You are on page 1of 10

Energy & Environmental Science

Cite this: Energy Environ. Sci., 2012, 5, 6012 www.rsc.org/ees

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

Dynamic Article Links <

PERSPECTIVE

Catalysts made of earth-abundant elements (Co, Ni, Fe) for water splitting: Recent progress and future challenges
Pingwu Du* and Richard Eisenberg*
Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

Received 22nd November 2011, Accepted 6th January 2012 DOI: 10.1039/c2ee03250c This article reviews recent signicant advances in the eld of water splitting. Catalysts play very important roles in two half reactions of water splitting - water reduction and water oxidation. Considering potential future applications, catalysts made of cheap and earth abundant element(s) are especially important for economically viable energy conversion. This article focuses only on catalysts made of cobalt (Co), nickel (Ni) and iron (Fe) elements for water reduction and water oxidation. Different series of catalysts that can be applied in electrocatalytic and photocatalytic water spitting are discussed in detail and their catalytic mechanisms are introduced. Finally, the future outlook and perspective of catalysts made of earth abundant elements will be discussed.

1. Introduction
Our current energy resources are mainly from fossil fuels (oil, coal and gas), in which the oil supplies are assumed to be depleted in 50150 years,1 and the use of the fossil fuels has resulted in numerous environmental problems, such as the increasing emissions of greenhouse gas, most notably carbon dioxide (CO2).2 In 2008, the total worldwide energy consumption was at a rate of 15 TW.3 This number, corresponding to a burn-rate of energy, is estimated to nearly double by 2050. Considering important factors including the economy, the environment and human health, the increased amount of energy must come from

Department of Chemistry, University of Rochester, Rochester, NY, 14627. E-mail: pdu@mail.rochester.edu; eisenberg@chem.rochester.edu Present address: Chemical Sciences & Engineering Division, Argonne National Laboratory, Bldg 200, E105, 9700 S. Cass Avenue, Argonne, IL 60439.

renewable and sustainable sources. Among all of the renewable sources (solar energy, wind, geothermal heat, tides, biomass, et al.), solar energy is the most feasible, and from a fundamental scientic viewpoint, the most attractive alternative energy source.4,5 In nature, photosynthesis plays a crucial role to convert solar energy into chemical energy, which fuels life on earth and provides carbohydrates (food), oxygen and even fossil fuels. By utilizing solar energy in plant and photosynthetic bacteria, CO2 and water (H2O) are consumed to produce oxygen and carbohydrates via a series of photochemical reactions and dark catalytic reactions.68 There are two related photosystems in photosynthesis7,9 (Scheme 1, Z-scheme): Photosystem I (P700) and Photosystem II (P680). When Photosystem I is excited by sun light, an electron transfer (ET) process is induced to reduce a series of electron acceptors, such as, phylloquinone and ferredoxin, ultimately reducing the cofactor NADP+ (nicotinamide adenine dinucleotide phosphate) to NADPH (biological

Broader context
Solar energy could be the future carbon-neutral energy resource to replace fossil fuels. The current goal is to develop efcient catalysts (materials) or systems to give us the ability to use solar energy on demand. Solar-driven water splitting for hydrogen production represents one fascinating pathway for clean energy conversion and energy storage from visible light irradiation into chemical bonds. During this process, catalysts play a crucial role for energy conversion through lowered energetic barriers of reaction, thus leading to lower energy loss in driving an uphill or energy-storing reaction. Most available water splitting catalysts are made of expensive noble metals (such as platinum, ruthenium, iridium and rhodium), but for real wide-spread application, catalysts made from elementally abundant and less expensive materials are urgently required. In this perspective, we present an overview of catalysts which are principally molecular and contain as metallic elements only cobalt, nickel or iron for electrocatalytic or photocatalytic water splitting. In addition, their catalytic mechanisms are also discussed. We hope this article will attract more attention in order to stimulate further work by others on the development of inexpensive, efcient and robust catalysts for water splitting in articial photosynthetic systems.
6012 | Energy Environ. Sci., 2012, 5, 60126021 This journal is The Royal Society of Chemistry 2012

View Article Online

Scheme 2

Scheme 1

hydrogen) and Photosystem I is oxidized. NADPH is both the electron and proton source for CO2 transformation into carbohydrates in the Calvin cycle.10 Oxidized Photosystem I is regenerated by obtaining electrons from Photosystem II through several electron relays, such as, plastoquinone, cytochrome f and plastocyanine. The oxidized Photosystem II mediates another important reaction - water oxidation to produce oxygen, catalyzed by a CaMn4O4 cluster.1113 Inspired by nature, the development of renewable carbon-free energy using the sun is a major scientic and technological challenge in meeting the energy needs of the future. Efcient solar cells (that is, solid state photovoltaic devices and dyesensitized solar cells) and solar-driven water splitting systems represent two major objectives in utilizing solar energy.14 While solar cells can directly convert solar radiation into electricity, they can not store energy.15 On the other hand, solar-driven water splitting can store energy in the form of fuel (H2) and oxidant (O2).1,5 Energy storage in chemical bonds is most efcient, and as a consequence, this strategy for solar energy storage has attracted great attention during the past three decades. Hydrogen is an environmentally-friendly energy carrier because its only oxidation product is H2O. Through widespread use of H2 as a fuel, generated either directly in an articial photosynthetic system or by electrolysis using solar-generated electricity,

dependence on fossil fuels can be decreased, with a corresponding decrease in CO2 emissions. Generally, overall water splitting consists of two half reactions - water reduction and water oxidation. Since the water oxidation reaction requires a four-electron transfer process coupled to the removal of four protons from water molecules to form oxygenoxygen bond (Scheme 2),16 this half reaction is considered as the obstacle to achieve the overall water splitting. In the literature, many homogeneous and heterogeneous catalysts made of noble metals (such as, ruthenium, iridium and rhodium) have been developed and several review articles have been published to summarize the recent progress.1721 However, these catalysts made of noble metals could signicantly increase the cost for real application. Since nature uses manganese as one of the key metal ions to construct the Mn4CaO4 ap-opened cube water oxidation catalyst, attention has been directed to various manganese complexes to mimic the function of Photosystem II, which has been reviewed in the literature.2224 As for water reduction, platinum has been widely used as an efcient water reduction catalyst. It is widely accepted that platinum must be replaced by cheaper materials because of its high price and very low earth abundance.25 Hence, the development of efcient water oxidation catalysts (WOCs) and water reduction catalysts (WRCs) made of cheap, earth-abundant elements is an immensely important subject in the eld of solar energy conversion. In this Perspective, we briey review the recent progress on the catalysts made of cheap and earth-abundant elements, including cobalt, nickel and iron (Co, Ni and Fe) elements for both water oxidation and water reduction. A review article focusing on cobalt has just been published.26 To differentiate this Perspective

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

Pingwu Du

Pingwu Du received a B.S. degree from Wuhan University (2001) and a Ph.D. degree from University of Rochester (2009) under the direction of Professor Richard Eisenberg. After two-year postdoc training at MIT, he joined Argonne National Laboratory as a Directors Fellow. His current research focuses on photocatalysis and inorganic chemistry. He is a recipient of Young Investigator Award (Inorganic Division, 2008) and Nobel Laureate Signature Award (2011) by the American Chemical Society.

A native New Yorker, Rich Eisenberg moved to the University of Rochester in 1973 where he was Tracy Harris Professor of Chemistry and Chair of the Department during 199194. Richs research interests are in inorganic and organometallic chemistry, photochemistry relating to solar energy conversion, and catalysis. He is the Editor-in-Chief of Inorganic Chemistry, and has served on Richard Eisenberg the editorial advisory boards of other ACS journals including JACS and Accounts of Chemical Research. He has been the recipient of a several ACS awards and in 2010, he received the University of Rochester Lifetime Achievement Award in Graduate Education and was elected to the U. S. National Academy of Sciences.
Energy Environ. Sci., 2012, 5, 60126021 | 6013

This journal is The Royal Society of Chemistry 2012

View Article Online

from this recent review, we will only concentrate on the literature for hydrogen production from water (the protons are directly from water), rather than hydrogen production from organic acids in non-aqueous media. Hematite-based materials for water oxidation have also been reviewed recently,27 and will not be discussed here. Different series of the Co, Ni and Fe catalysts that can be employed in both electrocatalytic and photocatalytic water splitting are discussed in detail, and their catalytic mechanisms are briey described. Finally, an outlook and perspective of catalysts made from earth abundant elements will be presented.
Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

2. Catalysts made of cobalt for water splitting


2.1 Water reduction Cobalt hydride was discovered to evolve hydrogen in the early 1970s.28 Since then, numerous cobalt complexes have been reported as WRCs, including systems containing the following ligands: [14]diene-N4 and [14]tetraene-N4 macrocycles,29,30 porphyrins,31 biand polypyridines,32,33 h5-cyclopentadienides,34,35 phosphines36 and dimethyl glyoximes and related derivatives.37 Most of these Co WRCs were used for electrocatalytic hydrogen production in which the proton source was organic acid rather than water. In this section, we focus only on the catalysts for hydrogen production from aqueous protons. In 1980, Fisher and Eisenberg reported two early examples using homogeneous cobalt macrocyclic complexes (Scheme 3, 1 and 2) for simultaneously electrocatalytic reduction of CO2 and water to produce CO and H2.29 These complexes were able to reduce CO2 and water at potentials ranging from 1.5 V to 1.6 V vs. SCE, with up to 93% current efciency. However, total turnover numbers (TONs) and turnover frequencies (TOF) were low, and there was a large overpotential for reaction. Control experiments showed that either catalyst is crucial for CO and H2 production upon electrolysis, and that water was also an essential component of the system, as no CO or H2 was detected when electrolysis was done in dry Me2SO in the presence of CO2. In 1981, Krishnan and Sutin reported the rst photocatalytic system using cobalt bipyridine complexes as the catalyst, together with Ru(bpy)32+ (bpy bipyridine) as the photosensitizer and ascorbate as the sacricial electron donor.32 During the reaction, Co(I) bipyridine species was identied as the key intermediate via laser ash photolysis. The optimum pH for H2 formation was 5.0, which was probably favorable for formation of the cobalt hydride intermediate. An isotope experiment conrmed that the hydrogen generated originated from protons in the water. The same catalyst was later used for CO2 and water reduction to produce CO and H233 that found with Co tetraazamacrocycle complexes29 but the driving force was visible light rather than electrochemical potential. In that system, Ru(bpy)32+ was used as

the photosensitizer and amines (triethylamine or triethanolamine, TEA or TEOA) were the sacricial electron donor. The system produced CO and H2 simultaneously with the relative amounts of CO and H2 proportional to the concentration of the dissolved CO2. More recently, Bernhard and co-workers have reported a series of iridium based photosensitizers together with Co(bpy)32+ as the catalyst (Scheme 3, 4) and TEOA as the sacricial electron donor for the visible light-driven production of hydrogen from water.38 Compared with the classical Ru(bpy)32+ photosensitizer, the iridium photosensitizers have higher activity, with up to 37 times greater quantum efciency for hydrogen production. Up to 9000 TONs (based on the photosensitizer) and 74 TONs (based on the catalyst) have been achieved. One early example of a cobaloxime-based catalyst for water reduction was reported in 1983 by Lehn and co-workers (Scheme 4, 5).39 The system contains Ru(bpy)32+ as the photosensitizer and TEOA as the sacricial donor. While H2 was formed at the rate of 16 turnovers per hour in a DMF/H2O mixture at pH $ 9.0, the instability and air-sensitivity of the Co(II) dimethylglyoximate catalyst limited further study at the time. The use of the more stable BF2-annulated cobaloxime 12 for hydrogen production was initially reported by Espenson40 in 1986, and has been developed during the past decade by Peters,30,41 Artero and Fontacave42,43 for electrocatalytic proton reduction. Related systems for the photochemical generation of H2 based on related cobaloxime catalysts were also reported by Artero and Fontecave,44 but in all of these systems, the catalysts were operating in non-aqueous media with added acids or organic ammonium salts as the proton sources. In 2008, Du and Eisenberg45 reported the application the Hbridged Co(III) bis(dimethylglyoximate) complex 6 bearing an axial pyridine ligand as the catalyst for visible light-driven water reduction in organic/aqueous media (mixture of MeCN and H2O), in combination with a platinum chromophore and TEOA as the sacricial electron donor. Complex 610 in Scheme 4 were all found to be active with catalytic activity varying in the visible

Scheme 3

Scheme 4

6014 | Energy Environ. Sci., 2012, 5, 60126021

This journal is The Royal Society of Chemistry 2012

View Article Online

light-driven production of H2 from aqueous protons. Interestingly, complex 11 was not an active catalyst for photocatalysis in this system. For 6, 1000 TONs was obtained in 10 h when the ratio of MeCN/H2O was 3 : 2 (v/v) and the efciency increased to 2150 TONs when the ratio was changed to 24 : 1. The rate of hydrogen production was found to be sensitive to different factors including the nature of the platinum chromophores, the redox properties of the cobaloxime catalysts, pH of the media, solvents, and the concentration of each reaction component. A mercury test was conducted to conrm that the catalytic reaction was really a homogeneous process.46 Spectroscopic and mechanistic studies of the system revealed an induction period to hydrogen evolution and an initial color change to bright yellow, characteristic of Co(III) reduction to Co(II). Hydrogen was usually produced after 12 h photolysis with l > 410 nm when [TEOA] 1.6 102 M. Mechanistically, H2 production occurs via further reduction of Co(II) to Co(I), followed by protonation process to form a Co(III) hydride intermediate. Variation of catalyst concentration reveals a linear relationship with H2 evolution that is consistent with a mechanism involving only a single Co for H2 generation. The mechanism is proposed to proceed by reduction of the Co(III) hydride intermediate and subsequent reaction with H+ to give H2 with regeneration of Co(II) catalyst. Interestingly, [Co(dmgBF2)2(Sol)2] (Scheme 4, 12), which is most effective in electrocatalytic reduction of H+ to H2, did not work in the this photosynthetic system, which may be due in part to the low reactivity of the corresponding Co(I) species. The reduction potential for 12 (ECo(II)/Co(I) 0.29 V vs. NHE) is relatively positive, indicating that the Co(I) species may not be basic enough to undergo protonation at pH 7.08.5. Sunsequently, Castellano and co-workers studied the relationship between the rate of hydrogen production and the inuence of the phenylacetylide p-conjugation length in the platinum based chromophore using cobaloximes as the WRCs.47 Following the success of application of Co(III) dimethylglyoximate complexes in visible light-driven hydrogen production systems, Eisenberg and co-workers discovered the rst homogeneous system for visible light-driven hydrogen production from water based on only earth-abundant elements,48 using organic dyes (Eosin Y or Rose Bengal) as the photosensitizers, Co(III) dimethylglyoximate complexes as the catalysts and TEOA as the sacricial donor. In 12 h of irradiation, 900 TONs were obtained in this system. Mechanistic studies revealed that the lowest lying triplet states in those organic dyes which were accessed through the with heavy atom effect play a very important role for electron transfer to achieve Co(III) reduction and subsequent hydrogen production involving Co(II), Co(I) and Co(III) hydride intermediates. Fluorescein, which has no heavy atom and therefore leads to minimal population of triplet states, yielded no hydrogen in an analogous system. In further studies, up to 9000 TONs were achieved when rhodamine dyes (with S or Se in place of O in the xanthene ring) were used as the photosensitizers.49 Sun and co-workers investigated a similar system containing Eosin Y or Rose Bengal as the photosensitizers, BF2annulated cobaloximes (Scheme 4, complex 1215) as the catalysts and TEOA as the sacricial donor, in which the highest TONs of 327 was achieved in 5 h.50 New catalysts based on cobalt have recently been reported. In 2011, McNamara, et al.,51 reported the rst example of a cobalt
This journal is The Royal Society of Chemistry 2012

dithiolene complex (Scheme 4, 16) for both photocatalytic and electrocatalytic hydrogen production from water. Complex 16 is a very active catalyst with up to 2700 TONs (based on the catalyst) and initial turnover rate of 880 mol H2/mol catalyst/h when employed with Ru(bpy)32+ as the photosensitizer and ascorbic acid as the sacricial donor. Complex 16 is also an active electrocatalyst in mixed MeCN/H2O in the presence of weak acids with a faradaic yield of >99% when the potential is held at 1.0 V vs. SCE. Stubbert, et al.,52 have found that a cobalt bis(iminopyridine) complex (Scheme 4, 17) is also a highly active electrocatalyst for aqueous proton reduction. Faradaic hydrogen yields up to 87 10% have been obtained under constant potential electrolysis at 1.4 V vs. SCE in aqueous buffered solution. The design involving a redox noninnocent bis(iminopyridine) ligand for cobalt is believed to be the key factor for the enhanced electrocatalysis without signicant ligand decomposition at moderate pH. A mechanism involving initial reduction of Co(II) and cobalt hydride formation is proposed. Another interesting approach to water reduction involves building molecular photochemical devices for visible light driven hydrogen production directly from water. The rst cobalt based photochemical molecular device (assembly of Ru(bpy)32+ derivatives and cobaloximes, Scheme 5, 18 as one example) was reported by Artero and Fontecave in 2008.44 In this system, ammonium salts (p-cyanoanilinium or trimethylammonium chloride) served as the proton source and even a small amount of water inhibited the systems activity. An iridium chromophore was later reported to replace Ru(bpy)32+ but the experiments were still conducted in the presence of ammonium salts as the proton source (Scheme 5, 19).53 Two other examples of photochemical devices were recently reported for visible light-driven hydrogen production directly using water as the proton source. One system contained a porphyrin as the photosensitizer and a H-bridged cobaloxime bearing an axial pyridine ligand as the catalyst (Scheme 5, 2022).54 Another was an iridium photosensitizer and a cobalt polypyridine complex as the catalyst (Scheme 5, 2326).55 However, both systems exhibited only modest turnovers for hydrogen production from water (TONs < 30). Besides molecular photochemical devices, a heterogeneous photochemical device was also developed based on TiO2 semiconductor. Lakadamyali and Reisner56 reported a photochemical device based on TiO2 anchored by functionalized Ru(bpy)32+ (Scheme 6, 27) as the photosensitizer and [Co(III)(dmgH)2(pyridyl-4-hydrophosphonate)Cl] (28) as the catalyst. It was quite

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

Scheme 5

Energy Environ. Sci., 2012, 5, 60126021 | 6015

View Article Online

Scheme 6

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

facile to assemble this system, which works under visible light irradiation in neutral water (no organic solvent). Although it seems the TOF of the system is low (0.005 s1), the modied TiO2 produced a very good level of activity based on semiconductor at about 400 mmol H2 h1(g TiO2)1. Phosphate-free cobaloxime (complex 6) was used as the catalyst in a control experiment and produce much less hydrogen, indicating the importance of the semiconductor as the electron relay. 2.2 Water oxidation Since the late 1970s, inorganic cobalt salts have been studied as water oxidation catalysts.5760 However, subsequent development of these cobalt salt catalysts was slow after initial discovery because the yields of oxygen production always dropped rapidly in the presence of a chemical oxidant. The problem was in the active cobalt species for water oxidation precipitating out of solution to form heterogeneous particles, which reduced the effective catalyst concentration and decreased O2 evolution. The precipitated particles were proposed to be cobalt oxides/ hydroxides (CoOx). In 2008, Kanan and Nocera used indium tin oxide (ITO) as the working electrode in a three-electrode system containing 0.05 M Co(II) and inorganic phosphate (Pi) as the electrolyte at pH 7.0.61 Surprisingly, a black thin lm was deposited on the surface of ITO (Scheme 7, 29) and bubbles could be seen when a potential of >1.2 V vs. NHE was applied in the system. The bubbles were conrmed to be oxygen. This cobalt-based water oxidation catalyst (Co-WOC) is a functional model of the oxygen-evolving complex of Photosystem II and demonstrates some fascinating properties,6267 including: (1) the lm is formed in situ under mild conditions, pH 7.0 and room temperature; (2) a variety of cobalt precursors (such as, CoCl2, Co(NO3)2, CoSO4, et al.) can be used for lm deposition; (3) the Co-WOC exhibits high activity in pH 7.0 water, river water, even sea water; (4) the catalyst lm is made of earth abundant cobalt; (5) the catalyst lm has high stability in the atmosphere; and (6) the catalyst lm is self-healing through a series of linked equilibria. More recently, Nocera and coworkers have interfaced the Co-WOC with light absorbing and charge separating materials (hematite semiconductor6870 and ptype/n-type silicon nanomaterials71,72) and a water reduction

catalyst to make an articial leaf with photoelectrochemical water splitting. X-ray absorption studies have shown that the Co-WOC has structural similarity to the active site of Photosystem II (CaMn4O4 cluster) (Scheme 7, 30).63,73 The Co-WOC is a hetergeneous catalyst at neutral pH with Co(IV) species likely important intermediates for water oxidation.74 A detailed mechanistic study for electrocatalytic water oxidation with Co-WOC has very recently been published.75 The results indicate that the Co-WOC has homogeneous water oxidation mechanism below pH 3.5 but switches to heterogeneous mechanism above pH 3.5. Besides the Co phosphate catalyst described above, heterogeneous cobalt oxide (Co3O4) was also reported as water oxidation catalyst, which likely involves similar active species to those obtained from the Co-WOC formed from simple cobalt salts.7678 So far, no detailed study of the mechanism has been reported for these materials.79 There are very few catalysts based on homogeneous cobalt complexes reported as active WOCs. In 2009, the Hill group reported a cobalt polyoxametalate (Co-POM) complex, B-type [Co4(H2O)2(a-PW9O34)2]10 (Scheme 7, 31) as an organic ligandfree, homogeneous water oxidation catalyst.80 The TOF for 31 was above 5 s1 and the highest TONs was >1000 in 3 min at pH 8.0 in the presence of a chemical oxidant (tris(2,20 -bipyridyl)triperchlororuthenium(III)). The efciency was described as higher than for previously reported heterogeneous cobalt-based catalysts. The yield of oxygen production was found to be highly dependent on different pH values and buffer systems. Meanwhile, to conrm that the Co-POM complex was a homogeneous catalyst in solution, the authors provided experimental data to conrm the catalyst as very stable in solution at pH 8.0, including UV-Vis spectra, 31P NMR of pre-reaction solution, 31P NMR of post-reaction solution, IR characterization, a poisoning experiment using a bipyridine ligand as an inhibitor, reuse of the catalyst after the initial catalytic run and voltammetric behavior of the catalyst in the presence of an oxidant. The same cobalt POM catalyst was later used in a photocatalytic water oxidation system by Hill and co-workers with sodium persulfate as the sacricial electron acceptor and tris(2,20 -bipyridyl)ruthenium(II) as the photosensitizer.81 A high quantum yield of 30% and >200 TONs was achieved at pH 8.0. The results showed that this complex was oxidatively stable, homogeneous and an efcient molecular catalyst for visible lightdriven catalytic oxidation of water. More recently, however, Finke and co-workers have published a detailed study of the mechanism of 31 for electrocatalytic water oxidation.82 In their study, they demonstrated that the Co-POM catalyst is just a catalyst precursor and decomposes during

Scheme 7

6016 | Energy Environ. Sci., 2012, 5, 60126021

This journal is The Royal Society of Chemistry 2012

View Article Online

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

electrocatalysis. The decomposed product, CoOx, is probably the real WOC. Four primary lines of evidence were presented to support this conclusion, including: (1) the instability of the oxidation wave from linear-sweep voltammetry; (2) bulk lm formation during electrocatalysis, that was characterized by SEM and EDX techniques; (3) UV-Vis spectra that revealed degradation of the Co-POM complex in solution during water oxidation; and (4) the determination that catalytic activity of a 500 mM solution of the Co-POM complex produced almost the same amount of oxygen as a 58 mM Co(NO3)2. All of this evidence supports the view that the Co-POM complex is simply the precursor of the real catalyst. The actual role Co-POM plays in this catalysis must await determination of the structure of the lm obtained during electrolysis by X-ray diffraction or X-ray absorption techniques. Another example of a Co-based homogeneous WOC was reported from Berlinguette and co-workers (Scheme 7, 32).83 An oxidatively stable pentadentate ligand of 2,6-(bis(bis-2-pyridyl) methoxy-methane)-pyridine (Py5) was used to synthesize the stable cobalt coordination compound, [Co(Py5)(OH2)](ClO4)2. This complex was found to undergo a proton-coupled electron transfer (PCET) step to stabilize a [Co(III)OH]2+ unit, and that further oxidation of this species renders a Co (IV) intermediate that reacts with water in the presence of a base to produce O2.

3. Catalysts made of nickel for water splitting


3.1 Water reduction Nickel is another interesting element for making WRCs. Nature uses nickel to build [NiFe] hydrogenases for proton reduction and hydrogen oxidation.84 The structure of the active site (Scheme 8, 33) of [NiFe] hydrogenase has been determined by single crystal X-ray diffraction.8587 Inspired by nature, many [NiFe] hydrogenase model complexes and nickel based functional complexes have been developed.8890 However, very few examples could be used for hydrogen production in a catalytic system. The Rauchfuss group recently reported several Ni-Fe complexes for electrocatalytic proton reduction from acids (Scheme 8, 34, 35), but to date, no model complex of [NiFe] hydrogenase has been reported for hydrogen production when water is the proton source. In this section, we only present a few

functional model complexes containing Ni for hydrogen production from water. Following the initial report by Fisher and Eisenberg of homogeneous nickel(II) macrocyclic catalysts (Scheme 8, 3638) for simultaneous electrocatalytic CO2 and water reduction to produce CO and H2,29 Collin and Sauvage described two related nickel(II) macrocyclic complexes, Ni(cyclam)2+ and Ni2(biscyclam)24+ (Scheme 8, 39,40), for the same electrocatalytic CO2 and water reduction.91 The activity of the complexes was compared at a xed potential (1.25 V vs. SCE). The results showed that complex 40 is three (at pH 8.0) to ten times (at pH 7.0) more active than complex 39. Turnover numbers of up to 100 were reached, and a reaction mechanism involving a nickel hydride intermediate was proposed. DuBois et al. recently designed a series of nickel based molecular catalysts,9294 [Ni(P2 ArN2Ar)2](BF4)2 (Scheme 8, 41), in which each ligand contains two phosphine coordination sites and two non-coordinating basic sites (amine groups), for electrocatalytic hydrogen production. It was thought that these basic sites, inspired by a similar pendant amine in [FeFe]-hydrogenases, play a very important role to achieve high catalytic activity. A very impressive nickel catalyst (Scheme 8, 42) with only two pendant amine groups was just reported for electrocatalytic proton reduction from ammonium salts with a huge turnover frequency based on catalytic current of 100,000 s1.95 In related work, Artero and co-workers reported a functionalized carbon nanotube electrode with DuBois rst generation nickel catalyst for electrocatalytic hydrogen production.96,97 None of above systems were reported to form H2 directly from water (the addition of other proton sources was needed), but some of them function for proton reduction in the presence a small amount of water (<0.5 M).93 One example utilizing DuBois nickel(II) based catalyst for the light-driven generation of hydrogen from water was reported by McLaughlin, et al.98 Upon visible light irradiation, the system produces hydrogen in the presence of Eosin Y or Ru(bpy)32+ as the photosensitizer and ascorbic acid as the sacricial electron donor in mixed MeCN/H2O (v/v 1 : 1). Up to 2700 TONs (based on the catalyst) were obtained when Eosin Y was the photosensitizer, which was the highest number reported in the literature up to then for a photocatalytic system consisting of noble metal-free materials. 3.2 Water oxidation To the best of our knowledge, there is no homogeneous nickel complex reported in the literature as a water oxidation catalyst. However, nickel based spinels and perovskites, as well as other non-noble metal oxides, have been used in commercial alkaline electrolyzers.99 Alkaline electrolyzers represent a very mature technology that is the current standard for large-scale electrolysis. These systems operate under harsh conditions (high KOH concentration, high pH). Key disadvantages of this technology are its highly caustic electrolyte and its inability to produce hydrogen at high pressures. The latter means that for hydrogen storage under any pressure, an external compressor is needed, adding signicant cost and complexity to the system. A nickel based oxide lm (Ni-WOC) made under mild conditions for catalytic water oxidation was recently reported.100
Energy Environ. Sci., 2012, 5, 60126021 | 6017

Scheme 8

This journal is The Royal Society of Chemistry 2012

View Article Online

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

The lm can be deposited onto ITO or FTO when a potential of 1.3 V vs. NHE was applied in a solution containing 1 mM Ni(NO3)2 and borate buffer at pH 9.2. This deposition strategy is similar to those employed for the formation of a related CoWOC. This Ni-based catalyst evolves oxygen in near-neutral conditions with an activity of 1 mA/cm2 at an overpotential of $425 mV, and exhibits long-term stability in water with no observed corrosion. Faradaic efciency for oxygen evolution, as measured by uorescence detection, is almost quantitative. More importantly, these simple operating conditions may permit energy storage to be performed with devices that are inexpensive and highly manufacturable. A recent publication101 using X-ray absorption spectroscopy technique reveals that an electrodeposited Ni catalyst lm is amorphous and its atomic structure is dominated by extensive dim-oxido bridging between Ni(III)/Ni(IV) ions, but no detectable mono-m-oxido bridging between the redox-active metal ions. Interestingly, this structure is similar to that of an analogous CoWOC lm and colloidal Mn-based catalysts.63,73,102 The structural similarity of Ni-WOC and Co-WOC is also surprisingly shared with the CaMn4O4 complex in Photosystem II.13

Scheme 10

4. Catalysts made of iron for water splitting


4.1 Water reduction Iron probably is the most abundant transition metal in nature and, from a biological point of view, the most important. All three classes of hydrogenases ([FeFe] hydrogenases, [NiFe] hydrogenases, [Fe]-only hydrogenases) contain iron as one of the active site components.84,103 The active site of an [FeFe]hydrogenase is shown in Scheme 9, 43. Hydrogenases catalyze the interconversion 2H+ + 2e $ H2 at rates up to 9000 molecules per second per site. To mimic the function of hydrogenase enzymes, scientists are synthesizing and studying model complexes of [FeFe], [NiFe] and [Fe]-only hydrogenase active sites. Some of these model complexes have been investigated for electrocatalytic proton production,90,104 but to date very few examples have been found for direct water reduction and all at potentials substantially more negative than the thermodynamic value for the equilibrium, which is pH dependent.105 In 2008, Sun and co-workers reported the rst system using model complexes of [FeFe]-hydrogenases for photocatalytic hydrogen production in aqueous solution.106 The structures of the complexes employed are given in Scheme 10, 4446. In combination with Ru(bpy)32+ type photosensitizers and ascorbic acid as the sacricial electron source, all three diiron complexes catalyze hydrogen production in mixed MeCN/H2O solution under visible light irradiation. Complex 45 showed the best activity with 86 TONs based on the photosensitizer and 4.3 TONs based on the catalyst in three hours of photolysis (l > 400

nm). Reductive quenching by ascorbate yielded Ru(bpy)3+ that was followed by electron transfer from to 45 or 46, as conrmed by laser ash photolysis. An HFe(II)Fe(I) intermediate was proposed as the key species for hydrogen production. The same multi-component approach was later applied for hydrogen production based on a cyclometalated iridium(III) photosensitizer, [Ir(ppy)2(bpy)]+, in the presence of TEA and diiron complex 47 in mixed acetone/H2O solution.107 $466 TONs was achieved using this iridium(III) photosensitizer (l > 400 nm), which functioned more active than Ru(bpy)32+ based photosensitizers. Beller and co-workers reported a catalytic system for lightdriven water reduction using simple iron carbonyl complexes.108 These iron carbonyl catalysts include [{CpFe(CO)2}2], [Fe(CO)5], [Fe2(CO)9], [(cot)Fe(CO)3] and [Fe3(CO)12] (Scheme 9, 4852). In this work, [Ir(bpy)(ppy)2]PF6 and TEA were used as the photosensitizer and sacricial electron donor, respectively. Upon irradiation, the system with [Fe3(CO)12] gave the highest activity for hydrogen production with 510 TONs in 6 h. An iron carbonyl hydride was proposed as the key reaction intermediate. The same system was later improved to yield up to 1500 TONs in 24 h and 13.4 incident photon to hydrogen efciency by the addition of different phosphine ligands,109 but their role in the catalysis is not fully understood. 4.2 Water oxidation Although iron is such an abundant and inexpensive element capable of facile oxidation and reduction, there are few reports of using iron materials for water oxidation in a catalytic system. In the literature, iron based complexes, for example, iron(III)-tetraamido macrocyclic ligand complexes (Fe(III)-TAMLs), have been reported to activate both dioxygen and peroxides rapidly.110 Fe(III)-TAMLs are able to split dioxygen to form m-oxo-diiron(IV) complexes, which involves in an equilibrium with a monomeric Fe(IV)-oxo species in water.111 Fe(III)-TAMLs can also be converted to a Fe(V)-oxo complex when oxidized by peracids in water.112 The electrochemical potentials and high reactivity of Fe(III)-TAMLs render them potentially able to catalyze directly water oxidation to form oxygen.113
This journal is The Royal Society of Chemistry 2012

Scheme 9

6018 | Energy Environ. Sci., 2012, 5, 60126021

View Article Online

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

In 2010, Bernhard and co-workers rst demonstrated that FeTAMLs (Scheme 10, 5357) can be used to catalyze efciently water oxidation reaction to generate oxygen. In the presence of excess ceric ammonium nitrate (CAN) in water at pH 0.7, complexes 5457 were examined for water oxidation. The results showed that complexes 5457 were active catalysts at varying rates, whereas 53 exhibited no appreciable activity. The highest TOF > 4680 h1 (TONs 16) was observed from 57 and the oxygen rate has a rst-order dependence on [57]. Oxygen production could also be observed when NaIO4 is used as the oxidant. Mechanistic studies revealed the formation of an Fe(IV)-O-Fe(IV) dimer, which is probably the key intermediate during water oxidation. Recently, the Costas group reported more efcient water oxidation catalysts based on iron with tetradentate nitrogenbased ligands.114 All of these complexes (Scheme 10, 5863) were known from the literature but no experiments had been done previously using them for water oxidation. The authors found that these complexes are all active catalysts in the presence of CAN or NaIO4 as the oxidant and TONs of >1000 were obtained using complex 59, which is the highest reported value for any homogeneous system based on a rst-row transition metal. Interestingly, only the complexes with two accessible coordination sites in cis-relative positions showed activity for water oxidation. The complexes (64, 65), which have only one and two trans available coordination site, respectively, are inactive for water oxidation. Isotope experiments conrmed that the produced oxygen was really from water, and not from the added oxidants. Furthermore, an iron(III) oxohydroxo diferric dimer was excluded as the key intermediate due to its low activity. Other mechanistic studies indicated that an Fe(IV) O species might be the key intermediate for this oxidation reaction, as was observed from the UV-Vis spectra during titration of CAN into iron complex solution in water (lmax 776 nm).

5. Conclusion and perspective


Catalysts are one of the key factors for efcient energy conversion and storage. Great progress has been achieved in the past several decades using materials made of earth abundant elements (Co, Ni and Fe) for water splitting. More than 9000 TONs have been reported for hydrogen production from water using cobaloxime WRCs and up to 1000 TONs has been found for oxygen production using cobalt based WOCs. For nickel based materials, more than 2700 TONs have been reached for hydrogen production and a nickel oxide lm has been developed for water oxidation. Iron is potentially the most interesting element with which to make catalysts for water splitting. In nature, hydrogenase and monooxygenase enzymes115 are two classes of ironcontaining proteins that use the redox capabilities of iron to activate efciently hydrogen and oxygen, respectively. For iron based catalysts, up to 1500 TONs have been reported for hydrogen production from water and more than 1000 TONs have been found for oxygen production from water. For all of the WRCs based on Co, Ni and Fe, metal hydrides are proposed as key reaction intermediates in the mechanisms of action. Some of these intermediates have been detected spectroscopically. In the water oxidation reaction, Co(IV) and Ni(III) species have been proposed as key intermediates during oxygen evolution using Co-WOCs and Ni-WOCs, respectively. While the key
This journal is The Royal Society of Chemistry 2012

intermediates in the water oxidation reaction based on Fe-WOCs are not established, the extensive literature on O2 activation by oxygenase enzymes point to the likely importance of Fe]O species in this reaction. Besides the above achievements in water splitting, it is still a crucial challenge for synthetic chemists and physical chemists to design and identify economically viable molecular and heterogeneous catalysts based on earth abundant elements, of Co, Ni and Fe. It is evident that considerably more work will be done in this area to extend the activity, efciency and robustness of catalysts based on these elements. A number of common problems still remain for most of molecular catalysts for water reduction and water oxidation described above. First, stability is a key concern, with reported lifetimes for molecular catalysts ranging from several minutes to several days during catalysis. The catalysts usually decompose during reaction. The practical lifetime needs to be long, accommodating millions to billions of turnovers. Second, air-sensitivity is another major problem. For WRCs, the reduced forms of these molecular catalysts, especially in organic media, are generally air-sensitive, with atmospheric O2 able to oxidize them rapidly. The ideal water reduction catalyst should promote hydrogen production much faster than oxygenation to avoid decomposition or deactivation pathways. Third, water tolerance is an important concern for many reported H2-generating catalysts. While some reported molecular catalysts work well in organic solvents, it is clear that water needs to be the solvent of choice for water splitting. To date, the efciencies of a number of the molecular catalysts that function well in organic solvents drop rapidly in the presence of even small amounts of water. Fourth, the working potential for electrocatalytic splitting of water is also very important. The ideal working potentials should be as close as possible to the thermodynamic potentials of the two half-reactions for water reduction and water oxidation. If the working potential has too large an overvoltage, some energy is lost in the process (that is, the difference between the working potential and the required minimal potential) and the half reaction loses efciency. How to decrease the overpotentials to the minimum values for water splitting is a continuing challenge for the design of better catalysts. For heterogeneous catalysts, a related problem is efciency. For example, neither cobalt based nor nickel based WOCs have shown better efciency to date than commercial electrolyzers or some homogeneous WOCs. For light-driven half-reactions, however, the concern about overpotential and reaction barriers changes. In the photogeneration of H2, the excited state photosensitizer PS* will undergo electron transfer quenching which can be either by electron transfer from PS* to the catalyst (oxidative quenching) or by electron transfer from the sacricial donor to PS* (reductive quenching). The feasibility of each step is determined by the redox potentials of the catalyst, the sacricial donor and the excited photosensitizer PS*, and the relative rates of the oxidative and reductive quenching paths will be controlled by the particular quenching rate constant and the quencher concentration. In systems such as these, overpotential of the catalyst for generating H2 is not a direct factor. Rather, it helps to determine which photochemical path may be followed. In the research described in this Perspective, we have focused mainly on molecularly-based systems for doing one of the
Energy Environ. Sci., 2012, 5, 60126021 | 6019

View Article Online

half-reactions of water splitting, either proton reduction to H2 or water oxidation to O2. The primary objective in these studies has been to develop the components of each half-reaction to be active, efcient and robust. Actual energy storage in studies dealing with the water splitting half-reactions is not, or should not be, the prime research objective, and in some cases involving light and a sacricial electron donor or acceptor, net energy storage is not even achieved. In fact, for proton reduction to H2, the oxidation of a sacricial donor such as a tertiary amine might provide most of the potential energy required to produce hydrogen, and for water oxidation to O2, electron acceptors such as persulfate are chemically strong enough oxidants so that there is no need for light to drive the reaction. While the consumption of a sacricial donor/acceptor may seem like a major disadvantage for a half-reaction system, it is through these studies of the water splitting half-reactions that actual systems for solar energy storage by water splitting can be developed. Once systems for performing the two half-reactions have been developed sufciently in terms of robustness and activity, they will need to be joined through a conductive membrane. The system will also need to allow protons to move from one half-reaction compartment to the other so that a signicant proton gradient does not occur. We end where we began. Whether water splitting is achieved by electrolysis using electricity generated by photovoltaic devices or dye-sensitized solar cells, or by articial photosynthesis remains at this point unclear. While technology for the former is at a better developed stage at this point, an articial photosynthetic system for water splitting has promise of greater simplicity and efciency. The durability, scalability, cost and efciency will in the long run decide which approach to water splitting and hydrogen generation will win this competition. Either way, however, solar energy must become humankinds principal source of energy for sustainable development. Science plays a key role in this challenge, with the design, synthesis and study of new catalysts, sensitizers and systems based on earth-abundant elements at its heart.

Acknowledgements
R.E. wishes to acknowledge support of research in this area over the years from the U. S. Department of Energy, Ofce of basic Chemical Sciences. We also thank the reviewers of this article for their constructive and detailed comments.

References
1 N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 1572915735. 2 Climate Change 2007: Mitigation of Climate Change, IPCC Working Group III Fourth Assessment Report; Intergovernmental Panel on Climate Change: Geneva, 2007. 3 Energy - Consumption Consumption by fuel, 19652008, Statistical Review of World Energy 2009, British Petroleum. 4 N. Armaroli and V. Balzani, Angew. Chem., Int. Ed., 2007, 46, 5266. 5 A. J. Esswein and D. G. Nocera, Chem. Rev., 2007, 107, 40224047. 6 R. P. F. Gregory, Biochemistry of photosynthesis, Wiley-Interscience, New York, 1977. 7 D. W. Lawlor, Photosynthesis, Springer-Verlag, New York, 2001. 8 R. E. Blankenship, Molecular Mechanisms of Photosynthesis, Blackwell Science, Oxford, U.K., 2002.

9 S. M. Danks, E. H. Evans, P. A. Whittaker, Photosynthetic Systems: Structure, Function, and Assmebly, John Wiley and Sons, New York, 1985. 10 M. Calvin, Science, 1974, 184, 375377. 11 K. N. Ferreira, T. M. Iverson, K. Maghlaoui, J. Barber and S. Iwata, Science, 2004, 303, 18311838. ller, M. Barra, M. Grabolle and 12 M. Haumann, P. Liebisch, C. Mu H. Dau, Science, 2005, 310, 10211019. 13 J. Barber, Inorg. Chem., 2008, 47, 17001710. 14 M. Gr atzel, Nature, 2001, 414, 338344. 15 M. Gr atzel, Inorg. Chem., 2005, 44, 68416851. 16 R. D. Britt, In Oxygenic Photosynthesis:The Light Reactions, Kluwer Academic Publishers, Dordrecht, The Netherlands, 1996. 17 Z. Deng, H.-W. Tseng, R. Zong, D. Wang and R. Thummel, Inorg. Chem., 2008, 47, 18351848. 18 F. Liu, J. J. Concepcion, J. W. Jurss, T. Cardolaccia, J. L. Templeton and T. J. Meyer, Inorg. Chem., 2008, 47, 17271752. 19 I. Romero, M. Rodr guez, C. Sens, J. Mola, M. R. Kollipara, L. Francas, E. Mas-Marza, L. Escriche and A. Llobet, Inorg. Chem., 2008, 47, 18241834. 20 R. Brimblecombe, C. Dismukes, G. F. Swiegersc and L. Spiccia, Dalton Trans., 2009, 93749384. 21 L. Duan, L. Tong, Y. Xu and L. Sun, Energy Environ. Sci., 2011, 4, 32963313. 22 J. P. McEvoy and G. W. Brudvig, Chem. Rev., 2006, 106, 44554483. 23 R. Tagore, R. H. Crabtree and G. W. Brudvig, Inorg. Chem., 2008, 47, 18151823. 24 G. C. Dismukes, R. Brimblecombe, G. A. N. Felton, R. S. Pryadun, J. E. Sheats, L. Spiccia and G. F. Swiegers, Acc. Chem. Res., 2009, 42, 19351943. 25 H. B. Gray, Nature, 2009, 1, 7. 26 V. Artero, M. Chavarot-Kerlidou and M. Fontecave, Angew. Chem., Int. Ed., 2011, 50, 72387266. 27 Y. Lin, G. Yuan, S. Sheehan, S. Zhou and D. Wang, Energy Environ. Sci., 2011, 4, 48624869. 28 G. N. Schrauzer and R. J. Holland, J. Am. Chem. Soc., 1971, 93, 15051506. 29 B. J. Fisher and R. Eisenberg, J. Am. Chem. Soc., 1980, 102, 7361 7363. 30 X. Hu, B. S. Brunschwig and J. C. Peters, J. Am. Chem. Soc., 2007, 129, 89888998. 31 R. M. Kellett and T. G. Spiro, Inorg. Chem., 1985, 24, 23732377. 32 C. V. Krishnan and N. Sutin, J. Am. Chem. Soc., 1981, 103, 2141 2142. 33 J.-M. Lehn and R. Ziessel, Proc. Natl. Acad. Sci. U. S. A., 1982, 79, 701704. lle and M. Gr 34 V. Houlding, T. Geiger, U. Ko atzel, Chem. Commun., 1982, 681683. lle and S. Ohst, Inorg. Chem., 1986, 25, 26892694. 35 U. Ko 36 E. S. Wiedner, J. Y. Yang, W. G. Dougherty, W. S. Kassel, R. M. Bullock, M. Rakowski DuBois and D. L. DuBois, Organometallics, 2010, 29, 53905401. 37 J. L. Dempsey, B. S. Brunschwig, J. R. Winkler and H. B. Gray, Acc. Chem. Res., 2009, 42, 19952004. 38 J. I. Goldsmith, W. R. Hudson, M. S. Lowry, T. H. Anderson and S. Bernhard, J. Am. Chem. Soc., 2005, 127, 75027510. 39 J. Hawecker, J. M. Lehn and R. Ziessel, Nouv J. Chim., 1983, 7, 271 277. 40 P. Connolly and J. H. Espenson, Inorg. Chem., 1986, 25, 26842688. 41 X. Hu, B. M. Cossairt, B. S. Brunschwig, N. S. Lewis and J. C. Peters, Chem. Commun., 2005, 47234725. 42 M. Razavet, V. Artero and M. Fontecave, Inorg. Chem., 2005, 44, 47864795. 43 V. C. A. Baffert and M. Fontecave, Inorg. Chem., 2007, 46, 1817 1824. 44 A. Fihri, V. Artero, M. Razavet, C. Baffert, W. Leibl and M. Fontecave, Angew. Chem., Int. Ed., 2008, 47, 564567. 45 P. Du, K. Knowles and R. Eisenberg, J. Am. Chem. Soc., 2008, 130, 1257612577. 46 D. R. Anton and R. H. O. Crabtree, Organometallics, 1983, 2, 855 859. 47 X. Wang, S. Goeb, Z. Ji, N. A. Pogulaichenko and F. N. Castellano, Inorg. Chem., 2011, 50, 705707. 48 T. Lazarides, T. McCormick, P. Du, G. Luo, B. Lindley and R. Eisenberg, J. Am. Chem. Soc., 2009, 131, 91929194.

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

6020 | Energy Environ. Sci., 2012, 5, 60126021

This journal is The Royal Society of Chemistry 2012

View Article Online


49 T. M. McCormick, B. D. Calitree, A. Orchard, N. D. Kraut, F. V. Bright, M. R. Detty and R. Eisenberg, J. Am. Chem. Soc., 2010, 132, 1548015483. 50 P. Zhang, M. Wang, J. Dong, X. Li, F. Wang, L. Wu and L. Sun, J. Phys. Chem. C, 2010, 114, 1586815874. 51 W. R. McNamara, Z. Han, P. J. Alperin, W. W. Brennessel, P. L. Holland and R. Eisenberg, J. Am. Chem.Soc., 2011, 133, 1536815371. 52 B. D. Stubbert, J. C. Peters and H. B. Gray, J. Am. Chem. Soc., 2011, 133, 1807018073. 53 A. Fihri, V. Artero, A. Pereira and M. Fontecave, Dalton Trans., 2008, 55675569. 54 P. Zhang, M. Wang, C. Li, X. Li, J. Dong and L. Sun, Chem. Commun., 2010, 46, 88068808. 55 S. Jasimuddin, T. Yamada, K. Fukuju, J. Otsuki and K. Sakai, Chem. Commun., 2010, 46, 84668468. 56 F. Lakadamyali and E. Reisner, Chem. Commun., 2011, 47, 1695 1697. 57 V. Y. Sharovich and V. V. Strelets, Nouv. J. Chim., 1978, 2, 199 201. 58 V. Y. Sharovich, N. K. Khannanov and V. V. Strelets, Nouv. J. Chim., 1980, 4, 8184. 59 G. L. Elizarova, L. G. Matvienko, N. V. Lozhkina, V. N. Parmonand and K. I. Zamaraev, React. Kinet. Catal. Lett., 1981, 16, 191194. 60 B. S. Brunschwig, M. H. Chou, C. Creutz, P. Ghosh and N. Sutin, J. Am. Chem. Soc., 1983, 105, 48324833. 61 M. W. Kanan and D. G. Nocera, Science, 2008, 321, 10721075. 62 M. W. Kanan, Y. Surendranath and D. G. Nocera, Chem. Soc. Rev., 2009, 38, 109114. 63 M. W. Kanan, J. Yano, Y. Surendranath, M. Dinca, V. K. Yachandra and D. G. Nocera, J. Am. Chem. Soc., 2010, 132, 1369213701. 64 D. A. Lutterman, Y. Surendranath and D. G. Nocera, J. Am. Chem. Soc., 2009, 131, 38383839. 65 Y. Surendranath, M. Dinca and D. G. Nocera, J. Am. Chem. Soc., 2009, 131, 26152620. 66 Y. Surendranath, M. W. Kanan and D. G. Nocera, J. Am. Chem. Soc., 2010, 132, 1650116509. 67 M. D. Symes, Y. Surendranath, D. A. Lutterman and D. G. Nocera, J. Am. Chem. Soc., 2011, 133, 51745177. 68 D. K. Zhong, J. Sun, H. Inumaru and D. R. Gamelin, J. Am. Chem. Soc., 2009, 131, 60866087. 69 D. K. Zhong and D. R. Gamelin, J. Am. Chem. Soc., 2010, 132, 42024207. 70 D. K. Zhong, S. Choi and D. R. Gamelin, J. Am. Chem. Soc., 2011, 133, 1837018377. 71 J. J. H. Pijpers, M. T. Winkler, Y. Surendranath, T. Buonassisi and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 1005610061. 72 S. Y. Reece, J. A. Hamel, K. Sung, T. D. Jarvi, A. J. Esswein, J. J. H. Pijpers and D. G. Nocera, Science, 2011, 334, 645648. 73 M. Risch, V. Khare, I. Zaharieva, L. Gerencser, P. Chernev and H. Dau, J. Am. Chem. Soc., 2009, 131, 69366937. 74 J. G. McAlpin, Y. Surendranath, M. Dinca, T. A. Stich, S. A. Stoian, W. H. Casey, D. G. Nocera and R. D. Britt, J. Am. Chem. Soc., 2010, 132, 68826883. 75 J. B. Gerken, J. G. McAlpin, J. Y. C. Chen, M. L. Rigsby, W. H. Casey, R. D. Britt and S. S. Stahl, J. Am. Chem. Soc., 2011, 133, 1443114442. 76 A. Harriman, I. J. Pickering, J. M. Thomas and P. A. Christensen, J. Chem. Soc., Faraday Trans. 1, 1988, 84, 27952806. 77 F. Jiao and H. Frei, Angew. Chem., Int. Ed., 2009, 48, 18411844. 78 Y. Liang, Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier and H. Dai, Nat. Mater., 2011, 10, 780786. 79 F. Jiao and H. Frei, Energy Environ. Sci., 2010, 3, 10181027. 80 Q. Yin, J. M. Tan, C. Besson, Y. V. Geletii, D. G. Musaev, A. E. Kuznetsov, Z. Luo, K. I. Hardcastle and C. L. Hill, Science, 2010, 328, 342345. 81 Z. Huang, Z. Luo, Y. V. Geletii, J. W. Vickers, Q. Yin, D. Wu, Y. Hou, Y. Ding, J. Song, D. G. Musaev, C. L. Hill and T. Lian, J. Am. Chem. Soc., 2011, 133. 82 J. J. Stracke and R. G. Finke, J. Am. Chem. Soc., 2011, 133, 14872 14875. 83 D. J. Wasylenko, C. Ganesamoorthy, J. Borau-Garcia and C. P. Berlinguette, Chem. Commun., 2011, 47, 42494251. 84 J. C. Fontecilla-Camps, A. Volbeda, C. Cavazza and Y. Nicolet, Chem. Rev., 2007, 107, 42734303. 85 A. Volbeda, M. H. Charon, C. Piras, E. C. Hatchikian, M. Frey and J. C. Fontecilla-Camps, Nature, 1995, 373, 580587. 86 Y. Higuchi, T. Yagi and N. Yasuoka, Structure, 1997, 5, 16711680. 87 Y. Higuchi, H. Ogata, K. Miki, N. Yasuoka and T. Yagi, Structure, 1999, 7, 549556. 88 Z. Li, Y. Ohki and K. Tatsumi, J. Am. Chem. Soc., 2005, 127, 8950 8951. 89 B. E. Barton, C. M. Whaley, T. B. Rauchfuss and D. L. Gray, J. Am. Chem. Soc., 2009, 131, 69426943. 90 C. Tard and C. J. Pickett, Chem. Rev., 2009, 109, 22452274. 91 J.-P. Collin, A. Jouaiti and J.-P. Sauvage, Inorg. Chem., 1988, 27, 19861990. 92 A. D. Wilson, R. H. Newell, M. J. McNevin, J. T. Muckerman, M. R. DuBois and D. L. DuBois, J. Am. Chem. Soc., 2006, 128, 358366. 93 U. J. Kilgore, J. A. S. Roberts, D. H. Pool, A. M. Appel, M. P. Stewart, M. R. DuBois, W. G. Dougherty, W. S. Kassel, R. M. Bullock and D. L. DuBois, J. Am. Chem. Soc., 2011, 133, 58615872. 94 M. R. Dubois and D. L. Dubois, Acc. Chem. Res., 2009, 42, 1974 1982. 95 M. L. Helm, M. P. Stewart, R. M. Bullock, M. R. DuBois and D. L. DuBois, Science, 2011, 333, 863866. 96 A. L. Goff, V. Artero, B. Jousselme, P. D. Tran, N. Guillet, R. M etay e, A. Fihri, S. Palacin and M. Fontecave, Science, 2009, 326, 13841387. 97 P. D. Tran, A. L. Goff, J. Heidkamp, B. Jousselme, N. Guillet, S. Palacin, H. Dau, M. Fontecave and V. Artero, Angew. Chem., Int. Ed., 2011, 50, 13711374. 98 M. P. McLaughlin, T. M. McCormick, R. Eisenberg and P. L. Holland, Chem. Commun., 2011, 47, 79897991. 99 R. N. Singh, J. P. Pandey, B. L. N. K. Singh, P. Chartier and J. F. Koenig, Electrochim. Acta, 2000, 45, 19111919. 100 M. Dinc a, Y. Surendranath and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 1033710341. 101 M. Risch, K. Klingan, J. Heidkamp, D. Ehrenberg, P. Chernev, I. Zaharieva and H. Dau, Chem. Commun., 2011, 47, 1191211914. 102 I. Zaharieva, M. M. Najafpour, M. Wiechert and M. Haumann, Energy Environ. Sci., 2011, 4, 24002408. 103 S. Shima, O. Pilak, S. Vogt, M. Schick, M. S. Stagni, W. MeyerKlaucke, E. Warkentin, R. K. Thauer and U. Ermler, Science, 2008, 321, 572575. 104 J. W. Tye, J. Lee, H.-W. Wang, R. Mejia-Rodriguez, J. H. Reibenspies, M. B. Hall and M. Y. Darensbourg, Inorg. Chem., 2005, 44, 55505552.  105 L. Sun, B. Akermark and S. Ott, Coord. Chem. Rev., 2005, 249, 16531663.  106 Y. Na, M. Wang, J. Pan, P. Zhang, B. Akermark and L. Sun, Inorg. Chem., 2008, 47, 28052810. 107 P. Zhang, M. Wang, Y. Na, X. Li, Y. Jiang and L. Sun, Dalton Trans., 2010, 39, 12041206. 108 F. G artner, B. Sundararaju, A.-E. Surkus, A. Boddien, B. Loges, H. Junge, P. H. Dixneuf and M. Beller, Angew. Chem., Int. Ed., 2009, 48, 99629965. 109 F. G artner, A. Boddien, E. Barsch, K. Fumino, S. Losse, H. Junge, ckner, R. Ludwig and M. Beller, Chem.Eur. D. Hollmann, A. Bru J., 2011, 17, 64256436. 110 A. D. Ryabov and T. J. Collins, Adv. Inorg. Chem., 2009, 61, 471 521. 111 A. Ghosh, F. Tiago de Oliveira, T. Yano, T. Nishioka, E. S. Beach, nck, A. D. Ryabov, C. P. Horwitz and I. Kinoshita, E. Mu T. J. Collins, J. Am. Chem. Soc., 2005, 127, 25052513. 112 F. Tiago de Oliveira, A. Chanda, D. Banerjee, X. Shan, S. Mondal, nck and T. J. Collins, Science, L. Que, Jr., E. L. Bominaar, E. Mu 2007, 315, 835838. 113 A. Chanda, X. Shan, M. Chakrabarti, W. C. Ellis, D. Popescu, F. Tiagode Oliveria, D. Wang, L. Que, Jr., T. J. Collins, nck and E. L. Bominaar, Inorg. Chem., 2008, 47, 3669 E. Mu 3678. mez, J. J. Pla and 114 J. L. Fillol, Z. Codol a, I. Garcia-Bosch, L. Go M. Costas, Nat. Chem., 2011, 3, 807813. 115 S. Harayama, M. Kok and E. L. Neidle, Annu. Rev. Microbiol., 1992, 46, 565601.

Downloaded by Ewha Womens University on 05/04/2013 09:51:35. Published on 06 January 2012 on http://pubs.rsc.org | doi:10.1039/C2EE03250C

This journal is The Royal Society of Chemistry 2012

Energy Environ. Sci., 2012, 5, 60126021 | 6021

You might also like