Anda di halaman 1dari 10

ChemComm

FEATURE ARTICLE
View Article Online
View Journal | View Issue

Cite this: Chem. Commun., 2013, 49, 218

Homogeneous water oxidation catalysts containing a single metal site


Derek J. Wasylenko, Ryan D. Palmer and Curtis P. Berlinguette*
The recent recognition that a single metal site is capable of mediating the multiple electron and proton transfer events associated with water oxidation represents a pivotal discovery for the field. This finding

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

Received 3rd August 2012, Accepted 12th October 2012 DOI: 10.1039/c2cc35632e www.rsc.org/chemcomm

has led to a remarkable expansion of known synthetic water oxidation catalysts, and has provided the means to gain unprecedented insight into the reaction steps involved with OO bond formation. This perspective reflects on the key studies that have advanced our understanding of water oxidation catalysis while summarizing molecular features that are integral to negotiating this complicated reaction pathway with the goal of helping identify new frontiers of discovery for the field.

1. Introduction
In nature, an inorganic oxygen evolution catalyst (OEC) within photosystem II is capable of mediating the formation of dioxygen from water at approximately the thermodynamic limit.13 Replicating the reactivity of the OEC represents a central scientific challenge for moving society towards a non-carbonaceous energy economy because one of the primary obstacles for splitting water into hydrogen fuels which is often invoked in renewable energy storage schemes46 is the water oxidation half-reaction (eqn (1)).7 This scenario has provided the impetus to develop molecular mimics of the OEC to gain a better fundamental understanding of the reaction pathways in pursuit of commercially viable catalysts capable of eciently mediating OO bond formation.811 2H2O - 4H+ + O2 + 4e (1)

The multi-metallic CaMn4O5 core of the OEC1,2 has long enforced the idea that molecular catalysts must also have multiple metal centers to accommodate the accumulation of the four oxidizing equivalents needed to oxidize water.8 The first confirmed synthetic water oxidation catalyst (WOC) was the so-called blue dimer, [(bpy)2(OH2)RuIIIORuIII(OH2)(bpy)2]4+, reported by Meyer et al. in 1982 (Fig. 1).12 Electronic and stability issues associated with this system have complicated mechanistic studies, however, leaving some of the critical reaction steps open to interpretation.1316 Unfortunately, the list of water oxidation catalysts reported over the next quarter-century was limited to roughly a handful of systems (Fig. 1).8 Documented examples of water oxidation catalysts included a series of homometallic
Department of Chemistry and Centre for Advanced Solar Materials, University of Calgary, 2500 University Drive N.W., Calgary, Canada T2N 1N4. E-mail: cberling@ucalgary.ca

systems each containing no less than two ruthenium1719 or manganese2022 centers. The notion that a single metal site could accommodate the four-electron and four-proton steps associated with splitting water was one that gained traction in 2008,2325 and ushered in an exciting new era of discovery for the field of homogeneous water oxidation catalysis.26 Not only are there now hundreds of WOCs in the literature, there is now a far clearer understanding of the mechanistic details surrounding water oxidation.25 (The coincident emergence of WOCs imbedded within polyoxometallate (POM) assemblies27,28 was also an important milestone for the field, but will not be detailed here.) This perspective will attempt to summarize the remarkably rapid progression of the field by outlining some of the key findings made in recent times. (Comprehensive accounts of the field can be found in two excellent recently published reviews.29,30) We will then seek to reduce this expansive body of work down to the key molecular design elements that appear to be critical for making eective WOCs, and will conclude by identifying some of the key questions remaining in the field, at least from our perspective.

2. Single-site water oxidation catalysts


The concept of a single metal site mediating the water splitting reaction represented a paradigm shift for the field. Recognition of this feature quickly expanded the catalog of known water oxidation catalysts, and has provided the opportunity to probe reaction pathways using much simpler platforms thus oering unprecedented resolution into the catalytic cycle. This section will summarize some key single-site catalysts that have shed light on the catalytic cycle.
This journal is
c

218

Chem. Commun., 2013, 49, 218--227

The Royal Society of Chemistry 2013

View Article Online

Feature Article

ChemComm

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

Fig. 1 Chronological listing of representative WOCs reported over the past three decades.1722

2.1

Ruthenium-based WOCs

The earliest documented case of a single ruthenium center being a WOC was a 1990 report outlining the Ce(IV)-driven oxidation of water catalyzed by trans-[RuIII(H-dmg)XY]n+ (H-dmg = dimethylglyoximato; X, Y = Cl or ClO4; and X = ClO4 and Y = imidazole or 2-methylimidazole).31 This was a preliminary finding, however, as certain claims were not validated by experiment (e.g. a Ru(VII) center was conjectured to be the active site, but the Ru(VII)/Ru(VI) couple of B0.811.03 V vs. NHE as defined in the paper is not appropriately matched to the thermodynamic potential needed to drive water oxidation at the pH levels tested). A 2000 study by Yagi et al. provided evidence suggesting that cis[RuIII(NH3)4Cl2]+ may act to split water in a unimolecular fashion because the catalyst was immobilized within a proton-conducting Nafion film along with some supporting kinetics data, but more comprehensive experiments qualifying the large amount of N2 evolved in the absence of decomposition were never provided.32 The first claim of water oxidation catalysis by a ruthenium complex that was coordinated by polypyridyl ligands was made in a 1997 study of [RuII(tpy)(pap)(OH2)]2+ (tpy = 2,2 0 ;6 0 ,200 terpyridine; pap = 2-(phenylazo)pyridine).33 While there remain open questions surrounding the thermodynamic feasibility of the assigned redox couples with respect to water oxidation, this system may be the first documentated example of a single-site ruthenium-based WOC. Notwithstanding, a 2005 study of a series of ruthenium complexes bearing oxidatively stable polypyridyl ligands (Fig. 2) by Zhong and Thummel was arguably the first credible paper showing that a single metal site could be a WOC.19 Mechanistic studies were not carried out to determine the order in [Ru], however, and thus it remained an open question of whether the complexes could split water in a unimolecular fashion. They later elaborated on their initial report by preparing an extensive series of complexes bearing pyridyl derivatives of diering denticities (Fig. 2).24 The scope
This journal is
c

Fig. 2 Representative examples of documented ruthenium-based WOCs.

of complexes in this paper capable of producing dioxygen provided irrefutable evidence that a single metal site could indeed mediate water oxidation catalysis. Almost coincident with this Thummel paper published in late 2008 was the emergence of a tremendously influential paper by Meyer et al. that disseminated not only two new single-site WOCs (Fig. 2), but, more importantly, provided a plausible mechanism for water oxidation by a single metal site.25 Using spectrochemical techniques, they were able to elucidate that a high valent [RuV QO]3+ species was reactive towards water and could form a peroxo complex prior to the liberation of dioxygen. This water nucleophilic attack (WNA) mechanism (also called acidbase mechanism) detailed how a single metal site could negotiate four-electron-transfer steps while requiring the metal to pass through four dierent redox levels. While the details of this cycle will be provided in the next section, this finding brought clarity to the OO bond formation step for the very first time for the field at a single metal centre.
Chem. Commun., 2013, 49, 218--227 219

The Royal Society of Chemistry 2013

View Article Online

ChemComm We read these two 2008 papers by Thummel and Meyer with considerable interest at the time because we had spent the year validating dioxygen evolution from [RuII(tpy)(bpy)(OH2)]2+ (and related dimers34),35 and it was stated in both of these papers that this specific complex was not a catalyst for water oxidation. It therefore brought some comfort to us when Masaoka and Sakai would subsequently show evidence that [RuII(tpy)(bpy)(OH2)]2+ was a WOC closely aligned with our results.36 In the meantime, and because it was still a contentious issue, we developed a series of complexes where electron donating and electron withdrawing groups were placed about the periphery of the [RuII(tpy)(bpy)(OH2)]2+ platform (Fig. 2) to show that the catalytic activity and TONs correlates to electron density at the metal oxo unit to unequivocally prove that [RuII(tpy)(bpy)(OH2)]2+ was indeed a WOC.35 The debate would soon lose significance, however, as a myriad of ruthenium systems were disclosed by numerous groups thus bringing incontrovertible evidence that a single metal center was enough to catalyze water oxidation. In early 2009, Sun, Llobet and coworkers reported complex [Ru(4-picoline)2(2,20 -bipyridine-6,6 0 -dicarboxylic acid]37 (Fig. 2) and a related derivative that were distinctive because they contained an anionic chelate that could make higher oxidation levels more accessible.38 They demonstrated that the complex could catalyze water oxidation when driven by either Ce(IV), a terminal oxidant widely used in the field, or by a photo-initiated system using a sensitizer such as [Ru(bpy)3]2+ and S2O82 or [CoIII(NH3)5Cl]2+ as sacrificial electron scavengers.39 The latter process was the first example of water oxidation driven by light using a single-site catalyst (noting that the Ru-based WOC still acts merely as an electrocatalyst). Subsequent studies on this same family of compounds by Sun, Privalov and co-workers has not only revealed a distinctive radical coupling reaction pathway (vide infra), but also remarkably high TOF of B300 s1 the highest value reported to date to our knowledge.40 They have also explored the use of these WOCs in photoelectrochemical devices where the catalyst is deposited in a Nafion membrane on a TiO2 substrate bearing a photosensitizer,41 and an electrocatalytic device where the WOC is immobilized in carbon nanotubes.42 Another class of single-site ruthenium WOCs consists of systems where the catalytically active metal is embedded within an oxidatively stable POM. The strategy was recently outlined by Fukuzumi and coworkers, complete with a full mechanistic study, for two heteroundecatungstates, [RuIII(OH2)XW11O39]5 (where X = Ge or Si).43 This approach overcomes the problems inherent to conventional complexes where the organic ligands are susceptible to deactivation or decomposition, as POMs are composed of mineral-like silicates or tungstates that are essentially impervious to oxidative degradation. Support for this claim was provided by quantitative yields of O2 being produced based on the amount of oxidant (20 or 50 TON) added to the reaction mixture. It should be noted, however, that many conventional WOCs exhibit commensurate TONs without displaying a loss in activity, but this study did not include a higher excess of oxidant to probe the relative stabilities of the POM complexes. The POM complexes were also found to go through the identical acidbase mechanism despite the dramatic change in ligand architecture.
220 Chem. Commun., 2013, 49, 218--227

Feature Article

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

Fig. 3 A listing of selected iridium-based homogeneous water oxidation catalysts.

2.2

Iridium-based WOCs

Another key paper that appeared in 2008 was one authored by Bernhard et al. showing that complexes consisting of a single iridium center could also act as eective WOCs.23 This seminal report revealed that a family of cyclometalated iridium complexes bearing two substituted 2-phenylpyridine (ppy) ligands with two solvent water molecules could mediate dioxygen formation (Fig. 3). This report triggered the emergence of numerous molecular iridium WOCs. The most extensively studied iridum-based WOCs to date are the half-sandwich complexes bearing cyclopentadiene (Cp) or pentamethylcyclopentadiene (Cp*) ligands a series of compounds first reported in 2009 by Crabtree and Brudvig (Fig. 3).44 They demonstrated that three new complexes of the form [IrIII(Cp*)(N4C)X], where the cyclometalated ligands were either ppy ligands or phenylpyrimidine (ppm) and X was either Cl or CF3SO3 (triflate; OTf), were fast and robust catalysts when using Ce(IV) as the chemical oxidant. Macchioni and coworkers45 would show that an aquated version of the Ir half-sandwich complexes (Fig. 3) exhibited faster catalytic rates than those bearing a second chelating ligand,45 a finding that was supported by a subsequent report by Crabtree and Brudvig.46 This same study showed that the complexes could rapidly lose ligands, such as tmeda (tetramethylethylenediamine), PPh3 and CO to form [Ir(Cp*)(H2O)3]2+ in situ. To overcome ligand loss, Crabtree and Brudvig reported a homogenous Cp*Ir complex with a strongly s-donating N-heterocyclic carbene (NHC) ligand, [Ir(Cp*)(NHC)Cl]+, that could stabilize an Ir(IV) intermediate to an extent that it could be observed by EPR spectroscopy.47 Bernhard and Albrecht utilized mesoionic carbene ligands that could facilitate the facile formation of high-valent states a strategy that resulted in higher activities than most ruthenium-based catalysts while also providing a new benchmark for TONs: 10 000 after 5 days.48 Hetterscheid and Reek showed that a series of iridium complexes with NHC ligands to be one of the first WOCs that could exhibit both high TOF (1.5 s1) and TON (>2000).49 2.3 WOCs containing a 3d transition metal

Evidence of water oxidation catalyzed by a manganese complex was provided by Sun and coworkers in 2007, where a xanthene backbone functionalized with mono- and bis-corrole units was used to bind the metal.50 Water oxidation experiments were
This journal is
c

The Royal Society of Chemistry 2013

View Article Online

Feature Article

ChemComm

Fig. 4 Single-site first-row transition metal complexes capable of catalytically oxidizing water.

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

conducted by controlled-potential electrolysis in a 3 : 2 CH3CN/ CH2Cl2 solution containing 30 mL of an aqueous 10% nBu4NOH solution, where oxygen was observed via electrochemical detection at B0.75 V after scanning to B1.5 V. In the absence of mechanistic details, it was not clear whether the monometallic complex acts in a unimolecular fashion, especially because a related bimetallic Mn complex exhibits much higher activity. A subsequent paper by the same authors investigated the OO bond forming reaction between a high- valent Mn(V)QO with hydroxide and found it was not catalytic.51 In 2010, Bernhard, Collins and coworkers reported a series of Fe-based WOCs bearing a tetra-amido macrocyclic ligand (TAML) (Fig. 4), a ligand with proven utility in harshly oxidizing conditions.52 While many of the complexes studied were highly active (e.g. TOF >1.3 s1) in the early stages of the reaction, deactivation of the catalysts occurred in t o 20 s (dioxygen continues to form at a very slow rate, but this may originate from the hydrolytic decomposition of the oxidant). A more recent report by Costas et al. contained a suite of six related iron-based WOCs bearing tetradentate ligands that situated the vacant coordination sites in a cis arrangement (Fig. 4).53 Related iron complexes with the coordination sites positioned trans to each other as well as one containing a pentatdentate ligand were not deemed to be WOCs prompting the authors to propose a mechanism where the incoming water substrate attacks the high-valent [Fe(V)QO] unit with concomitant stabilization and deprotonation by the adjacent water molecule situated at the cis position. These studies were carried out in acidic media, therefore precluding the nucleophilic attack by a hydroxide substrate. Our program reported the first example of a WOC containing a single cobalt center (Fig. 4).54 We had tried a number of dierent chelating ligands, but found that the vast majority of ligands were incapable of preventing decomposition into catalytically active CoOx films or nanoparticles. Our systematic analysis of the ruthenium WOCs provided a clue that a primary decomposition pathway was the labilization of the bond trans to the active MO vector. Building on this information, we inserted a cobalt(II) center into the tetrapodal pentadentate ligand, 2,6-(bis(bis-2-pyridyl)methoxymethane)pyridine (PY5),55,56 and found that it acted as a remarkably active WOC. It was also a rare example of welldefined PCET at a single MOH2 site where M is a first row transition metal.57 Nocera and coworkers subsequently reported a series of cobalt-corrole WOCs the best of which
This journal is
c

being a perfluorinated hangman complex (Fig. 4).58 Finally, a copper-bipyridine complex formed by a self assembly reaction was recently reported by Mayer and Goldberg to catalytically oxidize water at one of the highest rates reported to date.59

3. Mechanistic details
The details surrounding the reaction steps needed to bring two water molecules together to form an OO bond remained a source of considerable debate until the advent of single-site WOCs. Complicating the issue was the lack of synthetically accessible WOCs available to researchers prior to the finding that a single metal site could mediate water oxidation, and those that were known were rife with issues that complicated their study (e.g., solubility, stability). Consequently, the vast majority of studies have hitherto been confined to the blue dimer, a system that is still, in large part, much more complicated to study than many of the WOCs presented in the previous section. The catalysts now available to researchers has led to an accumulation of knowledge that has brought clarity to the reaction details associated with water oxidation. There are two generally accepted mechanisms for the key OO bond formation step: (i) an acidbase or water nucleophilic attack (WNA) pathway, where the substrate attacks an oxo ligand bound to a high valent metal;25,26,6062 and (ii) a radical coupling pathway that involves the coupling of two metaloxo units bearing radical character (Scheme 1).6365 The WNA pathway requires a metaloxo fragment that is suciently electrophilic to undergo attack by a nucleophilic substrate (e.g., H2O, OH) to form the requisite OO bond. The highest-occupied molecular

Scheme 1 Proposed OO bond forming mechanisms at high-valent metaloxo complexes.

The Royal Society of Chemistry 2013

Chem. Commun., 2013, 49, 218--227

221

View Article Online

ChemComm orbital (HOMO) of water (or hydroxide) is of s symmetry and interacts with the lowest-unoccupied molecular orbital (LUMO) of the pseudo-octahedral metaloxo complex that is of dp* character. This interaction results in the formation of an OO s bond with the concomitant cleavage of an MO p bond leading to the formal two-electron reduction of the metal.13 Alternatively, a bimolecular RC mechanism is possible in cases where the metaloxo bears significant radical character (i.e. [RuIVQO ]) and the steric bulk of the ligands is accommodating to such an interaction. In this case, a singly occupied molecular orbital (SOMO) with dp* character couple and form a [MOOM] complex that may then undergo further oxidative transformations to release O2. (It has been argued that the electronic structure of certain Ru(V)QO fragments are better described by the resononance partner [RuIVQO ].64) Other mechanisms have been invoked, such as one that involves an expanded coordination sphere,24,53 but they will not be addressed in detail here. 3.1 Water nucleophilic attack (WNA) pathways

Feature Article This study showed that a [RuOH2] complex can undergo consecutive PCET steps to aord a [RuIV QO]2+ complex at relatively low potentials, thereby allowing access to the high-valent [RuV QO]3+ complex that can undergo nucleophilic attack by a solvent water molecule to form a [RuIIIOOH]2+ (the high energy intermediate, [RuIIIO2H2]3+, is bypassed due to spontaneous deprotonation by water; Scheme 2). In our view, the most defining feature of this study was the perceptive observation of the spectral changes that occurred after adding one equiv. of Ce(IV) to the [RuIV QO]2+ species: the two resultant temporal domains could then be assigned as the sequential formation of [RuV QO]3+ and [RuIIIOOH]2+. Following this step, another series of electron- and proton-transfer steps aords the [RuIVOO]2+ complex that spontaneously loses dioxygen upon the rearrangement of the p-bound [RuIVZ2-O2]2+ form the bent, end-on [RuIVZ1-O2]2+ with subsequent coordination of a water molecule. The liberation of oxygen was assigned as the ratedetermining step for a series of ruthenium complexes under certain conditions, but OO bond formation appears to be the slow step in the majority of cases.26,62,66,67 Experiments carried out in concentrated acids can enhance the oxidizing power of Ce(IV) and enhance reactivity, but the RDS can actually be the further oxidation of the [RuIVOO]2+ to [RuVOO]3+ that will, in

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

The seminal 2008 report by Meyer et al. detailed all of the key reaction steps for a catalytic water oxidation cycle for the first time.25

Scheme 2 Summary of WNA-type mechanisms for single-site catalysts based on ruthenium, iridium, and cobalt.

222

Chem. Commun., 2013, 49, 218--227

This journal is

The Royal Society of Chemistry 2013

View Article Online

Feature Article turn, rapidly undergo reductive elimination of O2 to form [RuIIIOH]2+ closing a slightly dierent catalytic cycle (Scheme 2). The proposed mechanism was also investigated independently by DFT methods with results that largely supported the so-called acidbase mechanism.61 We were also able to identify some of the key intermediates including those late in the cycle by massspectrometric techniques.62 There are some intriguing variabilities to this reaction pathway that could be extracted because of the basic structure of the catalyst. One of the key findings determined independently by us62,67 and the Meyer group68 was that the presence of a base can accelerate the OO bond formation process by bypassing high energy [RuIIIO2H2]3+ intermediate via a concerted baseassisted atom-proton-transfer (APT) process. Meyer demonstrated this eect by showing that the addition of bases, such as acetate, accelerated reaction rates.68 We later reached the same conclusion after documenting much faster reaction rates in HNO3 relative to other commonly used acids (e.g., triflic acid, perchloric acid).67 This finding was surprising because there was a long-standing notion in the field that cerium is a stronger oxidant in stronger acids and thus the corresponding reactivity should be faster in, for example, HClO4 than HNO3.13,16 We therefore rationalized that because the NO3 anion was a stronger base than the other acid anions and water, it therefore assisted the APT pathway. It should be cautioned, however, that we also found the NO3 anion not to be an innocent anion: an O-atom derived from NO3 can be incorporated into the final dioxygen product (albeit to a small extent).62 Another intriguing observation we made was that the [RuIVQO]2+ site can be made suciently electrophilic to react with water,62 a feature later corroborated by Thummel and Fujita for a related system.69 Similar reactivity has been documented for WOCs containing other transition metals. Mechanistic studies on iridium-based catalysts, for example, have shown that the metal typically passes through the III/IV/V redox levels, where the latter [IrVQO]3+ motif is ostensibly the species that is reactive towards water.44,46 The OO bond formation step is invoked as the RDS in most cases, but the oxidation of the [IrIVOH]2+ intermediate has also been documented to be the slowest step when, for example, low concentrations of the oxidant are used.46 Mechanistic studies of the few iron complexes in the literature invoke the generation of a [FeVQO]3+ moiety70 that is poised for nucleophilic attack by the substrate.53 Our study of cobalt-based catalyst in basic media suggests a catalytic cycle where the complex undergoes a well-defined PCET step to furnish [CoIIIOH]2+, which undergoes a second oxidation event to form a [CoIVOH]3+ intermediate (Scheme 2).54 This species is characterized by a large LUMO situated primarily on the O-atom that is poised for nucleophilic attack by an incoming water and/or hydroxide substrate.71 Using the well-defined reactivity of single-site Ru catalysts as a guide, we postulated that the presence of a base assists reactivity by bypassing a peroxo species after the OO bond formation step, and that subsequent electrontransfer and deprotonation steps presumably furnish a [CoIII OO]2+ complex. The exclusion of dioxygen from this species and subsequent substrate binding replenishes the starting [CoIIOH2]2+ species to close the catalytic cycle. At this early stage,
This journal is
c

ChemComm however, we do not rule out other resonance forms of the CoO fragment being the catalytically species as a means of suppressing the electrostatic repulsion between the substrate and the catalyst; studies are underway to examine these possibilities. 3.2 Radical coupling pathways

The elaborate work of Sun, Privalov and coworkers over the past few years has unravelled a bimolecular pathway consonant with the radical coupling mechanism (Scheme 2). The distinguishing features of their catalytic cycle is the formation of the [RuIVO ]3+ oxyl radical,64 a resonance form of the aforementioned [RuVQO]3+. Two of these units couple to yield a [RuIV (O2)RuIV] complex that subsequently liberates dioxygen (note that we still classify these unimolecular complexes here as single-site catalysts despite the bimolecular pathway). Support for this pathway was provided by a rate law that was secondorder in [Ru], a reasonably small calculated activation energy of 12 kcal mol1 for the radical coupling step, as well as the isolation of a [RuIVOH]2 dimer characterized by single crystal X-ray crystallography.37 A hallmark of complexes that have been shown to proceed through this pathway is an expansion of the primary coordination sphere. Consequently, these systems require a ligand with a suciently large bite angle (e.g., 1231 for 2,2 0 -bipyridine-6,6 0 -dicarboxylate ligand) and enough flexibility to accommodate seamless OO bond formation. A more rigid ligand backbone (i.e., 1,10-phenanthroline-2,9-dicarboxylate), for example, is unable to accommodate reorganizational energies associated with various oxidation steps thus forcing the complex to pass through the unimolecular WNA pathway.63 They have also shown that the chelating axial ligands can participate in intermolecular pp interactions that serve to accelerate reactivity. Very recently, Sun and coworkers reported a derivative that displays remarkably high TOF of B300 s1 that are comparable to the performance of the OEC in PSII.40 3.3 Other issues relevant to mechanistic studies

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

Experimental conditions play a significant role in governing the reaction pathways and catalytic output. Indeed, the lack of coherence in the field can often lead to misleading or conflicting claims. We therefore set out to try to establish some guidelines surrounding the cerium-driven oxidation of ruthenium single site catalysts,67 but we came to appreciate that establishing a standard set of reaction conditions for all WOCs is a huge challenge due to solubility problems and a host of other issues inherent to each particular catalyst or oxidant. For example, Ce(IV) is only stable in an acidic medium, yet some catalysts are operative only in basic media. We also found, as mentioned above, that reactivity can be aected by the acid medium, and that the anion of CAN, a widely used terminal oxidant, can be activated during catalysis and incorporated into the dioxygen product thus complicating mechanistic studies.62 We maintain that when possible, Ce(IV) salts should at the very least be dissolved in a strongly acidic medium (and preferably not HNO3) prior to being introduced to the solution containing the catalyst to avoid hydrolysis and potentially misleading OO bond formation steps.
Chem. Commun., 2013, 49, 218--227 223

The Royal Society of Chemistry 2013

View Article Online

ChemComm There is also the lingering issue of whether the true source of catalysis is the homogeneous complex of interest versus a metaloxide solid or nanoparticle present on the electrode or in solution.72,73 There seems to be some consensus that the majority of reported ruthenium WOCs are authentic homogeneous catalysts given their reasonably high stability under oxidizing conditions, and the large volume of self-consistent data correlating structure to reactivity.35 Catalysts with all other metals reported to date, however, have been the subject of much controversy given that they are more susceptible to decomposition to highly active solids. For example, the stability of some iridium complexes has been called into question due to underlying concerns of IrOx nanoparticles or films are the true source of catalytic behaviour.7477 Crabtree and Brudvig have used a electrochemical quartz crystal nanobalance to show that some complexes do decompose under anodic potentials (>1 V vs. NHE at pH 6) to deposit a highly active and robust film of IrOx on a variety of anode substrates,78 yet others retain their structural integrity for meaningful time periods. This important issue is one that has unfortunately consumed much of our time when studying WOCs containing first-row transition metals71 and is worthy of an entire review,73 and therefore will not be discussed in detail here. We cannot emphasize enough, however, that experimental conditions do matter and there must remain some consistency on this front if studies are going to provide anything of substance for the community.67

Feature Article

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

Fig. 5 Qualitative metaloxo molecular orbital diagram in a pseudo-octahedral environment around the metal assuming C4v symmetry (derived from ref. 81).

platforms under the harshly oxidizing conditions associated with these experiments. 4.3 A reactive, high-valent metaloxo unit

4. Design elements of a homogeneous WOC


This extensive catalog of WOCs oers the opportunity to assess the structural and electronic properties of transition metal complexes that are integral to successfully generating dioxygen from water. The preeminent features of these systems are described in sequence below. 4.1 Metal with multiple, accessible redox levels

All WOCs reported to date contain a metal that passes through at least three dierent oxidation levels. It is currently dicult to envisage a cycle where a metal undergoes a less extreme redox variance in the absence of a redox-active ligand scaold that participates in the reaction or a neighboring metal. Notwithstanding, all of the generally accepted mechanisms to date confine reactivity primarily to the metal and oxygen atoms of interest. It is also worth noting that the metal must be at least tetravalent to react with the substrate, although the most common catalysts proceed through a pentavalent species. 4.2 Oxidatively stable ligand environment

The vast majority of the complexes contain oxidatively stable polypyridyl, carbene, and/or Cp ligands. An alternative approach is to use POMs as a support because they already exist in a fully oxidized form; however, one must still be careful that the catalytically active metal is not leached from the POM assembly72 (a cautionary tale that exists for WOCs of all types). Nonetheless, ligands that would have been viewed with scepticism (e.g., contain saturated carbon atoms53,79) a few years ago have been shown to be perfectly viable ligand
224 Chem. Commun., 2013, 49, 218--227

Imperative to the design of WOCs is an electronic structure that renders a reactive metaloxygen complex necessary for water activation.80 In a pseudo-octahedral environment, coordination of an oxo (O2) ligand to the metal causes distortion along the principle axis arising from destabilization of the dz2 orbital induced by the relatively short metaloxygen bond (Fig. 5).81 This feature aects the choice of transition metals that can be used to balance the correct amalgamation of stability and reactivity to carry out oxidative reactions. For example, late transition metals with oxidation states less than 4+ have filled dp orbitals and are unable to accept electrons from the oxygen 2px and 2py orbitals, thereby making a multiple metaloxo bond inaccessible. (Other geometries have been proposed to sidestep these barriers; e.g., a square-planar PtIVQO complex stabilized by a PNC pincer ligand framework was indirectly characterized and displayed interesting oxidative capabilities.82,83) Early transition metaloxo complexes tend to be too stable owing to their empty dp orbitals that are eectively filled by electrons from the oxo ligand that form relatively inert metaloxygen triple bond.81 Consequently, the most widely utilized transition metals reside in groups 79 because of their propensity to form metal oxygen double bonds that strike a balance of stability and reactivity upon donation of electrons from the oxygen 2px and 2py orbitals to the partially vacant dp orbitals. Oxo-complexes of ruthenium and osmium have been proven to be particularly eective oxidation catalysts owing to their stability in a polypyridyl environment, and highly reactive oxo unit.84 Modifications of the ligand scaold can have a profound eect on the reactivity and stability of the metaloxo unit.35,62 4.4 Accommodate proton-coupled electron-transfer events

In order for water to be oxidized eciently, the transfer of protons and electrons must be coupled to avoid high-energy intermediates and to enable access to high redox levels stabilized by multiple MO bonding and redox potential levelling.85,86 Reactions classified under PCET are distinguished by proton
This journal is
c

The Royal Society of Chemistry 2013

View Article Online

Feature Article

ChemComm

Scheme 3 A square scheme for cis-[RuII(bpy)2(py)(OH2)]2+.85,87

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

transfer (PT) and electron transfer (ET) events occurring in a stepwise or concerted mechanism, resulting in diering proton and electron content of reactants and products. This phenomenon is common for many redox processes in aqueous media where there are dissociable protons, and is an essential aspect of many biochemical reactions (e.g., provides the necessary driving force for shuttling protons over large distances within the photosynthetic construct85). As a result of the proton dependencies, PCET redox processes are aected by the Nernstian relationship governed by eqn (2), where m is the number of protons transferred and n is the number of electrons transferred under standard ambient temperatures and pressures (SATP) conditions. E = E0 0.059(H+/e) (2)

A square scheme outlines the dierences in these pathways with an archetypal example of a redox-active metal complex bearing a coordinated water molecule, cis-[RuII(bpy)2(py)(OH2)]2+ (bpy = 2,2 0 -bipyridine; py = pyridine) that undergoes PCET chemistry (Scheme 3).87 Beginning with cis-[RuII(bpy)2(py)(OH2)]2+ (abbreviated [RumOHn]x+ with m designating the oxidation state of Ru; n designating the proton content of the bound oxygen ligand; and x denoting the charge of the complex),

it can get oxidized to generate [RuIIIOH2]3+ at 1.02 V at pH o 3. Deprotonation of [RuIIOH2]2+ at pH > 10.6 enables a much more accessible oxidation event at Eo 0 = 0.46 V, as expected for a complex bearing an anionic ligand and a lower charge. An EPT mechanism would involve initial horizontal movement followed by vertical transition and vice versa for a PET mechanism. At pH 7, for example, PCET becomes operative and consequently the redox process occurs at an intermediary potential of 0.66 V. Much greater dierences in redox potentials and pKa values are observed for the Ru(IV)/Ru(III) redox chemistry in the upper left corner of Fig. 6, where the pKas associated with [RuIVOH]3+ and [RuIIIO]+ are outside the accessible region in aqueous solutions and the reduction potentials associated with ET in the absence of PT are separated by >1.2 V. This observation is due to the stabilization of the Ru(IV) state through formation of a stable RuO multiple bond.84,88 An implication of both the stabilization imparted by the formation of the oxo complex and PCET is the small potential dierence of 0.08 V between adjacent valent states (DEo 0 ) of Ru observed at pH 7 between the [RuIVQO]2+/ [RuIIIOH]2+ and [RuIIIOH]2+/[RuIIOH2]2+ redox couples. This phenomenon is known as redox potential levelling whereby preservation of the charge on the complexes (i.e. [RuIIOH2]2+; [RuIIIOH]2+; and [RuIVQO]2+) minimizes electrostatic eects connected with further oxidation.85 In the absence of oxo bond formation and PCET, DEo 0 are typically >1 V as demonstrated with the related example of cis-[RuII(bpy)2Cl2] with DEo 0 = 1.7 V between the Ru(III)/Ru(II) and Ru(IV)/Ru(III) redox couples.89 A drastic example of this stabilization phenomenon by PCET is demonstrated by the redox chemistry exhibited by cis-[RuII(bpy)2(OH2)2]2+ and trans-[RuII(bpy)2(OH2)2]2+ where Ru can be oxidized from Ru(II) to Ru(VI) over a potential range of only 0.6 V at pH 6. High oxidation states are supported by significant donation of electron density from the oxo(s) to the Ru through p interactions. The important implication of redox potential levelling is that multiple oxidizing equivalents can be accumulated over a small potential range without changing the overall charge of the species, eectively minimizing the energy required for multiple electron redox catalysis such as in water oxidation.

5. Summary and outlook


Fig. 6 Eo 0 vs. pH (Pourbaix) diagram for a prototypical octahedral ruthenium complex with a [RuIIOH2]2+ fragment, [RuII(bpy)2(py)(OH2)]2+.84

At the time of our entry into the field of homogeneous water oxidation catalysis in 2007 there were very few established
Chem. Commun., 2013, 49, 218--227 225

This journal is

The Royal Society of Chemistry 2013

View Article Online

ChemComm catalysts in the literature, and a clear understanding of the catalytic water oxidation cycle was decidedly lacking. Today, the number of purported WOCs has grown to >100, and the vast majority of these systems bear a single metal site. This statistic indicates that the syntheses of molecular WOCs can be reasonably accessible, which raises the question: where does the field go from here? From a fundamental perspective, we hold the view that we need to further explore how the electronic geometry and d electron count aect the reaction of the metaloxo unit with water and/or hydroxide. There will inevitably be WOCs utilizing new metal identities in the coming years, and we will follow with keen interest how the reactive fragment of the molecule interacts with the substrate. Indeed, a clear understanding of how each specific metal interacts with the incoming substrate, the intermediates, and the products will be of tremendous value to the field of heterogeneous water oxidation catalysis, which has experienced nominal progress over the course of the last three decades with the exception of the recent interest in amorphous materials.90 We hold the view that a diverse composition of metals will ultimately be needed in order for electrolytically generated hydrogen to be of significance in large-scale energy storage schemes, and thus mechanistic insight provided by the field of homogeneous water oxidation catalysis will almost certainly oer valuable guidance for these researchers within such a huge parameter space (other approaches includes the SHArK Project led by Parkinson,91 high throughput assay by Stahl92). One of the penultimate goals of many researchers in this field is to establish an infinitely stable complex under operating conditions that are relevant to a commercial device. Thus, it is somewhat surprising to see the lack of attention devoted to identifying how catalysts degrade, and shunting such pathways. We addressed this particular issue by examining how TONs were aected by EDGs/EWGs positioned about the periphery of the ligand scaold for a single-site ruthenium complex,35 and the information we gained from these studies was hugely informative. One of the important findings of this work was identifying the inherent need to stabilize the metal-ligand bond trans to the aqua ligand that is destabilized upon formation of a high valent metaloxo. It is for this reason that we set out to stabilize this particular bond in complexes bearing more kinetically labile first-row transition metals.54 Another really important contribution to the field would be to tune the electronic structure of a WOC in such a way where light could be used to overcome a large activation barrier associated with an individual reaction step. For example, if the activation barrier associated with the rate-determining step was one that could be overcome with the assistance of light, the over-potentials associated with the chemical transformation would be reduced. If this goal could be achieved with a single metal site, the resultant photoelectrocatalyst could be easily attached to a mesoporous layer of titania on conducting glass enabling the ecient production of gaseous fuels within a Gratzel cell type of arrangement. Given the lack of progress in developing high eciency photoanodes for splitting water, there is significant insight to be gained by systematically
226 Chem. Commun., 2013, 49, 218--227

Feature Article developing photo responsive molecular systems capable of negotiating the water oxidation. Another area of enormous opportunity available to researchers is the development of a high throughput method for discerning whether a homogeneous WOC is indeed an authentic catalyst, or a precursor to a catalytically active decomposition product. This particular issue presents a significant challenge for the field because every new catalyst will be viewed with scepticism even after a diverse array of control experiments is carried out. This reality provides the impetus to develop new techniques that can rapidly scan for metal oxide solids deposited on the surface of the electrode probe as well as nano-particulate solids formed in solution. The challenge is further compounded by the need for these methods to be capable of scanning a diverse set of reaction conditions. Some researchers have already started to oer solutions to these issues,78 but the further expansion of these techniques is absolutely required if the rapid development of water oxidation catalysis is to persist into the foreseeable future.

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

Acknowledgements
Luvdeep Bhandari and Dr Matthew Dewit are gratefully acknowledged for their assistance with the construction of this manuscript.

References
1 K. N. Ferreira, T. M. Iverson, K. Maghlaoui, J. Barber and S. Iwata, Science, 2004, 303, 1831. 2 J. Yano and V. K. Yachandra, Inorg. Chem., 2008, 47, 1711. 3 J. P. McEvoy and G. W. Brudvig, Chem. Rev., 2006, 106, 4455. 4 T. R. Cook, D. K. Dogutan, S. Y. Reece, Y. Surendranath, T. S. Teets and D. G. Nocera, Chem. Rev., 2010, 110, 6474. 5 N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 15729. 6 V. Balzani, A. Credi and M. Venturi, ChemSusChem, 2008, 1, 26. 7 T. J. Meyer, Nature, 2008, 451, 778. 8 M. Yagi and M. Kaneko, Chem. Rev., 2001, 101, 21. 9 L. Hammarstrom and S. Hammes-Schier, Acc. Chem. Res., 2009, 42, 1859. 10 D. Gust, T. A. Moore and A. L. Moore, Acc. Chem. Res., 2009, 42, 1890. 11 J. H. Alstrum-Acevedo, M. K. Brennaman and T. J. Meyer, Inorg. Chem., 2005, 44, 6802. 12 S. W. Gersten, G. J. Samuels and T. J. Meyer, J. Am. Chem. Soc., 1982, 104, 4029. 13 F. Liu, J. J. Concepcion, J. W. Jurss, T. Cardolaccia, J. L. Templeton and T. J. Meyer, Inorg. Chem., 2008, 47, 1727. 14 R. A. Binstead, C. W. Chronister, J. Ni, C. M. Hartshorn and T. J. Meyer, J. Am. Chem. Soc., 2000, 122, 8464. 15 X. Yang and M.-H. Baik, J. Am. Chem. Soc., 2004, 126, 13222. 16 J. K. Hurst, J. L. Cape, A. E. Clark, S. Das and C. Qin, Inorg. Chem., 2008, 47, 1753. 17 T. Wada, K. Tsuge and K. Tanaka, Angew. Chem., Int. Ed., 2000, 39, 1479. 18 C. Sens, I. Romero, M. Rodriguez, A. Llobet, T. Parella and J. BenetBuchholz, J. Am. Chem. Soc., 2004, 126, 7798. 19 R. Zong and R. Thummel, J. Am. Chem. Soc., 2005, 127, 12802. 20 J. Limburg, G. W. Brudvig and R. H. Crabtree, J. Am. Chem. Soc., 1997, 119, 2761. 21 Y. Naruta, M. Sasayama and T. Sasaki, Angew. Chem., Int. Ed., 1994, 33, 1839. 22 K. Beckmann, H. Uchtenhagen, G. Berggren, M. F. Anderlund, A. Thapper, J. Messinger, S. Styring and P. Kurz, Energy Environ. Sci., 2008, 1, 668. 23 N. D. McDaniel, F. J. Coughlin, L. L. Tinker and S. Bernhard, J. Am. Chem. Soc., 2008, 130, 210.
This journal is
c

The Royal Society of Chemistry 2013

View Article Online

Feature Article
24 H.-W. Tseng, R. Zong, J. T. Muckerman and R. P. Thummel, Inorg. Chem., 2008, 47, 11763. 25 J. J. Concepcion, J. W. Jurss, J. L. Templeton and T. J. Meyer, J. Am. Chem. Soc., 2008, 130, 16462. 26 J. J. Concepcion, J. W. Jurss, M. K. Brennaman, P. G. Hoertz, A. O. T. Patrocinio, N. Y. Murakami Iha, J. L. Templeton and T. J. Meyer, Acc. Chem. Res., 2009, 42, 1954. 27 A. Sartorel, M. Carraro, G. Scorrano, R. De Zorzi, S. Geremia, N. D. McDaniel, S. Bernhard and M. Bonchio, J. Am. Chem. Soc., 2008, 130, 5006. gerler, D. A. Hillesheim, D. G. Musaev 28 Y. V. Geletii, B. Botar, P. Ko and C. L. Hill, Angew. Chem., Int. Ed., 2008, 47, 3896. 29 R. Cao, W. Lai and P. Du, Energy Environ. Sci., 2012, 5, 8134. 30 D. G. H. Hetterscheid and J. N. H. Reek, Angew. Chem., Int. Ed., 2012, 51, 9740. 31 M. M. T. Khan, G. Ramachandraiah, S. H. Mehta, S. H. R. Abdi and S. Kumar, J. Mol. Catal., 1990, 58, 199. 32 M. Yagi, E. Tomita, S. Sakita, T. Kuwabara and K. Nagai, J. Phys. Chem. B, 2005, 109, 21489. 33 N. C. Pramanik and S. Bhattacharya, Transition Met. Chem., 1997, 22, 524. 34 D. J. Wasylenko, C. Ganesamoorthy, B. D. Koivisto and C. P. Berlinguette, Eur. J. Inorg. Chem., 2010, 3135. 35 D. J. Wasylenko, C. Ganesamoorthy, B. D. Koivisto, M. A. Henderson and C. P. Berlinguette, Inorg. Chem., 2010, 49, 2202. 36 S. Masaoka and K. Sakai, Chem. Lett., 2009, 38, 182. 37 L. L. Duan, A. Fischer, Y. H. Xu and L. C. Sun, J. Am. Chem. Soc., 2009, 131, 10397. 38 L. Duan, Y. Xu, P. Zhang, M. Wang and L. Sun, Inorg. Chem., 2010, 49, 209. 39 L. Duan, Y. Xu, M. Gorlov, L. Tong, S. Andersson and L. Sun, Chem. Eur. J., 2010, 16, 4659. 40 L. Duan, F. Bozoglian, S. Manda, B. Stewart, T. Privalov, A. Llobet and L. Sun, Nat. Chem., 2012, 4, 418. 41 L. Li, L. Duan, Y. Xu, M. Gorlov, A. Hagfeldt and L. Sun, Chem. Commun., 2010, 46, 7307. 42 F. Li, B. Zhang, X. Li, Y. Jiang, L. Chen, Y. Li and L. Sun, Angew. Chem., Int. Ed., 2011, 50, 12276. 43 M. Murakami, D. Hong, T. Suenobu, S. Yamaguchi, T. Ogura and S. Fukuzumi, J. Am. Chem. Soc., 2011, 133, 11605. 44 J. F. Hull, D. Balcells, J. D. Blakemore, C. D. Incarvito, O. Eisenstein, G. W. Brudvig and R. H. Crabtree, J. Am. Chem. Soc., 2009, 131, 8730. 45 A. Savini, G. Bellachioma, G. Ciancaleoni, C. Zuccaccia, D. Zuccaccia and A. Macchioni, Chem. Commun., 2010, 46, 9218. 46 J. D. Blakemore, N. D. Schley, D. Balcells, J. F. Hull, G. W. Olack, C. D. Incarvito, O. Eisenstein, G. W. Brudvig and R. H. Crabtree, J. Am. Chem. Soc., 2010, 132, 16017. 47 P. B. Brewster, J. D. Blakemore, N. D. Schley, C. D. Incarvito, N. Hazari, G. W. Brudvig and R. H. Crabtree, Organometallics, 2011, 30, 965. 48 R. Lalrempuia, N. D. McDaniel, H. Muller-Bunz, S. Bernhard and M. Albrecht, Angew. Chem., Int. Ed., 2010, 49, 9765. 49 D. G. H. Hetterscheid and J. N. H. Reek, Chem. Commun., 2011, 47, 2712. 50 Y. Gao, J. Liu, M. Wang, Y. Na, B. Akermark and L. Sun, Tetrahedron, 2007, 63, 1987. 51 Y. Gao, T. Aakermark, J. Liu, L. Sun and B. Aakermark, J. Am. Chem. Soc., 2009, 131, 8726. 52 W. C. Ellis, N. D. McDaniel, S. Bernhard and T. J. Collins, J. Am. Chem. Soc., 2010, 132, 10990. `, I. Garcia-Bosch, L. Go mez, J. J. Pla and 53 J. L. Fillol, Z. Codola M. Costas, Nat. Chem., 2011, 3, 807. 54 D. J. Wasylenko, C. Ganesamoorthy, J. Borau-Garcia and C. P. Berlinguette, Chem. Commun., 2011, 47, 4249. 55 R. T. Jonas and T. D. P. Stack, J. Am. Chem. Soc., 1997, 119, 8566. 56 M. E. de Vries, R. M. La Crois, G. Roelfes, H. Kooijman, L. A. Spek, R. Hage and B. L. Feringa, Chem. Commun., 1997, 1549. 57 D. J. Wasylenko, H. M. Tatlock, J. R. Gardinier and C. P. Berlinguette, Manuscript in revision.

ChemComm
58 D. K. Dogutan, R. McGuire and D. G. Nocera, J. Am. Chem. Soc., 2011, 133, 9178. 59 S. M. Barnett, K. I. Goldberg and J. M. Mayer, Nat. Chem., 2012, 4, 498. 60 Z. Chen, J. J. Concepcion, J. W. Jurss and T. J. Meyer, J. Am. Chem. Soc., 2009, 131, 15580. 61 L.-P. Wang, Q. Wu and T. Van Voorhis, Inorg. Chem., 2010, 49, 4543. 62 D. J. Wasylenko, C. Ganesamoorthy, M. A. Henderson, B. D. Koivisto, H. D. Ostho and C. P. Berlinguette, J. Am. Chem. Soc., 2010, 132, 16094. 63 L. Tong, L. Duan, Y. Xu, T. Privalov and L. Sun, Angew. Chem., Int. Ed., 2011, 50, 445. n, L. Duan, B. Akermark, L. Sun and T. Privalov, Angew. 64 J. Nyhle Chem., Int. Ed., 2010, 49, 1773. 65 S. Romain, L. Vigara and A. Llobet, Acc. Chem. Res., 2009, 42, 1944. 66 A. M. Angeles-Boza and J. P. Roth, Inorg. Chem., 2012, 51, 4722. 67 D. J. Wasylenko, C. Ganesamoorthy, M. A. Henderson and C. P. Berlinguette, Inorg. Chem., 2011, 50, 3662. 68 Z. Chen, J. J. Concepcion, X. Hu, W. Yang, P. G. Hoertz and T. J. Meyer, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 7225. 69 D. E. Polyansky, J. T. Muckerman, J. Rochford, R. Zong, R. P. Thummel and E. Fujita, J. Am. Chem. Soc., 2011, 133, 14649. 70 F. Tiago de Oliveira, A. Chanda, D. Banerjee, X. Shan, S. Mondal, L. Que Jr, E. L. Bominaar, E. Munck and T. J. Collins, Science, 2007, 315, 835. 71 D. J. Wasylenko, R. D. Palmer, E. Schott and C. P. Berlinguette, Chem. Commun., 2012, 48, 2107. 72 J. J. Stracke and R. G. Finke, J. Am. Chem. Soc., 2011, 133, 14872. 73 R. H. Crabtree, Chem. Rev., 2012, 112, 1536. 74 J. D. Blakemore, N. D. Schley, G. W. Olack, C. D. Incarvito, G. W. Brudvig and R. H. Crabtree, Chem. Sci., 2011, 2, 94. 75 D. B. Grotjahn, D. B. Brown, J. K. Martin, D. C. Marelius, M. C. Abadjian, H. N. Tran, G. Kalyuzhny, K. S. Vecchio, Z. G. Specht, S. A. Cortes-Llamas, V. Miranda-Soto, C. van Niekerk, C. E. Moore and A. L. Rheingold, J. Am. Chem. Soc., 2011, 133, 19024. 76 N. Marquet, F. Gartner, S. Losse, M. M. Pohl, H. Junge and M. Beller, ChemSusChem, 2011, 4, 1598. 77 U. Hintermair, S. M. Hashmi, M. Elimelech and R. H. Crabtree, J. Am. Chem. Soc., 2012, 134, 9785. 78 N. D. Schley, J. D. Blakemore, N. K. Subbaiyan, C. D. Incarvito, F. DSouza, R. H. Crabtree and G. W. Brudvig, J. Am. Chem. Soc., 2011, 133, 10473. 79 N. Yukie, J. J. Concepcion, J. W. Jurss, M. Iha, M. K. Brennaman, P. G. Hoertz, A. O. T. Patrocinio, N. Y. Murakami Iha, J. L. Templeton and T. J. Meyer, Acc. Chem. Res., 2009, 42, 1954. 80 R. Eisenberg and H. B. Gray, Inorg. Chem., 2008, 47, 1697. 81 T. A. Betley, Q. Wu, T. Van Voorhis and D. G. Nocera, Inorg. Chem., 2008, 47, 1849. 82 E. Poverenov, I. Efremenko, A. I. Frenkel, Y. Ben-David, L. J. W. Shimon, G. Leitus, L. Konstantinovski, J. M. L. Martin and D. Milstein, Nature, 2008, 455, 1093. 83 I. Efremenko, E. Poverenov, J. M. L. Martin and D. Milstein, J. Am. Chem. Soc., 2010, 132, 14886. 84 T. J. Meyer and M. H. V. Huynh, Inorg. Chem., 2003, 42, 8140. 85 M. H. V. Huynh and T. J. Meyer, Chem. Rev., 2007, 107, 5004. 86 D. R. Weinberg, C. J. Gagliardi, J. F. Hull, C. F. Murphy, C. A. Kent, B. C. Westlake, A. Paul, D. H. Ess, D. G. McCaerty and T. J. Meyer, Chem. Rev., 2012, 112, 4016. 87 B. A. Moyer and T. J. Meyer, J. Am. Chem. Soc., 1978, 100, 3601. ant and A.-L. Teillout, Proc. Natl. 88 C. Costentin, M. Robert, J.-M. Save Acad. Sci. U. S. A., 2009, 106, 11829. 89 G. A. Heath, K. A. Moock, D. W. A. Sharp and L. J. Yellowlees, J. Chem. Soc., Chem. Commun., 1985, 1503. 90 M. W. Kanan and D. G. Nocera, Science, 2008, 321, 1072. 91 B. A. Parkinson and J. Schuttlefield, The SHArK Project (Solar Hydrogen Activity Research Kit), http://www.thesharkproject.org. , A. B. Powell and S. S. Stahl, 92 J. B. Gerken, J. Y. C. Chen, R. C. Masse Angew. Chem., Int. Ed., 2012, 51, 6676.

Downloaded by Ewha Womens University on 05/04/2013 09:43:23. Published on 07 November 2012 on http://pubs.rsc.org | doi:10.1039/C2CC35632E

This journal is

The Royal Society of Chemistry 2013

Chem. Commun., 2013, 49, 218--227

227

Anda mungkin juga menyukai