Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

View Online / Journal Homepage / Table of Contents for this issue

Energy & Dynamic Article Links < C


Environmental Science
Cite this: Energy Environ. Sci., 2012, 5, 9269
www.rsc.org/ees REVIEW
Metal organic frameworks for electrochemical applications
Adina Morozan and Frederic Jaouen*
Received 26th July 2012, Accepted 13th September 2012
DOI: 10.1039/c2ee22989g
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

Metal–organic frameworks (MOFs) have received huge attention in the last years as promising porous
materials with unrivalled degree of tunability for a wide range of applications including gas storage or
separation, catalysis, drug delivery and imaging. The present review appraises the application of MOFs
in the field of electrochemistry. From materials for rechargeable batteries, supercapacitors and fuel cells
Downloaded by Stanford University on 19 October 2012

to electrocatalysis or corrosion inhibition, MOFs or MOF-derived materials are gaining momentum in


this field. For real breakthroughs, combining their electrochemical properties with appropriate
electronic and ionic conductivity will be required.

I. Introduction of synthetically versatile imidazolate groups coordinating either


ZnII (ZIF 1 to 4, 6 to 8, and 10 to 11) or CoII (ZIF  9
First defined by Yaghi and co-workers in 1995 (ref. 1 and 2) and 12).12
following the pioneering work of Hoskins and Robson3–5 metal– MOFs can be obtained using inexpensive and high-yield
organic frameworks (MOFs) are crystalline porous hybrid synthesis methods at low temperature (250  C). Synthesis
materials comprising coupling units (metal ions or metal–oxo under hydro(solvo)thermal conditions is the most frequent
units) coordinated by electron-donating organic ligands. MOFs method, but microwave, ultrasonic, electrochemical and diffu-
have the highest specific surface areas reported to date for porous sional approaches have been investigated as well.13,14
materials (up to 6200 m2 g1)6,7 and large internal pore volumes MOFs can be prepared from relatively cheap precursors.
with well defined pore sizes and apertures. The term coordination Inorganic salts (nitrates, sulfates, and chlorides) are typically
polymer is also used and is often a synonym for MOF, except for used for the metal-ion precursor. The organic linkers are gener-
the metal-free structures that, by definition, cannot be included ally multidentate organic ligands such as carboxylates, azoles or
in the family of MOFs.8–10 nitriles.15,16 In many syntheses, structure-directing agents are also
MOFs represent a large class of materials in number, diversity employed to help assembling the building units into a MOF
of composition and structure. Examples of large families of without participating in the formula of the final compound.13,14
MOFs are (i) metal–organic polyhedras wherein transition metal In contrast to inorganic microporous materials with high
ions are coordinated by either nitrogen or carboxylate electron- specific surface area, several MOFs present a flexible structure
donor organic units,11 and (ii) zeolitic imidazolate frameworks that behaves dynamically and responds to external factors such
(ZIFs) which possess the topology of zeolites but are composed as guest molecules, temperature or pressure. The entire frame-
work of MOFs is supported by coordination bonds and/or other
weak cooperative interactions such as H-bonding, p–p stacking,
Institut Charles Gerhardt UMR CNRS 5253, Agr egats, Interfaces et
Mat eriaux pour l’Energie, Universit
e Montpellier II, Place Eug ene
and van der Waals interaction. Thus, structural flexibility is often
Bataillon, 34095 Montpellier cedex 5, France. E-mail: frederic.jaouen@ expected even under mild conditions,17,18 which allows these
univ-montp2.fr; Fax: +33 467143304; Tel: +33 467143211 MOFs to reversibly modulate their pore size depending on

Broader context
The chemical and structural diversity offered by metal–organic frameworks (MOFs) presents unique opportunities for catalysis, gas
separation, drug storage and delivery, imaging and sensing, optoelectronics or energy storage. Much on-going research in the field of
electrochemistry is involved in efficient and clean energy conversion and storage. Breakthroughs with novel materials are required to
improve the performance or lower the cost of electrochemical power sources. Compared to inorganic high-surface-area materials
that are the working principle of many current industrial electrocatalysts or electrode materials, MOFs offer the theoretical
advantage of 100% utilization and improved accessibility of the metal-cation centers due to their regular 3D dispersion in an open
structure. Moreover, MOFs are an outstanding platform to create bio-inspired enzyme mimics with high electrocatalytic activity
based on abundant transition metals.

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9269
View Online

external factors. This flexibility offers new possibilities with II. MOFs as electrode materials for rechargeable
exploitable properties that are important for catalysis,19 gas batteries
separation,20 drug storage and delivery,13,21 imaging and
sensing,22 optoelectronics23 or energy storage.24–26 II.1 Positive electrode in Li-ion battery
While the synthesis, chemistry and structure of MOFs and
The rechargeable lithium-ion (Li-ion) battery is the most
their application in various technological areas have been
advanced electrochemical energy conversion device today, and
described in comprehensive review papers,7,14,27–29 their appli-
much used for portable electronics. During discharge, Li inter-
cations in the field of electrochemistry have not been reviewed
calated/alloyed in the anode material is oxidized to lithium ions
previously. The investigation of MOFs in this area is quite
(Li+) that migrate to the cathode (positive electrode) where they
recent but expanding, as the present review will show. Impor-
insert. The main challenges are (i) electrode integrity over many
tant applications of electrochemistry are energy storage and
discharge–recharge cycles and (ii) improved capacity. The latter
conversion (supercapacitors, batteries, fuel cells), corrosion
target can only be reached with novel electrode materials. The
inhibition and reduction of highly oxidizing and toxic
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

low energy density of the positive electrode is a main factor


compounds. Electrochemistry, by definition, involves the
limiting the overall energy density of those batteries,37 mostly due
transfer or storage of electrons at the electrode–electrolyte
to the use of metal oxides with high atomic number relative to the
interface. The poor electron-conductive properties of most
anode material. Current standard materials are LiCoO2 at the
MOFs would a priori exclude them from being used as electrode
cathode and graphite at the anode with capacities of 148 and
materials or electrocatalysts. Nevertheless, MOFs have been
372 mA h g1, respectively.
Downloaded by Stanford University on 19 October 2012

successfully used as electrode materials for rechargeable


The first successful use of a MOF as a rechargeable inter-
batteries30,31 and strategies have been envisaged to overcome
calation electrode was reported by Ferey et al.30 Before that
their insulating character.32–35 The redox behaviour of metal
report, two other attempts had been made but Li-insertion led
cations inside MOFs could provide a pathway for electrons.33
to the transformation of the host material (Ni-based micro-
Alternatively, the tuning of the linker structure may lead to
porous phosphate or Zn-based MOF) into a Li2O matrix and
better charge transfer inside the framework.34 A bolder but
Ni nanoparticles or Zn-based nanocomposite, respectively.38,39
efficient strategy is to mix MOFs with conductive phases (metal
The groups of Ferey and Tarascon realized that a better
nanocrystals, carbon nanostructures, conductive polymers,
stability might be achieved through stronger metal–oxygen
etc.).32,35,36 This will however provide electron conduction at a
bonds, using 3d transition metals with fewer electrons in the 3d
macroscopic level, but not necessarily locally inside the MOF
orbitals than Ni or Zn, i.e. metals in a higher oxidation state
particles due to size-exclusion effects defined by the pore
(VIII, CrIII, FeIII). Also, to facilitate electron storage or release,
apertures.
a mixed-valence state of the metal was proposed in order to
The present review gives an account of the previous investi-
achieve long-range electron delocalization. Based on their
gations of MOFs as (i) electrode materials for batteries, (ii)
previous studies,40 the authors investigated the material
electrocatalysts for important reactions taking place in fuel cells
FeIII(OH)0.8F0.2(bdc)$H2O. This material was mixed with 15
or electrolyzers, (iii) electrode materials for supercapacitors, (iv)
wt% carbon to form the positive electrode for a Li-ion battery,
Li+ or H+ conductive materials for batteries or fuel cells, and (v)
while Li-metal was used as the negative electrode.30 With the
as surface films for corrosion inhibition. The review also includes
scalar x being defined by LixFe(OH)0.8F0.2(bdc)$H2O, a
the high surface-area carbons or carbon-based electrocatalysts
maximum lithium uptake of x ¼ 0.6 could be reached upon
resulting from the pyrolysis of MOFs. For every application, the
discharge (i.e. Li insertion in the cathode). At higher x-values,
state-of-the-art and challenges to be overcome are summarized
the discharge led to an irreversible behaviour, unsuitable for a
first. Performances obtained with MOFs or MOF-derived
rechargeable battery. The value of x ¼ 0.6 gives a gravimetric
materials are discussed in comparison to those obtained with
capacity of 75 mA h g1. This value is quite low due to the
state-of-the-art materials. For better readability, the acronyms
limited Li-insertion and/or the low density of the material.
used for the ligands are not explained within the text, but are
Li-insertion up to x ¼ 0.6 was fully reversible for at least 50
listed at the end.
cycles.

Fr
ederic Jaouen is a CNRS scientist at the Charles Gerhardt
Adina Morozan received her PhD from the University of Bucharest Institute for Molecular Chemistry and Material Sciences in
in 2007 on the topic of novel functional materials for fuel cells. Montpellier (France) since 2011. He was awarded the same year
After a two-year postdoctoral project on ‘‘Enzymes and organo- by the French ANR a chair of excellence at the University of
metallic catalysts in hydrogen fuel cells’’ at the laboratory of Montpellier II. He received his PhD in chemical engineering from
Chemistry of Surfaces and Interfaces at CEA Saclay, she joined in the Royal Institute of Technology of Stockholm in 2003 under the
2012 the Charles Gerhardt Institute for Molecular Chemistry and supervision of Prof. G€
oran Lindbergh. He then moved to Canada to
Material Sciences in Montpellier. She is currently a postdoctoral work with Prof. Jean-Pol Dodelet as a research associate at INRS.
researcher and her research activity focuses on synthesis, charac- His current research focuses on non-precious metal catalysts for
terization and degradation studies of non-precious metal catalysts proton exchange membrane fuel cells.
for the oxygen reduction reaction in a fuel cell environment.

9270 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

Changes in the coordination of the Fe ion have been followed discharge (250–400%) lead to fast capacity decay.44 Nano-
by M€ ossbauer spectroscopy. At x ¼ 0, a pure high-spin state of composite materials have been investigated to reduce this
FeIII was observed while at x ¼ 0.3 and 0.5, an increasing signal capacity decay.46 In the other approach, Li storage by the
of a high-spin FeII appeared. Quantitative analysis of the spectra reversible conversion mechanism is possible for a range of metal
gave, at x ¼ 0.5, the composition Li0.5FeII0.45FeIII0.55, i.e. a oxides (Cr, Mn, Fe, Co, Ni, Cu) and has led to experimental
mixed-valence material. The same group also investigated capacities of 700–900 mA h g1 with good cyclability.46,47
changes in that MOF structure upon Li insertion/extraction by These transition metal oxides reversibly react with Li+ according
in situ X-ray absorption fine structure (XAFS).41 The average to eqn (1):
oxidation state of Fe was found to decrease below +3 upon Li-
insertion, with a maximum molar ratio of 0.5 Li per Fe. While a MxOy + 2yLi+ + 2ye 4 xM + yLi2O (1)
local atomic reversible rearrangement around Fe was revealed
upon electrochemical cycling, the closest Fe–Fe distance was The formed Li2O matrix is active and metal nanoparticles are
conserved, showing the flexible structure locally around the Fe re-oxidized into transition metal oxides during the next
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

centres inside this MOF. The local flexibility of MOF materials discharge. Note that the conversion-reaction concept also applies
in general might ultimately be advantageous vs. the more rigid to metal nitrides, sulfides, etc.45
structures of today’s battery materials. Several MOFs have been used in innovative ways to form
More recently, the detailed mechanism of Li-insertion in the anodes that store Li+ via the conversion reaction mechanism. We
same material was investigated with density functional theory first review the use of MOFs investigated as such, and then
Downloaded by Stanford University on 19 October 2012

(DFT).42,43 Four different sites for Li insertion were identified MOFs that are used as a sacrificial framework to form transition
(Fig. 1). metal oxide nanoparticles that are subsequently used as anode
Upon discharge between 0 < x < 0.25, the A and B sites were materials.
predicted to be equally possible (A slightly more favorable), An as-synthesized MOF supporting the conversion-reaction
while for 0.25 < x < 0.5, the second Li insertion into the cell (4 Fe mechanism with Li was first reported by Li et al.39 The
atoms) was of similar energy for (i) 2 sites A occupied, (ii) 2 sites Zn4O(btb)2 MOF was synthesized with different morphologies
B occupied or (iii) 1 site A occupied and 1 site B occupied. The (Scanning Electron Microscope (SEM) images, Fig. 2), depend-
site A lies close to the hydroxyl bridging ligand of the inorganic ing on the reaction conditions used during the solvo-thermal
chains while site B lies close to the carboxylic group of the bdc- step.
linkers. Samples (c) and (d), in spite of their different morphologies,
yielded the same electrochemical response for Li storage. The
II.2 Negative electrode in Li-ion battery materials showed a large irreversible loss during the first cycle,
but retained the same capacity during cycles 2 to 50. The gravi-
Currently, anode materials (i.e. anode during the discharge) used metric capacity was however limited to 100–110 mA h g1 in
in commercial Li-ion batteries are graphitic carbon materials cycles 2–50, compared to 400 mA h g1 observed during the first
owing to their high stability and low cost. However, the inter- charge. An obvious decomposition of MOF particles during the
calation mechanism of Li+ in graphite defines a maximum first charge was observed in situ by Transmission Electron
gravimetric capacity of 372 mA h g1.44 The Li alloying mech- Microscopy (TEM). In 2010, Saravanan et al. reported on a more
anism or conversion mechanism represents alternative successful application of formate MOFs for Li-storage.31
approaches.45 Sn and Si are the two prominent candidates for the Zn3(HCOO)6, Co3(HCOO)6 and Zn1.5Co1.5(HCOO)6 were
alloying approach and form the SnLi4.4 and SiLi4.4 alloys with
theoretical gravimetric capacities of 993 and 4200 mA h g1,
respectively. However, their large volume changes upon charge/

Fig. 2 SEM images of samples of Zn4O(btb)2 obtained under different


synthesis conditions: (a) 100  C then natural cooling, (b) 100  C then
Fig. 1 Illustration of the four crystallographic sites A to D for lithium 0.5  C min1 cooling rate, (c) 120  C then 0.1  C min1 cooling rate, (d)
insertion in the MOF FeIII(OH)0.8F0.2(bdc). [Reprinted with permission 90  C then 0.1  C min1 cooling rate. [Reprinted with permission from
from ref. 42. Copyright 2011 Elsevier.] ref. 39. Copyright 2006 Elsevier.]

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9271
View Online

synthesized and investigated. The active materials were mixed on deep charge. Upon the subsequent discharge, the alloy is
with 15 wt% carbon black and 15 wt% binder for electrochemical transformed back into metal nanoparticles and then into zinc-
evaluation. The best results were obtained with Zn3(HCOO)6 formate MOF. The Co3(HCOO)6 MOF was also investigated. In
(Fig. 3). This MOF also shows a large irreversible loss during the contrast to Zn, Co does not form an alloy with Li, thereby the
first cycle, but interesting stability afterwards. After 60 cycles, the second step in the mechanism proposed above does not apply to
capacity of 560 mA h g1 was retained, corresponding to 9.6 Li Co3(HCOO)6. Its capacity stabilized to 410 mA h g1 after 60
atoms per repetitive unit of the MOF. The authors noted that cycles, which is equivalent to 6.8 Li atoms per repetitive unit of
these results are superior to those obtained with ZnO, showing the MOF. This is coherent with Co metal and Li-formate
typically only 200 mA h g1 after 10 cycles.48 Changes in the formation. Last, the mixed-metal MOF Zn1.5Co1.5(HCOO)6
structure and composition of the MOF during charge/discharge showed a capacity of 510 mA h g1 after 60 cycles.
were investigated by X-ray diffraction (XRD), Fourier-Trans- These works show that MOFs are promising as negative
form Infrared Spectroscopy (FT-IR) and TEM. Diffraction electrode materials for Li-ion batteries with performance
peaks were absent after the Zn3(HCOO)6 MOF had been elec- matching those obtained with the other most recent approaches.
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

trochemically charged or discharged. We now turn our attention to the use of MOFs as sacrificial
However, formate ions were still present after charge/ precursors to form transition metal oxide nanoparticles upon
discharge, as indicated by FT-IR, while no signal for Li2O was calcination. In the conversion reaction mechanism approach for
observed. TEM showed that the initial morphology of the MOF novel anode materials, controlled structure and texture of tran-
was retained with crystal lattice fringes observed by HR-TEM sition metal oxide nanoparticles is required.49 This is where
Downloaded by Stanford University on 19 October 2012

and the electron diffraction pattern corresponding to those of the MOFs can play a key role. Recently, several research groups
initial MOF structure. Based on these results, the following have synthesized transition metal oxide nanoparticles by using
conversion mechanism was proposed (eqn (2) and (3)): sacrificial MOFs, and investigated such oxides with a controlled
structure/morphology as anode materials.50,51 Liu et al. synthe-
Zn3(HCOO)6 + 6Li+ + 6e 4 3Zn + 6HCOOLi (2) sized Co3O4 nanoparticles by calcination at 600  C of a cobalt-
based MOF.50 The MOF was Co3(ndc)3(dmf)4. A pure Co3O4
3Zn + 3Li+ + 3e 4 3LiZn (3) phase was obtained, with an agglomerate structure of 250 nm,
composed of smaller primary particles of 25 nm. The measured
The MOF reacts first with Li+ to form Zn nanoparticles and Li specific surface area (5.3 m2 g1) corresponded closely to the
formate during the first charge, followed by LiZn alloy formation diameter of 250 nm (3.9 m2 g1 calculated for a bulky sphere of
Co3O4 with a diameter of 250 nm). This material was mixed with
5 wt% carbon black and 10 wt% polyvinylidene fluoride (PVDF)
to form an electrode that was subsequently tested electrochemi-
cally. The capacity was stable over 50 cycles, and even slightly
increased with cycling. A reversible capacity of 965 mA h g1 was
obtained after 50 cycles.50 In spite of the agglomerate structure
and low surface area of this cobalt oxide material, its capacity for
Li storage compares favourably to non-agglomerated cobalt
oxide nanostructures investigated by others.49,52 Thus, the
primary–secondary Co3O4 structure obtained from a cobalt-
based MOF via pyrolysis is a unique feature leading to improved
performance for cobalt oxide materials in Li-ion batteries.
In yet another approach for novel anode materials, mixed
transition metal oxide materials with a spinel structure are
considered. The series AFe2O4 (A ¼ Co, Ni and Zn) and ACo2O4
(A ¼ Zn or Cu) have shown capacities of 600–900 mA h g1.53,54
ZnMn2O4 has also been recently investigated.44 It showed a high
initial capacity of 1100 mA h g1 and a retained capacity of 600
mA h g1 after 20 cycles, much higher than the capacity after 20
cycles for ZnO or Mn2O3 taken separately (50 and 200 mA h g1,
respectively).44 The reversible overall reaction (eqn (4)) leads to a
theoretical capacity of 1008 mA h g1.

ZnMn2O4 + 9Li+ + 9e / ZnLi + 2Mn + 4Li2O (4)

ZnMn2O4 was obtained by co-precipitation of metal acetates


followed by calcination at 600  C, leading to spheroidal particles
Fig. 3 Electrochemical evaluation of Zn3(HCOO)6 MOF. (a) First
of 30–40 nm diameter. By pyrolysing appropriate MOF precur-
charge/discharge curve, (b) 2nd to 60th cycle, (c) capacity vs. cycle number. sors, Zhao et al. synthesised the spinel oxides AMn2O4 (A ¼ Zn,
The constant current for charge/discharge was 60 mA g1, corresponding Co or Ni) resulting in a unique gypsum-flower-like morphology
to 0.11 C rate. [Reproduced from ref. 31.] that is derived from the MOF.51 Other oxide structures obtained

9272 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

comprise a composite of mesoporous carbon and sulphur, so as


to efficiently deliver the electrons throughout the electrode and
minimize the diffusion of soluble polysulfide in the electrolyte.56
The capacity retention is still insufficient though, and it is critical
to further improve the positive electrode.
The idea behind the work developed by Tarascon’s group
relies on the use of a MOF as an improved matrix for sulphur
confinement.57 The investigated chromium trimesate MOF was
built up from oxocentered trimers of chromium octahedra linked
with btc-ligands (Cr3O(btc)2X, with X ¼ OH or F). This struc-
ture defines two types of mesoporous cages (25–29 A)  connected
through microporous windows (5–9 A).  The drawback of this
MOF is its insulating nature. Previous work showed however
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

that this is not redhibitory provided the insulating material is


mixed with conductive carbon to form the electrode.58 Thus, S
was impregnated within the pores of the MOF by melt-diffusion
at 155  C which is the melting point of S. The ratio between MOF
and S was adjusted to leave some free space for volume expan-
Downloaded by Stanford University on 19 October 2012

sion upon sulphur lithiation. The investigated MOF–S


composite contained 48 wt% S, and was subsequently mixed with
a high surface area carbon black by ball milling. Cells were
Fig. 4 Illustration of the ZnMn2–ptcda MOF precursor. (a) SEM image
assembled from this composite, a Li foil, a 1 M Li–organic
of ZnMn2–ptcda; (b) SEM and (c) TEM images of ZnMn2O4. [Repro-
duced from ref. 51.] electrolyte and a glass fiber separator. Melt-diffusion of the
MOF–S composite resulted in no crystalline phase for S (from
XRD analysis), attesting both a strong interaction between S and
by calcination of MOFs are discussed in Section IV.2. The the MOF surface, or the impossibility to create an ordered
authors started by synthesizing mixed-metal MOFs comprising structure within the confined space of the MOF. No amorph-
two metal cations using the corresponding metal acetates and ization of the MOF was noticed after ball milling. The specific
ptcda as a ligand. The synthesized MOFs, ZnMn2–ptcda, present surface area of 360 m2 g1 for the MOF–S composite indicated
a gypsum-flower-like morphology with a size of ca. 3 mm the filling of the majority of the pores by S, while the specific
(Fig. 4a). These MOFs were then calcinated in air at 450  C, surface area for the MOF itself is 1485 m2 g1. The best
resulting in powders of ZnMn2O4, CoMn2O4 or NiMn2O4, compromise between high initial capacity and capacity retention
respectively, as well as crystals of ptcda. The spinels retained the upon cycling was obtained with the MOF–S composite with 50%
morphology of the precursor MOFs, showing gypsum-flower- carbon additive (Fig. 5a and d-blue triangles). This material was
like morphology of 2–3 mm comprising 60 nm thick platelets as
shown in Fig. 4b and c. The specific surface area of this
ZnMn2O4 material was 42 m2 g1 and it has a mesoporous pore
structure, as shown by its type-IV adsorption isotherm.51
The spinel structure of ZnMn2O4 was investigated as an anode
material for Li-ion battery. It was mixed with 15 wt% acetylene
black and 10 wt% polyvinylidene fluoride (PVDF) to form an
electrode. An initial capacity of 800 mA h g1 (second cycle) was
measured, and it decreased slowly to ca. 500 mA h g1 after 30
cycles. This is comparable to, but not better than, results
obtained with the ZnMn2O4 material obtained by the co-
precipitation method and having a crystallite size in the range
50–150 nm depending on the calcination temperature.44

II.3 Positive electrode in Li–S battery


The rechargeable non-aqueous Li–S battery operates on
discharge by reduction of S at the cathode and Li oxidation to
Fig. 5 Electrochemical behavior during the first reduction of (a) MOF–
Li+ at the anode. The reduction of S leads to various poly-
S + 50% C composite, (b) mesoporous carbon–S, and (c) SBA-15–S +
sulphides before ending with the formation of Li2S.55 This type of
55% C composite in 1 M Li organic electrolyte at a C/20 discharge rate.
battery has a high theoretical specific energy (2567 W h kg1) but The right handside Y-axis in a, b and c is proportional to the amount of
suffers from several practical problems, i.e. capacity fading as a polysulfides released in the electrolyte. (d) Cycling performance of the
result of soluble polysulfide intermediates formed at the cathode composites in 1 M Li organic electrolyte at a C/10 discharge rate.
(Li2Sn with 3 # n # 6), and the insulating nature of S, Li2S and [Reprinted with permission from ref. 57. Copyright 2011 American
Li2S2 phases. The positive electrodes recently developed Chemical Society.]

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9273
View Online

further compared to other S-host materials investigated for the almost reversible on Pt, in alkaline medium, the same reaction is
same applications, i.e. a mesoporous silica (SBA-15, 7 nm pore 2–3 orders of magnitude slower and the associated voltage loss is
aperture, 740 m2 g1) and a mesoporous carbon obtained from not negligible even with Pt, the best known catalyst for that
SBA-15 by the template method (1160 m2 g1).59 The initial reaction.60 More efficient electrocatalysts in alkaline medium or
capacity of 52 wt% S infiltrated into SBA-15 and mixed with 50% cheaper electrocatalysts for the acidic medium are therefore
carbon is higher than that of the MOF/S–carbon composite but desirable.61
its capacity fading is faster (Fig. 5c and d-green circles). Finally, A new family of hybrid framework materials with uncommon
the mesoporous carbon obtained by replication of the meso- 3D structures such as polyoxometalate-based metal organic
porous SBA-15 consists of carbon nanorods of 6–7 nm diameter frameworks (POMOFs) emerged recently from the connection of
spaced by ca. 2 nm hexagonal arrangement (Fig. 5b). The initial polyoxometalates (POMs) to one another through bridging
capacity of the mesoporous carbon–S composite is between those organic linkers.62,63 Polyoxometalates are a vast class of clusters
of the two others but its capacity fading is the fastest, leading to made of oxyanions of transition metals in high oxidation states
the smallest capacity after only 32 cycles (Fig. 5d-red squares). such as WVI, MoVI or VV, linked together by oxo-ligands. The
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

Hence, the MOF–S composite resulted in the highest capacity extensive electrochemical activity of POMs in important elec-
retention which was ascribed to its unique pore shape. Moreover, trocatalytic reactions such as NOx, nitrate and nitrite reduction,
the existence of polar parts due to coordinatively unsaturated H2O2 oxidation and O2 reduction64–66 may be further exploited if
metal cations in MOF may further help in trapping the highly they are inserted into a rigid framework with a well-defined
polar polysulfides. The amount of polysulfide species released porous network.
Downloaded by Stanford University on 19 October 2012

into the electrolyte during the first discharge was quantified by a A series of POMOFs containing POMs building blocks con-
four-electrode cell (right hand-side Y-axis in Fig. 5a–c). It clearly nected by btc-linkers was synthesized and revealed promising
demonstrates that the MOF–S composite released the smallest catalytic activity for the HER at pH 1.67 These POMOFs were
amount of polysulfides. In conclusion, the S confinement is more developed from 3-Keggin’s type POMs with chemical formula
important than active material conductivity, since the latter can 3-PMoV8MoVI4O40x(OH)xM4, where M ¼ ZnII or LaIII. This
be imparted by additional carbon to form the electrode. structure features an 3-Keggin core surrounded by four metal
Reducing or controlling the MOF particle size may further cations. In their paper, the authors investigated the structure
alleviate the disadvantage induced by its insulating or poorly with M ¼ ZnII. In ref. 67, the compound (tba)3[PMoV8MoVI4O36-
conductive character. (OH)4Zn4](btc)4/3$6H2O (labelled 3(trim)4/3) was synthesized
and characterized. Two other compounds (labelled 32(trim)2
and [3(trim)]N) with different stoichiometries and topologies
III. MOFs for electrocatalysis were also obtained upon slight changes in the synthesis
Before going into the review of the results, it is emphasized here conditions (Fig. 6). It was shown that tba cations not only act
that, while in the reviewed literature many studies reported their as a simple counterion but also as a structure-directing agent,
data against a practical reference electrode (Ag/AgCl, saturated since attempts to replace those with smaller cations failed.
calomel, etc.), we discuss here and give electric potentials vs. the
Reversible Hydrogen Electrode (RHE), which is more mean-
ingful for reactions involving protons or hydroxyl-anions and
having therefore a pH-dependent thermodynamic potential. The
change of scale was done according to eqn (5):

E(V vs. RHE) ¼ E(V vs. ref.) + E(V of ref. vs. NHE)
+ 0.06  pH (5)

where pH is the pH of the electrolyte used for the experiments


and NHE is the Normal Hydrogen Electrode defined by aH+ ¼ 1
and aH2 ¼ 1.

III.1 Hydrogen evolution reaction


The hydrogen evolution reaction (HER) in acidic or alkaline
medium is important for electrolyzers. Large production of H2 at
the lowest possible cell voltage is paramount in setting low price
electrolytically-produced hydrogen. While most H2 is currently
produced by steam reforming of methane (with high CO
content), in the future, only H2 produced from water is desirable,
Fig. 6 POM building blocks of 3(trim)4/3, 32(trim)2, and [3(trim)]N, a
for environmental and sustainability reasons, as well as for the view of their unit cell, and their schematic representation. Comparison of
sake of the H2 purity, which is paramount for low-temperature 3(trim)4/3, 32(trim)2, and [3(trim)]N at a scan rate of 2 mV s1, highlighting
fuel cells. The HER is also the cathode reaction in the membrane- the electrocatalytic process of HER at pH 1 (1 M LiCl + HCl). [Reprinted
and diaphragm chlor-alkali processes used to produce Cl2 and with permission from ref. 67. Copyright 2011 American Chemical
where H2 is a by-product. While in acidic medium, the HER is Society.]

9274 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

After synthesis, the tba cations were exchanged for smaller molecule instead of only 2 electrons for the HER. Only a few
cations, but this resulted in a partial decomposition of the materials combine sufficient electrocatalytic activity and stability
POMOFs. at the high electric potential experienced during electrolysis
Cyclic voltammograms in Ar-purged 1 M LiCl + HCl (pH 1) (1.6–2 V vs. RHE). The current state-of-the-art catalyst is
electrolyte showed two redox waves characteristic of the revers- considered to be IrO2, but Ir is a very expensive and rare metal,
ible oxido-reduction of Mo centers. The redox waves were even when compared to Pt.68,69 In alkaline medium, recent
centered at ca. +0.42 V vs. RHE (+0.12 V vs. SCE) for 3(trim)4/3 perovskites based on more abundant elements have shown mass
and 32(trim)2 and +0.20 to +0.18 V vs. RHE (0.1 to 0.12 V vs. activities similar to that of IrO2 and one-order of magnitude
SCE) for 32(trim)2 and 3(trim)4/3. An electrocatalytic reduction higher surface-specific activity.70 Nickel-based catalysts71,72 or
process was revealed with an onset potential of ca. +0.3 V vs. mixed oxides73,74 are other alternative electrocatalysts for the
RHE at pH 1 (0 V vs. SCE) and ascribed to HER (Fig. 6). H2 OER in alkaline medium.
production was verified by online mass spectroscopy with the The group of Wenbin Lin synthesized the MOF
electrode under potentiostatic control at 0.1 V vs. RHE. Zr6O4(OH)4(bpdc)6 doped with various metal complexes (Ir-,
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

However, the value of +0.3 V vs. RHE for the onset potential for Re-, or Ru-based) and investigated their activity for a range of
HER is very intriguing since it is thermodynamically impossible. reactions related to solar energy utilization (water oxidation,
One possible explanation for this onset potential significantly photocatalytic CO2 reduction, and organic photocatalysis).75
above 0 V vs. RHE is a local pH within the POMOFs being lower The metal complexes investigated for OER were: (complex 1)
than that in the bulk electrolyte. Hence, this abnormal behaviour [IrIII(Cp*)-(dcppy)] (Cp* ¼ pentamethylcyclopentadienyl);
Downloaded by Stanford University on 19 October 2012

makes it difficult to interpret the curve and separate the electro- (complex 2) [IrIII(Cp*)(dcbpy)]+ and (complex 3) [IrIII(dcp-
catalytic and local pH effects. Electrolytes other than 1 M LiCl + py)2(H2O)]+. These three Ir-based complexes were evaluated as
HCl (pH 1) were investigated, by replacing Li+ with smaller such for the OER, or after incorporation in the MOF
cations (Na+, K+, Cs+). The catalytic process for HER was Zr6O4(OH)4(bpdc)6 (labelled MOFs 1–3, with specific surface
decreased by 2.5–3 when Li+ was replaced by Na+ or K+, and area 1250–1500 m2 g1). Water electro-oxidation to O2 was
replacement by Cs+ resulted in a drastic alteration of the CV and investigated in homogeneous solution at pH 1 with Ce4+ as an
HER as well as material instability. The following mechanism for oxidant, resulting in a total reaction described by eqn (9), being
HER on this POMOF was proposed (eqn (6)–(8)): the sum of two electrochemical reactions (the OER and Ce4+ to
Ce3+ electro-reduction).
POMOF + 2e + 2H+ ¼ H2POMOF (6)
4Ce4+ + 2H2O / 4Ce3+ + O2 + 4H+ (9)
H2POMOF + 2e+2(Li(H2O)n)+ ¼ [Li2(H2O)2nPOMOF]
+ H2 (7) The amount of O2 produced was followed with time and the
three Ir-complexes as well as the three resulting MOFs 1–3
[Li2(H2O)2nPOMOF] + 2H+ ¼ H2POMOF + 2(Li(H2O)n)+ (8) showed some activity. The order of activity was 1 > 2 > 3, both
for the complexes and the resulting MOF–complex composites.
where eqn (6) is the first redox wave (highest potential) and
About 6–8 wt% of the complexes were incorporated into the
corresponds to POMOF reduction; eqn (7) is the second redox
Zr6O4(OH)4(bpdc)6 structure, resulting in an apparent activity
wave and corresponds to the replacement of H+ by Li+ in the
being lower for the MOFs 1–3 vs. the pure complexes 1–3.
POMOF and the concurrent H+ reduction from H2POMOF to
However, with respect to the Ir mass, the activity was equal or
H2; eqn (8) corresponds to the replacement of Li+ by H+ without
higher for the MOF–complex composites. As a control experi-
any reduction or oxidation of species. Clearly, this mechanism
ment, the parent MOF Zr6O4(OH)4(bpdc)6 was also investigated
involves the alkali cation, which ties in with experimental
but showed no activity for the OER, demonstrating the impor-
observations. In eqn (8), H2POMOF is regenerated, which allows
tance of Ir-based active centers.
a cycle of reactions from eqn (7) and (8) to proceed and H2 to be
In summary, activity for the OER at pH 1 was demonstrated
produced in an electrocatalytic process. According to this
by these Ir-complexes and MOFs incorporating such complexes,
mechanism, the onset potential for HER would be set by the
but since the OER was investigated without performing direct
H2POMOF/H2 redox potential, which might be different from
electrochemistry, the lack of control of the electric potential does
that of the H+/H2 couple.
not allow one to draw a clear conclusion about the absolute
In summary, the H2 production starting at around +0.3 V vs.
activity of such catalysts. Electrochemical investigation of such
RHE would present a great advantage for an electrolysis cell, and
Ir-complex–MOF hybrid materials would be interesting.
could in principle lower the cell voltage and result in lower
The electrochemical reactivity of the commercially available
electrical energy required to produce H2. It remains to be verified
Fe(btc) MOF (Basolite F300) was investigated by Babu et al. in
if such materials can sustain the high current densities required
alkaline medium (pH 11.3–12.2).76 At potentials above +1.7 V vs.
for an industrial application for water electrolysis.
RHE, an oxidation current was observed and ascribed to the
OER (Fig. 7a). However, this oxidation current was not observed
if the electrode potential had not been previously scanned at
III.2 Oxygen evolution reaction
potentials lower than +1.5 V vs. RHE (ca. 0.4 V vs. SCE)
The Oxygen Evolution Reaction (OER), the anodic reaction in (Fig. 7a). The authors noticed that this potential corresponds to
water electrolyzers and recharging aqueous metal–air batteries, is the FeII/FeIII redox potential in this MOF and they discuss a
a very sluggish reaction due to the 4 electrons involved per water possible surface modification of the MOF at low potential that

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9275
View Online

cells.37 Recently, Fe(Co)–N–C catalysts with molecular active


sites of the type metal–Nx obtained by pyrolysis at high
temperature have shown interesting activity and power perfor-
mance in PEM fuel cells.80–82 Pyrolyzed and non-pyrolyzed
transition metal macrocycles and complexes have shown inter-
esting ORR activity in alkaline medium83–85 but had, until
recently, failed to show sufficient ORR activity in acid medium.
Due to the accessibility of metal cations in MOF structures, the
high volumetric density of metal-ion sites and high micropore
surface area, MOFs are interesting as such or as sacrificial
materials to be pyrolyzed, in order to synthesize non-precious
metal catalysts (NPMC) for the ORR.
Liu and co-workers from Argonne National Laboratory pio-
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

neered the use of MOF as a sacrificial precursor for preparing


ORR catalysts.86,87 The cobalt imidazolate framework was
selected as a precursor for the synthesis of ORR catalysts
through pyrolysis because of its analogous structure with the
Co–N4 (or Fe–N4) macromolecules previously used as precursors
Downloaded by Stanford University on 19 October 2012

for the preparation of NPMCs.83 The original MOF structure


was not retained after the thermal activation above 600  C, but it
played a key role in generating a high density of catalytic-sites
with proper accessibility for oxygen. An optimal stability and
Fig. 7 (a) Cyclic voltammograms at 50 mV s1 for the Fe(btc) MOF activity for the ORR at pH 1 was achieved after pyrolysis of the
immobilised on a platinum-disk electrode immersed in aqueous 0.1 M MOF in inert atmosphere at 750  C, with an onset potential of
KCl with (i) 2, (ii) 4, (iii) 6, (iv) 8, (v) 10, and (vi) 15 mM NaOH. The inset 0.83 V vs. RHE and the ORR mechanism characterized by 3.2–
shows the peak current vs. hydroxide concentration. (b) Cyclic voltam- 3.5 electrons per O2 molecule. These are close to state-of-the-art
mograms at 50 mV s1 with a start potential of (i) 0.5, (ii) 0.4, (iii) 0.3, (iv) results for cobalt-based NPMCs. The specific surface area
0.1, and (v) 0.0 V vs. SCE for the Fe(btc) MOF immobilised on the
decreased from 300 to 264 m2 g1 after pyrolysis at 750  C, but
platinum disk electrode immersed in aqueous 0.1 M KCl + 5 mM NaOH.
increased to 434 m2 g1 after removal of cobalt inactive species
(c) Schematic drawing of MOF attached to the working electrode surface
before and after reductive catalyst formation. [Reprinted with permission by acid-washing of the pyrolysed product.
from ref. 76. Copyright 2010 Elsevier.] Another MOF, a ZnII zeolitic-imidazolate-framework was
selected for its high nitrogen content and high microporous
surface area (1700 m2 g1) in order to form an ORR-catalyst
might create iron hydroxides (Fig. 7c). The discussion was precursor when mixed with 1,10-phenanthroline and ferrous
however not supported by experimental facts at this stage. The acetate.82 The MOF (Basolite Z1200, or ZIF-8) is composed of
response of the clean Pt electrode gave a smaller oxidation Zn(MeIm)2 primary units and shows a sodalite-type structure,
current than observed for the MOF immobilized on Pt, exhibiting an interesting nanopore topology formed by bridging
demonstrating a role for the MOF. The phenomenon occurring the ZnII centres to N-atoms of imidazolate ligands. An optimized
during the low potential sweep and the exact nature of the active Fe/N/C-catalyst was prepared from the pyrolysis in argon at
sites need to be identified in the future. 1050  C, then in ammonia at 950  C, of the catalyst precursor
obtained by low-energy planetary ballmilling of ZIF-8, phe-
nanthroline and ferrous acetate mixed in optimized ratios. After
III.3 Oxygen reduction reaction
the heat treatments, the catalyst precursor including the MOF is
The slow kinetics of the Oxygen Reduction Reaction (ORR) is transformed into a nitrogenated microporous carbon structure
responsible for the major source of voltage loss in low temper- hosting FeNx active sites. The microporous surface area of the
ature H2/air fuel cells, even when the best Pt-based catalysts are optimized catalyst was ca. 1000 m2 g1. The activity and power
used.77 In alkaline fuel cells, a variety of catalysts is available performance of the optimized catalyst were measured in a H2–O2
(platinum-group metals (PGM), gold, silver, transition metal PEM fuel cell and yielded a current density of 1.2–1.3 A cm2 at
oxides and chalcogenides, perovskites and spinels of rather 0.6 V (Fig. 8) and a peak power density of 0.91 W cm2. This is
abundant elements, as well as transition metal macrocycles) with 2–2.5 times higher than the best previously reported performance
ORR activity similar to that obtained with platinum or plat- for NPMC obtained with the same synthesis procedure but using
inum-group-metals, while in acidic fuel cells, the choice of cata- a high surface-area carbon black as a microporous host80 instead
lysts is much more restricted due to the poor activity for the ORR of the MOF (Fig. 8). These new results bring hope that Pt could
or poor durability at low pH of otherwise active materials.78,79 be ultimately replaced by NPMCs in PEM fuel cells. Stability of
Today, supported Pt nanoparticles, alloyed or not with more the optimized NPMC catalyst is now the main barrier for an
abundant transition metals, is the state-of-the-art catalyst for industrial application.
proton exchange membrane (PEM) fuel cells. Due to the low More recently, modifications of this synthesis method were
availability and high cost of Pt, alternative catalysts for the ORR explored by changing the MOF (Basolite F300, a commercially
are highly desirable in order to decrease the cost of PEM fuel available Fe–btc MOF with a specific area of 1500 m2 g1),

9276 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

Fig. 8 H2/O2 PEM fuel cell polarisation curves for a state-of-the-art Pt-
based cathode (green squares, 0.3 mgPt cm2), a Fe-based cathode
catalyst synthesized using the MOF ZIF-8 (blue stars) and a Fe-based
Downloaded by Stanford University on 19 October 2012

cathode catalyst synthesized using a high surface area carbon black


instead of ZIF-8. [Reprinted by permission from Macmillan Publishers Fig. 9 Scheme of the chemical structures of (a) reduced graphene oxide
Ltd: Nat. Commun. (ref. 82), copyright 2011.] (r-GO), (b) G-dye, (c) H2–porphyrin, (d) (Fe–P)n MOF, (e) (G-dye–FeP)n
MOF, and (f) magnified view of layers inside the framework of (G-dye–
FeP)n MOF showing how graphene sheets intercalated between
changing the Fe precursor (Fe–phthalocyanine) or changing the porphyrin networks. [Reprinted with permission from ref. 89. Copyright
organic precursor (polyaniline).88 Only minor changes in elec- 2012 American Chemical Society.]
trochemical performance were observed as long as the catalyst
precursor contained ZIF-8 as the MOF, but major negative auxiliary ligand. The latter stabilizes the structure in aqueous
changes were observed when ZIF-8 was substituted for F300. In medium, since the Cu–btc MOF is not stable enough.90 In
contrast to ZIF-8 where Zn native of the MOF conveniently Cu–bipy–btc, each CuII center is coordinated with a carboxyl
forms volatile products at high temperature, Fe native of F300 group from btc, one O atom from water, and two chelated N
only accumulates during heat treatment and catalyzes graphiti- atoms from bipy (Fig. 10b). This coordination type shows a
zation, resulting in excess Fe and low microporous surface area better structural stability in aqueous media relative to the
after pyrolysis. Moreover, and in contrast to ZIF-8, F300 does Cu-coordination in Cu–btc (Fig. 10a).
not contain nitrogen, which means that the only source of N in The Cu–bipy–btc-modified electrode shows a redox wave at
the catalyst precursor is phenanthroline or polyaniline. In ca. +0.40 V vs. RHE (0.15 V vs. Ag/AgCl) (Fig. 10c) ascribed to
summary, the MOF chemistry obviously plays an important role
for sacrificial MOFs to be used for the synthesis of NPMCs by
pyrolysis.
Some ORR catalysts with the preserved MOF structure (no
pyrolysis step) have also been recently reported.89,90 It is well
known that porphyrins may be used as building blocks to form
2D or 3D MOFs.91–93 The introduction of metallo-porphyrins in
MOFs is attractive because of the extraordinary catalytic prop-
erties of metal-porphyrins for biological reactions. Kian Ping
Loh’s group reported in 2012 on a composite MOF structure
(G-dye–FeP)n comprising pyridine-functionalized graphene
(reduced graphene oxide sheets functionalized with donor–p–
acceptor dye containing pyridinium moieties, and labelled
G-dye) and an iron porphyrin (5,10,15,20-tetrakis(4-carboxyl)-
iron-porphyrin; labelled FeP). G-dye was used as a building
block for the assembly of MOFs. The composite graphene–
metalloporphyrin metal–organic framework formed by the
combination of G-dye and FeP (Fig. 9) exhibits interesting
Fig. 10 (a) Cu coordination in Cu–btc; (b) Cu coordination in Cu–bipy–
catalytic activity towards the ORR in alkaline medium with an
btc; (c) typical CVs obtained at the Cu–bipy–btc-modified glassy carbon
onset potential of ca. 0.93 V vs. RHE. The larger bond polarity
electrode in N2- (dotted curve) or O2-saturated electrolyte (solid curve);
due to nitrogen ligands in the G-dye, the catalytically active iron– (d). typical rotating-ring disk electrode voltammograms obtained with
porphyrin and the framework porosity act synergistically to bare (black curve) and Cu–bipy–btc-modified (red curve) glassy carbon
afford a near 4-electron ORR pathway.89 electrode (solid curves) and platinum-ring current (dotted curves) under
Another example of a MOF-based NPMC for the ORR is O2-saturated electrolyte. The electrolyte pH was 6. [Reprinted with
offered by CuII–bipy–btc with btc as a main ligand and bipy as an permission from ref. 90. Copyright 2012 Elsevier.]

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9277
View Online

the CuI/II oxidation states. A large increase in the reduction ammonium are 1.51 and 0.9 V vs. SHE, respectively. Bromate
current in O2-saturated medium attests the electrocatalysis of the (BrO3) is a highly toxic and carcinogenic compound that is
ORR. A stable electrocatalytic behavior of the Cu–bipy–btc formed upon water disinfection by ozone as a side reaction
MOF was obtained at pH 6 with an onset potential for the ORR between ozone and bromide anions present in water.97
of ca. +0.5 V vs. RHE (ca. 50 mV vs. Ag/AgCl at pH 6) The BrO3/Br2 and BrO3/Br standard potentials are 1.48 and
(Fig. 10d), much higher than that obtained with the bare glassy 1.42 V vs. SHE, respectively.
carbon electrode (Fig. 10c). The electron number involved in the Electrocatalytic reduction of NO2 to N2O by CuII complexes
Cu–bipy–btc-catalyzed ORR was 3.8 (calculated from Fig. 10d), through one-electron reduction of CuI/CuII set the basis for the
which is close to 4 expected for the full electro-reduction of O2 synthesis of a novel MOF with similar Cu-coordination.98 The
to H2O. MOF [Cu2(bpdc)2(dpq)2(H2O)]$H2O was obtained from the
hydrothermal reaction of copper chloride with the mixed ligands
bpdc and dpq.99 Due to its insolubility in water, the MOF could
III.4 Alcohol oxidation
be employed to fabricate MOF/carbon/paraffin paste electrodes,
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

While hydrogen is by far the most easily electro-oxidized fuel for which were then investigated for the electrocatalytic reduction of
low temperature fuel cells, the challenges associated with its nitrite. The catalytic currents corresponding to the reduction of
storage, especially at small scale, leave a door open to the use of nitrite in pH 2 phosphate buffer solution displayed a linear
alcohols as a liquid fuel with high theoretical energy density for relation with NO2 concentration (0–4 mM range) at a potential
fuel cells. The activity and stability of Pt or PGM, especially of ca. +0.27 V vs. RHE. The electrode also showed a high
Downloaded by Stanford University on 19 October 2012

under acidic environment, make it a suitable anode electro- stability that could be attributed primarily to the insolubility of
catalyst for electro-oxidation in low temperature fuel cells. this MOF. The same MOF was investigated for the bromate
However, and besides the restrictive aspects of cost and avail- electroreduction (see later).
ability, Pt or PGM are readily poisoned by the strongly- As previously discussed in Section III.1, MOFs incorporating
adsorbing intermediates of alcohol oxidation, including CO. POMs can be synthesized to take advantage of the extensive
Therefore, the search for an efficient, cheap and durable anode electrochemical activity of polyoxometallate compounds, if these
catalyst is important. soluble compounds can be retained in insoluble MOFs.67 A 3D
The first example of the utilization of a noble-metal-free MOF- Cu-based MOF incorporating the small polyoxomolybdate
based material as a catalyst for ethanol electro-oxidation was {Mo2} has been synthesized via an in situ reaction under
recently reported by Kitagawa’s group.94 The MOF was N,N0 - hydrothermal conditions.100 In this novel MOF [CuI(pci)
bis(2-hydroxyethyl)dithiooxamidato-copper(II) [(HOC2H4)2d- Mo2VO5], the (Mo2VO5N)n dimeric chains alternately connect to
toaCu], a 2D MOF composed of dimeric Cu units with a Cu–Cu copper chains via corner-sharing mode to form a bimetallic
bond distance of 2.60 A  and bridging ligands (HOC2H4)2dtoa2. (CuIMo2VO5N)n layer. These adjacent layers are linked by pci
Each Cu ion is linked to two N, two S and one Cu atom. ligands. Both the polyoxomolybdate oligomer and copper ions
Previously, Kitagawa et al. had reported a proton conductivity of have catalytic properties, and the [CuI(pci)Mo2VO5]-modified
104 S m1 for this MOF, at 75% relative humidity and room carbon paste electrode has a good electrocatalytic activity
temperature and an electron conductivity of 106 S m1.95 This toward the reduction of nitrite in acidic medium.100
MOF was coated on a glassy-carbon electrode, and subsequently Another POMOF with electrocatalytic activity for the nitrite
investigated in 0.5 M H2SO4 solution. Cyclic voltammetry reduction was recently reported.101 The Wells–Dawson-type
showed one redox peak centered at 0.35 V vs. Ag/AgCl and polyoxometalate (As2W18O62)6 was used as a template to build
ascribed to CuI/CuII. Addition of ethanol resulted (i) in an the POMOF Co2(bpy)3(ox)(As2W18O62)$2(H2bpy)$2H2O.
increase of the redox peak, ascribed to some ethanol oxidation Analysis by XRD showed that this compound consists of a non-
and charge transfer to Cu atoms, and (ii) a sustainable ethanol coordinating POM template and 3D host framework. The elec-
oxidation current starting at 0.84 V vs. RHE. The product of trocatalytic activity of this compound was investigated in 1 M
ethanol oxidation was acetaldehyde (2 electron oxidation), H2SO4 aqueous solution with carbon paste electrodes. In the
instead of the full six-electron oxidation to CO2. This means that nitrite-free electrolyte, three stable redox pairs were visible and
the overpotential for ethanol oxidation is very high, since the ascribed to the tungsten centers of the POM. As expected due to
ethanol–acetaldehyde redox potential is 0.21 V vs. RHE. the known activity of POMs for nitrite reduction, this hybrid
compound also showed electrocatalytic activity toward the
reduction of nitrite with an onset potential of +0.28 V vs. RHE
III.5 Nitrite and bromate reduction
(SI of ref. 101).
Nitrate and nitrite contamination from fertilizers in groundwater Yet another POMOF based on the 3-Keggin anion and a Zn-
is a major concern.96 Removal of these highly oxidizing species based MOF was synthesized and demonstrated electrocatalytic
could be performed by electrolysis, the ultimate goal being their reduction of bromate (Fig. 11).102 Going from modeling methods
reduction to the inert gas N2 but many intermediate products to construct and evaluate the stability of several hypothetical
exist (NO2, NO, N2O). The nitrate or nitrite electro-reduction to POMOFs, the synthesis of a viable and stable structure was
N2 is a complex multi-electronic reaction (5 and 3 electrons to achieved, namely [NBu4]3[PMoV8MoVI4O36(OH)4Zn4(bdc)2]$
be reduced to N2 for nitrate and nitrite, respectively), and 2H2O. Electrochemistry at pH 2 was carried out by entrapping
hence very sluggish. The standard thermodynamic potentials for the POMOF in a carbon paste or in polyvinyl alcohol, to analyze
NO3/N2 and NO3/NO2 are 1.25 V and 0.835 V vs. SHE, the redox waves and the electro-reduction of BrO3. A linear
respectively. Standard potentials for nitrite reduction to N2 and correlation was found between BrO3 concentration (range

9278 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

incorporating pseudo-capacitive redox centers are therefore


potential candidates as electrode materials for EDLCs with
improved energy density. This is discussed in Section IV.1.
Alternatively, MOFs can also be used as a sacrificial micropo-
rous template for the preparation of either nano–micro scale
metal-oxide materials (Section IV.2) or microporous carbon
materials (Section IV.3) and both may find application for
EDLC or for other electrochemical applications.

IV.1 As-synthesized MOFs

The introduction of as-synthesized MOFs in EDLCs has been


reported only recently.105,106 Since cobalt oxides, and especially
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

Co3O4, are known to be an efficient material for electrodes in


supercapacitors,107 Lee et al. synthesized from cobalt nitrate and
terephthalic acid a Co-based MOF and investigated its capacitive
behaviour in various 1 M aqueous electrolytes, namely LiOH,
KCl, LiCl and KOH.105 Only in LiOH-electrolyte did the Co-
Downloaded by Stanford University on 19 October 2012

based MOF show an interesting pseudo-capacitive behaviour


Fig. 11 Cyclic voltammetry of [NBu4]3[PMoV8MoVI4O36(OH)4- (Fig. 12). A gravimetric capacitance of 150–200 F g1 was
Zn4(bdc)2]$2H2O at 50 mV s1 and pH 2 (1 M LiCl in HCl) in the absence
reported for the MOF film on an ITO glass substrate. Interest-
of bromate and upon increasing concentrations of bromate (0.2–40 mM).
ingly, no additional carbon or other conductive phase was mixed
[Reprinted with permission from ref. 102. Copyright 2009 American
Chemical Society.] with the MOF. The capacitance was retained with a loss of only
1.5% for 1000 cycles. This capacitance is already comparable to
that reported for Co3O4 nanoparticles, 200–290 F g1.107
0.2–40 mM) and the peak current at +0.17 V vs. RHE (0.2 V vs. Diaz et al. investigated a Zn-based MOF (MOF-5,
SCE), thus opening the way for its detection and reduction by Zn4O(bdc)3) partially substituted with Co (Zn3.68
this POMOF (Fig. 11).102 The onset potential is ca. +0.36 V vs. Co0.32O(bdc)3(def)0.75).106 Electrodes were prepared by mixing
RHE (+0.1 V vs. SCE at pH 2), and corresponds to the second MOF : carbon black : PTFE in 75 : 15 : 10 wt%. From cyclic
redox wave observed in the bromate-free electrolyte. The reac- voltammetry or chronoamperometry in 0.1 M tetrabutylammo-
tion product is believed to be bromide, based on previous reports nium hexafluorophosphate (tbaPF6) in acetonitrile, a gravimetric
on POMs-catalyzed bromate reaction. capacitance of only 0.5 F g1 MOF was estimated. This low
The MOF [Cu2(bpdc)2(dpq)2(H2O)]$H2O previously discussed capacitance might be due to incompatibility between that
for the nitrite reduction99 was also investigated for the electro- particular MOF and this electrolyte, since Fig. 12 clearly shows
catalytic reduction of BrO3 at pH 2. Some activity was observed huge differences in capacitance depending on the electrolyte.
below +0.41 V vs. RHE (+0.05 V vs. SCE). Better activity for
bromate reduction was observed with the other Cu-based MOF
IV.2 Sacrificial MOFs for metal or metal-oxides
previously synthesized by the same group, with an onset poten-
tial of +0.56 V vs. RHE.98 Controlled synthesis of the size and morphology of metal or
metal oxides is paramount for catalytic applications, including

IV. MOFs as electrode materials for supercapacitors


Electrochemical double layer capacitors (EDLCs) are very effi-
cient for the storage or release of electrical energy at very high
rates due to the absence of Faradaic reactions in this device.
Electrical charges are stored at the electrode–electrolyte interface
and the gravimetric capacitance, Cm, of a given porous material
is the mathematical product of the average surface specific
capacitance, Cs, and the total surface area, S. The surface specific
capacitance is however a complex function of the pore size, with,
for carbonaceous materials, a decreasing trend with decreasing
pore size down to a critical size of ca. 1.5 nm, below which value
the specific capacitance increases as the inverse of the pore
size.103,104 The latter anomalous increase occurs both in aqueous
and organic electrolytes and is due to the distortion of the
solvation shell of ions. Microporous carbons with high specific
area, short pores and pore size less than 1 nm are therefore Fig. 12 Cyclic voltammetry in various 1 M electrolytes of a Co-based
expected to be ideal materials for EDLCs. MOFs with control- MOF synthesized from cobalt nitrate and terephthalic acid. [Reprinted
lable pore size in the micropore range (0.6–2 nm) and with permission from ref. 105. Copyright 2012 Elsevier.]

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9279
View Online

electrochemical reactions. Reduced particle size means more used as a precursor.119 Thermogravimetric analysis of the MOF
reactive surface per mass of metal, but particle size effect may in inert gas showed a loss of the crystalline structure at 365  C.
also lead to changes in the surface-specific activity while crys- Calcination of the MOF was carried out at 450  C, and resulted
tallographic orientation also tunes the rate of reactions. in the pure phase Co3O4. The resulting material shows a 3D
A general strategy for the synthesis of carbon-supported metal organisation of mm size, composed of spheroidal primary parti-
or metal-oxide nanoparticles starting from MOFs was reported cles of 30–100 nm size (Fig. 13). A specific surface area of 21 m2
by Poddar’s group.108 By heat treatment of various MOFs at g1 was measured, in agreement with the primary particle size.
900  C in N2 or in air, these authors observed the formation of Cyclic voltammetry in 6 M KOH showed two redox pairs
metal or metal oxide nanoparticles embedded in a carbon matrix. ascribed to the Co3O4–CoOH and CoOH–CoO2 couples.
For Cu- and Co-based MOFs, metal nanoparticles were Gravimetric capacitances of 210 and 179 F g1 were reported at
obtained after heat treatment in N2 while metal oxide nano- scan rates of 5 and 30 mV s1, respectively. Stability over 1000
particles were obtained after calcination in air. For Zn, Mg and cycles was investigated and the capacitance increased from 206 to
Cd based MOFs, pure metal oxide nanoparticles were obtained 210 F g1 over the first 130 cycles to subsequently decline to 204
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

after heat-treatment in air or in N2. The authors found a corre- F g1 after 1000 cycles. Higher calcination temperatures (550 and
lation between the redox potential of metaln+/metal and the 650  C) lead to decreased capacitance, probably due to increased
ability of a metaln+-based MOF to form metal nanoparticles. particle size of the resulting materials. This capacity is compa-
Cu2+ and Co2+ have redox potentials of +0.34 and 0.28 V vs. rable to that obtained with state-of-the-art Co3O4 materials,
SHE, while Zn2+, Mn2+ and Mg2+ have much lower potentials of 200–290 F g1.107
Downloaded by Stanford University on 19 October 2012

0.76, 1.18 and 2.37 V vs. SHE, respectively. It was


concluded that only metal cations with a potential above a
IV.3 Sacrificial MOFs for microporous carbons
certain value could react with the organic ligands of the MOFs to
be reduced to zero oxidation state during pyrolysis in inert MOFs have also been investigated extensively as a thermally
atmosphere. Synthesis of zinc-oxide nanostructures from Zn- decomposable template for synthesizing carbon materials with
based MOFs has interested several research groups.109–114 The specific surface area up to 3447 m2 g1. The carbon materials
intrinsic shape of the starting MOF is sometimes retained after its were obtained by thermal decomposition of MOFs120–123 or
total transformation into an inorganic structure.109,114–116 Nano- MOF–organic precursor composites prepared by vapour
wires of g-MnO2 have been synthesized from a Mn–bipyridine phase124 or incipient wetness techniques.120,125–127 Furfuryl
MOF.117 The precursor MOF was boiled in an air-saturated alcohol has been the reference organic precursor124–127 while two
NaOH solution, resulting in the displacement of the ligands by Zn-based MOFs, MOF-5 and ZIF-8, have been investigated as
oxygen and MnO2 nanowire formation, with diameters of 20–40 major sacrificial MOFs with a conveniently low boiling temper-
nm grown in the [002] crystalline direction. Pure hematite or ature of Zn metal, 907  C.82,120,121,123–127
magnetite nanorods with 70 nm width and 800 nm length were Carbon materials obtained from MOF-5 (Zn4O(bdc)3) are
obtained from the pyrolysis in air, or in air and then in N2, of Fe- discussed first. The first work using a MOF to prepare nano-
MIL-88B (a 3D Fe–bdc MOF).115 In2O3 nanorods with different porous carbons124 revealed the importance of the pyrolysis
shapes were also synthesized by calcination of In-based MOFs temperature for the structural evolution into carbon. First,
synthesized with various ligands.116 Last, the perovskite structure MOF-5 was impregnated by furfuryl alcohol through vapour
GdCoO3 was obtained by calcination of a mixed metal-ion deposition. The authors reported a specific surface area of 2872
MOF.118 m2 g1 for the carbon obtained after pyrolysis at 1000  C in
Nano/micro Co3O4 materials for EDLCs were prepared by argon. On the contrary, the sample obtained at 530  C had a
Zhang et al. by calcination of a cobalt(bdc)2 MOF which was specific surface area of only 217 m2 g1 and the material con-
tained much ZnO. After removal of the latter by acid washing,
the specific surface area rose to 1732 m2 g1, showing that pores
formed independently of ZnO evaporation occurring at higher
temperature. The same group applied the incipient wetness
technique for the impregnation of furfuryl alcohol into MOF-
5.125 The specific surface areas for samples obtained at temper-
atures ranging from 530 to 1000  C, and after acid washing, fall
into the range from 1140 to 3040 m2 g1, with the highest surface
area obtained by carbonizing at 530  C. This carbon is micro–
mesoporous with an average pore size of 3.7 nm. Hu et al.
demonstrated that MOF-5 can be converted into a micro-mes-
oporous carbon without any additional organic precursor.120
MOF-5 was simply pyrolysed at 600 and then 900  C in N2,
resulting in an amorphous carbon (no diffraction peaks) having a
specific surface area of 1812 m2 g1. Other porous carbons were
prepared by pyrolysis of composites of MOF-5 impregnated or
Fig. 13 SEM images (a and b) and TEM images (c and d) of a Co3O4 covered with additional carbon precursors (phenolic resin,
material obtained by calcination of a Co(bdc)2 MOF. [Reprinted with carbon tetrachloride + ethylenediamine). The resulting carbons
permission from ref. 119. Copyright 2012 ESG.] were subsequently activated by KOH at 500  C to further control

9280 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

the microstructure.120 KOH activation resulted in a negative


effect on the surface area when phenolic resin was used (decrease
from 1543 to 1271 m2 g1), while the contrary was obtained when
the mixture carbon tetrachloride + ethylenediamine was used
(increase from 348 m2 g1 to 2222 m2 g1). Yang et al. synthe-
sized two Zn-based MOFs with two different bdc-linkers: ami-
nobenzene dicarboxylate and naphthalene dicarboxylate
(labelled IRMOF-3 and IRMOF-8, respectively, in their work).
After pyrolysis at 900  C in N2, the carbon materials showed
specific surface areas of 1678 and 1978 m2 g1, respectively.121
We now discuss the second most employed MOF as a ther-
mally decomposable template, namely ZIF-8 (Zn(N2C4H6)2).
After impregnation with furfuryl alcohol and pyrolysis in Ar at
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

1000  C, a carbon material with a high surface area of 3405 m2


g1 was reported.126 The pyrolysis temperature was important
here as well, since a pyrolysis at 800  C resulted in a lower surface Fig. 14 Comparison of the galvanostatic charge/discharge curves at the
current density of 0.1 A g1 for five carbon samples obtained after
area of 2169 m2 g1. Using furfuryl-alcohol-impregnated ZIF-8,
pyrolysis at different temperatures of a MOF-5–furfuryl alcohol
another group reported however a surface area (mostly micro-
Downloaded by Stanford University on 19 October 2012

composite. [Reprinted with permission from ref. 125. Copyright 2010


porous) of only 933–1130 m2 g1 after pyrolysis at 900–1100  C Elsevier.]
in Ar.127 After activation by KOH at 700  C, areas of 1425–3188
m2 g1 were reached, with the highest area obtained by subse-
quent KOH activation of the sample obtained at 900  C. Last, (measured under galvanostatic control at 0.25 A g1), the highest
pyrolysis of ZIF-8 at 700–1000  C in N2 yielded carbon materials capacitance being obtained with the 650  C sample (195 F g1 for
with areas of 520–1110 m2 g1;123 confirming the results of 1521 m2 g1) while the sample with a higher area (2524 m2 g1,
Almasoudi and Mokaya.127 obtained at 1000  C) showed a capacitance of only 129 F g1
Park and co-workers synthesized a third Zn-based MOF with (Fig. 14). In acidic medium, capacitances of 200 F g1 at 0.25 A
terephthalic acid, the simplest bdc-type linker.121 Pyrolysis of the g1 were obtained with two samples having surface areas of 2169
resulting MOF (labelled IRMOF-1 in ref. 121) at 900  C in N2 and 3405 m2 g1.126 These samples were obtained from the
yielded a carbon material with an area of 3174 m2 g1, the highest pyrolysis of ZIF-8–furfuryl alcohol composites. Due to the high
in that study. This demonstrates the impact, for a given metal, of N-content of ZIF-8, these carbons certainly contain nitrogen
the functionalization of the ligand used to build the MOF since heteroatoms, however the N-content and possible role in the
the two functionalized bdc-type ligands (aminobenzene and capacitance was not measured or discussed. Last, from cyclic
naphthalene dicarboxylate) resulted in carbon materials with voltammetry at 50 mV s1, a capacitance of 110–130 F g1 was
specific surface areas of only 1750–2350 m2 g1.121 measured for samples obtained from the direct pyrolysis of ZIF-
Apart from Zn-based MOFs, only a few MOFs based on other 8.123 No loss of capacitance was observed after 250 cycles.
metal-cations have been investigated as sacrificial MOFs for the The results obtained with MOF-derived carbons compare well
production of carbon materials. A cobalt-based imidazolate with typical values of gravimetric capacitance obtained with
MOF was synthesized and heat treated at 750  C, resulting in a activated carbons, which are 100–120 F g1 in organic electro-
carbon material with a surface area of 264 m2 g1 embedding lytes and 150–300 F g1 in aqueous electrolytes.129 However, no
Co–Nx active sites for the ORR.87 Pyrolysis of iron-based or clear advantage is tangible and activated carbons may be cheaper
nickel-based MOFs seems to result either in more graphitic and easier to mass-produce compared to MOF-derived carbons.
carbon with low surface area, or in carbon nanotubes.122,128 The
pyrolysis of [Ni3(btc)2$12H2O] at 500  C was shown to result in
multiwall carbon nanotubes.122 In this case, the Ni-based MOF V. MOFs for battery or fuel-cell electrolytes
has a dual purpose: it serves as a carbon source but also provides
V.1 MOFs for lithium-battery electrolyte
the catalyst (Ni) required for the nanotube growth.
The charge-storage properties for EDLCs of the resulting Currently, macroporous polymer membranes swelled with
carbon materials are now discussed. These materials showed lithium salts dissolved in organic carbonates are utilized as a
excellent performance and stability both in aqueous (6 M KOH separator in Li-ion batteries.130 The liquid electrolyte restricts the
or 1 M H2SO4) and organic electrolytes (1.5 M NEt4BF4 in battery shape while the lack of rigidity for the current separator
acetonitrile).120,123–126 The largest capacitances reported by Hu precludes the use of solid lithium as an anode material because of
et al. in basic and organic electrolytes from galvanostatic charge– the possible formation of lithium dendrites upon repeated
discharge curves at 0.25 A g1 were 274 and 168 F g1, respec- cycling. An alternate solution to avoid those features would be
tively. They were obtained with a KOH-activated MOF-derived the use of rigid solid separators.131
carbon, which was the sample with the highest area in that study The use of MOFs as a solid separator was recently investi-
(2222 m2 g1). In acidic medium, samples obtained after different gated.132 The evaluation started by investigating three different
pyrolysis temperatures (from 530 to 1000  C) showed different MOF structures: MOF-177 (a benzenetribenzoate bridged
EDLC behaviour.125 However, there was no direct correlation framework with ZnII sites),133 Cu–BTTri (a triazolate-bridged
between the surface area of the samples and their capacitance framework with CuII sites),134 and Mg2(dobdc).135 The ionic

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9281
View Online

Interfacial contact resistances (ca. 25–50 mU cm2) always add to


the pure membrane resistance, making the search for even higher
membrane conductivities of little significance. However, more
robust or cheaper materials with similar conductivity are sought
after. In contrast, novel proton-conducting membranes or
materials with higher conductivity than the current status at
temperatures above 100  C (anhydrous proton conduction) are
required. PEMFCs operating in the range 120–180  C would
present several technical advantages: (a) better tolerance to CO
and other contaminants in the H2 fuel, (b) reasonable perfor-
mance with methanol or ethanol as fuels due to improved
kinetics and CO-tolerance, and (c) easier water and heat
Fig. 15 Structure of Mg2(dobdc) and its modification by a Li-alkoxide management.
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

to form the solid electrolyte. [Reprinted with permission from ref. 132. Since 2003, proton-conducting materials based on MOFs have
Copyright 2011 American Chemical Society.] been investigated, both in hydrated and dehydrated condi-
tions.140 The different approaches investigated so far can be
classified into three types: (i) as-synthesized MOFs,141–146 (ii)
conductivities were measured after uptake of a common elec- MOFs loaded with guest molecules or incorporating inorganic
Downloaded by Stanford University on 19 October 2012

trolyte solution (1 M LiBF4 in a 1 : 1 mixture of ethylene nano-structures,147–151 and (iii) MOFs that are chemically modi-
carbonate (EC) and diethyl carbonate (DEC)) by the MOFs and fied in a post-synthesis step.152
pressing of the resulting powder into pellets. The conductivity In approach (i), as-synthesized MOFs, the vast family of MOF
values ranged from 107 to 104 S m1. The Mg2(dobdc) was crystalline structures can provide size- and shape-tunable pores
found to be the most promising MOF with room-temperature for proton conduction, with controllable hydrophilicity and
conductivity of 1.8  104 S m1. The Mg2(dobdc) structure acidity. Carboxylic, sulfonic or phosphonic acid functionalities
consists of 1D hexagonal channels with diameters around 14 A 
can be introduced in the ligand chemistry of the framework. In
that are lined with Mg2+ cations (Fig. 15). This conductivity is this approach, proton conduction is observed only in a hydrated
still three orders of magnitude below the level desired for a environment (Table 1). Kitagawa’s group controlled the hydro-
battery electrolyte (101 S m1).136 philicity of MOFs to achieve high proton conductivity in oxalate-
Further, the MOF Mg2(dobdc) was functionalized with a Li- bridged layered MOFs, {NR3(CH2COOH)}[MCr(ox)3]$nH2O
alkoxide (i.e. Li isopropoxide, LiOPr), following the assumption (R ¼ methyl, ethyl or n-butyl, and M ¼ Mn or Fe.143 They are
that alkoxide anions preferentially bind Mg2+ while leaving Li+ labelled R–MCr. The carboxylic groups provide proton
free to move. The functionalized MOF (Fig. 15) was soaked in conductivity while the hydrophilicity of the cations
the same electrolyte as before: 1 M LiBF4 in a 1 : 1 mixture of EC {NR3(CH2COOH)}+ is tuned by the NR3 residue. The most
and DEC. The resulting compound formula was found to be hydrophilic sample, methyl–FeCr, showed the highest proton
Mg2(dobdc)$0.35LiOPr$0.25LiBF4$EC$DEC. At this stage, conductivity in that study, 102 S m1, at ambient temperature
XRD analysis showed that the initial MOF structure was and 65% RH. This MOF shows a 2D layer structure with an
conserved. The conductivity (3.1  102 S m1) was more than  enough to host two water
interlayer periodic distance of 6.9 A,
two orders of magnitude that of the Mg2(dobdc)$ molecules. Previously, the same group reported a proton
0.05LiBF4$EC$DEC material obtained after soaking conductivity of 7  104 S m1 at 70  C and 95% RH for the
Mg2(dobdc) in the same electrolyte (1.8  104 S m1). This total MOF Fe(OH)(bdc–(COOH)2.142 It was the highest conductivity
increase in conductivity was partly due to increased lithium obtained in that work, which investigated several MOFs of the
electrolyte concentration in the MOF (factor ca. 7) and increased type M(OH)(bdc–R) with M ¼ Al, Fe; R ¼ H, NH2, OH, or
Li+ mobility (factor ca. 25). These results are very promising and (COOH)2. Increased proton conductivity correlated well over
represent a new approach for creating solid lithium electrolyte two orders of magnitude with a decreased pKa-value of the
materials. However, the mechanism for improved Li mobility functional group, R. The highest conductivity reported by that
remains unclear. group (Table 1) was however obtained with another MOF, a
ferrous oxalate dihydrate MOF.141 It shows a 1D structure with
two water molecules coordinated to every ferrous cation,
V.2 MOFs for proton-conductive membranes
resulting in a 1D ordered array of water molecules. Sahoo et al.
Currently, state-of-the-art materials for low and high-tempera- investigated the role of halogens (Br or Cl) on the proton
ture PEMFCs are sulfonated fluoro-polymers (e.g. Nafion) and conductivity of chiral Zn-based MOFs.145 Four MOFs were
phosphoric-acid doped polybenzimidazole membranes, respec- synthesized, being structural isomers (Cl or Br as the anion
tively.137–139 Proton conductivities reached with such membrane coordinating the metal atoms) or enantiomers with respect to the
materials are typically 10 S m1 for Nafion at 80  C and 100% ligand backbone (d or l amino-acid derivative). Only the two
Relative Humidity (RH), and 5–20 S m1 at 150  C and 0% RH MOFs involving Cl-atoms showed proton conductivity
for high-temperature membranes. This results in only 25 mV (Table 1). The role of halogen atoms with different electro-
ohmic loss across the membrane at 1 A cm2 for a 25 mm thick negativities was thus evidenced. The stronger OH/Cl–Zn
Nafion membrane such as Nafion 211 or 212. Such high interaction resulted in increased acidity, which is a major driving
conductivities are necessary for high power density PEMFCs. force for proton conduction. The group of Prof. Shimizu

9282 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

Table 1 Proton conductivities of various metal organic frameworks

Proton conductivity
(S m1) T ( C) RH (%) MOF structure and (when applicable) loaded molecule Ref.

8  103 25 65 Methyl–FeCr$3H2O 149


7  104 70 95 Fe(OH)(bdc–(COOH)2) 142
4  103 30 98 [Zn(l-LCl)(Cl)](H2O)2 145
4  103 30 98 [Zn(d-LCl)(Cl)](H2O)2
1.3  101 25 98 Ferrous oxalate dihydrate 141
3  103 25 98 Zn3(L)(H2O)2$2H2O; L ¼ [1,3,5-benzenetriphosphonate]6 144
3  102 150 0 Na3(2,4,6-trihydroxy-1,3,5-benzene-trisulfonate) loaded with 1,2,4-triazole 147
2.2  103 120 0 [Al(m2-OH)(1,4-ndc)]n loaded with imidazole 150
0.8 25 98 (NH4)2[Zn2(ox)3]$3H2O loaded with adipic acid 149
1 200 0 Cr-MIL-101 loaded with CsHSO4 148
0.3 20 100 Sulphated MIL-53(Al) 152
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

reported a proton conductivity of ca. 3  103 S m1 at 25  C H2/air fuel cell, which is close to the theoretical voltage and
and 98% RH for a Zn-based phosphonate MOF.144 Metal- demonstrates the impermeability to H2. The proton mobility in
phosphonate MOFs with a coordinatively unsaturated metal and N-heterocycles such as imidazole or triazole and the electro-
Downloaded by Stanford University on 19 October 2012

ligand are interesting for proton conduction since uncoordinated chemical stability of the latter were previously shown.153 This is
oxygen atoms from phosphonates may act as H-bond acceptors an interesting approach for high-temperature proton conducting
while unsaturated metal cations impart water retention to the materials but conductivities are still quite low compared to
MOF. The Zn-based phosphonate MOF investigated in ref. 144 current membrane materials operating at 150  C.
shows a layered structure with 11.5 A interlayer distance. In a related work, aluminium-based porous frameworks were
In approach (ii), MOFs loaded with amphoteric guest mole- reported to conduct protons above 100  C through the incor-
cules (imidazole or triazole) or incorporating inorganic nano- poration of imidazole (Im) within the pores.150 The study of two
structures (polyoxometalates) have been investigated for proton compounds, [Al(m2-OH)(1,4-ndc)]n and [Al(m2-OH)(bdc)]n,
conduction under poorly hydrated or fully dehydrated condi- showed that proton conductivity varies because of the different
tions. Shimizu and co-workers synthesized a MOF based on pore shapes of those structures and of the different host–guest
sodium cations and benzene–trisulfonate ligand that was subse- interactions. The highest conductivity of 2.2  103 S m1 was
quently loaded with the guest molecule 1,2,4-triazole (Fig. 16).147 obtained at 120  C with [Al(m2-OH)(1,4-ndc)]n–Im.
While the as-synthesized MOF labelled b-PCMOF2 showed A highly proton-conductive MOF was obtained by intro-
poor proton conduction (3–5 105 S m1 at room temperature ducing adipic acid in a layered oxalate-bridged framework
and 0% RH, and two orders of magnitude lower conductivity at (NH4)2[Zn2(ox)3]$3H2O.149 The conductivity reached 0.8 S m1
80  C), the triazole (Tz) loaded MOF (b-PCMOF2(Tz)x) showed at 98% RH at 25  C (Table 1). The carboxylic acidic groups of
similar poor proton conductivity at room temperature but a adipic acid are situated in the interlayer space of the framework
much increased conductivity with increasing temperature. At as well as in the voids of the layer. The water content was shown
80  C, the conductivity was ca. 3  103 S m1 and, at 150  C, to be crucial for high proton conductivity, since the dihydrate
3  102 S m1 (Table 1).147 This demonstrates the importance of material obtained at 70% RH showed a proton conductivity of
triazole in the anhydrous proton-conduction mechanism. In only 6  104 S m1.
addition, the authors prepared a membrane-electrode-assembly Ponomareva et al. introduced the acid salt CsHSO4 into a Cr-
by pressing b-PCMOF2(Tz)0.45 in-between two Pt/C electrodes. based MOF devoid of proton-conductive properties.148 Solid
An open circuit voltage of 1.18 V was measured at 100  C for a acids of the MmHn(XO4)p family (M ¼ Li, Na, K, Rb, Cs, NH4;
X ¼ S, Se, P, As; n, m and p between 1 and 5) undergo a
superprotonic phase transition at 50–230  C and are interesting
as anhydrous proton conductors at high temperature with
proton conductivities in the range of 0.1–1 S m1. The chosen
MOF was Cr3O(H2O)Y(bdc)0.15HNO3$13H2O, where Y ¼ F or
NO3. This MOF is also known under the label Cr-MIL-101. The
results of that study revealed proton conductivity of 105 to 1 S
m1 at temperatures of 50–200  C for these nanocomposites with
the formula (1  x)CsHSO4-(x)Cr-MIL-101, and chemical
stability up to 190  C (Fig. 17). As the amount of MOF in the
composite was increased from 0.01 to 0.07, the phase transition
temperature was decreased by 10–14  C and, more importantly,
the low temperature proton conductivity was increased by up to
Fig. 16 The structure of the metal organic framework b-PCMOF2
2 orders of magnitude. However, the high temperature proton
including guest triazole molecules and the trisulfonated benzene-deriva- conductivity decreased at the same time for x > 0.02. The best
tive ligand. [Reprinted by permission from ref. 154. Macmillan Publishers compromise was obtained for x ¼ 0.02, with 2 orders of
Ltd: Nat. Chem., copyright 2009.] magnitude higher conductivity relative to CsHSO4 below the

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9283
View Online

the metallic surface needs to be protected in order to avoid severe


corrosion and metal loss.155 The inhibitor is dispersed in the
pickling solution and reduces the metal corrosion by a blocking
and/or electrocatalytic action. In an adsorption–blocking
method, the adsorption strength of the inhibitor on the metal
surface is important. It depends on physico-chemical properties,
functional groups, steric factors, molecular size and structure,
presence of electron donor atoms and p-orbital character of
donating electrons. Organic molecules with heteroatoms such as
N, P or S are well-known inhibitors for many metals and alloys in
acidic media. Coordination complexes have also been more
recently investigated as inhibitors, and possess different cyclic
and heterocyclic ligands. MOFs, that may combine metal–ligand
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

complexes and a continuous 2D or 3D structure, topologies rich


in p-systems and hetero-aromatic moieties, are potential mate-
rials as corrosion inhibitors. A few studies have investigated
this effect.156–158
The inhibitor role of the MOF (AgCN)4$(qox)2 for carbon
Downloaded by Stanford University on 19 October 2012

steel in 1 M HCl was investigated.156 This MOF was chosen for


its ease of synthesis and presence of two aromatic heterocycles in
its unit cell. Its topology is a 2D sheet with deformed six-
Fig. 17 Schematic structure of Cr3O(H2O)Y(bdc)yHNO3$nH2O: Cr3O membered rings and these sheets are close-packed in parallel via
clusters are connected to each other through terephthalic linkers; the H-bonds and p–p stacking (Fig.18a). The nitrogen heteroatoms
supertetrahedra formed are joined by vertexes and construct a zeotypic are important for strong adsorption of the MOF to a metal
structure with two types of mesocages: small (yellow) and large (blue). surface. Corrosion potential and corrosion current densities
Temperature dependencies of proton conductivity for (1  x)CsHSO4x were measured for various concentrations of the MOF (from 0 to
[Cr3O(H2O)Y(bdc)yHNO3$nH2O] (x ¼ 0.01–0.07) in comparison with 104 M) in 1 M HCl. The working electrode was a polished, flat
CsHSO4. [Reprinted with permission from ref. 148. Copyright 2012 surface of carbon steel. With addition of the MOF in 1 M HCl
Elsevier.]
solution, a suppression of both anodic and cathodic branches of
transition temperature and equal conductivity to CsHSO4 above the polarization curves was reported, but was more pronounced
the transition temperature. The large energy of adhesion of for the anodic branch (Fig. 18b). Clearly, the MOF adsorption
CsHSO4 to the MOF surface and disordered state, evidenced by
XRD, of the acid salt in the nanopores resulted in the formation
of a highly-conductive CsHSO4 phase at low temperature.
Last, hybrid materials comprising a MOF and Strandberg-
type polyoxometalate were investigated.151 Strandberg-type
polyoxometalates are heteropoly anions in which organic groups
are covalently bonded to a phosphorus heteroatom. Their hybrid
material [Mo5P2O23] [Cu(phen)(H2O)]3$5H2O(Cu3Mo5P2)
showed a proton conductivity of 2  103 S m1 at 25  C and
98% RH.
In approach (iii), Goesten et al. functionalized the MOF MIL-
53(Al) (aluminium terephthalate MOF) in a post-synthesis step
by reacting it at room temperature in a solution of H2SO4–Tf2O
in nitromethane (Tf2O ¼ trifluoromethanesulfonic anhydride).152
Sulfoxy-acid moieties covalently bonded to the aryl carbons of
the organic linker were generated for approximately 50% of
the terephthalate linking units. A proton conductivity of ca.
0.3 S m1 at 20  C and 100% RH was observed for the sulphated
MIL-53(Al) material. The proton conductivity decreased with
Fig. 18 (a) 2D-fused deformed six-membered rings in (AgCN)4$(qox)2,
increasing temperature, but that mostly reflected the decrease in (b) potentiodynamic polarization curves for corrosion of C-steel in the
RH since the atmosphere was N2 saturated in water at room absence and presence of different concentrations of the MOF
temperature. (AgCN)4$(qox)2 at 25  C. [Reprinted with permission from ref. 156.
Copyright 2011 Elsevier.] (c) Extended structure of Ag(qox)(4-ab)
showing the four parallel molecules surrounding the fifth one, (d)
VI. MOFs for corrosion inhibition potentiodynamic polarization curves for corrosion of C-steel in the
absence and presence of different concentrations of the inhibitor
Corrosion inhibitors are required in industrial steps of chemical Ag(qox)(4-ab) at 25  C. [Reprinted with permission from ref. 158.
cleaning and pickling necessary to remove the oxide layer when Copyright 2011 Springer.]

9284 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

on the metal surface retarded both the H2 evolution reaction and blank solution to 17, 14 mV dec1 and 32, 32 mV dec1
the metal dissolution. The corrosion current density was sup- respectively for the two MOFs at 103 M concentration. The
pressed by almost one order of magnitude from the blank to 104 simultaneous decrease of both Tafel slopes in the presence of
M MOF in solution (4.13 to 0.44 mA cm2), and the anodic and increased concentration of either MOF suggests that the inhibi-
cathodic Tafel slopes changed from 211, 224 mV dec1 to 109, tion acts on both processes, but with a dominant anodic inhibi-
129 mV dec1, respectively, for the blank and 104 M MOF in tory effect (metal dissolution). These inhibition efficiencies are
acidic solution. The parallel change of the anodic and cathodic comparable to those previously reported for 4-amino piperidine
Tafel slopes with increased MOF concentration supports a and piperazine free ligands which were 63% and 62% in
blocking action of the MOF and adsorption of the MOF on the perchloric acid.159 The important input of those works results
metal surface. The adsorption energy was estimated to be ca. from the opening of new possibilities for AgI complexes to be
30 kJ mol1 and ascribed to mixed physi- and chemisorption. used as corrosion inhibitors and more specifically as microbial
The inhibition efficiency (% of corrosion current reduction vs. the corrosion inhibitors since the anti-microbial effect of Ag is
blank) at 105 M MOF was 67% at 25  C but decreased with well known.
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

increasing temperature (42% at 55  C). From the adsorption


process investigated by electrochemical impedance spectroscopy
VII. Discussion and conclusions
(EIS), the charge transfer was found to increase with increased
concentration of MOF while the double layer capacitance The crystalline but porous nature of MOFs combined with their
decreased with increased MOF concentration. The latter vast diversity of chemical functionalization and metal-cation
Downloaded by Stanford University on 19 October 2012

suggests that the MOF is strongly adsorbed and changes the coordination presents a huge potential for catalysis and for
electronic structure of the steel surface. The authors concluded electrocatalysis in particular. Such materials are an excellent
that the MOF network seems to be oriented parallel to the platform to mimic the complex structure and high catalytic
surface via adsorption through non-binding electron doublets activity of active sites found in enzymes, based on abundant
from N atoms. transition metals. Simultaneously, MOFs may show the stability
The same authors performed a similar study using another of inorganic nano- or microparticles that are today the working
MOF, Ag(qox)(4-ab).158 The unit cell of this MOF comprises a principle of industrial electrocatalysts or active electrode mate-
silver atom coordinated to one N-atom of the qox ligand, two rials. In contrast to pore-free nano- or microparticles where only
O-atoms of the two carboxylate groups and a second silver atom. surface metal atoms in contact with the electrolyte can be elec-
The extended structure comprises infinite discrete building trochemically active for surface reactions, the metal-ion utiliza-
blocks, which are arranged in a unique way so that four parallel tion in electrochemically active MOFs can theoretically reach
molecules surround the fifth one (Fig. 18c). H-bonds and p–p 100% when the framework provides sufficient electron and ionic
stacking ensure 3D network cohesion. Electrochemical experi- conductivity. In addition, for electrode materials for batteries,
ments were performed as above156 (Fig. 18d), and the corrosion the use of MOFs would change the mechanism of charge/
current density was suppressed by 84% from the blank to 104 M discharge from a slow insertion mechanism (governed by solid
MOF in solution (4.13 to 0.65 mA cm2). The inhibition effi- phase diffusion and, often, complex phase changes) to a fast
ciency at 105 M MOF decreased with increasing temperature surface reaction mechanism similar to that found in EDLCs.
(from 97% to 56% as temperature increased from 25 to 55  C). As Ionic conductivity can be imparted either by the MOF itself or by
before, impedance spectroscopy showed increased charge trans- the electrolyte or guest molecules introduced in the pores of the
fer resistance and decreased double layer capacitance in the framework. While values of ionic conductivity for Li-ion
presence of the MOF in solution, suggesting that adsorption is batteries and PEM fuel cells have been reported and discussed in
the corrosion inhibition process. Sections V.1 and V.2, a mathematical analysis is necessary to
Massoud et al. investigated the use of two silver pyrazine understand the implications of such numbers.
MOFs as corrosion inhibitors for mild steel in acidic medium.157 While for a battery separator or fuel cell membrane, the
These two MOFs have the composition [Ag2(ampyz)(NO3)2]n voltage drop across the separator or membrane can be easily
and [Ag(2,3-pyzdic)(NO3)]n. [Ag2(ampyz)(NO3)2]n has a 2D calculated from the material’s conductivity, the current density
sheet structure via NO3 bridging and pyrazino bridging and 3D and the separator thickness, such a calculation is not straight-
structure through extensive hydrogen bonding between the forward for MOFs inserted in porous electrodes as electro-active
amino function of the ampyz and the uncoordinated O from materials. The requirements on MOF-conductivity (electronic or
NO3. [Ag(2,3-pyzdic)(NO3)]n has a 1D chain structure and ionic) are lower in porous electrodes than in a membrane/sepa-
strong hydrogen bonds further connect these chains to form a 3D rator since electro-catalytically active MOFs can be mixed with
network. The AgI ion is linearly coordinated to two pyzdic other phases presenting a much higher electron and ionic
molecules. In contrast to [Ag2(ampyz)(NO3)2]n, this 3D network conductivity. Charge transport within the MOF then only needs
does not reveal any easily recognizable 3D-net. Corrosion to take place at the level of the MOF particle size, d, which is
potential and corrosion current densities were measured for often a much smaller distance than that of the porous electrode
various concentrations of the two MOFs (from 0 to 103 M) in thickness, L. Fig. 19 is a scheme of an idealized porous electrode
0.1 M nitric acid. The working electrode was a polished, flat with a volume split in 1/3 of a MOF active material and 2/3 of a
surface of mild steel. For [Ag2(ampyz)(NO3)2]n and [Ag(2,3- mixed phase with high electron and ionic conductivities.
pyzdic)(NO3)]n, the inhibition efficiency was 64% and 50% On the basis of this scheme, one may calculate Ohmic losses
respectively at 103 M MOF-concentration, and the anodic and due to electron or ion transport from the interface between the
cathodic Tafel slopes changed from 78, 78 mV dec1 for the MOF and the mixed phase and into the core of the MOF phase.

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9285
View Online

Table 2 Ohmic losses within the MOF phase for the ideal porous elec-
trode schemed in Fig. 19

Iv (A m3) d (nm) s (S m1) DE (mV)

‘‘Battery electrode’’ 106 200 105 1.5


106 15
1000 104 3.75
106 375
‘‘PEM fuel cell cathode 108
200 104 15
with non-precious 105 150
metal catalyst’’ 1000 102 3.75
104 375
Fig. 19 Scheme of a porous electrode comprising a MOF active material ‘‘PEM fuel cell cathode 109 200 103 15
with Pt-like activity’’ 104 150
(1/3rd of volume) and a mixed electron and ion conductive phase (2/3rd of
1000 102 37.5
volume). In this example, a reduction reaction takes place.
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

103 375

Two values of electrode current densities that are typical for


batteries and fuel cells are considered, i.e. 10 mA cm2 and 1 A complex function of x, with higher currents close to the MOF-
cm2, respectively. Assuming an electrode thickness L of 100 mm mixed phase interface, and smaller currents at the core of the
Downloaded by Stanford University on 19 October 2012

(discussed later), this gives average volumetric currents in the MOF phase. A more complex mathematical treatment would
porous electrode, Iv, of 106 and 108 A m3, respectively. Since then be needed to find the exact solution. However, our present
the electrochemical reaction would take place in the MOF only goal is only to find the order of magnitude required for
(the mixed phase is assumed to be electrochemically inactive), the conductivities in order to avoid unacceptable Ohmic losses
average volumetric current is thus, in our example, 3  Iv since within the MOF phase.
only 1/3rd of the electrode volume is occupied by the MOF. For a battery application, it is seen that a conductivity as low
From this, one can apply Ohm’s law at position x in the MOF as 106 S m1 is sufficient if the MOF particles are 200 nm in size
phase (Fig. 19, right), eqn (10): or less. For larger particles (1 mm), a conductivity of at least 105
S m1 is preferable. In Section V.1, the ionic conductivities
dE ¼ rI(x)dx (10) reported for MOFs impregnated with typical organic Li+-con-
taining electrolytes ranged from 107 to 2  104 S m1,133–135
where E is the electric potential at position x (either the potential
and even reached 3  102 S m1 for one material.135 Therefore,
of the electronic or ionic phase), r the resistivity (U m) and I the
it seems that most MOFs designed with that purpose would fulfil
electron or ion current density at position x (A m2). Assuming
the requirement on ionic conductivity for a battery application.
that the local volumetric current in the MOF-phase is
This threshold of ionic conductivity must however also be met by
uniform, the current density I(x) will be a linear function of x,
MOFs designed in first instance as an active electrode material
with I(x ¼ 0) ¼ 0 (by plane symmetry) and I(x ¼ d/2) ¼ 3  Iv 
for battery, and not as an ion-conductive MOF. Both properties
d/2. The expression of I(x) is then (eqn (11)):
can certainly be met simultaneously, as is de facto shown by the
interesting experimental results obtained by MOFs in battery
I(x) ¼ 3Ivx (11)
electrodes.30,31,57 The measured electron conductivities of MOFs
Replacing I(x) by its expression in eqn (10), and integrating are discussed later.
between x ¼ 0 and x ¼ d/2, one obtains the Ohmic loss (eqn (12)): For PEM fuel cell application, two cases are considered. The
first one assumes a 100 mm thick cathode for a MOF-based non-
DE ¼ 3/8rIvd2 ¼ 3/8(1/r)Ivd2 (12) precious metal catalyst (NPMC) with a volumetric activity
similar to that obtained with a state-of-the-art NPMC.82 The
where s is the ionic or electron conductivity of the MOF (S m1) second one assumes a MOF-based NPMC with Pt-like activity,
and DE the ionic or electronic potential drop from x ¼ d/2 to x ¼ for which an electrode current density of 1 A cm2 with a 10 mm
0. Table 2 presents values of DE for several combinations of thick electrode would become possible.77 In the first case, the
volumetric current Iv, MOF particle size d, and conductivity s. If volumetric current in the active material is smaller and hence
both the ionic and electronic potentials would show a loss, the requirement on the MOF conductivity is lower. In the first case
total practical loss of the electric potential will be the sum of both and with small MOF particles (200 nm in size), a conductivity of
losses since an electrode potential is defined by the electric 104 S m1 is sufficient while for larger particles a conductivity of
potential in the electronic phase minus that in the electrolytic at least 2  103 S m1 is necessary. For the second case,
phase. required conductivities are higher by one order of magnitude, i.e.
The calculated values of DE are in the range of 1.5–375 mV. Of 103 and 2  102 S m1 for small and large MOF particles,
course, when the DE-value is in the order of a Tafel slope respectively. Proton conductivities experimentally measured at
(typically 60 mV loss in electric potential either in the ionic or high relative humidity for most MOFs designed for that purpose
electronic phase results in a one order of magnitude smaller are in the range 7  104 to 8  103 S m1 (Table 1), with three
reaction rate), the volumetric current will no longer be uniform exceptions showing 0.13, 0.3 and 0.8 S m1.141,149,152 Other MOFs
within the MOF active material, as was assumed to perform this conducting protons at high temperature and low relative
simple calculation. The volumetric current would become a humidity show conductivities of 2  103 to 3  102 S m1

9286 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

(Table 1) with the exception of the CsHSO4-loaded MOF with a or cost compared to other materials. The very large spectrum of
conductivity of 1 S m1.148 As for batteries, if MOFs designed for ionic and electronic conductivities shown by MOFs makes the
a fuel cell electrocatalytic process can also combine the proton analysis of their electrochemical properties extremely compli-
conduction reported for MOFs designed for the latter purpose, cated. A MOF with intrinsically good electrocatalytic properties
then it seems that the requirements on ionic conductivity can be may show very poor performance if either its ionic or electronic
met. Perhaps, the most flexible approach to provide a MOF conductivity is poor. For real breakthroughs, MOFs will prob-
electrocatalyst with proton conductivity would be a modification ably need to be designed and optimized with respect to all these
approach, so that the MOF can be designed in first instance for criteria at once.
specific electrocatalytic properties, then functionalized for
proton conduction by a post-synthesis treatment.152 List of acronyms for the ligands, by order of
The topic of electron conductivity of MOFs has not been appearance
discussed so far. Kitagawa et al. reported an electron conduc-
tivity of 106 S m1 for a MOF that was also a proton conductor
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

(104 S m1).94 The conductivity of MOFs has been recently bdc 1,4-Benzenedicarboxylate
reviewed by Givaja et al.160 Electron-conductivities of MOFs btb 1,3,5-Benzenetribenzoate
measured at room temperature and 1 bar span from 1012 to 104 ndc 2,6-Naphthalene-dicarboxylate
S m1. Cu- or Ag-based MOFs were shown to switch from 1011 dmf N,N0 -Dimethyl-formamide
to 109 S m1 with cyanopyridine as a ligand to 7  104 to 4  ptcda Perylene-3,4,9,10-tetracarboxylic dianhydride
Downloaded by Stanford University on 19 October 2012

103 S m1 with pyridine-3,4-dicarbonitrile as a ligand.161 btc Benzene-1,3,5-tricarboxylate


General guiding lines for electron conductivity in MOFs are the tba Tetrabutylammonium
mixed-valence oxidation state of metal ions and metal-to-metal bpdc para-Biphenyldicarboxylate
direct interaction.160,162 In summary, many MOFs show electron dcppy 2-Phenylpyridine-5,40 -dicarboxylic acid
conductivity in the range of 102 to 101 S m1, while conduc- dcbpy 2,20 -Bipyridine-5,50 -dicarboxylic acid
tivities at or above 1 S m1 are rarely encountered.160 Such MeIm 2-Methylimidazolate
conductivities are predicted to be acceptable for local transport bipy 2,20 -Bipyridine
within MOF particles of 200–1000 nm in size, even for high dtoa N,N0 -Dithiooxamidato disubsituted
power density applications such as batteries or PEM fuel cells dpq Dipyrido[3,2-d:20 ,30 -f]quinoxaline
(Table 2). However, not all MOFs show conductivities of at least pci Pyridine-4-carboxylic ion
102 S m1 and some are truly insulators. Therefore, the design bpy 4,40 -Bipyridine
of MOF materials combining electro-catalytic activity for a ox Oxalate
specific reaction and the associated necessary electron conduc- def N,N0 -Diethylformamide
tivity will be important in the future. dobdc 1,4-Dioxido-2,5-benzenedicarboxylate
In conclusion, the analysis of the conductivity requirements 1,4-ndc 1,4-Naphthalenedicarboxylate
tells us that optimized MOFs are suitable as electrode materials qox Quinoxaline
in electrochemical applications, including high power density 4-ab 4-Amino-benzoic acid
applications, if they are seconded by other materials for the long ampyz Aminopyrazine
range transport (1–100 mm) of ions and electrons across porous 2,3-pyzdic 2,3-Pyrazinedicarboxamide
electrodes. Optimized MOFs could therefore be the ideal mate-
rial for electrochemists, with uniformly dispersed active sites with
a well-defined structure, 100% utilisation of active sites, and
mixed electron and ionic conduction, a combination of require- Acknowledgements
ments that is specific to electrochemical reactions. Successful
examples of MOF materials in batteries and EDLCs have This work was supported by the Agence Nationale de la
already been reported.30,31,57,105 Recherche under contract ANR 2011 CHEX 004 01.
MOFs have also been used as a sacrificial template for the
synthesis of highly microporous carbons for EDLCs or cathode Notes and references
catalysts for PEM fuel cells.82,86,88,120,123,125 Currently, the highest 1 O. M. Yaghi and H. L. Li, J. Am. Chem. Soc., 1995, 117, 10401–
performance reported for a NPMC at the cathode of a PEM fuel 10402.
cell involves a Zn-based MOF.82 MOFs have also been used as a 2 O. M. Yaghi, G. M. Li and H. L. Li, Nature, 1995, 378, 703–706.
3 B. F. Hoskins and R. Robson, J. Am. Chem. Soc., 1989, 111, 5962–
sacrificial template for the synthesis of nano- or micrometric
5964.
metal oxides for EDLCs119 for rechargeable Li-battery 4 B. F. Hoskins and R. Robson, J. Am. Chem. Soc., 1990, 112, 1546–
anodes.50,51 1554.
The investigation of MOFs for electrochemical applications is 5 B. F. Abrahams, B. F. Hoskins, D. M. Michail and R. Robson,
Nature, 1994, 369, 727–729.
a recent field and some results obtained with MOFs or MOF- 6 H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi,
derived materials are at par with the state-of-the-art obtained A. O. Yazaydin, R. Q. Snurr, M. O’Keeffe, J. Kim and
either with different materials, or with similar materials obtained O. M. Yaghi, Science, 2010, 329, 424–428.
via other synthesis approaches. However, to date, it is difficult to 7 W. Xuan, C. Zhu, Y. Liu and Y. Cui, Chem. Soc. Rev., 2012, 41,
1677–1695.
find examples where the use of MOFs for electrode or electrolyte 8 C. Janiak, Dalton Trans., 2003, 2781–2804.
materials resulted in a clear advantage in performance, durability 9 R. Robson, Dalton Trans., 2008, 5113–5131.

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9287
View Online

10 K. Biradha, A. Ramana and J. J. Vittal, Cryst. Growth Des., 2009, 9, 46 A. S. Arico, P. Bruce, B. Scrosati, J.-M. Tarascon and W. Van
2969. Schalwijk, Nat. Mater., 2005, 4, 366–377.
11 A. C. Sudik, A. R. Millward, N. W. Ockwig, A. P. Cote, J. Kim and 47 P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J. M. Tarascon,
O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 7110–7118. Nature, 2000, 407, 496–499.
12 K. S. Park, Z. Ni, A. P. Cote, J. Y. Choi, R. Huang, F. J. Uribe- 48 H. Wang, Q. Pan, Y. Cheng, J. Zhao and G. Yin, Electrochim. Acta,
Romo, H. K. Chae, M. O’Keeffe and O. M. Yaghi, Proc. Natl. 2009, 54, 2851–2855.
Acad. Sci. U. S. A., 2006, 103, 10186–10191. 49 X. W. Lou, D. Deng, J. Y. Lee and L. A. Archer, J. Mater. Chem.,
13 G. Ferey, Chem. Soc. Rev., 2008, 37, 191–214. 2008, 18, 4397–4401.
14 N. Stock and S. Biswas, Chem. Rev., 2011, 112, 933–969. 50 B. Liu, X. Zhang, H. Shioyama, T. Mukai, T. Sakai and Q. Xu, J.
15 S. Qiu and G. Zhu, Coord. Chem. Rev., 2009, 253, 2891–2911. Power Sources, 2010, 195, 857–861.
16 O. M. Yaghi, M. O’Keeffe, N. W. Ockwig, H. K. Chae, 51 J. Zhao, F. Wang, P. Su, M. Li, J. Chen, Q. Yang and C. Li, J.
M. Eddaoudi and J. Kim, Nature, 2003, 423, 705–714. Mater. Chem., 2012, 22, 13328–13333.
17 S. Keskin and S. Kizilel, Ind. Eng. Chem. Res., 2011, 50, 1799–1812. 52 N. Du, H. Zhang, B. Chen, J. Wu, X. Ma, Z. Liu, Y. Zhang,
18 J.-R. Li, R. J. Kuppler and H.-C. Zhou, Chem. Soc. Rev., 2009, 38, D. Yang, X. Huang and J. Tu, Adv. Mater., 2007, 19, 4505–4509.
1477–1504. 53 P. Lavela and J. L. Tirado, J. Power Sources, 2007, 172, 379–387.
19 A. Corma, H. Garcia and F. X. Llabres i Xamena, Chem. Rev., 2010, 54 Y. Sharma, N. Sharma, G. V. S. Rao and B. V. R. Chowdari, Adv.
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

110, 4606–4655. Funct. Mater., 2007, 17, 2855–2861.


20 A. U. Czaja, N. Trukhan and U. Mueller, Chem. Soc. Rev., 2009, 38, 55 P. G. Bruce, S. A. Freunberger, L. J. Hardwick and J.-M. Tarascon,
1284–1293. Nat. Mater., 2012, 11, 19–29.
21 P. Horcajada, C. Serre, G. Maurin, N. A. Ramsahye, F. Balas, 56 X. Ji, K. T. Lee and L. F. Nazar, Nat. Mater., 2009, 8, 500–506.
M. Vallet-Regi, M. Sebban, F. Taulelle and G. Ferey, J. Am. 57 R. Demir-Cakan, M. Morcrette, F. Nouar, C. Davoisne, T. Devic,
Chem. Soc., 2008, 130, 6774–6780. D. Gonbeau, R. Dominko, C. Serre, G. Ferey and
22 P. Horcajada, T. Chalati, C. Serre, B. Gillet, C. Sebrie, T. Baati, J.-M. Tarascon, J. Am. Chem. Soc., 2011, 133, 16154–16160.
Downloaded by Stanford University on 19 October 2012

J. F. Eubank, D. Heurtaux, P. Clayette, C. Kreuz, J.-S. Chang, 58 R. Dominko, M. Gaberscek, J. Drofenik, M. Bele, S. Pejovnik and
Y. K. Hwang, V. Marsaud, P.-N. Bories, L. Cynober, S. Gil, J. Jamnik, J. Power Sources, 2003, 119, 770–773.
G. Ferey, P. Couvreur and R. Gref, Nat. Mater., 2010, 9, 172–178. 59 X. Ji, S. Evers, R. Black and L. F. Nazar, Nat. Commun., 2011, 2,
23 L. E. Kreno, K. Leong, O. K. Farha, M. Allendorf, R. P. Van Duyne 325.
and J. T. Hupp, Chem. Rev., 2011, 112, 1105–1125. 60 W. Sheng, H. A. Gasteiger and Y. Shao-Horn, J. Electrochem. Soc.,
24 H. Wu, W. Zhou and T. Yildirim, J. Am. Chem. Soc., 2007, 129, 2010, 157, B1529–B1536.
5314–5315. 61 R. Subbaraman, D. Tripkovic, D. Strmcnik, K.-C. Chang,
25 N. L. Rosi, J. Eckert, M. Eddaoudi, D. T. Vodak, J. Kim, M. Uchimura, A. P. Paulikas, V. Stamenkovic and
M. O’Keeffe and O. M. Yaghi, Science, 2003, 300, 1127–1129. N. M. Markovic, Science, 2011, 334, 1256–1260.
26 F. Vilela, K. Zhang and M. Antonietti, Energy Environ. Sci., 2012, 5, 62 C. Streb, C. Ritchie, D.-L. Long, P. Koegerler and L. Cronin,
7819–7832. Angew. Chem., Int. Ed., 2007, 46, 7579–7582.
27 S. T. Meek, J. A. Greathouse and M. D. Allendorf, Adv. Mater., 63 S.-T. Zheng, H. Zhang and G.-Y. Yang, Angew. Chem., Int. Ed.,
2011, 23, 249–267. 2008, 47, 3909–3913.
28 R. J. Kuppler, D. J. Timmons, Q.-R. Fang, J.-R. Li, T. A. Makal, 64 B. Keita and L. Nadjo, J. Mol. Catal. A: Chem., 2007, 262, 190–215.
M. D. Young, D. Yuan, D. Zhao, W. Zhuang and H.-C. Zhou, 65 S. Caudo, G. Centi, C. Genovese, G. Giordano, T. Granato,
Coord. Chem. Rev., 2009, 253, 3042–3066. A. Katovic and S. Perathoner, in From Zeolites to Porous Mof
29 U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt Materials: the 40th Anniversary of International Zeolite Conference,
and J. Pastre, J. Mater. Chem., 2006, 16, 626–636. ed. R. Xu, Z. Gao, J. Chen and W. Yan, Elsevier, Amsterdam,
30 G. Ferey, F. Millange, M. Morcrette, C. Serre, M.-L. Doublet, 2007, pp. 2054–2062.
J.-M. Greneche and J.-M. Tarascon, Angew. Chem., Int. Ed., 2007, 66 I. M. Mbomekalle, B. Keita, L. Nadjo, P. Berthet, K. I. Hardcastle,
46, 3259–3263. C. L. Hill and T. M. Anderson, Inorg. Chem., 2003, 42, 1163–1169.
31 K. Saravanan, M. Nagarathinam, P. Balaya and J. J. Vittal, J. 67 B. Nohra, H. El Moll, L. M. Rodriguez Albelo, P. Mialane,
Mater. Chem., 2010, 20, 8329–8335. J. Marrot, C. Mellot-Draznieks, M. O’Keeffe, R. N. Biboum,
32 M. D. Allendorf, A. Schwartzberg, V. Stavila and A. A. Talin, J. Lemaire, B. Keita, L. Nadjo and A. Dolbecq, J. Am. Chem.
Chem.–Eur. J., 2011, 17, 11372–11388. Soc., 2011, 133, 13363–13374.
33 Y. Kobayashi, B. Jacobs, M. D. Allendorf and J. R. Long, Chem. 68 E. Slavcheva, I. Radev, S. Bliznakov, G. Topalov, P. Andreev and
Mater., 2010, 22, 4120–4122. E. Budevski, Electrochim. Acta, 2007, 52, 3889–3894.
34 J. N. Behera, D. M. D’Alessandro, N. Soheilnia and J. R. Long, 69 W. Hu, Y. Wang, X. Hu, Y. Zhou and S. Chen, J. Mater. Chem.,
Chem. Mater., 2009, 21, 1922–1926. 2012, 22, 6010–6016.
35 M. Jahan, Q. Bao, J.-X. Yang and K. P. Loh, J. Am. Chem. Soc., 70 J. Suntivich, K. J. May, H. A. Gasteiger, J. B. Goodenough and
2010, 132, 14487–14495. Y. Shao-Horn, Science, 2011, 334, 1383–1385.
36 C. Petit and T. J. Bandosz, Adv. Mater., 2009, 21, 4753–4757. 71 W. M. Martinez, A. M. Fernandez, U. Cano and A. Sandoval, Int. J.
37 F. T. Wagner, B. Lakshmanan and M. F. Mathias, J. Phys. Chem. Hydrogen Energy, 2010, 35, 8457–8462.
Lett., 2010, 1, 2204–2219. 72 B. S. Yeo and A. T. Bell, J. Phys. Chem. C, 2012, 116, 8394–8400.
38 P. Tran-Van, K. Barthelet, M. Morcrette, M. Herlem, 73 X. Wu, J. Tayal, S. Basu and K. Scott, Int. J. Hydrogen Energy,
J. M. Tarascon, A. K. Cheetham and G. Ferey, J. New Mater. 2011, 36, 14796–14804.
Electrochem. Syst., 2003, 6, 29–31. 74 A. Garsuch, A. Panchenko, C. Querner, A. Karpov, S. Huber and
39 X. Li, F. Cheng, S. Zhang and J. Chen, J. Power Sources, 2006, 160, R. Oesten, Electrochem. Commun., 2010, 12, 1642–1645.
542–547. 75 C. Wang, Z. Xie, K. E. deKrafft and W. Lin, J. Am. Chem. Soc.,
40 K. Barthelet, J. Marrot, D. Riou and G. Ferey, Angew. Chem., Int. 2011, 133, 13445–13454.
Ed., 2002, 41, 281–284. 76 K. F. Babu, M. A. Kulandainathan, I. Katsounaros, L. Rassaei,
41 G. de Combarieu, S. Hamelet, F. Millange, M. Morcrette, A. D. Burrows, P. R. Raithby and F. Marken, Electrochem.
J.-M. Tarascon, G. Ferey and R. I. Walton, Electrochem. Commun., 2010, 12, 632–635.
Commun., 2009, 11, 1881–1884. 77 H. A. Gasteiger, S. S. Kocha, B. Sompalli and F. T. Wagner, Appl.
42 C. Combelles, M. Ben Yahia, L. Pedesseau and M. L. Doublet, J. Catal., B, 2005, 56, 9–35.
Power Sources, 2011, 196, 3426–3432. 78 A. Morozan, B. Jousselme and S. Palacin, Energy Environ. Sci.,
43 C. Combelles, M. Ben Yahia, L. Pedesseau and M. L. Doublet, J. 2011, 4, 1238–1254.
Phys. Chem. C, 2010, 114, 9518–9527. 79 F. Jaouen, E. Proietti, M. Lefevre, R. Chenitz, J.-P. Dodelet, G. Wu,
44 F. M. Courtel, H. Duncan, Y. Abu-Lebdeh and I. J. Davidson, J. H. T. Chung, C. M. Johnston and P. Zelenay, Energy Environ. Sci.,
Mater. Chem., 2011, 21, 10206–10218. 2011, 4, 114–130.
45 J. Cabana, L. Monconduit, D. Larcher and M. Rosa Palacin, Adv. 80 M. Lefevre, E. Proietti, F. Jaouen and J.-P. Dodelet, Science, 2009,
Mater., 2010, 22, E170–E192. 324, 71–74.

9288 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012
View Online

81 G. Wu, K. L. More, C. M. Johnston and P. Zelenay, Science, 2011, 117 Y. J. Xiong, Y. Xie, Z. Q. Li and C. Z. Wu, Chem.–Eur. J., 2003, 9,
332, 443–447. 1645–1651.
82 E. Proietti, F. Jaouen, M. Lefevre, N. Larouche, J. Tian, J. Herranz 118 P. Mahata, T. Aarthi, G. Madras and S. Natarajan, J. Phys. Chem.
and J.-P. Dodelet, Nat. Commun., 2011, 2, 416. C, 2007, 111, 1665–1674.
83 J. P. Dodelet, in N4-Macrocyclic Metal Complexes, ed. J. H. Zagal, 119 F. Zhang, L. Hao, L. Zhang and X. Zhang, Int. J. Electrochem. Sci.,
F. Bedioui and J. P. Dodelet, Springer, New York, 2006, pp. 83–148. 2011, 6, 2943–2954.
84 G. Dong, M. Huang and L. Guan, Phys. Chem. Chem. Phys., 2012, 120 J. Hu, H. Wang, Q. Gao and H. Guo, Carbon, 2010, 48, 3599–
14, 2557–2559. 3606.
85 A. Morozan, P. Jegou, M. Pinault, S. Campidelli, B. Jousselme and 121 S. J. Yang, T. Kim, J. H. Im, Y. S. Kim, K. Lee, H. Jung and
S. Palacin, ChemSusChem, 2012, 5, 647–651. C. R. Park, Chem. Mater., 2012, 24, 464–470.
86 G. Goenaga, S. Ma, S. Yuan and D.-J. Liu, ECS Trans., 2010, 33, 122 L. Chen, J. Bai, C. Wang, Y. Pan, M. Scheer and X. You, Chem.
579–586. Commun., 2008, 1581–1583.
87 S. Ma, G. A. Goenaga, A. V. Call and D.-J. Liu, Chem.–Eur. J., 123 W. Chaikittisilp, M. Hu, H. Wang, H.-S. Huang, T. Fujita,
2011, 17, 2063–2067. K. C. Wu, L.-C. Chen, Y. Yamauchi and K. Ariga, Chem.
88 M. Lefevre and J.-P. Dodelet, 221st ECS Meeting Seattle, Commun., 2012, 48, 7259–7261.
Washington, 2012. 124 B. Liu, H. Shioyama, T. Akita and Q. Xu, J. Am. Chem. Soc., 2008,
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G

89 M. Jahan, Q. Bao and K. P. Loh, J. Am. Chem. Soc., 2012, 134, 130, 5390–5391.
6707–6713. 125 B. Liu, H. Shioyama, H. Jiang, X. Zhang and Q. Xu, Carbon, 2010,
90 J. Mao, L. Yang, P. Yu, X. Wei and L. Mao, Electrochem. Commun., 48, 456–463.
2012, 19, 29–31. 126 H.-L. Jiang, B. Liu, Y.-Q. Lan, K. Kuratani, T. Akita, H. Shioyama,
91 X. Feng, L. Chen, Y. Dong and D. Jiang, Chem. Commun., 2011, 47, F. Zong and Q. Xu, J. Am. Chem. Soc., 2011, 133, 11854–11857.
1979–1981. 127 A. Almasoudi and R. Mokaya, J. Mater. Chem., 2012, 22, 146–152.
92 O. K. Farha, A. M. Shultz, A. A. Sarjeant, S. T. Nguyen and 128 H.-K. Youn, J. Kim and W.-S. Ahn, Mater. Lett., 2011, 65, 3055–
Downloaded by Stanford University on 19 October 2012

J. T. Hupp, J. Am. Chem. Soc., 2011, 133, 5652–5655. 3057.


93 T. Ohmura, A. Usuki, K. Fukumori, T. Ohta, M. Ito and 129 P. Simon and Y. Gogotsi, Nat. Mater., 2008, 7, 845–854.
K. Tatsumi, Inorg. Chem., 2006, 45, 7988–7990. 130 P. Arora and Z. M. Zhang, Chem. Rev., 2004, 104, 4419–4462.
94 L. Yang, S. Kinoshita, T. Yamada, S. Kanda, H. Kitagawa, 131 C. Monroe and J. Newman, J. Electrochem. Soc., 2005, 152, A396–
M. Tokunaga, T. Ishimoto, T. Ogura, R. Nagumo, A. Miyamoto A404.
and M. Koyama, Angew. Chem., Int. Ed., 2010, 49, 5348–5351. 132 B. M. Wiers, M.-L. Foo, N. P. Balsara and J. R. Long, J. Am. Chem.
95 H. Kitagawa, Y. Nagao, M. Fujishima, R. Ikeda and S. Kanda, Soc., 2011, 133, 14522–14525.
Inorg. Chem. Commun., 2003, 6, 346–348. 133 H. K. Chae, D. Y. Siberio-Perez, J. Kim, Y. Go, M. Eddaoudi,
96 A. M. Fan, in Encyclopedia of Environmental Health, ed. J. O. A. J. Matzger, M. O’Keeffe and O. M. Yaghi, Nature, 2004, 427,
Nriagu, Springer, Burlington, 2011, pp. 137–145. 523–527.
97 K. C. M. Campbell, Toxicology, 2006, 221, 205–211. 134 A. Demessence, D. M. D’Alessandro, M. L. Foo and J. R. Long, J.
98 X. Wang, H. Zhao, H. Lin, G. Liu, J. Fang and B. Chen, Am. Chem. Soc., 2009, 131, 8784–8786.
Electroanalysis, 2008, 20, 1055–1060. 135 S. R. Caskey, A. G. Wong-Foy and A. J. Matzger, J. Am. Chem.
99 H. Lin, X. Wang, H. Hu, B. Chen and G. Liu, Solid State Sci., 2009, Soc., 2008, 130, 10870–10871.
11, 643–650. 136 K. Xu, Chem. Rev., 2004, 104, 4303–4417.
100 W. Wei, K. Yu, Z.-h. Su, Y. Yu, B.-b. Zhou and C.-c. Zhu, Inorg. 137 B. Smitha, S. Sridhar and A. A. Khan, J. Membr. Sci., 2005, 259, 10–
Chem. Commun., 2012, 17, 21–25. 26.
101 Y. Yang, S. Liu, C. Li, S. Li, G. Ren, F. Wei and Q. Tang, Inorg. 138 Q. Li, J. O. Jensen, R. F. Savinell and N. J. Bjerrum, Prog. Polym.
Chem. Commun., 2012, 17, 54–57. Sci., 2009, 34, 449.
102 L. Marleny Rodriguez-Albelo, A. Rabdel Ruiz-Salvador, 139 J. Roziere and D. J. Jones, Annu. Rev. Mater. Res., 2003, 33, 503–
A. Sampieri, D. W. Lewis, A. Gomez, B. Nohra, P. Mialane, 555.
J. Marrot, F. Secheresse, C. Mellot-Draznieks, R. N. Biboum, 140 Y. Nagao, M. Fujishima, R. Ikeda, S. Kanda and H. Kitagawa,
B. Keita, L. Nadjo and A. Dolbecq, J. Am. Chem. Soc., 2009, 131, Synth. Met., 2003, 133, 431–432.
16078–16087. 141 T. Yamada, M. Sadakiyo and H. Kitagawa, J. Am. Chem. Soc.,
103 J. Chmiola, G. Yushin, Y. Gogotsi, C. Portet, P. Simon and 2009, 131, 3144–3145.
P. L. Taberna, Science, 2006, 313, 1760–1763. 142 A. Shigematsu, T. Yamada and H. Kitagawa, J. Am. Chem. Soc.,
104 G. Lota, T. A. Centeno, E. Frackowiak and F. Stoeckli, Electrochim. 2011, 133, 2034–2036.
Acta, 2008, 53, 2210–2216. 143 M. Sadakiyo, H. Okawa, A. Shigematsu, M. Ohba, T. Yamada and
105 D. Y. Lee, S. J. Yoon, N. K. Shrestha, S.-H. Lee, H. Ahn and H. Kitagawa, J. Am. Chem. Soc., 2012, 134, 5472–5475.
S.-H. Han, Microporous Mesoporous Mater., 2012, 153, 163– 144 J. M. Taylor, R. K. Mah, I. L. Moudrakovski, C. I. Ratcliffe,
165. R. Vaidhyanathan and G. K. H. Shimizu, J. Am. Chem. Soc.,
106 R. Diaz, M. Gisela Orcajo, J. A. Botas, G. Calleja and J. Palma, 2010, 132, 14055–14057.
Mater. Lett., 2012, 68, 126–128. 145 S. C. Sahoo, T. Kundu and R. Banerjee, J. Am. Chem. Soc., 2011,
107 C. D. Lokhande, D. P. Dubal and O.-S. Joo, Curr. Appl. Phys., 2011, 133, 17950–17958.
11, 255–270. 146 N. C. Jeong, B. Samanta, C. Y. Lee, O. K. Farha and J. T. Hupp, J.
108 R. Das, P. Pachfule, R. Banerjee and P. Poddar, Nanoscale, 2012, 4, Am. Chem. Soc., 2012, 134, 51–54.
591–599. 147 J. A. Hurd, R. Vaidhyanathan, V. Thangadurai, C. I. Ratcliffe,
109 C. Y. Su, A. M. Goforth, M. D. Smith, P. J. Pellechia and H. C. zur I. L. Moudrakovski and G. K. H. Shimizu, Nat. Chem., 2009, 1,
Loye, J. Am. Chem. Soc., 2004, 126, 3576–3586. 705–710.
110 X. Liu, Angew. Chem., Int. Ed., 2009, 48, 3018–3021. 148 V. G. Ponomareva, K. A. Kovalenko, A. P. Chupakhin,
111 Z. Q. Li, Y. J. Xiong and Y. Xie, Nanotechnology, 2005, 16, 2303– E. S. Shutova and V. P. Fedin, Solid State Ionics, 2012, DOI:
2308. 10.1016/j.ssi.2012.01.044.
112 H. Thakuria, B. M. Borah and G. Das, Eur. J. Inorg. Chem., 2007, 149 M. Sadakiyo, T. Yamada and H. Kitagawa, J. Am. Chem. Soc.,
524–529. 2009, 131, 9906–9907.
113 H.-Y. Shi, B. Deng, S.-L. Zhong, L. Wang and A.-W. Xu, J. Mater. 150 S. Bureekaew, S. Horike, M. Higuchi, M. Mizuno, T. Kawamura,
Chem., 2011, 21, 12309–12315. D. Tanaka, N. Yanai and S. Kitagawa, Nat. Mater., 2009, 8, 831–
114 J. Zhao, M. Li, J. Sun, L. Liu, P. Su, Q. Yang and C. Li, Chem.–Eur. 836.
J., 2012, 18, 3163–3168. 151 C. Dey, T. Kundu and R. Banerjee, Chem. Commun., 2012, 48, 266–
115 W. Cho, S. Park and M. Oh, Chem. Commun., 2011, 47, 4138– 268.
4140. 152 M. G. Goesten, J. Juan-Alcaniz, E. V. Ramos-Fernandez,
116 W. Cho, Y. H. Lee, H. J. Lee and M. Oh, Chem. Commun., 2009, K. B. S. S. Gupta, E. Stavitski, H. van Bekkum, J. Gascon and
4756–4758. F. Kapteijn, J. Catal., 2011, 281, 177–187.

This journal is ª The Royal Society of Chemistry 2012 Energy Environ. Sci., 2012, 5, 9269–9290 | 9289
View Online

153 S. W. Li, Z. Zhou, Y. L. Zhang, M. L. Liu and W. Li, Chem. Mater., 158 S. E.-d. H. Etaiw, A. E.-A. S. Fouda, S. A. Amer and M. M. El-
2005, 17, 5884–5886. bendary, J. Inorg. Organomet. Polym. Mater., 2011, 21, 327–335.
154 H. Kitagawa, Nat. Chem., 2009, 1, 689–690. 159 K. Babic-Samardzija, K. F. Khaled and N. Hackerman, Anti-
155 M. Elayyachy, B. Hammouti and A. El Idrissi, Appl. Surf. Sci., 2005, Corros. Methods Mater., 2005, 52, 11–21.
249, 176–182. 160 G. Givaja, P. Amo-Ochoa, C. J. Gomez-Garcia and F. Zamora,
156 S. E.-d. H. Etaiw, A. E.-A. S. Fouda, S. N. Abdou and M. M. El- Chem. Soc. Rev., 2012, 41, 115–147.
bendary, Corros. Sci., 2011, 53, 3657–3665. 161 P. Lin, R. A. Henderson, R. W. Harrington, W. Clegg, C. D. Wu and
157 Aa. A. Massoud, A. Hefnawy, V. Langer, M. A. Khatab, X. T. Wu, Inorg. Chem., 2004, 43, 181–188.
L. Ohrstrom and M. A. M. Abu-Youssef, Polyhedron, 2009, 28, 162 D. M. D’Alessandro, J. R. R. Kanga and J. S. Caddy, Aust. J.
2794–2802. Chem., 2011, 64, 718–722.
Published on 05 October 2012 on http://pubs.rsc.org | doi:10.1039/C2EE22989G
Downloaded by Stanford University on 19 October 2012

9290 | Energy Environ. Sci., 2012, 5, 9269–9290 This journal is ª The Royal Society of Chemistry 2012

You might also like