction between the chemical hydride molecules and the pores of
MOFs. Therefore, the rational selection of MOFs with pore sizes and interaction sites
matching chemical hydride molecules is crucial for strengthening the pore-substrate
interaction. In addition to effective nanoconfinement, immobilization of catalytically
active metal NPs is also a good strategy to suppress the generation of volatile
byproducts.

Catalysis of Liquid-Phase Chemical Hydrides for Hydrogen Generation


MOFs participate in chemical hydrogen storage as supports, catalysts, or sacrificial
precursors of catalysts to enhance hydrogen liberation kinetics.2,39–41 In 2011, bime-
tallic AuPd NPs were embedded in MIL-101 by a conventional impregnation method
to dehydrogenate formic acid at a convenient temperature.39 After grafting ethyle-
nediamine (ED) onto MIL-101, the hydrogen liberation rate over the AuPd@PED-
MIL-101 catalyst was significantly improved as a result of the incorporation of a
larger amount of Pd precursor during the impregnation process and the formation
of smaller metal particles inside MOF pores. Many groups have endeavored
improving the dehydrogenation of chemical hydrides by using MOF-supported
metal NP catalysts.25,40–42 Moreover, some new synthetic strategies have been es-
tablished to effectively embed metal NPs with controllable particle size, ingredients,
and location of MOF pores. For example, the double-solvent method,38,41 which in-
volves a large amount of hydrophobic solvent to disperse MOFs and a small amount
(less than the total MOF pore volume) of hydrophilic solvent to dissolve metal pre-
cursors, enables MOFs to completely encapsulate the metal precursors into hydro-
philic pores. A liquid concentration-controlled reduction, in cooperation with the
double-solvent method, has been conducted under mild conditions to control the
location of AuNi alloy NPs in the interior pores of MIL-101 (Figure 3).41 The MOF-
supported AuNi alloy NP catalyst synergistically promotes the hydrolysis of AB. Its
turnover frequency (TOF) value of 66.2 molH2 molcat min1 is much higher than
that of non-noble-metal catalysts and even higher than that of some noble-metal

58 Chem 2, 52–80, January 12, 2017


Figure 4. Photocatalytic H2 Evolution over a POM-Implanted MOF Catalyst
(A) Schematic illustration showing synergistic visible-light excitation of the MOF framework and
multi-electron injection into the encapsulated POMs.
(B) Time-dependent HER TONs of a MOF catalyst with methanol as the sacrificial electron donor.
Reprinted with permission from Zhang et al. 46 Copyright 2015 American Chemical Society.

catalysts. Thus far, a large number of different MOF-supported metal NPs have been
used as catalysts for the dehydrogenation of various liquid-phase chemical
hydrides.42

SOLAR ENERGY CONVERSION


Photocatalytic Hydrogen Production
The evolution of hydrogen from water in the presence of a photocatalyst to facilitate
the displacement of fossil fuel is a hot topic in the research community. In 2009, the
first MOF photocatalyst, {Ru2(1,4-BDC)2}n (1,4-BDC = 1,4-benzenedicarboxylic acid),
was used to induce hydrogen evolution from water under the irradiation of visible
light with the help of Ru(bpy)32+ (as a photosensitizer), MV2+ (as an electron relay),
and EDTA-Na2 (as a sacrificial donor),43 where bpy, MV, and EDTA are 2,20 -bipyri-
dine, N,N0 -dimethyl-4,40 -bipyridinium, and EDTA, respectively. This MOF photoca-
talyst displays a turnover number of 8.16 and a quantum yield of 4.82%. Since this
discovery, the photolytic activity of various MOFs with catalytically active sites has
been examined.4,44

Initially, a new strategy in the photocatalysis of water splitting was envisaged by


impregnating MOFs with noble-metal NPs to enhance the efficiency of hydrogen
generation.14,45 Subsequently, an extended methodology, namely the implanta-
tion of functional materials into photosensitizer-based MOFs constructed from
[Ir(ppy)2(bpy)]+- (ppy = 2-phenylpyridine) or [Ru(bpy)3]2+-derived organic ligands,
has been developed.4,4648 Single or multiple catalytic modules, such as amor-
phous C3N4, metals, metal oxides, metal complexes, and polyoxometalates
(POMs), have been integrated into MOFs by either direct synthesis or PSM for
the photocatalytic reaction (Figure 4). An interesting composite, which is made
up of water-soluble [Ru(bpy)3]2+-based supramolecular MOF (SMOF-1) absorbing
anionic Wells-Dawson-type polyoxometalate (WD-POM) [P2W18O62]6– clusters, is
capable of boosting photo-induced H2 evolution.47 The corresponding catalytic
performance is about four times higher than that of heterogeneous WD-POM@
[Ru(bpy)3]2+ MOF catalysts. The excellent catalytic activity is attributed to the
uniform impregnation of the pores of SMOF-1 with WD-POM clusters, which is
beneficial to mass and electron transport in the H2 evolution process. However,
molecular catalysts, even immobilized in the pores of MOFs, are often prone to
transform into metal colloids, which decreases their catalytic efficiency in artificial

Chem 2, 52–80, January 12, 2017 59


photosynthesis. As a result, it is of great significance to tackle the stability of
molecular catalysts. In this context, a water-stable bipyridine-decorated UiO-67
has been selected as a self-healing scaffold,48 and it displaying outstanding self-
healing behavior.

MOF derivatives have been applied in the evolution of photo-induced hydrogen


from water. Nanoscale MIL-101 octahedral particles have been coated with amor-
phous titania shells in a controllable thickness.34 After calcination, the resulting par-
ticles have been converted to core-shell nanostructures with an iron oxide (a-Fe2O3)
core and a TiO2 (anatase phase) shell. Under the irradiation of visible light and with
the help of K2PtCl4, the amount of H2 generated from water is increased in linear pro-
portion to time, reaching 30.0 mmol mg1 after 48 hr. The present synergistically
photophysical properties may originate from the p-n heterojunction effect. Almost
all these photocatalytic reactions have taken place in the presence of noble-metal
particles, ions, or complexes, increasing the cost of hydrogen production. These
non-precious-metal photocatalysts are therefore highly desirable. A non-noble-
metal catalyst, which is made up of a cobalt complex based on N2,N20 -propanediyl-
bis(2,3-butanedione-2-imine-3-oxime) ligand encapsulated in the pores of a
photoactive MOF NH2-MIL-125(Ti), exhibit excellent photocatalytic activity for
hydrogen generation from water.49 In comparison with the homogeneous Co-based
molecular catalyst and the pristine NH2-MIL-125(Ti) catalyst, the photocatalytic
activity of the new Co@MOF composite is remarkably enhanced. Moreover, a
constant TOF of 0.8 has been maintained by Co@MOF for 65 hr, confirming its
high stability under the present conditions. In this artificial Co@MOF catalysis sys-
tem, diffusion limitations are not observed, and the charge transfer is still efficient
even at high molecular catalyst loadings, which is responsible for the high photoca-
talytic activity.

Photocatalytic Carbon Dioxide Reduction


CO2 reduction is of great interest because the products, such as CO, HCOOH,
MeOH, and CH4, are regarded as energy carriers and are important for industrial
applications. In addition, the conversion of carbon dioxide into fuels is a promising
solution to the environment problem caused by excessive CO2 emission. In this
respect, researchers are making use of solar energy and electrical energy to induce
CO2 activation and conversion into fuels in the presence of MOF-based catalysts.4,13
Here, we review the main progress achieved by using MOF materials to promote
photo-driven CO2 reduction. The first MOF-based catalyst for photochemical CO2
reduction was synthesized by direct incorporation of a catalytic component
[ReI(CO)3(dcbpy)Cl] (dcbpy = 2,20 -bipyridine-5,50 -dicarboxylic acid) into highly
stable porous UiO-67.50 The photocatalytic CO2 reduction performance of
[ReI(CO)3(dcbpy)Cl]-doped UiO-67 is much higher than that of the homogeneous
[ReI(CO)3(dcbpy)Cl] catalyst. The photo-induced CO2 reduction over the MOF cata-
lyst occurs exclusively via a unimolecular mechanism as a result of the uniform
dispersion of the [ReI(CO)3(dcbpy)Cl] active sites in the MOF framework. In order
to efficiently capture and use light energy, an amino-functionalized MOF, NH2-
MIL-125(Ti), has been designed and prepared.51 Comparison of the photocatalytic
activity of NH2-MIL-125(Ti) and the parent MIL-125(Ti) under similar conditions cer-
tifies that the present visible-light-driven photocatalysis is associated with amino
functionalization. A porphyrin-based MOF catalyst, PCN-222, shows excellent activ-
ity in CO2 reduction under visible-light irradiation (Figure 5).15 Mechanistic informa-
tion collected from ultrafast transient absorption spectroscopy and time-resolved
photoluminescence spectroscopy has revealed the presence of long-lived electron
traps in PCN-222, which effectively inhibits the recombination of electrons and holes

60 Chem 2, 52–80, January 12, 2017


Figure 5. Photo-Induced CO2 Reduction over a MOF Catalyst
(A) 3D structure of PCN-222.
(B) Gas sorption isotherms of PCN-222.
(C) Mott-Schottky plots for PCN-222.
(D) Time-dependent amount of HCOO  under visible-light irradiation.
Reprinted with permission from Xu et al.15 Copyright 2015 American Chemical Society.

and therefore greatly enhances the photocatalytic conversion of CO2 into formate
anion. Such a deep understanding of the interior mechanism of CO2 reduction is a
key step for obtaining high-efficiency catalysts. These three examples involve
direct integration of catalytically active sites into MOFs. Alternatively, Wang
et al.52 established an artificial CO2 photoreduction system by using Co-ZIF-9 and
[Ru(bpy)3]Cl2$6H2O as a cocatalyst and a photosensitizer, respectively. Although
these heterogeneous MOF-based catalysts exhibit striking photo-induced CO2
reduction performance, they are usually implemented at high temperature and
high CO2 pressure, which is unfavorable for practical use. An interesting zirconium
MOF based on {[RuII(H2bpydc) (terpy) (CO)](PF6)2}n (bpydc = 2,20 -bipyridine-5,50 -di-
carboxylate and terpy = 2,20 :60 ,200 -terpyridine) exhibits the highest photocatalytic
activity among all MOF catalysts.53 Even when the CO2 concentration is gradually
decreased to 5%, the MOF-based catalyst maintains constant catalytic efficiency.
This investigation highlights the potential of MOF-based catalysts for practical
photo-induced reduction.

The photoactivities of MOF catalysts in photo-induced CO2 reduction are still not
comparable with those of inorganic semiconductor species, which motivates us to
further develop durable, low-cost, and high-performance MOF catalysts. The inte-
gration of semiconductors and MOFs is another promising solution to enhance the
efficiency of exciton generation and charge separation. In comparison with bare
TiO2, a {Cu3(BTC)2}n@TiO2 (BTC = 1,3,5-benzenetricarboxylate) hybrid with a
core-shell structure displays both improved CO2 conversion efficiency and
outstanding stability.54 The high selectivity of carbon dioxide conversion to
methane implies that the hybrid core-shell material harvests much higher electron

Chem 2, 52–80, January 12, 2017 61


density than bare TiO2 and thus exhibits improved electron-hole separation. The
rational assembly of light-harvesting and catalytically active functional materials
generates a new catalytic system, in which the thermal effect acquired from
the intense absorption of visible light and infrared radiation is used to
promote the reaction. The synergistic photocatalytic and thermocatalytic process
paves the way for light-driven CO2 conversion and is expected to trigger a new
photocatalyst design strategy. Also, the utilization of H2 to execute the photocata-
lytic CO2 conversion, which is denoted as CO2 hydrogenation, is very effective and
feasible. The photocatalytic activity of a MOF-derived Fe@C catalyst has been
evaluated in a batch-type reaction system using a mixture of CO2 and H2 in a
1:1 molar ratio as the reactant.55 Compared with Fe@SiO2 and Fe@CNT catalysts,
the amount of CO over Fe@C catalyst is greatly enhanced under the same exper-
imental conditions. The corresponding light-driven CO2 reaction rate is much
higher than that of other CO2 hydrogenation systems using H2O or H2 as hydrogen
sources.

The above two sections describe different artificial photocatalytic systems for
hydrogen evolution and CO2 reduction based on MOFs, MOF composites, and
MOF derivatives as catalysts, co-catalysts, or sensitizers. The corresponding photo-
catalytic activities have been evaluated, and the intrinsic mechanisms have been
studied in depth. However, the necessary addition of additives, including sensitizers
and sacrificial reagents, and need for noble-metal species increase the cost of
hydrogen production and CO2 reduction, limiting practical use. In addition, the
sacrificial reagents of organic amines will shorten the lifetime of MOF catalysts.
Therefore, stable high-efficiency non-precious-metal catalysts and simple artificial
photocatalytic systems involving only the reaction substrates and catalysts should
be developed.

Photoelectric Conversion
Solar cells, also known as photovoltaic cells, are electrical devices for direct conver-
sion of solar energy to electricity, which is another intriguing application of MOFs.
The regular arrangement of photoactive molecules is an important factor in deter-
mining solar cell performance. A highly porous, crystalline, and monolithic
porphyrin-based MOF thin film of Zn-SURMOF2 has been fabricated on a conduc-
tive fluorine-doped tin oxide (FTO) substrate (as the bottom anode) by a liquid-
phase epitaxy method.56 The MOF film was covered by an iodine and triiodine
electrolyte (I– and I3– in acetonitrile) for assembling a solar cell by using Pt as a
cathode. The assembled photovoltaic cell shows a current-voltage (I-V) curve with
Voc = 0.57 V, Jsc = 0.45 mA cm2, and an efficiency (h) of 0.2%. After immobilization
of Pd(II) ions in the porphyrin cores of Zn-SURMOF2, the photophysical property
of the modified solar cell was significantly improved with Voc = 0.70 V, Jsc =
0.71 mA cm2, and h = 0.45%. This result is very impressive among single-compo-
nent photovoltaic solar cells. The electronic structure calculations indicate that the
Zn-SURMOF-2 arrays can be classified as an indirect band-gap semiconductor. This
inhibits direct electron-hole recombination, leading to superior photophysical
properties. MOF applications in photovoltaic solar cells have been limited mainly
by their poor semiconductive properties. Notably, MOFs have another interesting
application in solar cells. MOF-525 nanocrystals about 140 nm in size have been
added to the precursor solution of CH3NH3PbI3-xClx perovskite for thin film depo-
sition.57 The MOF@perovskite solar cell shows enhanced photovoltaic activity in
comparison with that of the pristine perovskite thin film cell, which is attributed
to the better morphology or crystallinity of the perovskite phase induced by the
MOF additive.

62 Chem 2, 52–80, January 12, 2017


Figure 6. MOF-Based Proton Conductors
(A) Honeycomb layer structure of {(NH 4 ) 2 (adp)[Zn 2 (ox) 3 ]$3H2 O} n without showing NH 4 +, solvent
molecules, or hydrogen atoms. Reprinted with permission from Sadakiyo et al.58 Copyright 2009
American Chemical Society.
(B) The crystal structure of MIL-53(Fe and Al). Reprinted with permission from Shigematsu et al. 60
Copyright 2011 American Chemical Society.
C, light blue; O, red. The blue atoms show functional groups (–NH 2 , –OH, or –COOH).

ELECTRICAL ENERGY STORAGE AND CONVERSION


Fuel Cells
Fuel cells are of great importance among energy storage and conversion technolo-
gies, serving as electrochemical devices to convert fuels (e.g., hydrogen, natural gas,
and methanol) to electricity for powering vehicles, stationary facilities, and portable
appliances. Nevertheless, the current fuel cells are still some distance away from the
price, efficiency, and durability targets proposed by the US DOE in 2016 for the
development of fuel-cell technology. This requires that we develop the present
electrode catalysts and electrolyte membranes further and optimize the ancillary
facilities of fuel cells. In the field of fuel cells, MOFs, MOF composites, and MOF
derivatives are potential electrolyte materials and electrode catalysts.

Proton Conduction at Low Temperature


Polymer electrolyte membrane fuel cells (PEMFCs) have attracted increased atten-
tion in view of their promising on-board applications. Nafion (a sulfonated fluoropol-
ymer) is a popular PEM material as a result of its high proton conductivity and
chemical and thermal stability. However, the corresponding operating conditions
require a temperature below 80 C and a humidified environment, leading to limita-
tions in efficiency improvement and utilization.26 As a consequence, the develop-
ment of new electrolyte materials with high proton conductivity is critical. The
emerging crystalline MOFs with high proton conductivities and low-cost syntheses
are promising for electrolytes in fuel cells. There are two types of proton conduction
of MOFs: low temperature (below 100 C and under humid conditions) and high tem-
perature (above 100 C and under anhydrous conditions).

In the low-temperature region, the first proton-conducting MOF, {(NH4)2(adp)


[Zn2(ox)3]$3H2O}n (ox = oxalic acid, adp = adipic acid), with a performance compa-
rable with that of Nafion, was reported in 2009; water molecules, NH4+ ions, and
carboxyl groups of adipic acid served as conducting media (Figure 6A).58 This
MOF shows proton conductivity as high as 8 3 103 S cm1 at 25 C under 98% rela-
tive humidity (RH). This MOF has a higher activation energy (Ea) of 0.63 eV than
Nafion (0.22 eV), indicating that the proton conduction originates from both Grot-
thus and vehicle mechanisms. Using two-dimensional (2D) porous oxalate-based

Chem 2, 52–80, January 12, 2017 63


frameworks as hosts had enabled implantation of different proton carriers into the
pores of MOFs for exploration of high proton conductors. Among these oxalate-
based MOF proton conductors, {[(Me2NH2)3(SO4)]2[Zn2(ox)3]}n is a rare example dis-
playing high conductivity in both low-temperature and high-temperature regions.59
In addition, in order to control the pore environment of MOFs, the introduction of
different substituents provides another way to design and prepare proton-con-
ducting MOFs (Figure 6B).60 This method allows the species, concentration, and
mobility of proton carriers to be adjusted effectively, which is beneficial for proton
conduction. As shown in the above-mentioned examples, coordinated water, guest
molecules, and functional groups of ligands play important roles in proton conduc-
tion. A new strategy has been envisaged for assembling proton-conducting MOFs
with organic phosphonic acid and sulfonic acid.26,61 Abundant oxygen atoms of
organic phosphonate or sulfonate ligands are bridged by solvent water molecules
via hydrogen bonds, providing a facile pathway for proton conduction. With the
inspiration of these advances, new MOFs with potential proton-conducting path-
ways have been investigated.62

Both defect and highly oriented film of MOFs have been studied on proton conduc-
tion. A zirconium-sulfoterephthalate MOF has been used as an example to illustrate
the effect of MOF defect on the proton-conducting behavior. Defect-containing
samples were prepared by introducing excess acetic acid or sulfoacetic acid in the
preparation process or by soaking the as-prepared sample in 0.1 M H2SO4.63
Compared with as-synthesized zirconium-sulfoterephthalate, all three treated
samples had predominantly improved conductivity in the range of 2.4 3 103 to
5.6 3 103 S cm1 at 95% RH and 65 C. These preliminary results on MOF defects
are significant for the design and application of MOFs as proton conductors. The
highly oriented MOF nanofilm made up of {Cu(TCPP)}n (H2TCPP = 5,10,15,20-tetra-
kis(4-carboxyphenyl)porphyrin) was prepared by depositing the nanostructured
sample between two Cr-Au electrodes on an SiO2 (300 nm thickness)@Si wafer
(Figure 7).64 The abnormally high conductivity of 3.9 3 103 S cm1 at 98% RH for
this thin film may be related to the numerous dangling groups, including acidic
coordinated water and noncoordinated carboxyl groups, decorating on the nano-
sheet surface, and the highly oriented crystalline morphology of the thin film.
Although MOFs conduct protons in the pristine or film formation with very high con-
ductivity at low temperature, there have been few investigations on assembling
MOF proton conductors into fuel cells.

Proton Conduction at High Temperature


In the high-temperature regime, the most successful strategy is to embed nonvola-
tile guests, such as triazole, imidazole, histamine, benzimidazole, and even H 2SO4
and H3PO4, into the pores of MOFs to set up facile proton delocalization pathways
for promoting proton conduction at temperatures above 100 C and in anhydrous
conditions.26 It is reported that, by encapsulating imidazole (Im) molecules into
aluminum MOFs, Im@{Al(m2-OH) (1,4-NDC)}n (1,4-NDC = 1,4-naphthalenedicarbox-
ylate) and Im@{Al(m2-OH) (1,4-BDC)}n present enhanced conductivities of 2.2 3 105
and 1.0 3 107 S cm1, respectively, in comparison with those of their parent MOFs
at 120 C.65 The difference in the proton conductivities of these two species is
attributed to the different Im dynamic motion in these two MOFs. A sulfonate-based
MOF, b-PCMOF2, encapsulating 1H-1,2,4-triazole (Tz) in the pores, conducts
protons at conductivity between 2 3 104 and 5 3 104 S cm1 at 150 C and in
anhydrous conditions.66 In the series of b-PCMOF2(Tz)x (x = 0.30, 0.45, and
0.60), b-PCMOF2(Tz)0.45, showing the highest proton conductivity, has been
further integrated into a membrane-electrode assembly to fabricate an H2-air cell.

64 Chem 2, 52–80, January 12, 2017


Figure 7. MOF Film for Proton Conduction
(A) Schematic illustration of the nanosheet-constructed MOF nanofilm.
(B and C) Structure model of the MOF nanofilm.
(D) Proton conductivity of the MOF nanofilm under various RH conditions. Inset: proton
conductivity versus adsorbed water of the MOF nanofilm under 95% RH.
(E) Least-squares fitting is shown as a solid line. Inset: typical Nyquist plot of the MOF nanofilm
measured under 98% RH at room temperature.
Reprinted with permission from Xu et al.64 Copyright 2013 American Chemical Society.

Electromotive force measurements on this cell clearly recorded a stable open-circuit


voltage of 1.18 V for 72 hr at 100 C. This result not only illustrates the effectiveness of
the guest-impregnation approach for exploring new proton conductors under anhy-
drous conditions but also paves the way for operating MOF-based fuel cells above
100 C. Besides the immobilization of inorganic phosphoric acid molecules to the
pores of MOFs as proton-hopping sites, inorganic phosphoric acid has been used
directly as the starting material to form MOFs with the help of a secondary organic
ligand.67 A dry H2-air cell has been manufactured by a membrane-electrode assem-
bly of {Zn(H2PO4)2(HTz)2}n (HTz = 1,2,4-triazole), which shows proton conductivity
of 104 S cm1 at 150 C.67 The open-circuit voltage of the fuel cell was measured
as 0.65 and 0.50 V at 25 C and 130 C, respectively. This strategy offers a low-cost
and facile way of synthesizing new proton conductors.

Oxygen Reduction Reaction


Operation of current PEMFCs requires platinum catalysts to boost the cathodic
oxygen reduction reaction (ORR), which is also important for metal-air batteries.
It is necessary to develop highly efficient and cheap cathode ORR catalysts to
replace expensive Pt catalysts. Non-precious-metal catalysts, such as various
carbon materials, transition-metal-based inorganic materials, and their composites,
are very promising for practical fuel-cell applications.24 With the exception of
Ni3(HITP)2 (HITP = 2,3,6,7,10,11-hexaiminotriphenylene), which display a striking

Chem 2, 52–80, January 12, 2017 65


onset potential of 0.82 V (versus a reversible hydrogen electrode [RHE]) in a 0.10 M
KOH solution,68 pure MOFs usually show low ORR performance as a result of their
insulating nature and poor stability. The addition of pyridinium-dye-functionalized
graphene oxide (GO) sheets to a metalloporphyrin-based MOF in an appropriate
amount is a good example of enhancing the ORR electrocatalytic properties of
MOFs.69 Recently, extensive studies have shown that MOF carbonization offers an
effective approach for fabricating desired structures for ORR as a result of the unique
feature of MOFs and MOF composites with rich metal and organic components as
well as tunable microstructures.21

In 2011, a cobalt imidazolate framework was used as a precursor to produce an ORR


catalyst by high-temperature treatment. This ORR catalyst exhibits an onset poten-
tial of 0.83 V (versus RHE) and a half-wave potential of 0.68 V in acidic solution.70
Metal-free N-doped porous carbon catalysts have been prepared from N-rich
ZIF-8 as both precursor and template and furfuryl alcohol (FA) and NH4OH as sec-
ondary carbon and nitrogen sources, respectively.71 It is revealed that the electro-
chemical reduction pathways of catalysts are determined at different carbonization
temperatures. There is an interesting report that bimetallic Zn/Co-ZIF, exhibiting an
isostructure with ZIF-8 and ZIF-67, has been used as template and precursor to
generate porous carbon materials. Co single atoms (SAs) over 4 wt % have been suc-
cessfully loaded in nitrogen-doped porous carbon.72 In the pyrolysis process, the
amount of Zn ions plays an important role in the formation of Co SAs. This atomically
dispersed material derived from the carbonization of bimetallic ZIF at 900 C displays
an outstanding ORR performance with E1/2 = 0.881 V (versus RHE), much more pos-
itive than that of commercial Pt/C (0.811 V). Although ZIF series are the most popular
precursors for ORR catalysts, the utilization of other heteroatom-containing MOFs as
precursors, such as zirconium-porphyrin frameworks, has been reported to provide
ORR catalysts with high catalytic activity.73 MOF composites have also been used as
precursors to afford effective ORR catalysts.74,75 An aluminum-based MOF, MIL-
101-NH2, has been utilized to encapsulate thiourea and cobalt chloride in a facile
double-phase encapsulation approach.74 An interesting guest-induced morphology
control in carbonization gives unique honeycomb-like N and S dual-doped porous
carbon nanostructures with the immobilization of Co9S8 NPs (Figure 8). This material
exhibits compelling catalytic activity, superb stability, and methanol tolerance for
oxygen reduction in alkaline medium. Moreover, 2D inorganic nanoplates based
on Co- and Al-doped layered double hydroxides (LDHs) have been used as scaffolds
and directing agents for epitaxial growth of ZIF-67 arrays.75 After the pyrolysis of
CoAl-LDH@ZIF-67, the carbon-based network obtained delivers superior ORR activ-
ity (Eonset = 0.94 V versus RHE) and stability relative to the commercial Pt/C material.

Supercapacitors
Supercapacitors (SCs), also called electrochemical capacitors or ultracapacitors,
have attracted intense research attention because of their high power density and
long lifecycle. Despite their high porosities, pure MOF SCs usually exhibit low spe-
cific capacitance as a result of poor conductivity. In order to enhance MOF conduc-
tivity, the incorporation of MOFs with conductive materials, such as graphene and
conductive polymers, has created a new strategy to improve SC performance. Suc-
cessful doping of 23 crystalline MOFs with graphene for the fabrication of SCs has
been accomplished by Choi et al.76 A large number of nanocrystalline MOFs with
different metals, ligands, structures, and pore volumes have been examined. A
supercapacitor based on {Zr6O4(OH)4(BPYDC)6}n (BPYDC = 2,20 -bipyridine-5,50 -di-
carboxylate) nanocrystals manifests the highest SC performance with stack and areal
capacitance of 0.64 and 5.09 mF cm2, respectively, and exhibits superior durability

66 Chem 2, 52–80, January 12, 2017


Figure 8. Schematic Illustration of Pyrolysis of the MOF Composite Comprising MIL-101-NH2
Encapsulating Thiourea and Cobalt Chloride in the Nanopores to Yield the Honeycomb-like
Porous Co9S8@CNST Catalysts
Reprinted with permission from Zhu et al. 74 Copyright 2016 WILEY-VCH Verlag GmbH & Co. KGaA,
Weinheim.

over at least 10,000 cycles (Figure 9). Under these circumstances, the large electro-
chemical double-layer capacitor (EDLC)-type capacitance of MOFs is mainly associ-
ated with their extremely large surface area and unique porous structure, which allow
diffusion of electrolyte ions through the micro- and mesopores.77 ZIF-67 nanocrys-
tals have been electrochemically interweaved with a conductive polyaniline (PANI)
to increase the conductivity. The deposited PANI,77 which directly interconnects
the external circuit and the MOF interior surface to support electron transportation,
induces a high areal capacitance of 2,146 mF cm2 at 10 mV s1. This example also
represents another effective strategy by combining MOFs and pseudocapacitor ma-
terials, including organic polymers and metal oxides.77,78 In a recent report, a high
conductive MOF, {Ni3(HITP)2}n (HITP = 2,3,6,7,10,11-hexaiminotriphenylene), has
been investigated for EDLC behavior (Figure 9).79 A MOF-based SC has been fabri-
cated with this MOF as a single-component electrode material in the absence of any
conductive additives and organic binders; it showed a gravimetric capacitance of
111 F g1 at a low discharge rate of 0.05 A g1. This value is higher than that of car-
bon nanotubes. Noticeably, this MOF-based supercapacitor exhibits high-capacity
retention over 10,000 cycles and a very high surface area normalized capacitance
of 18 mF cm2, superior to most carbon-based materials except holey graphene.79

Carbon SCs possess a variety of advantages, including low weight, fast charge and
discharge rates, and bipolar operational flexibility, which inspire capacitive applica-
tions of MOF-derived nanoporous carbons. In 2008, Xu’s group used MOF-5 as a
template and precursor and FA as a secondary carbon source to prepare NPCs for
the first time,27 and a capacitance of 204 F g1 has been observed for the as-synthe-
sized carbon electrode (electrolyte, 1.0 M H2SO4) at a sweep rate of 5 mV s1, indi-
cating that MOF-derived NPCs are excellent electrode materials for EDLCs. This
result also helps shed light on the preparation of NPCs by the MTC method.21
Carbonization of core-shell MOFs (ZIF-8@ZIF-67) leads to the desired nanoporous
core-shell carbon materials made up of nitrogen-doped carbon (NC) cores and
highly graphitic carbon (GC) shells (Figure 10).32 The results of capacitance assess-
ment reveal that the NC@GC electrode has a high specific capacitance of
270 F g1 at a current density of 2 A g1. The excellent capacitance of the

Chem 2, 52–80, January 12, 2017 67


Figure 9. MOF-Based Supercapacitors
(A) Schematic construction of nMOF supercapacitors. Reprinted with permission from Choi et al. 76
Copyright 2014 American Chemical Society.
(B) Structure and nanocrystal morphology of nMOF-867 and comparison of stack capacitances
among various EDLC materials. Reprinted with permission from Sheberla et al. 79 Copyright 2016
Nature Publishing Group.
(C) A space-filling diagram of idealized Ni 3 (HITP) 2 (Ni, green; F, lime; N, blue; C, gray; B, brown;
H, white).
(D) Comparison of BET-surface-area-normalized areal capacitances among various EDLC
materials.

NC@GC electrode is attributed to the synergetic effect of the high surface area,
hierarchical micro- and mesoporous structure, N-dopant composition, and a high
degree of graphitization of the GC shell. Two-dimensional carbon materials pre-
pared by the MTC method show outstanding SC performance. In a recent report,
a pure carbon material, in the form of two- to six-layered graphene nanoribbons,
was synthesized from as-synthesized non-hollow (solid) carbon nanorods and
showed excellent EDLC behavior.33 The SC investigation based on 2D CoS1.097/
NC nanocomposites derived from 2D porphyrin MOF nanosheets shows that the
morphology of nanocomposite is important for electron and ion transport and
thus for electrochemical performance.80

Lithium-Based Batteries
Among various battery technologies, lithium-ion batteries (LIBs) are a great com-
mercial success. The working principle of LIBs is that the lithium ions, which are
generated through the oxidation of lithium-based anode materials, migrate from
the anode to the cathode during discharge and return when charged.2 LIBs can po-
wer portable electronic devices because of their many advantages, such as high en-
ergy density, long lifespan, environmental friendliness, and no memory effect in the
charge and discharge processes. However, the energy and power densities of LIBs

68 Chem 2, 52–80, January 12, 2017


Figure 10. Synthetic Scheme for Preparation of MOFs and Their Derivatives
(A) ZIF-8 crystals and NC.
(B) ZIF-67 crystals and GC.
(C) Bimetallic ZIF (BMZIF) and N/Co-doped GC.
(D) Core-shell ZIF-8@ZIF-67 crystals and NC@GC.
Reprinted with permission from Tang et al.32 Copyright 2015 American Chemical Society.

hinder their application for large-scale electrical energy storage. As a result, it is


crucial to develop electrode and electrolyte materials for LIBs to improve their per-
formance. MOFs, MOF composites, and MOF derivatives provide an exciting oppor-
tunity to develop new electrode and electrolyte materials for LIBs and for the next
generation of power devices, such as lithium-oxygen (Li-O2) and lithium-sulfur
(Li-S) batteries.

LIBs
Pristine MOFs have been exploited as alternatives to conventional graphite anode
materials, because they offer a high surface area and permanent pores for Li+ ion
storage and migration during charge and discharge processes. Many initial studies
show that pristine MOFs, such as Zn-based MOF-177 and Mn-LCP (LCP = layered co-
ordination polymer), do not favor reversible Li+ ion transport because of their poor
cycling capacity.2 However, their high discharge capacity provides hope that new
MOFs can be developed as LIB anode materials. In 2014, an intercalated MOF
(iMOF) composed of 2,6-naphthalene dicarboxylate and lithium ions was revealed

Chem 2, 52–80, January 12, 2017 69


Figure 11. MOF Derivative and Corresponding Performance as a LIB Anode
(A) Schematic illustration of the synthesis procedure of N-doped graphene analogous particles and
model of N doping.
(B) Cycling performance at a current density of 100 mA g1 .
(C) Rate performance at different current densities from 100 to 1,600 mA g 1 .
Reprinted with permission from Zheng et al. 82 Copyright 2014 Nature Publishing Group.

to behave as an anode material for Li intercalation.81 At a flat potential of 0.8 V, this


iMOF presents a reversible two-electron-transfer for Li intercalation. Most impor-
tantly, the first MOF-based 4 V single cell has been fabricated with iMOF as the
anode and LiNi0.5Mn1.5O4 as the cathode and displayed high cell voltage, excellent
cyclability, high specific energy, and high specific power. This demonstrates that
MOFs are capable of serving as the anode materials of LIBs. However, it is a chal-
lenge to develop robust MOFs with reversible formation or regeneration to achieve
high-performance Li storage in excellent cyclability.

Various MOF derivatives composed of mono-, bi-, or multipodal metal oxides and
porous carbons, which possess much higher electrochemical performance than their
conventional counterparts, have been tested in LIB anode applications.82–85 For
instance, a ZIF-8-derived nitrogen-doped graphene material manifests a predomi-
nantly reversible specific capacity of 2,132 mAh g1 at a current density of
100 mA g1 after 50 cycles (Figure 11).82 As the current density is increased to
5 A g1, the electrode fabricated from N-C-800 maintains a capacity of
785 mAh g1 even after 1,000 cycles, representing a promising candidate as anode
material for high-performance LIBs. This noticeable electrochemical performance is
attributed to the unique N-doped carbon nanostructure, namely a large number of
highly N-doped graphene nanocrystallites located on the hexagonal lattice and
edges. A spindle-like mesoporous a-Fe2O3 material has been derived from MIL-
88-Fe through a two-step pyrolysis method.83 After 50 cycles at 0.2 A g1, the

70 Chem 2, 52–80, January 12, 2017


as-synthesized spindle-like a-Fe2O3 still shows a high charge capacity of
911 mAh g1. As a good comparison, the a-Fe2O3 anode fabricated by a one-
step method delivers a lower charge capacity of 772 mAh g1 after 50 cycles at
the same current density. This illustrates that the pyrolysis process is very important
for the electrochemical performance of MOF derivatives. These MOF-derived
metal oxides usually show poor cyclability in comparison with pure metal oxides
obtained by traditional methods. A possible solution is the use of porous carbon
nanostructures to coat or encapsulate metal oxides.24,84 A MOF-derived multipodal
composite of NiO/Ni/graphene, in the form of a hierarchical hollow ball-in-ball
nanostructure, displays a high reversible specific capacity of 1,144 mAh g1 and
excellent cyclability with nearly 100% capacity retention after 1,000 cycles.84 In
particular, a sodium-ion battery has been assembled from a NiO/Ni/graphene elec-
trode, showing outstanding cyclability and an excellent ratability of 207 mAh g1 at a
current density of 2 A g1.

Research on LIBs based on MOFs and MOF-derived cathode materials has devel-
oped rapidly since 2007.2,85–87 An electrochemical cell has been assembled with a
mixture of desolvated MIL-53(Fe) and 15 wt % conductive carbon as a cathode
and Li as an anode. The MIL-53(Fe) cathode material utilizes mixed-valence states
of metals during charge and discharge processes to perform electrochemical Li stor-
age.85 This cathode shows a gravimetric capacity of 75 mAh g1 and a volumetric
capacity of 140 mAh L1, which is determined by the limited number of Li ions in-
serted and the low density of the material. Although this MOF displays a relatively
low capacity, this work opens up a new avenue for MOF cathode materials to store
Li ions. In order to enhance the theoretical capacity, active anthraquinone (AQ) units
have been incorporated into a MOF, {Cu(2,7-AQDC)}n, where 2,7-AQDC is 2,7-an-
thraquinonedicarboxylic acid (Figure 12).87 In a voltage window between 4.0 and
1.7 V with a scan current of 1 mA, this resultant MOF-based cathode shows an initial
specific capacity of 147 mAh g1 and reversible capacities of 105 mAh g1 in
50 cycles. The electrochemical activity may originate from the participation of
both Cu2 paddlewheels and anthraquinone groups in the discharge reduction pro-
cess. Although the theoretical capacity has been increased to 162 mAh g1 based on
a facile synthetic strategy, the areal capacity is constrained by the exposed surface
areas and effective charge transfer rates of MOF materials. Thus far, it is still a
challenge to tackle those issues, including low capacity and capacity retention.
Moreover, MOFs and MOF-derived materials have been used to modify well-known
cathode materials.2

Li-O2 and Li-S Batteries


Li-O2 and Li-S batteries, as next-generation electric power sources, have attracted
considerable attention as a result of their high theoretical energy density of 3,500
and 2,500 Wh kg1, respectively. In comparison with matured and commercialized
LIBs, Li-O2 and Li-S batteries are still in development. To improve Li-oxygen
batteries, it is vital to mitigate the ORR and oxygen evolution reaction (OER) overpo-
tentials and optimize the transport kinetics of mass and electron. The first MOF-
based Li-O2 batteries were established in 2014.88 A series of MOFs, including
MOF-5, HKUST-1, and M-MOF-74 (M = Mg, Mn, Co) were judiciously selected to
correlate the relationship between the structure and performance (Figure 2). At an
applied current density of 50 mA g1 and operating voltages of 2.6–2.7 V,
Mn-MOF-74@carbon-black exhibits the most conspicuous discharge capacity of
9,420 mAh g1 among all cells. This study clearly demonstrates the important role
of OMSs on the performance of MOF-based Li-O2 batteries. In addition, a MOF-
derived ORR catalyst has been used as a cathode in a non-aqueous Li-O2 battery.

Chem 2, 52–80, January 12, 2017 71


Figure 12. Crystal Structure of MOF and Corresponding Performance as a LIB Cathode
(A) Crystal structure of {Cu(2,7-AQDC)} n .
(B) Charge-discharge profile of MOF battery.
(C) Cyclic performance of battery in 50 cycles.
(D) Cyclic voltammetry plot of battery.
Reprinted with permission from Zhang et al. 87 . Copyright 2014 American Chemical Society.

The N-Fe-MOF ORR catalyst with graphene and graphene tube structures was pre-
pared from Co-based MOF polyhedral cages as sacrificial templates and immobi-
lized dicyandiamide (DCDA) and iron acetate as an N/C source and a catalyst,
respectively.89 The resultant N-Fe-MOF catalyst displays excellent ORR catalytic ac-
tivity in both acidic and basic electrolytes, and the corresponding Li-O2 cell provides
a very high initial discharge capacity of 5,300 mAh g1.

The drawbacks of Li-S batteries include the use of Li metal as an anode, the forma-
tion of soluble polysulfides in the reduction process, and the required addition of
conductive additives. To tackle these issues, an Ni-MOF, {Ni6(BTB)4(BP)3}n (BTB =
benzene-1,3,5-tribenzoate and BP = 4,40 -bipyridyl), was selected to confine sulfur
(Figure 13).90 Galvanostatic charge and discharge measurements revealed the
typical charge and discharge behaviors for the above Ni-MOF/S electrode. At a
current density of 0.1 C (168 mA g1), the Ni-MOF/S composite offers a high-ca-
pacity retention of 89% after 100 cycles. The hierarchical porous structure of Ni-
MOF and strong interactions between Ni metals and polysulfides prevent the
dissolution and the shuttle effect of polysulfides. This investigation represents
the design of Li-S electrodes through the confinement of sulfur in the pores of
mesoporous MOFs to enhance the cyclability of Li-S batteries. However, the insu-
lating nature of MOFs gives rise to low sulfur utilization and a weak performance
rate. In order to mitigate the shuttling problem, an HKUST-1@GO separator works
as an ionic sieve in Li-S batteries, which selectively permits Li+ ion transport but
suppresses polysulfide migration. A Li-S battery based on the HKUST-1@GO

72 Chem 2, 52–80, January 12, 2017


Figure 13. Crystal Structure of MOF and Corresponding Performance in a Li-S Battery
(A) Crystal structure of Ni-MOF.
(B) Cycling performance of Ni-MOF/S composite at 0.1, 0.2, and 0.5 C rates at a voltage range of
1.5–3.0 V and schematic illustration of the interaction between polysulfides and Ni-MOF.
Reprinted with permission from Zheng et al. 90 Copyright 2014 American Chemical Society.

separator exhibits low capacity-fading rates of approximately 0.019% per cycle


over 1,500 cycles.91

Electrolysis of Water
Electrocatalytic Hydrogen Evolution Reaction
Electrolysis of water, including the hydrogen evolution reaction (HER) and the OER,
has been intensively investigated in recent years. Catalysts are required to decrease
the overpotential of these reactions, driving them at desired high catalytic current
densities. Efficient hydrogen evolution from electrolytic water splitting with the
help of MOFs, MOF composites, and MOF derivatives has emerged as another
important strategy for providing clean and renewable fuel. The integration of HER
catalytically active units into MOFs is an effective route to explore high-performance
electrodes. 2D ordered MOF films (MOS 1 and MOS 2) consisting of cobalt dithio-
lene catalytic sites and exhibiting significant HER catalytic activity have been de-
signed and constructed based on conjugated ligands of benzenehexathiol (BHT)
and triphenylene-2,3,6,7,10,11-hexathiolate (THT), respectively,92. ε-Keggin poly-
oxometalate-based metal-organic frameworks (PMOFs) are robust HER electrocata-
lysts under acidic conditions.16 Because polyoxometalate (POM) is located in
different environments in these PMOFs, their microenvironment effects are vital to
influence the corresponding catalytic activity and recyclability. Most importantly,
the electrode of ε(trim)4/3/CPE (tricarboxylate linker and carbon paste are denoted
as trim and CPE, respectively) shows an onset potential of about 20 mV (versus a
saturated calomel electrode) for HER, which is much more active than that of the
platinum counterpart (242 mV).16 In addition, highly porous MOFs have been
used as supports for the impregnation of electrocatalyst to decrease the required ki-
netic overpotential. For example, NU-1000 has been chosen for the immobilization
of Ni-S (Ni-S denotes a material composed of Ni and S elements with unspecified
stoichiometry).93 Instead of the surface deposition on MOF rods, Ni-S has been filled
in the bottom of MOF rods and coated on bare FTO. In aqueous HCl with a pH value
of 1, the NU-1000@Ni-S electrode exhibits an overpotential of 238 mV (versus RHE)
to afford a catalytic current of 10 mA cm2, which is obviously less than that of Ni-S
electrodeposited directly on FTO (560 mV).

Considerable efforts have been devoted to using MOF precursors to structure electro-
catalysts for HER.94,95 The decomposition of as-prepared NENU-5 nano-octahedra at

Chem 2, 52–80, January 12, 2017 73


Figure 14. Preparation Process of the Porous MoCx Nano-octahedra Used as Electrocatalysts for
the HER
Reprinted with permission from Wu et al.94 Copyright 2015 Nature Publishing Group.

800 C under N2 affords MoCx nano-octahedra (Figure 14).94 The MoCx electrocatalyst
shows a smaller Tafel slope (59 mV dec1) than that of the Pt/C catalyst (113 mV dec1) in
alkaline aqueous solution, confirming the excellent electrocatalytic activity of the MoCx
electrode for HER. In addition, superior molybdenum-oxide-based HER electrocatalysts
have been prepared from a hybrid precursor of NENU-5 and GO with different
loadings.95 The MoO2@PC-RGO electrode derived from the pyrolysis of PMOFs/GO
with 8 wt % GO at 800 C displays a very positive onset potential of ca. 0 V (versus
RHE), a low Tafel slope of 41 mV dec1, and a high exchange current density of 4.8 3
104 A cm2 to drive HER, exhibiting the most excellent electrocatalytic activity among
all non-precious-metal catalysts in acidic solution. In addition, this composite displays
long-term stability. These outstanding electrocatalytic properties of the hybrid elec-
trode are assigned to the synergistic effect of the conductive GO support, nanostruc-
tured MoO2 particles, and P-doped carbon substrate as well as the unique structure.
Comparison of these two examples illustrates that judicious selection of precursors is
critical to obtain different catalysts.

Earth-abundant 3d transition metals dispersed on MOF-derived porous carbons in


an atomically isolated form have been used to boost ORR and HER.72,96 For
example, an A-Ni-C catalyst with single Ni atoms dispersed on a carbon matrix
has been produced by successive pyrolysis of Ni-MOF, HCl leaching, and electro-
chemical treatments.96 The stable A-Ni-C electrode exhibits a very small Tafel slope
of 41 mV dec1 and overpotentials of 34, 48, and 112 mV (versus RHE) to deliver
abundant catalytic current density of 10, 20, and 100 mA cm2, respectively. These
overpotentials represent the best HER performance among the reported doped gra-
phene-supported earth-abundant electrocatalysts in acidic medium. The noticeable
activity of the A-Ni-C catalyst has been attributed to the unique chemical and elec-
tronic coupling between interior components. The current integration of graphitized
carbon materials and single-atom metals further highlights the excellent electroca-
talytic activity of atomic-metal-doped carbon catalysts and is creating a new area
of metal-carbon-based electrocatalysts.

Electrocatalytic OER
Although the above strategies provide innovative materials for efficient implemen-
tation of hydrogen evolution via electrolytic water splitting, HER efficiency and

74 Chem 2, 52–80, January 12, 2017


required overpotential are determined by the OER as a rate-limiting step. A four-
electron process involved in the OER undergoes slow kinetics and high-performance
catalysts are a prerequisite to reduce the operational overpotential. The cost and
scarcity of the most efficient noble-metal oxide OER catalysts such as IrO2 and
RuO2, hinder their commercialization. As a consequence, development of highly
active non-noble-metal OER catalysts for energy conversion is essential.17,97–100
A transition-metal-organic framework, Co-WOC-1, shows excellent OER electroca-
talytic activity in an alkaline solution.17 Mononuclear moieties of {Co(H2O)4(DMF)2}2+
are trapped in the pores of a 3D framework, which are the active sites for OER. The
corresponding Tafel slope of 128 mV dec1, overpotential of 390 mV (versus a
normal hydrogen electrode [NHE]), and TOF of 0.05 s1 reveal that Co-WOC-1 is
an efficient electrocatalyst for OER in alkaline medium.

MOF-derived heteroatom-doped nanocarbon materials are promising OER electro-


catalysts. An N-doped carbon nanotube framework (NCNTF) catalyst has been
prepared by the pyrolysis of ZIF-67 as a precursor at 700 C in a mixed Ar and H2
atmosphere.97 This electrode displays exciting electrocatalytic ORR performance
with a half-wave potential of 0.87 V (versus RHE) and a Tafel slope of 64 mV dec1
in KOH solution, surpassing the performance of its Pt/C counterpart. Likewise, this
electrode exhibits a smaller Tafel slope of 93 mV dec1 to drive the OER in compar-
ison with that of the Pt/C electrode (118 mV dec1). The durability and methanol
tolerance of the NCNTF electrode for both ORR and OER are higher than those of
the commercial Pt/C electrode under the same conditions. The excellent activity is
associated with the hollow nanostructure and unique component made up of crys-
talline NCNTs and Co nanoparticles wrapped by a few layered carbon shells.
Another interesting ZIF-67-derived composite, consisting of core-shell Co@Co3O4
nanoparticles embedded in N-doped carbon nanotube (CNT)-grafted carbon poly-
hedra, has been realized by reductive carbonization in an H2 atmosphere and
subsequently controlled oxidation.99 This composite is one of the best bifunctional
catalysts for both ORR and OER in 0.1 M KOH. More importantly, this catalyst is re-
garded as one of the best non-precious-metal electrocatalysts for reversible oxygen
electrodes given that it shows a smaller reversible overvoltage of 0.85 V (versus RHE)
between ORR and OER than that of Pt/C, IrO2, and RuO2. Comparative analysis of
these two examples further highlights that the MTC process plays an important
role in affording active ingredients for electrocatalysis.

Recently, a strategy has been developed to directly grow catalytically active sites to
well-aligned nanowire arrays on current collectors for fabricating a new generation
of effective electrodes for direct use in electrochemical OER. Co3O4-C porous nano-
wire arrays, Co3O4C-NA, have been synthesized by thermolysis of a Co-naphthalene
dicarboxylate MOF precursor to directly grow nanowire arrays on Cu foil (Fig-
ure 15).100 The whole Co3O4C-NA assembly grown on Cu foil, denoted as
Co3O4C-NA/Cu, has been used directly as the working electrode to promote the
OER in an alkaline electrolyte. The Co3O4C-NA/Cu electrode exhibits a sharp onset
potential at 1.47 V (versus RHE), which is slightly smaller than the 1.45 V for an IrO2/C
noble-metal catalyst. However, the former electrocatalyst displays a much larger
OER current density than that of the latter electrode at the same mass loading. In or-
der to supply the same current density of 10.0 mA cm2, Co3O4C-NA/Cu and IrO2/C
electrodes require operating potentials of 1.52 and 1.54 V, respectively. The Tafel
slope value of 70 mV dec1 for Co3O4C-NA/Cu electrocatalyst is much lower than
that of IrO2/C (97 mV dec1). For this electrode, the narrow gaps among nanowire
arrays are regarded as benefiting electrolyte and proton transport in the OER pro-
cess. In addition, the direct and close contact of nanowires and conductive current

Chem 2, 52–80, January 12, 2017 75


Figure 15. Fabrication of Hybrid Co3O4-Carbon Porous Nanowire Arrays
The scale bar represents 5 mm.
Reprinted with permission from Ma et al.100 Copyright 2014 American Chemical Society.

collectors promote the enhancement of electrical conductivity and durability of the


electrocatalyst, providing a new strategy for directly affording OER electrodes for
practical applications.

Conclusions and Perspectives


The utilization of MOFs, MOF composites, and MOF derivatives (including porous
carbons, metals, metal oxides, metal sulfide, and their composites) for a wide range
of energy applications has emerged in the drive to find more innovative energy tech-
nologies. Various MOFs, MOF composites, and MOF derivatives play important
roles in photo- and electrochemical energy storage and conversion, in terms of
storing gas molecules, enhancing gas diffusion, facilitating mass, electron, and
charge transportation, harvesting exoteric energy, promoting reactant activation,
enhancing conductivity and durability, and so on. These aspects are crucial for the
improvement of material performance and practical applications. The advantages
of MOF-based materials for energy storage and conversion are still being high-
lighted by some fundamental breakthroughs achieved worldwide. Moreover,
many synthetic and applicative approaches have been established to precisely
control material structure and effectively improve material properties. By virtue of
functional MOF hybridization and MOF carbonization methods, the utilization of
MOF has been further strengthened and extended in the energy area. In particular,
the confinement effect and synergetic functions of MOF composites and multicom-
ponent MOF derivatives are able to improve material cyclability and promote the
reaction kinetics involved in energy storage and conversion processes. In addition,
the rational combination and fabrication of MOF precursors as well as the optimal
calcination process work synergistically to enhance the performance of MOF
derivatives.

This review summarizes the recent progress in the use of MOFs, MOF composites,
and MOF derivatives for energy applications. After nearly a decade of intense
research by different research groups, tremendous advances in energy applications
based on MOFs and their composites and derivatives have been achieved in many
fields. However, there are still challenges to improve the performance in energy stor-
age and conversion to achieve the energy targets (Table 1). (1) The conductivities of
MOFs, the biggest obstacle for photo- and electrochemical energy applications,
should be further improved for practical use. (2) Many more MOFs should be

76 Chem 2, 52–80, January 12, 2017


Table 1. The Selected Targets of Some Energy Storage and Conversion Systems
Systems Selected Targets of Energy Storage and Organizations
Conversion System
Hydrogen storage gravimetric capacity: 5.5 wt % US DOE101
volumetric capacity: 0.040 kg/L
cost: $10/kWh ($333/kg stored hydrogen
capacity)
Methane storage gravimetric capacity: 0.5g/g US DOE102
cost of sorbent: $10/kg
Hydrogen production cost: $3.10/kg for central hydrogen plants US DOE103
and $3.70/kg for distributed hydrogen plants
Photovoltaic cells cost: $0.06/kWh for utility-scale photovotaic US DOE104
cells
Fuel cells power density: 650 W/L US DOE105
specific power: 650 W/kg
cost: 40 $/kWnet
durability cycle: 5,000
Ultracapacitors for electric discharge pulse: 4.2 kW in 2 s for 12 V start- USCAR106
drive vehicles stop and 8 kW in 6 s for 42 V start-stop
operating temperature: 30 C–52 C
Batteries for electric drive gravimetric density: 250 Wh/kg US DOE107
vehicles
volumetric density: 400 Wh/L
cost: $125/kwh

Abbreviations are as follows: DOE, Department of Education; and USCAR, US Council for Automotive
Research.

introduced in composites as well as derivatives for energy applications to explore


the effects of different electronic and chemical environment or structure effects on
the function of materials. (3) Design strategies and preparation methods for MOF
composites and derivatives should be developed to afford more advanced mate-
rials. For example, PSM of as-prepared MOF derivatives with the use of chemical
and electrochemical methods is a promising direction. (4) Application conditions,
especially harsh conditions such as high temperature, high pressure, and corrosion,
should be evaluated for practical use. (5) Composition, structure, morphology,
preparation technology, and intrinsic reaction mechanisms should be considered
simultaneously to decrease costs, increase specific performance, and extend usage
lifetime. (6) The interior mechanism involving individual ingredients, surfaces, and
interfaces should be pursued in depth with advanced characterization tools in com-
bination with theoretical calculations. New insights should be provided into the
synergetic effect of multicomponent composites to recognize the most important
factors. (7) Effective and facile screening methods to identify high-performance
materials should be established to save the time involved in the design and prepa-
ration of MOF-related materials and identify their rational use. (8) Much more effort
should be devoted to device fabrication. As research enthusiasm for using MOFs in
energy applications increases, further sustained and intense efforts in this field will
increase the perspective of MOFs, MOF composites, and MOF derivatives for real
energy applications.

AUTHOR CONTRIBUTIONS
Q.X. proposed the topic of the review. H.W. investigated the literature and wrote the
manuscript. Q.X., H.W., Q.-L.Z., and R.Z. discussed and revised the manuscript.

Chem 2, 52–80, January 12, 2017 77


ACKNOWLEDGMENTS
We are grateful to the editor for the kind invitation. We are pleased to acknowledge
the fine work of the talented and dedicated graduate students, postdoctoral fellows,
and colleagues who have worked with us in this area and whose names can be found
in the references. We would like to thank AIST and the Japan Society for the Promo-
tion of Science (JSPS) for financial support (KAKENHI no. 26289379). H.W. acknowl-
edge JSPS for a postdoctoral fellowship.

REFERENCES AND NOTES


1. Schoedel, A., Ji, Z., and Yaghi, O.M. (2016). intrinsic thermal management. Nature 527, 23. Li, J.-R., Sculley, J., and Zhou, H.-C. (2012).
The role of metal–organic frameworks in a 357–361. Metal–organic frameworks for separations.
carbon-neutral energy cycle. Nat. Energy 1, Chem. Rev. 112, 869–932.
16034–16046. 13. Kornienko, N., Zhao, Y., Kley, C.S., Zhu, C.,
Kim, D., Lin, S., Chang, C.J., Yaghi, O.M., and 24. Xia, W., Mahmood, A., Zou, R., and Xu, Q.
2. Li, S.-L., and Xu, Q. (2013). Metal–organic Yang, P. (2015). Metal–organic frameworks for (2015). Metal–organic frameworks and their
frameworks as platforms for clean energy. electrocatalytic reduction of carbon dioxide. derived nanostructures for electrochemical
Energy Environ. Sci. 6, 1656–1683. J. Am. Chem. Soc. 137, 14129–14135. energy storage and conversion. Energy
Environ. Sci. 8, 1837–1866.
3. Cook, T.R., Dogutan, D.K., Reece, S.Y., 14. Wang, C., deKrafft, K.E., and Lin, W. (2012). Pt
Surendranath, Y., Teets, T.S., and Nocera, nanoparticles@photoactive metal–organic 25. Zhu, Q.-L., and Xu, Q. (2016). Immobilization
D.G. (2010). Solar energy supply and storage frameworks: efficient hydrogen evolution via of ultrafine metal nanoparticles to high-
for the legacy and nonlegacy worlds. Chem. synergistic photoexcitation and electron surface-area materials and their catalytic
Rev. 110, 6474–6502. injection. J. Am. Chem. Soc. 134, 7211–7214. applications. Chem 1, 220–245.

4. Dhakshinamoorthy, A., Asiri, A.M., and Garc, 26. Ramaswamy, P., Wong, N.E., and Shimizu,
15. Xu, H.-Q., Hu, J., Wang, D., Li, Z., Zhang, Q.,
H. (2016). Metal–organic framework (MOF) G.K.H. (2014). MOFs as proton conductors-
Luo, Y., Yu, S.-H., and Jiang, H.-L. (2015).
compounds: photocatalysts for redox challenges and opportunities. Chem. Soc.
Visible-light photoreduction of CO2 in a
reactions and solar fuel production. Angew. Rev. 43, 5913–5932.
metal–organic framework: boosting electron-
Chem. Int. Ed. Engl. 55, 5414–5445.
hole separation via electron trap states. J. Am. 27. Liu, B., Shioyama, H., Akita, T., and Xu, Q.
5. Winter, M., and Brodd, R.J. (2004). What are Chem. Soc. 137, 13440–13443. (2008). Metal–organic framework as a
batteries, fuel cells, and supercapacitors? template for porous carbon synthesis. J. Am.
Chem. Rev. 104, 4245–4269. 16. Nohra, B., Moll, H.E., Albelo, L.M.R., Mialane, Chem. Soc. 130, 5390–5391.
P., Marrot, J., Mellot-Draznieks, C., O’Keeffe,
6. Furukawa, H., Cordova, K.E., O’Keeffe, M., M., Biboum, R.N., Lemaire, J., Keita, B., et al. 28. Liu, B., Shioyama, H., Jiang, H.-L., Zhang, X.,
and Yaghi, O.M. (2013). The chemistry and (2011). Polyoxometalate-based metal organic and Xu, Q. (2010). Metal–organic framework
applications of metal–organic frameworks. frameworks (POMOFs): structural trends, (MOF) as a template for syntheses of
Science 341, 974–986. energetics, and high electrocatalytic nanoporous carbons as electrode materials
efficiency for hydrogen evolution reaction. for supercapacitor. Carbon 48, 456–463.
7. Li, B., Wen, H.-M., Cui, Y., Zhou, W., Qian, G., J. Am. Chem. Soc. 133, 13363–13374.
and Chen, B. (2016). Emerging multifunctional 29. Jiang, H.-L., Liu, B., Lan, Y.-Q., Kuratani, K.,
metal–organic framework materials. Adv. 17. Manna, P., Debgupta, J., Bose, S., and Das, Akita, T., Shioyama, H., Zong, F., and Xu, Q.
Mater. 28, 8819–8860. S.K. (2016). A mononuclear CoII coordination (2011). From metal–organic framework to
complex locked in a confined space and nanoporous carbon: toward a very high
8. Rosi, N.L., Eckert, J., Eddaoudi, M., Vodak, acting as an electrochemical water-oxidation surface area and hydrogen uptake. J. Am.
D.T., Kim, J., O’Keeffe, M., and Yaghi, O.M. catalyst: a ‘‘ship-in-a-bottle’’ approach. Chem. Soc. 133, 11854–11857.
(2003). Hydrogen storage in microporous Angew. Chem. Int. Ed. Engl. 55, 2425–2430.
metal–organic frameworks. Science 300, 30. Hu, J., Wang, H., Gao, Q., and Guo, H. (2010).
1127–1129. 18. Zhang, Z., Chen, Y., Xu, X., Zhang, J., Xiang, Porous carbons prepared by using metal–
G., He, W., and Wang, X. (2014). Well-defined organic framework as the precursor for
9. Furukawa, H., Ko, N., Go, Y.B., Aratani, N., supercapacitors. Carbon 48, 3599–3606.
metal–organic framework hollow nanocages.
Choi, S.B., Choi, E., Yazaydin, A.O., Snurr,
Angew. Chem. Int. Ed. Engl. 53, 429–433.
R.Q., O’Keeffe, M., Kim, J., and Yaghi, O.M. 31. Yang, S.J., Kim, T., Im, J.H., Kim, Y.S., Lee, K.,
(2010). Ultrahigh porosity in metal–organic Jung, H., and Park, C.R. (2012). MOF-derived
19. Lu, G., Li, S., Guo, Z., Farha, O.K., Hauser,
frameworks. Science 329, 424–428. hierarchically porous carbon with exceptional
B.G., Qi, X., Wang, Y., Wang, X., Han, S., Liu,
porosity and hydrogen storage capacity.
10. Lin, J.-M., He, C.-T., Liu, Y., Liao, P.-Q., Zhou, X., et al. (2012). Imparting functionality to a
Chem. Mater. 24, 464–470.
D.-D., Zhang, J.-P., and Chen, X.-M. (2016). metal–organic framework material by
A metal–organic framework with a pore size/ controlled nanoparticle encapsulation. Nat. 32. Tang, J., Salunkhe, R.R., Liu, J., Torad, N.L.,
shape suitable for strong binding and close Chem. 4, 310–316. Imura, M., Furukawa, S., and Yamauchi, Y.
packing of methane. Angew. Chem. Int. Ed. (2015). Thermal conversion of core-shell
Engl. 55, 4674–4678. 20. Zhu, Q.-L., and Xu, Q. (2014). Metal–organic metal–organic frameworks: a new method for
framework composites. Chem. Soc. Rev. 43, selectively functionalized nanoporous hybrid
11. Li, B., Wen, H.-M., Wang, H., Wu, H., Tyagi, 5468–5512. carbon. J. Am. Chem. Soc. 137, 1572–1580.
M., Yildirim, T., Zhou, W., and Chen, B. (2014).
A porous metal–organic framework with 21. Sun, J.-K., and Xu, Q. (2014). Functional 33. Pachfule, P., Shinde, D., Majumder, M., and
dynamic pyrimidine groups exhibiting record materials derived from open framework Xu, Q. (2016). Fabrication of carbon nanorods
high methane storage working capacity. templates/precursors: synthesis and and graphene nanoribbons from a metal–
J. Am. Chem. Soc. 136, 6207–6210. applications. Energy Environ. Sci. 7, 2071– organic framework. Nat. Chem. 8, 718–724.
2100.
12. Mason, J.A., Oktawiec, J., Taylor, M.K., 34. deKrafft, K.E., Wang, C., and Lin, W. (2012).
Hudson, M.R., Rodriguez, J., Bachman, J.E., 22. Hu, Z., Deibert, B.J., and Li, J. (2014). Metal–organic framework templated
Gonzalez, M.I., Cervellino, A., Guagliardi, A., Luminescent metal–organic frameworks for synthesis of Fe2O3/TiO2 nanocomposite for
Brown, C.M., et al. (2015). Methane storage in chemical sensing and explosive detection. hydrogen production. Adv. Mater. 24, 2014–
flexible metal–organic frameworks with Chem. Soc. Rev. 43, 5815–5840. 2018.

78 Chem 2, 52–80, January 12, 2017


35. Xia, B.Y., Yan, Y., Li, N., Wu, H.B., Lou, X.W., photocatalytic activity for H2 production. Nat. 59. Nagarkar, S.S., Unni, S.M., Sharma, A.,
and Wang, X. (2016). A metal–organic- Commun. 7, 11580–11588. Kurungot, S., and Ghosh, S.K. (2014). Two-in-
framework-derived bi-functional oxygen one: inherent anhydrous and water-assisted
electrocatalyst. Nat. Energy 1, 15006–15013. 48. Kim, D., Whang, D.R., and Park, S.Y. (2016). high proton conduction in a 3D metal–organic
Self-healing of molecular catalyst and framework. Angew. Chem. Int. Ed. Engl. 53,
36. Bhakta, R.K., Herberg, J.L., Jacobs, B., photosensitizer on metal–organic framework: 2638–2642.
Highley, A., Behrens, R., Ockwig, N.W., robust molecular system for photocatalytic H2
Greathouse, J.A., and Allendorf, M.D. (2009). evolution from water. J. Am. Chem. Soc. 138, 60. Shigematsu, A., Yamada, T., and Kitagawa, H.
Metal–organic frameworks as templates for 8698–8701. (2011). Wide control of proton conductivity in
nanoscale NaAlH4. J. Am. Chem. Soc. 131, porous coordination polymers. J. Am. Chem.
13198–13199. 49. Nasalevich, M.A., Becker, R., Ramos- Soc. 133, 2034–2036.
Fernandez, E.V., Castellanos, S., Veber, S.L.,
37. Li, Z., Zhu, G., Lu, G., Qiu, S., and Yao, X. Fedin, M.V., Kapteijn, F., Reek, J.N.H., van der 61. Taylor, J.M., Mah, R.K., Moudrakovski, I.L.,
(2010). Ammonia borane confined by a metal– Vlugt, J.I., and Gascon, J. (2015). Co@NH2- Ratcliffe, C.I., Vaidhyanathan, R., and Shimizu,
organic framework for chemical hydrogen MIL-125(Ti): cobaloxime-derived metal– G.K.H. (2010). Facile proton conduction via
storage: enhancing kinetics and eliminating organic framework-based composite for ordered water molecules in a phosphonate
ammonia. J. Am. Chem. Soc. 132, 1490–1491. light-driven H2 production. Energy Environ. metal–organic framework. J. Am. Chem. Soc.
Sci. 8, 364–375. 132, 14055–14057.
38. Aijaz, A., Karkamkar, A., Choi, Y.J., Tsumori,
N., Rönnebro, E., Autrey, T., Shioyama, H., 50. Wang, C., Xie, Z., deKrafft, K.E., and Lin, W. 62. Zhai, Q.-G., Mao, C., Zhao, X., Lin, Q., Bu, F.,
and Xu, Q. (2012). Immobilizing highly (2011). Doping metal–organic frameworks for Chen, X., Bu, X., and Feng, P. (2015).
catalytically active Pt nanoparticles inside the water oxidation, carbon dioxide reduction, Cooperative crystallization of heterometallic
pores of metal–organic framework: a double and organic photocatalysis. J. Am. Chem. indium-chromium metal–organic polyhedra
solvents approach. J. Am. Chem. Soc. 134, Soc. 133, 13445–13454. and their fast proton conductivity. Angew.
13926–13929. Chem. Int. Ed. Engl. 54, 7886–7890.
51. Fu, Y., Sun, D., Chen, Y., Huang, R., Ding, Z.,
39. Gu, X., Lu, Z.-H., Jiang, H.-L., Akita, T., and Xu, Fu, X., and Li, Z. (2012). An amine- 63. Taylor, J.M., Komatsu, T., Dekura, S., Otsubo,
Q. (2011). Synergistic catalysis of metal– functionalized titanium metal–organic K., Takata, M., and Kitagawa, H. (2015). The
organic framework-immobilized Au/Pd framework photocatalyst with visible-light- role of a three dimensionally ordered defect
nanoparticles in dehydrogenation of formic induced activity for CO2 reduction. Angew. sublattice on the acidity of a sulfonated
acid for chemical hydrogen storage. J. Am. Chem. Int. Ed. Engl. 51, 3364–3367. metal–organic framework. J. Am. Chem. Soc.
Chem. Soc. 133, 11822–11825. 137, 11498–11506.
52. Wang, S., Yao, W., Lin, J., Ding, Z., and Wang,
40. Li, P.-Z., Aranishi, K., and Xu, Q. (2012). ZIF-8 X. (2014). Cobalt imidazolate metal–organic 64. Xu, G., Otsubo, K., Yamada, T., Sakaida, S.,
immobilized nickel nanoparticles: highly frameworks photosplit CO2 under mild and Kitagawa, H. (2013). Superprotonic
effective catalysts for hydrogen generation reaction conditions. Angew. Chem. Int. Ed. conductivity in a highly oriented crystalline
from hydrolysis of ammonia borane. Chem. Engl. 53, 1034–1038. metal–organic framework nanofilm. J. Am.
Commun. 48, 3173–3175. Chem. Soc. 135, 7438–7441.
53. Kajiwara, T., Fujii, M., Tsujimoto, M.,
41. Zhu, Q.-L., Li, J., and Xu, Q. (2013). Kobayashi, K., Higuchi, M., Tanaka, K., and 65. Bureekaew, S., Horike, S., Higuchi, M.,
Immobilizing metal nanoparticles to metal– Kitagawa, S. (2016). Photochemical reduction Mizuno, M., Kawamura, T., Tanaka, D., Yanai,
organic frameworks with size and location of low concentrations of CO2 in a porous N., and Kitagawa, S. (2009). One-dimensional
control for optimizing catalytic performance. coordination polymer with a ruthenium(II)-CO imidazole aggregate in aluminium porous
J. Am. Chem. Soc. 135, 10210–10213. complex. Angew. Chem. Int. Ed. Engl. 55, coordination polymers with high proton
42. Zhu, Q.-L., and Xu, Q. (2015). Liquid organic 2697–2700. conductivity. Nat. Mater. 8, 831–836.
and inorganic chemical hydrides for high-
54. Li, R., Hu, J., Deng, M., Wang, H., Wang, X., 66. Hurd, J.A., Vaidhyanathan, R., Thangadurai,
capacity hydrogen storage. Energy Environ.
Hu, Y., Jiang, H.-L., Jiang, J., Zhang, Q., Xie, V., Ratcliffe, C.I., Moudrakovski, I.L., and
Sci. 8, 478–512.
Y., and Xiong, Y. (2014). Integration of an Shimizu, G.K.H. (2009). Anhydrous proton
43. Kataoka, Y., Sato, K., Miyazaki, Y., Masuda, K., inorganic semiconductor with a metal– conduction at 150 C in a crystalline metal–
Tanaka, H., Naito, S., and Mori, W. (2009). organic framework: a platform for enhanced organic framework. Nat. Chem. 1, 706–710.
Photocatalytic hydrogen production from gaseous photocatalytic reactions. Adv. Mater.
water using porous material [Ru2(p-BDC)2]n. 26, 4783–4788. 67. Inukai, M., Horike, S., Itakura, T., Shinozaki, R.,
Energy Environ. Sci. 2, 397–400. Ogiwara, N., Umeyama, D., Nagarkar, S.,
55. Zhang, H., Wang, T., Wang, J., Liu, H., Dao, Nishiyama, Y., Malon, M., Hayashi, A., et al.
44. Wu, Z.-L., Wang, C.-H., Zhao, B., Dong, J., Lu, T.D., Li, M., Liu, G., Meng, X., Chang, K., Shi, (2016). Encapsulating mobile proton carriers
F., Wang, W.-H., Wang, W.-C., Wu, G.-J., Cui, L., et al. (2016). Surface-plasmon-enhanced into structural defects in coordination
J.-Z., and Cheng, P. (2016). A semi-conductive photodriven CO2 reduction catalyzed by polymer crystals: high anhydrous proton
copper-organic framework with two types of metal–organic-framework-derived iron conduction and fuel cell application. J. Am.
photocatalytic activity. Angew. Chem. Int. Ed. nanoparticles encapsulated by ultrathin Chem. Soc. 138, 8505–8511.
Engl. 55, 4938–4942. carbon layers. Adv. Mater. 28, 3703–3710.
68. Miner, E.M., Fukushima, T., Sheberla, D., Sun,
45. Fateeva, A., Chater, P.A., Ireland, C.P., Tahir, 56. Coupry, D., Addicoat, M., Yoneda, S., Tsutsui, L., Surendranath, Y., and Dinc
a, M. (2016).
A.A., Khimyak, Y.Z., Wiper, P.V., Darwent, J.R., Y., Sakurai, T., Seki, S., Wang, Z., Lindemann, Electrochemical oxygen reduction catalysed
and Rosseinsky, M.J. (2012). A water-stable P., Redel, E., Heine, T., and Wöll, C. (2015). by Ni3(hexaiminotriphenylene)2. Nat.
porphyrin-based metal–organic framework Photoinduced charge-carrier generation in Commun. 7, 10942–10948.
active for visible light photocatalysis. Angew. epitaxial MOF thin films: high efficiency as a
Chem. Int. Ed. Engl. 51, 7440–7444. result of an indirect electronic band gap? 69. Bao, Q., and Loh, K.P. (2012).
Angew. Chem. Int. Ed. Engl. 54, 7441–7445. Electrocatalytically active graphene-
46. Zhang, Z.-M., Zhang, T., Wang, C., Lin, Z., porphyrin MOF composite for oxygen
Long, L.-S., and Lin, W. (2015). 57. Chang, T.-H., Kung, C.-W., Chen, H.-W., reduction reaction. J. Am. Chem. Soc. 134,
Photosensitizing metal–organic framework Huang, T.-Y., Kao, S.-Y., Lu, H.-C., Lee, M.-H., 6707–6713.
enabling visible-light-driven proton reduction Boopathi, K.M., Chu, C.-W., and Ho, K.-C.
by a Wells-Dawson-type polyoxometalate. (2015). Planar heterojunction perovskite solar 70. Ma, S., Goenaga, G.A., Call, A.V., and Liu,
J. Am. Chem. Soc. 137, 3197–3200. cells incorporating metal–organic framework D.-J. (2011). Cobalt imidazolate framework as
nanocrystals. Adv. Mater. 27, 7229–7235. precursor for oxygen reduction reaction
47. Tian, J., Xu, Z.-Y., Zhang, D.-W., Wang, H., Xie, electrocatalysts. Chem. Eur. J. 11, 2063–2067.
S.-H., Xu, D.-W., Ren, Y.-H., Wang, H., Liu, Y., 58. Sadakiyo, M., Yamada, T., and Kitagawa, H.
and Li, Z.-T. (2016). Supramolecular metal– (2009). Rational designs for highly proton- 71. Aijaz, A., Fujiwara, N., and Xu, Q. (2014). From
organic frameworks that display high conductive metal–organic frameworks. J. Am. metal–organic framework to nitrogen-
homogeneous and heterogeneous Chem. Soc. 131, 9906–9907. decorated nanoporous carbons: high CO2

Chem 2, 52–80, January 12, 2017 79


uptake and efficient catalytic oxygen (2013). Superior lithium storage properties of catalysts for highly efficient hydrogen
reduction. J. Am. Chem. Soc. 136, 6790–6793. a-Fe2O3 nano-assembled spindles. Nano evolution. Angew. Chem. Int. Ed. Engl. 54,
Energy 2, 890–896. 12928–12932.
72. Yin, P., Yao, T., Wu, Y., Zheng, L., Lin, Y., Liu,
W., Ju, H., Zhu, J., Hong, X., Deng, Z., et al. 84. Zou, F., Chen, Y.-M., Liu, K., Yu, Z., Liang, W., 96. Fan, L., Liu, P.F., Yan, X., Gu, L., Yang, Z.Z.,
(2016). Single cobalt atoms with precise Bhaway, S.M., Gao, M., and Zhu, Y. (2016). Yang, H.G., Qiu, S., and Yao, X. (2016).
N-coordination as superior oxygen reduction Metal organic frameworks derived Atomically isolated nickel species anchored
reaction catalysts. Angew. Chem. Int. Ed. hierarchical hollow NiO/Ni/graphene on graphitized carbon for efficient hydrogen
Engl. 55, 10800–10805. composites for lithium and sodium storage. evolution electrocatalysis. Nat. Commun. 7,
ACS Nano 10, 377–386. 10667–10672.
73. Lin, Q., Bu, X., Kong, A., Mao, C., Zhao, X., Bu,
F., and Feng, P. (2015). New heterometallic 85. Férey, G., Millange, F., Morcrette, M., Serre,
97. Lu, X.-F., Liao, P.-Q., Wang, J.-W., Wu, J.-X.,
zirconium metalloporphyrin frameworks and C., Doublet, M.-L., Grenèche, J.-M., and
Chen, X.-W., He, C.-T., Zhang, J.-P., Li, G.-R.,
their heteroatom-activated high-surface-area Tarascon, J.-M. (2007). Mixed-valence Li/Fe-
and Chen, X.-M. (2016). An alkaline-stable,
carbon derivatives. J. Am. Chem. Soc. 137, based metal–organic frameworks with both
metal hydroxide mimicking metal–organic
2235–2238. reversible redox and sorption properties.
framework for efficient electrocatalytic
Angew. Chem. Int. Ed. Engl. 46, 3259–3263.
74. Zhu, Q.-L., Xia, W., Akita, T., Zou, R., and Xu, oxygen evolution. J. Am. Chem. Soc. 138,
Q. (2016). Metal–organic framework-derived 86. Nagarathinam, M., Saravanan, K., Phua, 8336–8339.
honeycomb-like open porous nanostructures E.J.H., Reddy, M.V., Chowdari, B.V.R., and
as precious-metal-free catalysts for highly Vittal, J.J. (2012). Redox-active metal- 98. Wurster, B., Grumelli, D., Hötger, D., Gutzler,
efficient oxygen electroreduction. Adv. centered oxalato phosphate open framework R., and Kern, K. (2016). Driving the oxygen
Mater. 28, 6391–6398. cathode materials for lithium ion batteries. evolution reaction by nonlinear cooperativity
Angew. Chem. Int. Ed. Engl. 51, 5866–5870. in bimetallic coordination catalysts. J. Am.
75. Li, Z., Shao, M., Zhou, L., Zhang, R., Zhang, C., Chem. Soc. 138, 3623–3626.
Wei, M., Evans, D.G., and Duan, X. (2016). 87. Zhang, Z., Yoshikawa, H., and Awaga, K.
Directed growth of metal–organic (2014). Monitoring the solid-state 99. Aijaz, A., Masa, J., Rösler, C., Xia, W., Weide,
frameworks and their derived carbon-based electrochemistry of Cu(2,7-AQDC) (AQDC = P., Botz, A.J.R., Fischer, R.A., Schuhmann, W.,
network for efficient electrocatalytic oxygen anthraquinone dicarboxylate) in a lithium and Muhler, M. (2016). Co@Co3O4
reduction. Adv. Mater. 28, 2337–2344. battery: coexistence of metal and ligand Encapsulated in carbon nanotube-grafted
redox activities in a metal–organic framework. nitrogen-doped carbon polyhedra as an
76. Choi, K.M., Jeong, H.M., Park, J.H., Zhang, J. Am. Chem. Soc. 136, 16112–16115. advanced bifunctional oxygen electrode.
Y.-B., Kang, J.K., and Yaghi, O.M. (2014). Angew. Chem. Int. Ed. Engl. 55, 4087–4091.
Supercapacitors of nanocrystalline metal– 88. Wu, D., Guo, Z., Yin, X., Pang, Q., Tu, B.,
organic frameworks. ACS Nano 8, 7451–7457. Zhang, L., Wang, Y.-G., and Li, Q. (2014). 100. Ma, T.Y., Dai, S., Jaroniec, M., and Qiao, S.Z.
Metal–organic frameworks as cathode (2014). Metal–organic framework derived
77. Wang, L., Feng, X., Ren, L., Piao, Q., Zhong, J., materials for Li-O2 batteries. Adv. Mater. 26, hybrid Co3O4-carbon porous nanowire arrays
Wang, Y., Li, H., Chen, Y., and Wang, B. (2015). 3258–3262. as reversible oxygen evolution electrodes.
Flexible solid-state supercapacitor based on
J. Am. Chem. Soc. 136, 13925–13931.
a metal–organic framework interwoven by 89. Li, Q., Xu, P., Gao, W., Ma, S., Zhang, G., Cao,
electrochemically-deposited PANI. J. Am. R., Cho, J., Wang, H.-L., and Wu, G. (2014).
101. US Department of Energy Office of Energy
Chem. Soc. 137, 4920–4923. Graphene/graphene-tube nanocomposites
Efficiency & Renewable Energy. Hydrogen
templated from cage-containing metal–
78. Zhang, Y.-Z., Cheng, T., Wang, Y., Lai, W.-Y., Storage. http://energy.gov/eere/fuelcells/
organic frameworks for oxygen reduction in
Pang, H., and Huang, W. (2016). A simple hydrogen-storage.
Li-O2 batteries. Adv. Mater. 26, 1378–1386.
approach to boost capacitance: flexible
supercapacitors based on manganese 90. Zheng, J., Tian, J., Wu, D., Gu, M., Xu, W., 102. US Department of Energy Advanced
oxides@MOFs via chemically induced in situ Wang, C., Gao, F., Engelhard, M.H., Zhang, Research Projects Agency – Energy. MOVE
self-transformation. Adv. Mater. 28, 5242– J.-G., Liu, J., and Xiao, J. (2014). Lewis acid- Program Overview. https://arpa-e.energy.
5248. base interactions between polysulfides and gov/sites/default/files/documents/files/
metal organic framework in lithium sulfur MOVE_ProgramOverview.pdf.
79. Sheberla, D., Bachman, J.C., Elias, J.S., Sun, batteries. Nano Lett. 14, 2345–2352.
C.-J., Shao-Horn, Y., and Dinc
a, M. (2016). 103. US Department of Energy. (2015). Hydrogen
Conductive MOF electrodes for stable 91. Bai, S., Liu, X., Zhu, K., Wu, S., and Zhou, H. Production. http://www.energy.gov/sites/
supercapacitors with high areal capacitance. (2016). Metal–organic framework-based prod/files/2015/06/f23/fcto_myrdd_
Nat. Mater. http://dx.doi.org/10.1038/ separator for lithium-sulfur batteries. Nat. production.pdf.
NMAT4766. Energy 1, 16094–16099.
104. US Department of Energy Office of Energy
80. Cao, F., Zhao, M., Yu, Y., Chen, B., Huang, Y., 92. Clough, A.J., Yoo, J.W., Mecklenburg, M.H., Efficiency & Renewable Energy.
Yang, J., Cao, X., Lu, Q., Zhang, X., Zhang, Z., and Marinescu, S.C. (2015). Two-dimensional Photovoltaics. http://energy.gov/eere/
et al. (2016). Synthesis of two-dimensional metal–organic surfaces for efficient hydrogen sunshot/photovoltaics.
CoS1.097/nitrogen-doped carbon evolution from water. J. Am. Chem. Soc. 137,
nanocomposites using metal–organic 118–121. 105. US Department of Energy Office of Energy
framework nanosheets as precursors for Efficiency & Renewable Energy. DOE
supercapacitor application. J. Am. Chem. 93. Hod, I., Deria, P., Bury, W., Mondloch, J.E.,
Technical Targets for Fuel Cell Systems and
Soc. 138, 6924–6927. Kung, C.-W., So, M., Sampson, M.D., Peters,
Stacks for Transportation Applications. http://
A.W., Kubiak, C.P., Farha, O.K., and Hupp,
energy.gov/eere/fuelcells/doe-technical-
81. Ogihara, N., Yasuda, T., Kishida, Y., Ohsuna, J.T. (2015). A porous proton-relaying metal–
targets-fuel-cell-systems-and-stacks-
T., Miyamoto, K., and Ohba, N. (2014). organic framework material that accelerates
transportation-applications.
Organic dicarboxylate negative electrode electrochemical hydrogen evolution. Nat.
materials with remarkably small strain for Commun. 6, 8304–8312.
106. United States Council for Automotive
high-voltage bipolar batteries. Angew. Chem.
94. Wu, H.B., Xia, B.Y., Yu, L., Yu, X.-Y., and Lou, Research. Energy Storage System Goals:
Int. Ed. Engl. 53, 11467–11472.
X.W. (2015). Porous molybdenum carbide Ultracapacitor Requirements. http://www.
82. Zheng, F., Yang, Y., and Chen, Q. (2014). High nano-octahedrons synthesized via confined uscar.org/guest/article_view.php?
lithium anodic performance of highly carburization in metal–organic frameworks for articles_id=85.
nitrogen-doped porous carbon prepared efficient hydrogen production. Nat. Commun.
from a metal–organic framework. Nat. 6, 6512–6519. 107. US Department of Energy Office of Energy
Commun. 5, 5261–5270. Efficiency & Renewable Energy. Vehicle
95. Tang, Y.-J., Gao, M.-R., Liu, C.-H., Li, S.-L., Technologies Office: Batteries. http://energy.
83. Banerjee, A., Aravindan, V., Bhatnagar, S., Jiang, H.-L., Lan, Y.-Q., Han, M., and Yu, S.-H. gov/eere/vehicles/vehicle-technologies-
Mhamane, D., Madhavi, S., and Ogale, S. (2015). Porous molybdenum-based hybrid office-batteries.

80 Chem 2, 52–80, January 12, 2017

You might also like