Co-reporter:Shashi Bhushan Sinha, Dimitar Y. Shopov, Liam S. Sharninghausen, Christopher J. Stein, Brandon Q. Mercado, David Balcells, Thomas Bondo Pedersen, Markus Reiher, Gary W. Brudvig, and Robert H. Crabtree
Journal of the American Chemical Society July 19, 2017 Volume 139(Issue 28) pp:9672-9672
Publication Date(Web):June 24, 2017
DOI:10.1021/jacs.7b04874
Chemical and electrochemical oxidation or reduction of our recently reported Ir(IV,IV) mono-μ-oxo dimers results in the formation of fully characterized Ir(IV,V) and Ir(III,III) complexes. The Ir(IV,V) dimers are unprecedented and exhibit remarkable stability under ambient conditions. This stability and modest reduction potential of 0.99 V vs NHE is in part attributed to complete charge delocalization across both Ir centers. Trends in crystallographic bond lengths and angles shed light on the structural changes accompanying oxidation and reduction. The similarity of these mono-μ-oxo dimers to our Ir “blue solution” water-oxidation catalyst gives insight into potential reactive intermediates of this structurally elusive catalyst. Additionally, a highly reactive material, proposed to be a Ir(V,V) μ-oxo species, is formed on electrochemical oxidation of the Ir(IV,V) complex in organic solvents at 1.9 V vs NHE. Spectroelectrochemistry shows reversible conversion between the Ir(IV,V) and proposed Ir(V,V) species without any degradation, highlighting the exceptional oxidation resistance of the 2-(2-pyridinyl)-2-propanolate (pyalk) ligand and robustness of these dimers. The Ir(III,III), Ir(IV,IV) and Ir(IV,V) redox states have been computationally studied both with DFT and multiconfigurational calculations. The calculations support the stability of these complexes and provide further insight into their electronic structures.
Co-reporter:Bradley J. Brennan, Kevin P. Regan, Alec C. Durrell, Charles A. Schmuttenmaer, and Gary W. Brudvig
ACS Energy Letters - New in 2016 January 13, 2017 Volume 2(Issue 1) pp:
Publication Date(Web):December 19, 2016
DOI:10.1021/acsenergylett.6b00583
Lateral charge transport in a redox-active monolayer can be utilized for solar energy harvesting. A model porphyrin system was chosen to study the influence of the solvent on lateral hole hopping, which plays a crucial role in the charge-transfer kinetics. We examined the influence of water, acetonitrile, and propylene carbonate as solvents. Hole-hopping lifetimes varied by nearly 3 orders of magnitude among solvents, ranging from 3 ns in water to 2800 ns in propylene carbonate, and increased nonlinearly as a function of added acetonitrile in aqueous solvent mixtures. These results elucidate the important roles of solvation, molecular packing dynamics, and lateral charge-transfer mechanisms that have implications for all dye-sensitized photoelectrochemical device designs.
Co-reporter:Yueshen Wu, Jianbing Jiang, Zhe Weng, Maoyu Wang, Daniël L. J. Broere, Yiren Zhong, Gary W. Brudvig, Zhenxing Feng, and Hailiang Wang
ACS Central Science August 23, 2017 Volume 3(Issue 8) pp:847-847
Publication Date(Web):July 26, 2017
DOI:10.1021/acscentsci.7b00160
Transition-metal-based molecular complexes are a class of catalyst materials for electrochemical CO2 reduction to CO that can be rationally designed to deliver high catalytic performance. One common mechanistic feature of these electrocatalysts developed thus far is an electrogenerated reduced metal center associated with catalytic CO2 reduction. Here we report a heterogenized zinc–porphyrin complex (zinc(II) 5,10,15,20-tetramesitylporphyrin) as an electrocatalyst that delivers a turnover frequency as high as 14.4 site–1 s–1 and a Faradaic efficiency as high as 95% for CO2 electroreduction to CO at −1.7 V vs the standard hydrogen electrode in an organic/water mixed electrolyte. While the Zn center is critical to the observed catalysis, in situ and operando X-ray absorption spectroscopic studies reveal that it is redox-innocent throughout the potential range. Cyclic voltammetry indicates that the porphyrin ligand may act as a redox mediator. Chemical reduction of the zinc–porphyrin complex further confirms that the reduction is ligand-based and the reduced species can react with CO2. This represents the first example of a transition-metal complex for CO2 electroreduction catalysis with its metal center being redox-innocent under working conditions.
Co-reporter:Gary W. Brudvig;Sebastiano Campagna
Chemical Society Reviews 2017 vol. 46(Issue 20) pp:6085-6087
Publication Date(Web):2017/10/16
DOI:10.1039/C7CS90096A
A graphical abstract is available for this content
Co-reporter:Kelly L. Materna;Robert H. Crabtree
Chemical Society Reviews 2017 vol. 46(Issue 20) pp:6099-6110
Publication Date(Web):2017/10/16
DOI:10.1039/C7CS00314E
Surface anchoring groups are needed to attach molecular units to photoanodes for photocatalytic water oxidation. The anchoring group must be hydrolytically stable and oxidation resistant under a variety of pH conditions. They must sometimes be electrically conducting for efficient light-induced electron injection from a photosensitizer to a metal oxide, but other times not conducting for accumulation of oxidizing equivalents on a water-oxidation catalyst. Commonly used anchors such as carboxylic acids and phosphonic acids have limited stability in aqueous environments, leading to surface hydrolysis and loss of catalytic function. More hydrolytically stable anchors, such as silatranes and hydroxamic acids, which are oxidation resistant and stable under acidic, neutral, and basic conditions, are more suitable for photoanode applications. Hydroxamic acids can be incorporated into dye molecules to give high electron injection efficiency due to their electrical conductivity and strong electronic coupling to the metal oxide surface. In contrast, silatranes, once bound as siloxanes, have diminished electronic coupling making them useful as catalyst anchors.
Co-reporter:Dimitar Y. Shopov;Liam S. Sharninghausen;Shashi Bhushan Sinha;Julia E. Borowski;Brandon Q. Mercado;Robert H. Crabtree
New Journal of Chemistry (1998-Present) 2017 vol. 41(Issue 14) pp:6709-6719
Publication Date(Web):2017/07/10
DOI:10.1039/C7NJ01845B
We have prepared and characterized a series of novel polydentate N,O-donor ligands derived from our well-studied ligand 2-(2-pyridinyl)-2-propanol (pyalkH), having the general formula Me{C(OH)(2-py)CH2}nH, where n = 2 or 3. Like pyalkH, these analogues bind via N and O with deprotonation at the latter, thus extending the strongly donor pyridine-alkoxide chelation power of pyalkH to polydentate forms. The greater denticity allows for more effective binding and polynuclear cluster formation with first-row transition metals. Several stable alkoxo-bridged polynuclear clusters of these ligands with Mn, Cu, Co and Ni have been prepared; all reported ligands and complexes have been characterized, including by X-ray crystallography. We report a one-step synthesis of these ligands, alongside pyalkH, on a multi-gram scale from inexpensive starting materials. We have also developed a new scalable procedure for the isolation of pyalkH that avoids the need for chromatography, making large-scale production of this ligand commercially viable.
Co-reporter:Liam S. Sharninghausen;Shashi Bhushan Sinha;Dimitar Y. Shopov;Dr. Bron Q. Mercado;Dr. David Balcells; Gary W. Brudvig; Robert H. Crabtree
Angewandte Chemie 2017 Volume 129(Issue 42) pp:13227-13231
Publication Date(Web):2017/10/09
DOI:10.1002/ange.201707593
AbstractWe have prepared and fully characterized two isomers of [IrIV(dpyp)2] (dpyp=meso-2,4-di(2-pyridinyl)-2,4-pentanediolate). These complexes can cleanly oxidize to [IrV(dpyp)2]+, which to our knowledge represent the first mononuclear coordination complexes of IrV in an N,O-donor environment. One isomer has been fully characterized in the IrV state, including by X-ray crystallography, XPS, and DFT calculations, all of which confirm metal-centered oxidation. The unprecedented stability of these IrV complexes is ascribed to the exceptional donor strength of the ligands, their resistance to oxidative degradation, and the presence of four highly donor alkoxide groups in a plane, which breaks the degeneracy of the d-orbitals and favors oxidation.
Co-reporter:Liam S. Sharninghausen;Shashi Bhushan Sinha;Dimitar Y. Shopov;Dr. Bron Q. Mercado;Dr. David Balcells; Gary W. Brudvig; Robert H. Crabtree
Angewandte Chemie International Edition 2017 Volume 56(Issue 42) pp:13047-13051
Publication Date(Web):2017/10/09
DOI:10.1002/anie.201707593
AbstractWe have prepared and fully characterized two isomers of [IrIV(dpyp)2] (dpyp=meso-2,4-di(2-pyridinyl)-2,4-pentanediolate). These complexes can cleanly oxidize to [IrV(dpyp)2]+, which to our knowledge represent the first mononuclear coordination complexes of IrV in an N,O-donor environment. One isomer has been fully characterized in the IrV state, including by X-ray crystallography, XPS, and DFT calculations, all of which confirm metal-centered oxidation. The unprecedented stability of these IrV complexes is ascribed to the exceptional donor strength of the ligands, their resistance to oxidative degradation, and the presence of four highly donor alkoxide groups in a plane, which breaks the degeneracy of the d-orbitals and favors oxidation.
Co-reporter:Dr. Jianbing Jiang;Kelly L. Materna;Dr. Svante Hedström;Dr. Ke R. Yang; Robert H. Crabtree; Victor S. Batista; Gary W. Brudvig
Angewandte Chemie 2017 Volume 129(Issue 31) pp:9239-9243
Publication Date(Web):2017/07/24
DOI:10.1002/ange.201704700
AbstractMain-group complexes are shown to be viable electrocatalysts for the H2-evolution reaction (HER) from acid. A series of antimony porphyrins with varying axial ligands were synthesized for electrocatalysis applications. The proton-reduction catalytic properties of TPSb(OH)2 (TP=5,10,15,20-tetra(p-tolyl)porphyrin) with two axial hydroxy ligands were studied in detail, demonstrating catalytic H2 production. Experiments, in conjunction with quantum chemistry calculations, show that the catalytic cycle is driven via the redox activity of both the porphyrin ligand and the Sb center. This study brings insight into main group catalysis and the role of redox-active ligands during catalysis.
Co-reporter:Dr. Jianbing Jiang;Kelly L. Materna;Dr. Svante Hedström;Dr. Ke R. Yang; Robert H. Crabtree; Victor S. Batista; Gary W. Brudvig
Angewandte Chemie International Edition 2017 Volume 56(Issue 31) pp:9111-9115
Publication Date(Web):2017/07/24
DOI:10.1002/anie.201704700
AbstractMain-group complexes are shown to be viable electrocatalysts for the H2-evolution reaction (HER) from acid. A series of antimony porphyrins with varying axial ligands were synthesized for electrocatalysis applications. The proton-reduction catalytic properties of TPSb(OH)2 (TP=5,10,15,20-tetra(p-tolyl)porphyrin) with two axial hydroxy ligands were studied in detail, demonstrating catalytic H2 production. Experiments, in conjunction with quantum chemistry calculations, show that the catalytic cycle is driven via the redox activity of both the porphyrin ligand and the Sb center. This study brings insight into main group catalysis and the role of redox-active ligands during catalysis.
Co-reporter:Jessica M. Wiwczar;Amy M. LaFountain;Jimin Wang
Photosynthesis Research 2017 Volume 134( Issue 2) pp:175-182
Publication Date(Web):24 July 2017
DOI:10.1007/s11120-017-0425-4
Photosystem II (PSII) of oxygenic photosynthetic organisms normally contains exclusively chlorophyll a (Chl a) as its major light-harvesting pigment. Chl a canonically consists of the chlorin headgroup with a 20-carbon, 4-isoprene unit, phytyl tail. We have examined the 1.9 Å crystal structure of PSII from thermophilic cyanobacteria reported by Shen and coworkers in 2012 (PDB accession of 3ARC/3WU2). A newly refined electron density map from this structure, presented here, reveals that some assignments of the cofactors may be different from those modeled in the 3ARC/3WU2 structure, including a specific Chl a that appears to have a truncated tail by one isoprene unit. We provide experimental evidence using high-performance liquid chromatography and mass spectrometry for a small population of Chl a esterified to a 15-carbon farnesyl tail in PSII of thermophilic cyanobacteria.
Co-reporter:Wen Liu;Jianbing Jiang;Ke R. Yang;Yingying Mi;Piranavan Kumaravadivel;Yiren Zhong;Qi Fan;Zhe Weng;Zishan Wu;Judy J. Cha;Henghui Zhou;Victor S. Batista;Hailiang Wang
PNAS 2017 114 (14 ) pp:3578-3583
Publication Date(Web):2017-04-04
DOI:10.1073/pnas.1620809114
Lithium–sulfur batteries (Li–S batteries) have attracted intense interest because of their high specific capacity and low
cost, although they are still hindered by severe capacity loss upon cycling caused by the soluble lithium polysulfide intermediates.
Although many structure innovations at the material and device levels have been explored for the ultimate goal of realizing
long cycle life of Li–S batteries, it remains a major challenge to achieve stable cycling while avoiding energy and power
density compromises caused by the introduction of significant dead weight/volume and increased electrochemical resistance.
Here we introduce an ultrathin composite film consisting of naphthalimide-functionalized poly(amidoamine) dendrimers and graphene
oxide nanosheets as a cycling stabilizer. Combining the dendrimer structure that can confine polysulfide intermediates chemically
and physically together with the graphene oxide that renders the film robust and thin (<1% of the thickness of the active
sulfur layer), the composite film is designed to enable stable cycling of sulfur cathodes without compromising the energy
and power densities. Our sulfur electrodes coated with the composite film exhibit very good cycling stability, together with
high sulfur content, large areal capacity, and improved power rate.
Co-reporter:David J. Vinyard, Sahr Khan, Mikhail Askerka, Victor S. BatistaGary W. Brudvig
The Journal of Physical Chemistry B 2017 Volume 121(Issue 5) pp:
Publication Date(Web):January 12, 2017
DOI:10.1021/acs.jpcb.7b00110
The S2 redox intermediate of the oxygen-evolving complex in photosystem II is present as two spin isomers. The S = 1/2 isomer gives rise to a multiline electron paramagnetic resonance (EPR) signal at g = 2.0, whereas the S = 5/2 isomer exhibits a broad EPR signal at g = 4.1. The electronic structures of these isomers are known, but their role in the catalytic cycle of water oxidation remains unclear. We show that formation of the S = 1/2 state from the S = 5/2 state is exergonic at temperatures above 160 K. However, the S = 1/2 isomer decays to S1 more slowly than the S = 5/2 isomer. These differences support the hypotheses that the S3 state is formed via the S2 state S = 5/2 isomer and that the stabilized S2 state S = 1/2 isomer plays a role in minimizing S2QA– decay under light-limiting conditions.
Co-reporter:Liam S. Sharninghausen, Shashi Bhushan Sinha, Dimitar Y. Shopov, Bonnie Choi, Brandon Q. Mercado, Xavier Roy, David Balcells, Gary W. Brudvig, and Robert H. Crabtree
Journal of the American Chemical Society 2016 Volume 138(Issue 49) pp:15917-15926
Publication Date(Web):November 15, 2016
DOI:10.1021/jacs.6b07716
Co-reporter:Zhe Weng; Jianbing Jiang; Yueshen Wu; Zishan Wu; Xiaoting Guo; Kelly L. Materna; Wen Liu; Victor S. Batista; Gary W. Brudvig;Hailiang Wang
Journal of the American Chemical Society 2016 Volume 138(Issue 26) pp:8076-8079
Publication Date(Web):June 16, 2016
DOI:10.1021/jacs.6b04746
Exploration of heterogeneous molecular catalysts combining the atomic-level tunability of molecular structures and the practical handling advantages of heterogeneous catalysts represents an attractive approach to developing high-performance catalysts for important and challenging chemical reactions such as electrochemical carbon dioxide reduction which holds the promise for converting emissions back to fuels utilizing renewable energy. Thus, far, efficient and selective electroreduction of CO2 to deeply reduced products such as hydrocarbons remains a big challenge. Here, we report a molecular copper-porphyrin complex (copper(II)-5,10,15,20-tetrakis(2,6-dihydroxyphenyl)porphyrin) that can be used as a heterogeneous electrocatalyst with high activity and selectivity for reducing CO2 to hydrocarbons in aqueous media. At −0.976 V vs the reversible hydrogen electrode, the catalyst is able to drive partial current densities of 13.2 and 8.4 mA cm–2 for methane and ethylene production from CO2 reduction, corresponding to turnover frequencies of 4.3 and 1.8 molecules·site–1·s–1 for methane and ethylene, respectively. This represents the highest catalytic activity to date for hydrocarbon production over a molecular CO2 reduction electrocatalyst. The unprecedented catalytic performance is attributed to the built-in hydroxyl groups in the porphyrin structure and the reactivity of the copper(I) metal center.
Co-reporter:Kelly L. Materna, Benjamin Rudshteyn, Bradley J. Brennan, Morgan H. Kane, Aaron J. Bloomfield, Daria L. Huang, Dimitar Y. Shopov, Victor S. Batista, Robert H. Crabtree, and Gary W. Brudvig
ACS Catalysis 2016 Volume 6(Issue 8) pp:5371
Publication Date(Web):July 8, 2016
DOI:10.1021/acscatal.6b01101
A pentamethylcyclopentadienyl (Cp*) iridium water-oxidation precatalyst was modified to include a silatrane functional group for covalent attachment to metal oxide semiconductor surfaces. The heterogenized catalyst was found to perform electrochemically driven water oxidation at an overpotential of 462 mV with a turnover number of 304 and turnover frequency of 0.035 s–1 in a 0.1 M KNO3 electrolyte at pH 5.8. Computational modeling of the experimental IR spectra suggests that the catalyst retains its Cp* group during the first hour of catalysis and likely remains monomeric.Keywords: alternative energy; electrocatalysis; iridium; metal oxide; silatrane; surface binding; water oxidation
Co-reporter:Bradley J. Brennan, Jeffrey Chen, Benjamin Rudshteyn, Subhajyoti Chaudhuri, Brandon Q. Mercado, Victor S. Batista, Robert H. Crabtree and Gary W. Brudvig
Chemical Communications 2016 vol. 52(Issue 14) pp:2972-2975
Publication Date(Web):12 Jan 2016
DOI:10.1039/C5CC09857B
Hydroxamate binding modes and protonation states have yet to be conclusively determined. Molecular titanium(IV) phenylhydroxamate complexes were synthesized as structural and spectroscopic models, and compared to functionalized TiO2 nanoparticles. In a combined experimental–theoretical study, we find that the predominant binding form is monodeprotonated, with evidence for the chelate mode.
Co-reporter:Yuta Tsubonouchi, Shu Lin, Alexander R. Parent, Gary W. Brudvig and Ken Sakai
Chemical Communications 2016 vol. 52(Issue 51) pp:8018-8021
Publication Date(Web):31 May 2016
DOI:10.1039/C6CC02816K
A μ-oxido-bridged triruthenium complex (RuT2+), formed by air-oxidation of a previously reported monoruthenium water oxidation catalyst (WOC), serves as an efficient photochemical WOC with the turnover frequency (TOF) and turnover number (TON) 0.90 s−1 and 610, respectively. The crystal structures of RuT2+ and its one-electron oxidized RuT3+ are also reported.
Co-reporter:Jianbing Jiang, John R. Swierk, Svante Hedström, Adam J. Matula, Robert H. Crabtree, Victor S. Batista, Charles A. Schmuttenmaer and Gary W. Brudvig
Physical Chemistry Chemical Physics 2016 vol. 18(Issue 28) pp:18678-18682
Publication Date(Web):30 Jun 2016
DOI:10.1039/C6CP04377A
Interfacial electron transfer dynamics of a series of photosensitizers bound to TiO2via linkers of varying conjugation strength are explored by spectroscopic and computational techniques. Injection and recombination depend on the extent of conjugation in the linker, where the LUMO delocalization determines the injection dynamics but both the HOMO and HOMO−1 are involved in recombination.
Co-reporter:David J. Vinyard, Mikhail Askerka, Richard J. Debus, Victor S. Batista, and Gary W. Brudvig
Biochemistry 2016 Volume 55(Issue 31) pp:4432-4436
Publication Date(Web):July 19, 2016
DOI:10.1021/acs.biochem.6b00543
Ammonia binds to two sites in the oxygen-evolving complex (OEC) of Photosystem II (PSII). The first is as a terminal ligand to Mn in the S2 state, and the second is at a site outside the OEC that is competitive with chloride. Binding of ammonia in this latter secondary site results in the S2 state S = 5/2 spin isomer being favored over the S = 1/2 spin isomer. Using electron paramagnetic resonance spectroscopy, we find that ammonia binds to the secondary site in wild-type Synechocystis sp. PCC 6803 PSII, but not in D2-K317A mutated PSII that does not bind chloride. By combining these results with quantum mechanics/molecular mechanics calculations, we propose that ammonia binds in the secondary site in competition with D1-D61 as a hydrogen bond acceptor to the OEC terminal water ligand, W1. Implications for the mechanism of ammonia binding via its primary site directly to Mn4 in the OEC are discussed.
Co-reporter:Bradley J. Brennan
The Journal of Physical Chemistry C 2016 Volume 120(Issue 23) pp:12495-12502
Publication Date(Web):May 18, 2016
DOI:10.1021/acs.jpcc.6b02635
Hydroxamic acids chelate metals with high affinity and form hydrolytically stable complexes with metal oxides such as TiO2. However, these appealing binding properties can cause problems during the preparation and application of metallocatalysts with appended hydroxamate anchoring groups. Here we show that the tetrahydropyran (THP) O-protected hydroxamate group can be cleaved in situ on a TiO2 surface at room temperature, leading to the surface-bound species. Surface-mediated deprotection has several advantages over direct surface functionalization including increased hydrolytic stability of the covalent interaction with the metal oxide surface and decreased aggregation of the surface species. Application of the surface-mediated chelation method for dye-sensitized photoelectrochemical cells (DSPC) was examined using the organic dye MK-2. Results show that the surface-mediated deprotection led to improved DSPC performance attributed to a decrease in dye aggregation relative to a DSPC prepared using standard methods. This simplified approach using THP-protected hydroxamates provides a convenient new method for functionalizing metal oxides.
Co-reporter:Jianbing Jiang, Robert H. Crabtree, and Gary W. Brudvig
The Journal of Organic Chemistry 2016 Volume 81(Issue 19) pp:9483-9488
Publication Date(Web):September 19, 2016
DOI:10.1021/acs.joc.6b01883
Trialkylstannanes are good leaving groups that have been used for the formation of carbon–metal bonds to electrode surfaces for analyses of single-molecule conductivity. Here, we report the multistep synthesis of two amide-containing compounds that are of interest in studies of molecular rectifiers. Each molecule has two trimethylstannyl units, one linked by a methylene and the other by an ethylene group. To account for the very different reactivities of the parent halides, a new methodology for one-step trimethylstannylation was developed and optimized.
Co-reporter:Jianbing Jiang, John R. Swierk, Kelly L. Materna, Svante HedströmShin Hee Lee, Robert H. Crabtree, Charles A. Schmuttenmaer, Victor S. Batista, Gary W. Brudvig
The Journal of Physical Chemistry C 2016 Volume 120(Issue 51) pp:28971-28982
Publication Date(Web):December 3, 2016
DOI:10.1021/acs.jpcc.6b10350
We report CF3-substituted porphyrins and evaluate their use as photosensitizers in water-splitting dye-sensitized photoelectrochemical cells (WS-DSPECs) by characterizing interfacial electron transfer on metal oxide surfaces. By using (CF3)2C6H3 instead of C6F5 substituents at the meso positions, we obtain the desired high potentials while avoiding the sensitivity of C6F5 substituents to nucleophilic substitution, a process that limits the types of synthetic reactions that can be used. Both the number of CF3 groups and the central metal tune the ground and excited-state potentials. A pair of porphyrins bearing carboxylic acids as anchoring groups were deposited on SnO2 and TiO2 surfaces, and the interfacial charge-injection and charge-recombination kinetics were characterized by using a combination of computational modeling, terahertz measurements, and transient absorption spectroscopy. We find that both free-base and metalated porphyrins inject into SnO2 and that recombination is slower for the latter case. These findings demonstrate that (CF3)2C6H3-substituted porphyrins are promising photosensitizers for use in WS-DSPECs.
Co-reporter:James D. Blakemore, Robert H. Crabtree, and Gary W. Brudvig
Chemical Reviews 2015 Volume 115(Issue 23) pp:12974
Publication Date(Web):July 7, 2015
DOI:10.1021/acs.chemrev.5b00122
Co-reporter:Karin J. Young, Bradley J. Brennan, Ranitendranath Tagore, and Gary W. Brudvig
Accounts of Chemical Research 2015 Volume 48(Issue 3) pp:567
Publication Date(Web):March 2, 2015
DOI:10.1021/ar5004175
Catalysts for light-driven water oxidation are a critical component for development of solar fuels technology. The multielectron redox chemistry required for this process has been successfully deployed on a global scale in natural photosynthesis by green plants and cyanobacteria using photosystem II (PSII). PSII employs a conserved, cuboidal Mn4CaOX cluster called the O2-evolving complex (OEC) that offers inspiration for artificial O2-evolution catalysts.In this Account, we describe our work on manganese model chemistry relevant to PSII, particularly the functional model [MnIII/IV2(terpy)2(μ-O)2(OH2)2](NO3)3 complex (terpy = 2,2′;6′,2″-terpyridine), a mixed-valent di-μ-oxo Mn dimer with two terminal aqua ligands. In the presence of oxo-donor oxidants such as HSO5–, this complex evolves O2 by two pathways, one of which incorporates solvent water in an O–O bond-forming reaction. Deactivation pathways of this catalyst include comproportionation to form an inactive MnIVMnIV dimer and also degradation to MnO2, a consequence of ligand loss when the oxidation state of the complex is reduced to labile MnII upon release of O2. The catalyst’s versatility has been shown by its continued catalytic activity after direct binding to the semiconductor titanium dioxide. In addition, after binding to the surface of TiO2 via a chromophoric linker, the catalyst can be oxidized by a photoinduced electron-transfer mechanism, mimicking the natural PSII process.Model oxomanganese complexes have also aided in interpreting biophysical and computational studies on PSII. In particular, the μ-oxo exchange rates of the Mn–terpy dimer have been instrumental in establishing that the time scale for μ-oxo exchange of high-valent oxomanganese complexes with terminal water ligands is slower than O2 evolution in the natural photosynthetic system. Furthermore, computational studies on the Mn–terpy dimer and the OEC point to similar MnIV–oxyl intermediates in the O–O bond-forming mechanism.Comparison between the OEC and the Mn–terpy dimer indicates that challenges remain in the development of synthetic Mn water-oxidation catalysts. These include redox leveling to couple multielectron reactions with one-electron steps, avoiding labile MnII species during the catalytic cycle, and protecting the catalyst active site from undesired side reactions.As the first example of a functional manganese O2-evolution catalyst, the Mn–terpy dimer exemplifies the interrelatedness of biomimetic chemistry with biophysical studies. The design of functional model complexes enriches the study of the natural photosynthetic system, while biology continues to provide inspiration for artificial photosynthetic technologies to meet global energy demand.
Co-reporter:Dimitar Y. Shopov; Benjamin Rudshteyn; Jesús Campos; Victor S. Batista; Robert H. Crabtree
Journal of the American Chemical Society 2015 Volume 137(Issue 22) pp:7243-7250
Publication Date(Web):May 19, 2015
DOI:10.1021/jacs.5b04185
The preparation of the facial and meridional isomers of [Ir(pyalk)3] (pyalk = 2-(2-pyridyl)isopropanoate), as model complexes for a powerful water oxidation catalyst, is reported. The strongly donating N3O3 ligand set is very oxidation-resistant, yet promotes facile metal-centered oxidation to form stable Ir(IV) compounds. The IrIII/IV reduction potentials of the two isomers differ by 340 mV despite the identical ligand set. A ligand field rationalization is advanced and supported by DFT calculations.
Co-reporter:Shashi B. Sinha; Dimitar Y. Shopov; Liam S. Sharninghausen; David J. Vinyard; Brandon Q. Mercado; Gary W. Brudvig;Robert H. Crabtree
Journal of the American Chemical Society 2015 Volume 137(Issue 50) pp:15692-15695
Publication Date(Web):December 7, 2015
DOI:10.1021/jacs.5b12148
We describe facial and meridional isomers of [RhIII(pyalk)3], as well as meridional [RhIV(pyalk)3]+ {pyalk =2-(2-pyridyl)-2-propanoate}, the first coordination complex in an N,O-donor environment to show a clean, reversible RhIII/IV redox couple and to have a stable Rh(IV) form, which we characterize by EPR and UV–visible spectroscopy as well as X-ray crystallography. The unprecedented stability of the Rh(IV) species is ascribed to the exceptional donor strength of the ligands, their oxidation resistance, and the meridional coordination geometry.
Co-reporter:Sahr Khan, Ke R. Yang, Mehmed Z. Ertem, Victor S. Batista, and Gary W. Brudvig
ACS Catalysis 2015 Volume 5(Issue 12) pp:7104
Publication Date(Web):October 22, 2015
DOI:10.1021/acscatal.5b01976
The biomimetic oxomanganese complex [MnIII/IV2(μ-O)2(terpy)2(OH2)2](NO3)3 (1; terpy = 2,2′:6′,2″-terpyridine) catalyzes O2 evolution from water when activated by oxidants, such as oxone (2KHSO5·KHSO4·K2SO4). The mechanism of this reaction has never been characterized, due to the fleeting nature of the intermediates. In the present study, we elucidate the underlying reaction mechanism through experimental and theoretical analyses of competitive kinetic oxygen isotope effects (KIEs) during catalytic turnover conditions. The experimental 18O KIE is a sensitive probe of the highest transition state in the O2-evolution mechanism and provides a strict constraint for calculated mechanisms. The 18O kinetic isotope effect of 1.013 ± 0.003 measured using natural abundance reactants is consistent with the calculated isotope effect of peroxymonosulfate binding to the complex, as described by density functional theory (DFT). This provides strong evidence for peroxymonosulfate binding being both the first irreversible and rate-determining step during turnover, in contrast to the previously held assumption that formation of a high-valent Mn-oxo/oxyl species is the highest barrier step that controls the rate of O2 evolution by this complex. The comparison of the measured and calculated KIEs supplements previous kinetic studies, enabling us to describe the complete mechanism of O2 evolution, starting from when the oxidant first binds to the manganese complex to when O2 is released. The reported findings lay the groundwork for understanding O2 evolution catalyzed by other biomimetic oxomanganese complexes, with features common to those of the O2-evolving complex of photosystem II, providing experimental and theoretical diagnostics of oxygen isotope effects that could reveal the nature of elusive reaction intermediates.Keywords: density functional theory; manganese complex; oxygen evolution mechanism; oxygen isotope effects; peroxymonosulfate
Co-reporter:Bradley J. Brennan, Yick Chong Lam, Paul M. Kim, Xing Zhang, and Gary W. Brudvig
ACS Applied Materials & Interfaces 2015 Volume 7(Issue 29) pp:16124
Publication Date(Web):July 2, 2015
DOI:10.1021/acsami.5b05050
Organic dyes with their wide range of molecular structures and spectroscopic features show great promise for solar energy applications. Corroles, structural analogues to porphyrins, are highly fluorescent molecules with tunable properties. We have synthesized a series of structurally similar corroles chelating gallium and phosphorus, along with a β-chlorinated phosphorus corrole, and determined their photophysical and electrochemical properties. The electrochemical potentials to oxidize the corroles range from 0.78 V vs NHE for the gallium corrole to 1.42 V for the β-octachlorinated phosphorus corrole. We are interested in developing photosensitizers for water oxidation on a metal oxide-based photoanode, so the corroles were modified to contain a meso-phenyl-COOH substituent for binding to metal oxide surfaces. The ability of these corrole dyes to act as photosensitizers was assessed by comparing the corroles in a model dye sensitized solar cell design. Transient absorption spectroscopy was utilized to analyze recombination dynamics and determine the kinetics of iodide oxidation. The most efficient photoelectrochemical cell was achieved for the phosphorus corrole P-2 with electrochemical properties and kinetics suitable for both photoinduced electron injection into TiO2 and oxidation of iodide. This structure–function study highlights the wide window for tuning corrole electrochemical potentials while still maintaining desirable photophysical properties, important variables when designing dyes for applications in photoelectrochemical water-oxidation cells.Keywords: corroles; electrochemistry; molecular design; nanosecond transient absorption spectroscopy; photoelectrochemical cells;
Co-reporter:Wendu Ding, Matthieu Koepf, Christopher Koenigsmann, Arunabh Batra, Latha Venkataraman, Christian F. A. Negre, Gary W. Brudvig, Robert H. Crabtree, Charles A. Schmuttenmaer, and Victor S. Batista
Journal of Chemical Theory and Computation 2015 Volume 11(Issue 12) pp:5888-5896
Publication Date(Web):November 3, 2015
DOI:10.1021/acs.jctc.5b00823
We report a systematic computational search of molecular frameworks for intrinsic rectification of electron transport. The screening of molecular rectifiers includes 52 molecules and conformers spanning over 9 series of structural motifs. N-Phenylbenzamide is found to be a promising framework with both suitable conductance and rectification properties. A targeted screening performed on 30 additional derivatives and conformers of N-phenylbenzamide yielded enhanced rectification based on asymmetric functionalization. We demonstrate that electron-donating substituent groups that maintain an asymmetric distribution of charge in the dominant transport channel (e.g., HOMO) enhance rectification by raising the channel closer to the Fermi level. These findings are particularly valuable for the design of molecular assemblies that could ensure directionality of electron transport in a wide range of applications, from molecular electronics to catalytic reactions.
Co-reporter:Julianne M. Thomsen, Daria L. Huang, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2015 vol. 44(Issue 28) pp:12452-12472
Publication Date(Web):05 May 2015
DOI:10.1039/C5DT00863H
Organometallic Ir precatalysts have been found to yield homogeneous Ir-based water-oxidation catalysts (WOCs) with very high activity. The Cp*Ir catalyst series can operate under a variety of regimes: it can either act as a homogeneous or a heterogeneous catalyst; it can be driven by chemical, photochemical, or electrochemical methods; and the molecular catalyst can either act in solution or supported as a molecular unit on a variety of solid oxides. In addition to optimizing the various reaction conditions, work has continued to elucidate the catalyst activation mechanism and identify water-oxidation intermediates. This Perspective will describe the development of the Cp*Ir series, their many forms as WOCs, and their ongoing characterization.
Co-reporter:Wojciech T. Osowiecki, Stafford W. Sheehan, Karin J. Young, Alec C. Durrell, Brandon Q. Mercado and Gary W. Brudvig
Dalton Transactions 2015 vol. 44(Issue 38) pp:16873-16881
Publication Date(Web):02 Sep 2015
DOI:10.1039/C5DT02390D
Splitting water into hydrogen and oxygen is one of the most promising ways of storing energy from intermittent, renewable sources in the future. Toward this goal, development of inexpensive, stable, and non-toxic catalysts for water oxidation is crucial. We report that the electrodeposition of manganese oxide in the presence of sodium dodecyl sulfate (SDS) produces a material that is highly active for electrocatalytic water oxidation at pH near 7 and remains stable for over 24 hours of sustained electrolysis. Clark electrode measurements demonstrate more than 95% Faradaic efficiency for oxygen evolution after an initial charging period. We found that catalytic performance was optimized in films prepared by electrodeposition using a precursor solution containing moderate concentration of substrates, namely 25 mM Mn2+ and 25 mM SDS. Microstructure and elemental analyses revealed that the deposited material, a mixed-phase manganese oxide, is structurally similar to materials used for electrochemical capacitors and batteries, drawing a parallel between highly studied cathode materials for rechargeable batteries and heterogeneous catalysts for water oxidation.
Co-reporter:Bradley J. Brennan, Alec C. Durrell, Matthieu Koepf, Robert H. Crabtree and Gary W. Brudvig
Physical Chemistry Chemical Physics 2015 vol. 17(Issue 19) pp:12728-12734
Publication Date(Web):16 Apr 2015
DOI:10.1039/C5CP01683E
Current molecular water-oxidation photoelectrocatalytic cells have substantial kinetic limitations under normal solar photon flux where electron–hole recombination processes may outcompete charge buildup on the catalytic centers. One method of overcoming these limitations is to design a system where multiple light-harvesting dyes work cooperatively with a single catalyst. We report a porphyrin monomer/dyad array for analysis of lateral hole transfer on a SnO2 surface consisting of a free-base porphyrin that functions to absorb light and initiate charge injection into the conduction band of SnO2, which leaves a positive charge on the organic moiety, and a free-base porphyrin/Zn-porphyrin dyad molecule that functions as a thermodynamic trap for the photoinduced holes. By using transient absorption spectroscopy, we have determined that the holes on the surface-bound free-base porphyrins are highly mobile via electron self-exchange between close-packed neighbors. The lateral charge-transfer processes were modelled by treating the system statistically with a random-walk method that utilizes experimentally derived kinetic parameters. The results of the modelling indicate that each self-exchange (hop) occurs within 25 ns and that the holes are efficiently transferred to the Zn-porphyrin. This hole-harvesting scheme provides a framework for enhancing the efficiency of multielectron photoelectrocatalytic reactions such as the four-electron oxidation of water.
Co-reporter:Ravi Pokhrel, Richard J. Debus, and Gary W. Brudvig
Biochemistry 2015 Volume 54(Issue 8) pp:1663-1672
Publication Date(Web):February 13, 2015
DOI:10.1021/bi501468h
Efficient proton removal from the oxygen-evolving complex (OEC) of photosystem II (PSII) and activation of substrate water molecules are some of the key aspects optimized in the OEC for high turnover rates. The hydrogen-bonding network around the OEC is critical for efficient proton transfer and for tuning the position and pKa values of the substrate water/hydroxo/oxo molecules. The D1-N181 residue is part of the hydrogen-bonding network on the active face of the OEC. D1-N181 is also associated with the chloride ion in the D2-K317 site and is one of the residues closest to a putative substrate water molecule bound as a terminal ligand to Mn4. We studied the effect of the D1-N181A and D1-N181S mutations on the water oxidation chemistry at the OEC. PSII core complexes isolated from the D1-N181A and D1-N181S mutants have steady-state O2 evolution rates lower than those of wild-type PSII core complexes. Fourier transform infrared spectroscopy indicates slight perturbations of the hydrogen-bonding network in D1-N181A and D1-N181S PSII core complexes, similar to the effects of some other mutations in the same region, but to a lesser extent. Unlike in wild-type PSII core complexes, a g = 4 signal was observed in the S2-state EPR spectra of D1-N181A and D1-N181S PSII core complexes in addition to the normal g = 2 multiline signal. The S-state cycling of D1-N181A and D1-N181S PSII core complexes was similar to that of wild-type PSII core complexes, whereas the O2-release kinetics of D1-N181A and D1-N181S PSII core complexes were much slower than the O2-release kinetics of wild-type PSII core complexes. On the basis of these results, we conclude that proton transfer is not compromised in the D1-N181A and D1-N181S mutants but that the O–O bond formation step is retarded in these mutants.
Co-reporter:Mikhail Askerka, David J. Vinyard, Jimin Wang, Gary W. Brudvig, and Victor S. Batista
Biochemistry 2015 Volume 54(Issue 9) pp:1713-1716
Publication Date(Web):February 24, 2015
DOI:10.1021/acs.biochem.5b00089
A recent femtosecond X-ray diffraction study produced the first high-resolution structural model of the oxygen-evolving complex of photosystem II that is free of radiation-induced manganese reduction (Protein Data Bank entries 4UB6 and 4UB8). We find, however, that the model does not match extended X-ray absorption fine structure and QM/MM data for the S1 state. This is attributed to uncertainty about the positions of oxygen atoms that remain partially unresolved, even at 1.95 Å resolution, next to the heavy manganese centers. In addition, the photosystem II crystals may contain significant amounts of the S0 state, because of extensive dark adaptation prior to data collection.
Co-reporter:David J. Vinyard and Gary W. Brudvig
Biochemistry 2015 Volume 54(Issue 2) pp:622-628
Publication Date(Web):December 22, 2014
DOI:10.1021/bi5014134
Water oxidation in Photosystem II occurs at the oxygen-evolving complex (OEC), which cycles through distinct intermediates, S0–S4. The inhibitor ammonia selectively binds to the S2 state at an unresolved site that is not competitive with substrate water. By monitoring the yields of flash-induced oxygen production, we show that ammonia decreases the net efficiency of OEC turnover and slows the decay kinetics of S2 to S1. The temperature dependence of biphasic S2 decay kinetics provides activation energies that do not vary in control and ammonia conditions. We interpret our data in the broader context of previous studies by introducing a kinetic model for both the formation and decay of ammonia-bound S2. The model predicts ammonia binds to S2 rapidly (t1/2 = 1 ms) with a large equilibrium constant. This finding implies that ammonia decreases the reduction potential of S2 by at least 2.7 kcal mol–1 (>120 mV), which is not consistent with ammonia substitution of a terminal water ligand of Mn(IV). Instead, these data support the proposal that ammonia binds as a bridging ligand between two Mn atoms. Implications for the mechanism of O–O bond formation are discussed.
Co-reporter:Wei Li;Stafford W. Sheehan;Da He;Yumin He;Xiahui Yao;Dr. Ronald L. Grimm;Dr. Gary W. Brudvig;Dr. Dunwei Wang
Angewandte Chemie International Edition 2015 Volume 54( Issue 39) pp:11428-11432
Publication Date(Web):
DOI:10.1002/anie.201504427
Abstract
Solar water splitting in acidic solutions has important technological implications, but has not been demonstrated to date in a dual absorber photoelectrochemical cell. The lack of functionally stable water-oxidation catalysts (WOCs) in acids is a key reason for this slow development. The only WOCs that are stable at low pH are Ir-based systems, which are typically too expensive to be implemented broadly. It is now shown that this deficiency may be corrected by applying an ultra-thin monolayer of a molecular Ir WOC to hematite for solar water splitting in acidic solutions. The turn-on voltage is observed to shift cathodically by 250 mV upon the application of a monolayer of the molecular Ir WOC. When the molecular WOC is replaced by a heterogeneous multilayer derivative, stable solar water splitting for over 5 h is achieved with near-unity Faradaic efficiency.
Co-reporter:Sahr Khan, Jennifer S. Sun, and Gary W. Brudvig
The Journal of Physical Chemistry B 2015 Volume 119(Issue 24) pp:7722-7728
Publication Date(Web):February 26, 2015
DOI:10.1021/jp513035u
The normal pathway of electron transfer on the electron-acceptor side of photosystem II (PSII) involves electron transfer from quinone A, QA, to quinone B, QB. It is possible to redirect electrons from QA– to water-soluble CoIII complexes, which opens a new avenue for harvesting electrons from water oxidation by immobilization of PSII on electrode surfaces. Herein, the kinetics of electron transfer from QA– to [Co(III)(terpy)2]3+ (terpy = 2,2′;6′,2″-terpyridine) are investigated with a spectrophotometric assay revealing that the reaction follows Michaelis–Menten saturation kinetics, is inhibited by cations, and is not affected by variation of the QA reduction potential. A negatively charged site on the stromal surface of the PSII protein complex, composed of glutamic acid residues near QA, is hypothesized to bind cations, especially divalent cations. The cations are proposed to tune the redox properties of QA through electrostatic interactions. These observations may thus explain the molecular basis of the effect of divalent cations like Ca2+, Sr2+, Mg2+, and Zn2+ on the redox properties of the quinones in PSII, which has previously been attributed to long-range conformational changes propagated from divalent cations binding to the Ca(II)-binding site in the oxygen-evolving complex on the lumenal side of the PSII complex.
Co-reporter:Wei Li;Stafford W. Sheehan;Da He;Yumin He;Xiahui Yao;Dr. Ronald L. Grimm;Dr. Gary W. Brudvig;Dr. Dunwei Wang
Angewandte Chemie 2015 Volume 127( Issue 39) pp:11590-11594
Publication Date(Web):
DOI:10.1002/ange.201504427
Abstract
Solar water splitting in acidic solutions has important technological implications, but has not been demonstrated to date in a dual absorber photoelectrochemical cell. The lack of functionally stable water-oxidation catalysts (WOCs) in acids is a key reason for this slow development. The only WOCs that are stable at low pH are Ir-based systems, which are typically too expensive to be implemented broadly. It is now shown that this deficiency may be corrected by applying an ultra-thin monolayer of a molecular Ir WOC to hematite for solar water splitting in acidic solutions. The turn-on voltage is observed to shift cathodically by 250 mV upon the application of a monolayer of the molecular Ir WOC. When the molecular WOC is replaced by a heterogeneous multilayer derivative, stable solar water splitting for over 5 h is achieved with near-unity Faradaic efficiency.
Co-reporter:Julianne M. Thomsen ; Stafford W. Sheehan ; Sara M. Hashmi ; Jesús Campos ; Ulrich Hintermair ; Robert H. Crabtree
Journal of the American Chemical Society 2014 Volume 136(Issue 39) pp:13826-13834
Publication Date(Web):September 4, 2014
DOI:10.1021/ja5068299
Organometallic iridium complexes bearing oxidatively stable chelate ligands are precursors for efficient homogeneous water-oxidation catalysts (WOCs), but their activity in oxygen evolution has so far been studied almost exclusively with sacrificial chemical oxidants. In this report, we study the electrochemical activation of Cp*Ir complexes and demonstrate true electrode-driven water oxidation catalyzed by a homogeneous iridium species in solution. Whereas the Cp* precursors exhibit no measurable O2-evolution activity, the molecular species formed after their oxidative activation are highly active homogeneous WOCs, capable of electrode-driven O2 evolution with high Faradaic efficiency. We have ruled out the formation of heterogeneous iridium oxides, either as colloids in solution or as deposits on the surface of the electrode, and found indication that the conversion of the precursor to the active molecular species occurs by a similar process whether carried out by chemical or electrochemical methods. This work makes these WOCs more practical for application in photoelectrochemical dyads for light-driven water splitting.
Co-reporter:Christian F. A. Negre, Karin J. Young, Ma. Belén Oviedo, Laura J. Allen, Cristián G. Sánchez, Katarzyna N. Jarzembska, Jason B. Benedict, Robert H. Crabtree, Philip Coppens, Gary W. Brudvig, and Victor S. Batista
Journal of the American Chemical Society 2014 Volume 136(Issue 46) pp:16420-16429
Publication Date(Web):October 22, 2014
DOI:10.1021/ja509270f
We find that crystallographically resolved Ti17O24(OPri)20 nanoparticles, functionalized by covalent attachment of 4-nitrophenyl-acetylacetonate or coumarin 343 adsorbates, exhibit hole injection into surface states when photoexcited with visible light (λ = 400–680 nm). Our findings are supported by photoelectrochemical measurements, EPR spectroscopy, and quantum dynamics simulations of interfacial charge transfer. The underlying mechanism is consistent with measurements of photocathodic currents generated with visible light for thin layers of functionalized polyoxotitanate nanocrystals deposited on FTO working electrodes. The reported experimental and theoretical analysis demonstrates for the first time the feasibility of p-type sensitization of TiO2 solely based on covalent binding of organic adsorbates.
Co-reporter:Ravi Pokhrel and Gary W. Brudvig
Physical Chemistry Chemical Physics 2014 vol. 16(Issue 24) pp:11812-11821
Publication Date(Web):28 Mar 2014
DOI:10.1039/C4CP00493K
Water oxidation at the oxygen-evolving complex (OEC) of photosystem II (PSII) involves multiple redox states called Sn states (n = 0–4). The S1 → S2 redox transition of the OEC has been studied extensively using various forms of spectroscopy, including electron paramagnetic resonance (EPR) and Fourier transform infrared (FTIR) spectroscopy. In the S2 state, two isomers of the OEC are observed by EPR: a ST = 1/2 form and a ST = 5/2 form. DFT-based structural models of the OEC have been proposed for the two spin isomers in the S2 state, but the factors that determine the stability of one form or the other are not known. Using structural information on the OEC and its surroundings, in conjunction with spectroscopic information available on the S1 → S2 transition for a variety of site-directed mutations, Ca2+ and Cl− substitutions, and small molecule inhibitors, we propose that the hydrogen-bonding network encompassing D1-D61 and the OEC-bound waters plays an important role in stabilizing one spin isomer over the other. In the presence of ammonia, PSII centers can be trapped in either the ST = 5/2 form after a 200 K illumination procedure or an ammonia-altered ST = 1/2 form upon annealing at 273 K. We propose a mechanism for ammonia binding to the OEC in the S2 state that takes into account the hydrogen-binding requirements for ammonia binding and the specificity for binding of ammonia but not methylamine. A discussion regarding the possibility of spin isomers of the OEC in the S1 state, analogous to the spin isomers of the S2 state, is also presented.
Co-reporter:C. Koenigsmann, T. S. Ripolles, B. J. Brennan, C. F. A. Negre, M. Koepf, A. C. Durrell, R. L. Milot, J. A. Torre, R. H. Crabtree, V. S. Batista, G. W. Brudvig, J. Bisquert and C. A. Schmuttenmaer
Physical Chemistry Chemical Physics 2014 vol. 16(Issue 31) pp:16629-16641
Publication Date(Web):04 Jul 2014
DOI:10.1039/C4CP02405B
An efficient synthetic protocol to functionalize the cyanoacrylic acid anchoring group of commercially available MK-2 dye with a highly water-stable hydroxamate anchoring group is described. Extensive characterization of this hydroxamate-modified dye (MK-2HA) reveals that the modification does not affect its favorable optoelectronic properties. Dye-sensitized solar cells (DSSCs) prepared with the MK-2HA dye attain improved efficiency (6.9%), relative to analogously prepared devices with commercial MK-2 and N719 dyes. The hydroxamate anchoring group also contributes to significantly increased water stability, with a decrease in the rate constant for dye desorption of MK-2HA relative to MK-2 in the presence of water by as much as 37.5%. In addition, the hydroxamate-anchored dye undergoes essentially no loss in DSSC efficiency and the external quantum efficiency improves when up to 20% water is purposefully added to the electrolyte. In contrast, devices prepared with the commercial dye suffer a 50% decline in efficiency under identical conditions, with a concomitant decrease in external quantum efficiency. Collectively, our results indicate that covalent functionalization of organic dyes with hydroxamate anchoring groups is a simple and efficient approach to improving the water stability of the dye–semiconductor interface and overall device durability.
Co-reporter:Shashi Bhushan Sinha, Jesús Campos, Gary W. Brudvig and Robert H. Crabtree
RSC Advances 2014 vol. 4(Issue 90) pp:49395-49399
Publication Date(Web):25 Sep 2014
DOI:10.1039/C4RA10510A
An inexpensive protocol for the conversion of –C6H4R into –COOH groups using Co(II)–Oxone mixture as the catalytic system is described. A series of substrates containing substituted and non-substituted phenyl groups could be selectively converted into carboxylic acids. Initial mechanistic data have been provided.
Co-reporter:Mikhail Askerka, Jimin Wang, Gary W. Brudvig, and Victor S. Batista
Biochemistry 2014 Volume 53(Issue 44) pp:
Publication Date(Web):October 27, 2014
DOI:10.1021/bi5011915
The S1 → S2 transition of the oxygen-evolving complex (OEC) of photosystem II does not involve the transfer of a proton to the lumen and occurs at cryogenic temperatures. Therefore, it is commonly thought to involve only Mn oxidation without any significant change in the structure of the OEC. Here, we analyze structural changes upon the S1 → S2 transition, as revealed by quantum mechanics/molecular mechanics methods and the isomorphous difference Fourier method applied to serial femtosecond X-ray diffraction data. We find that the main structural change in the OEC is in the position of the dangling Mn and its coordination environment.
Co-reporter:Wendu Ding;Dr. Christian F. A. Negre;Dr. Julio L. Palma;Dr. Alec C. Durrell;Dr. Laura J. Allen;Dr. Karin J. Young;Rebecca L. Milot; Charles A. Schmuttenmaer; Gary W. Brudvig; Robert H. Crabtree; Victor S. Batista
ChemPhysChem 2014 Volume 15( Issue 6) pp:1138-1147
Publication Date(Web):
DOI:10.1002/cphc.201400063
Abstract
Linkers that favor rectification of interfacial electron transfer are likely to be required for efficient photo-driven catalysis of multi-electron reactions at electrode surfaces. Design principles are discussed, together with the synthesis and characterization of a specific pair of molecular linkers, related by inversion of the direction of an amide bond in the heart of the molecule. The linkers have a terpyridyl group that can covalently bind Mn as in a well-known water oxidation catalyst and an acetylacetonate group that allows attachment to TiO2 surfaces. The appropriate choice of the sense of the amide linkage yields directionality of interfacial electron transfer, essential to enhance electron injection and slow back-electron transfer. Support comes from electron paramagnetic resonance and terahertz spectroscopic measurements, as well as computational modeling characterizing the asymmetry of electron transfer properties.
Co-reporter:Katherine E. Shinopoulos;Jianfeng Yu;Peter J. Nixon
Photosynthesis Research 2014 Volume 120( Issue 1-2) pp:141-152
Publication Date(Web):2014 May
DOI:10.1007/s11120-013-9793-6
Secondary electron transfer in photosystem II (PSII), which occurs when water oxidation is inhibited, involves redox-active carotenoids (Car), as well as chlorophylls (Chl), and cytochrome b559 (Cyt b559), and is believed to play a role in photoprotection. CarD2 may be the initial point of secondary electron transfer because it is the closest cofactor to both P680, the initial oxidant, and to Cyt b559, the terminal secondary electron donor within PSII. In order to characterize the role of CarD2 and to determine the effects of perturbing CarD2 on both the electron-transfer events and on the identity of the redox-active cofactors, it is necessary to vary the properties of CarD2 selectively without affecting the ten other Car per PSII. To this end, site-directed mutations around the binding pocket of CarD2 (D2-G47W, D2-G47F, and D2-T50F) have been generated in Synechocystis sp. PCC 6803. Characterization by near-IR and EPR spectroscopy provides the first experimental evidence that CarD2 is one of the redox-active carotenoids in PSII. There is a specific perturbation of the Car∙+ near-IR spectrum in all three mutated PSII samples, allowing the assignment of the spectral signature of CarD2∙+; CarD2∙+ exhibits a near-IR peak at 980 nm and is the predominant secondary donor oxidized in a charge separation at low temperature in ferricyanide-treated wild-type PSII. The yield of secondary donor radicals is substantially decreased in PSII complexes isolated from each mutant. In addition, the kinetics of radical formation are altered in the mutated PSII samples. These results are consistent with oxidation of CarD2 being the initial step in secondary electron transfer. Furthermore, normal light levels during mutant cell growth perturb the shape of the Chl∙+ near-IR absorption peak and generate a dark-stable radical observable in the EPR spectra, indicating a higher susceptibility to photodamage further linking the secondary electron-transfer pathway to photoprotection.
Co-reporter:Alexander R. Parent, Robert H. Crabtree and Gary W. Brudvig
Chemical Society Reviews 2013 vol. 42(Issue 6) pp:2247-2252
Publication Date(Web):13 Sep 2012
DOI:10.1039/C2CS35225G
In this tutorial review, we compare chemical oxidants for driving water-oxidation catalysts, focusing on the advantages and disadvantages of each oxidant.
Co-reporter:Ulrich Hintermair ; Stafford W. Sheehan ; Alexander R. Parent ; Daniel H. Ess ; David T. Richens ; Patrick H. Vaccaro ; Gary W. Brudvig ;Robert H. Crabtree
Journal of the American Chemical Society 2013 Volume 135(Issue 29) pp:10837-10851
Publication Date(Web):July 3, 2013
DOI:10.1021/ja4048762
We present evidence for Cp* being a sacrificial placeholder ligand in the [Cp*IrIII(chelate)X] series of homogeneous oxidation catalysts. UV–vis and 1H NMR profiles as well as MALDI-MS data show a rapid and irreversible loss of the Cp* ligand under reaction conditions, which likely proceeds through an intramolecular inner-sphere oxidation pathway reminiscent of the reductive in situ elimination of diolefin placeholder ligands in hydrogenation catalysis by [(diene)MI(L,L′)]+ (M = Rh and Ir) precursors. When oxidatively stable chelate ligands are bound to the iridium in addition to the Cp*, the oxidized precursors yield homogeneous solutions with a characteristic blue color that remain active in both water- and CH-oxidation catalysis without further induction period. Electrophoresis suggests the presence of well-defined Ir-cations, and TEM-EDX, XPS, 17O NMR, and resonance-Raman spectroscopy data are most consistent with the molecular identity of the blue species to be a bis-μ-oxo di-iridium(IV) coordination compound with two waters and one chelate ligand bound to each metal. DFT calculations give insight into the electronic structure of this catalyst resting state, and time-dependent simulations agree with the assignments of the experimental spectroscopic data. [(cod)IrI(chelate)] precursors bearing the same chelate ligands are shown to be equally effective precatalysts for both water- and CH-oxidations using NaIO4 as chemical oxidant.
Co-reporter:Karin J. Young, Michael K. Takase, and Gary W. Brudvig
Inorganic Chemistry 2013 Volume 52(Issue 13) pp:7615-7622
Publication Date(Web):June 18, 2013
DOI:10.1021/ic400691e
Four manganese complexes of pentadentate ligands have been studied for their ability to act as oxygen evolution catalysts in the presence of Oxone or hydrogen peroxide. The complexes [Mn(PaPy3)(NO3)](ClO4) (1) (PaPy3H = N,N-bis(2-pyridylmethyl)-amine-N-ethyl-2-pyridine-2-carboxamide) and [Mn(PaPy3)(μ-O)(PaPy3)Mn](ClO4)2 (2) feature an anionic carboxamido ligand trans to the labile sixth coordination site, while [Mn(N4Py)OTf](OTf) (3) (N4Py = N,N-bis(2-pyridylmethyl)-N-bis(2-pyridyl)methylamine) and [Mn(PY5)(OH2)](ClO4)2 (4) (PY5 = 2,6-bis(bis(2-pyridyl)methoxymethane)-pyridine) have neutral ligands of varying flexibility. 1 and 2 are shown to evolve oxygen in the presence of either Oxone or hydrogen peroxide, but 3 evolves oxygen only in the presence of hydrogen peroxide. 4 is inactive. The activity of 1 and 2 with Oxone suggests that the presence of an anionic N-donor ligand plays a role in stabilizing putative high-valent intermediates. Anionic N-donor ligands may be viewed as alternatives to μ-oxo ligands that are prone to protonation in low-valent Mn species formed during a catalytic cycle, resulting in loss of catalyst structure.
Co-reporter:James D. Blakemore, Michael W. Mara, Maxwell N. Kushner-Lenhoff, Nathan D. Schley, Steven J. Konezny, Ivan Rivalta, Christian F. A. Negre, Robert C. Snoeberger, Oleksandr Kokhan, Jier Huang, Andrew Stickrath, Lan Anh Tran, Maria L. Parr, Lin X. Chen, David M. Tiede, Victor S. Batista, Robert H. Crabtree, and Gary W. Brudvig
Inorganic Chemistry 2013 Volume 52(Issue 4) pp:1860-1871
Publication Date(Web):February 5, 2013
DOI:10.1021/ic301968j
Upon electrochemical oxidation of the precursor complexes [Cp*Ir(H2O)3]SO4 (1) or [(Cp*Ir)2(OH)3]OH (2) (Cp* = pentamethylcyclopentadienyl), a blue layer of amorphous iridium oxide containing a carbon admixture (BL) is deposited onto the anode. The solid-state, amorphous iridium oxide material that is formed from the molecular precursors is significantly more active for water-oxidation catalysis than crystalline IrO2 and functions as a remarkably robust catalyst, capable of catalyzing water oxidation without deactivation or significant corrosion for at least 70 h. Elemental analysis reveals that BL contains carbon that is derived from the Cp* ligand (∼ 3% by mass after prolonged electrolysis). Because the electrodeposition of precursors 1 or 2 gives a highly active catalyst material, and electrochemical oxidation of other iridium complexes seems not to result in immediate conversion to iridium oxide materials, we investigate here the nature of the deposited material. The steps leading to the formation of BL and its structure have been investigated by a combination of spectroscopic and theoretical methods. IR spectroscopy shows that the carbon content of BL, while containing some C–H bonds intact at short times, is composed primarily of components with C═O fragments at longer times. X-ray absorption and X-ray absorption fine structure show that, on average, the six ligands to iridium in BL are likely oxygen atoms, consistent with formation of iridium oxide under the oxidizing conditions. High-energy X-ray scattering (HEXS) and pair distribution function (PDF) analysis (obtained ex situ on powder samples) show that BL is largely free of the molecular precursors and is composed of small, <7 Å, iridium oxide domains. Density functional theory (DFT) modeling of the X-ray data suggests a limited set of final components in BL; ketomalonate has been chosen as a model fragment because it gives a good fit to the HEXS-PDF data and is a potential decomposition product of Cp*.
Co-reporter:Maxwell N. Kushner-Lenhoff, James D. Blakemore, Nathan D. Schley, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2013 vol. 42(Issue 10) pp:3617-3622
Publication Date(Web):13 Dec 2012
DOI:10.1039/C2DT32326E
A thin layer of an amorphous, mixed-valence iridium oxide (electrodeposited from an organometallic precursor, [Cp*Ir(H2O)3]2+) is a heterogeneous catalyst among the most active and stable currently available for electrochemical water oxidation. We show that buffers can improve the oxygen-evolution activity of such thin-layer catalysts near neutral pH, but that buffer identity and concentration, as well as the solution pH, remain key determinants of long-term electrocatalyst activity and stability; for example, phosphate buffer can reduce the overpotential by up to 173 mV.
Co-reporter:Ravi Pokhrel and Gary W. Brudvig
Biochemistry 2013 Volume 52(Issue 21) pp:
Publication Date(Web):April 30, 2013
DOI:10.1021/bi400206q
The role of chloride in photosystem II (PSII) is unclear. Several monovalent anions compete for the Cl– site(s) in PSII, and some even support activity. NO2– has been reported to be an activator in Cl–-depleted PSII membranes. In this paper, we report a detailed investigation of the chemistry of NO2– with PSII. NO2– is shown to inhibit PSII activity, and the effects on the donor side as well as the acceptor side are characterized using steady-state O2-evolution assays, electron paramagnetic resonance (EPR) spectroscopy, electron-transfer assays, and flash-induced polarographic O2 yield measurements. Enzyme kinetics analysis shows multiple sites of NO2– inhibition in PSII with significant inhibition of oxygen evolution at <5 mM NO2–. By EPR spectroscopy, the yield of the S2 state remains unchanged up to 15 mM NO2–. However, the S2-state g = 4.1 signal is favored over the g = 2 multiline signal with increasing NO2– concentrations. This could indicate competition of NO2– for the Cl– site at higher NO2– concentrations. In addition to the donor-side chemistry, there is clear evidence of an acceptor-side effect of NO2–. The g = 1.9 Fe(II)-QA–• signal is replaced by a broad g = 1.6 signal in the presence of NO2–. Additionally, a g = 1.8 Fe(II)-Q–• signal is present in the dark, indicating the formation of a NO2–-bound Fe(II)-QB–• species in the dark. Electron-transfer assays suggest that the inhibitory effect of NO2– on the activity of PSII is largely due to the donor-side chemistry of NO2–. UV–visible spectroscopy and flash-induced polarographic O2 yield measurements indicate that NO2– is oxidized by the oxygen-evolving complex in the higher S states, contributing to the donor-side inhibition by NO2–.
Co-reporter:Ravi Pokhrel, Rachel J. Service, Richard J. Debus, and Gary W. Brudvig
Biochemistry 2013 Volume 52(Issue 28) pp:
Publication Date(Web):June 20, 2013
DOI:10.1021/bi301700u
The role of chloride in photosystem II (PSII) is unclear. Using structural information from PSII and a careful comparison with other chloride-activated enzymes, we proposed a role for chloride at the D2-K317 site in PSII [Pokhrel, R., et al. (2011) Biochemistry 50, 2725–2734]. To probe the role of chloride at this site, the D2-K317R, D2-K317A, D2-K317Q, and D2-K317E mutations were created in the cyanobacterium Synechocystis sp. PCC 6803. Purified PSII from the mutants was probed with Fourier transform infrared difference spectroscopy, demonstrating that compared to PSII from wild-type Synechocystis, PSII from all four mutants exhibit changes in the conformations of the polypeptide backbone and carboxylate groups. However, D2-K317R PSII exhibits minor changes, whereas D2-K317A, D2-K317Q, and D2-K317E PSII exhibit more substantial changes in polypeptide conformations. Steady-state oxygen-evolution measurements of purified PSII core complexes show that the oxygen-evolution activity of D2-K317A is independent of chloride. This is consistent with the loss of the chloride requirement when the charged K residue is replaced with an uncharged residue that no longer binds to an essential carboxylate (D1-D61) in the absence of chloride, analogous to observations in other chloride-activated enzymes. In contrast, the oxygen-evolution activity of D2-K317R is sensitive to the chloride concentration in the assay buffer; the effective KD for chloride binding is higher in D2-K317R than in wild-type PSII, possibly because of a less optimal binding site in the mutant. The S2 states of wild-type, D2-K317A, and D2-K317R PSII were probed using electron paramagnetic resonance spectroscopy. A g = 2 multiline signal, similar to the wild-type signal, was observed for D2-K317A and D2-K317R. However, a g = 4 signal was also observed for D2-K317R. Measurements of flash-dependent O2 yields showed that D2-K317A and D2-K317R have a higher miss factor than wild-type PSII. The oxygen-release kinetics of D2-K317A and D2-K317R were slower than those of the wild type, in the following order: D2-K317A < D2-K317R < wild type. These results collectively suggest that proton transfer is inefficient in D2-K317A and D2-K317R, thereby giving rise to a higher miss factor and slower oxygen-release kinetics.
Co-reporter:Rhitankar Pal, Christian F. A. Negre, Leslie Vogt, Ravi Pokhrel, Mehmed Z. Ertem, Gary W. Brudvig, and Victor S. Batista
Biochemistry 2013 Volume 52(Issue 44) pp:
Publication Date(Web):October 15, 2013
DOI:10.1021/bi401214v
The S0 → S1 transition of the oxygen-evolving complex (OEC) of photosystem II is one of the least understood steps in the Kok cycle of water splitting. We introduce a quantum mechanics/molecular mechanics (QM/MM) model of the S0 state that is consistent with extended X-ray absorption fine structure spectroscopy and X-ray diffraction data. In conjunction with the QM/MM model of the S1 state, we address the proton-coupled electron-transfer (PCET) process that occurs during the S0 → S1 transition, where oxidation of a Mn center and deprotonation of a μ-oxo bridge lead to a significant rearrangement in the OEC. A hydrogen bonding network, linking the D1-D61 residue to a Mn-bound water molecule, is proposed to facilitate the PCET mechanism.
Co-reporter:Meng Zhou, Ulrich Hintermair, Brian G. Hashiguchi, Alexander R. Parent, Sara M. Hashmi, Menachem Elimelech, Roy A. Periana, Gary W. Brudvig, and Robert H. Crabtree
Organometallics 2013 Volume 32(Issue 4) pp:957-965
Publication Date(Web):
DOI:10.1021/om301252w
Sodium periodate (NaIO4) is shown to be a milder and more efficient terminal oxidant for C–H oxidation with Cp*Ir (Cp* = C5Me5) precatalysts than ceric(IV) ammonium nitrate. Synthetically useful yields, regioselectivities, and functional group tolerance were found for methylene oxidation of substrates bearing a phenyl, ketone, ester, or sulfonate group. Oxidation of the natural products (−)-ambroxide and sclareolide proceeded selectively, and retention of configuration was seen in cis-decalin hydroxylation. At 60 °C, even primary C–H bonds can be activated: whereas methane was overoxidized to CO2 in 39% yield without giving partially oxidized products, ethane was transformed into acetic acid in 25% yield based on total NaIO4. 18O labeling was demonstrated in cis-decalin hydroxylation with 18OH2 and NaIO4. A kinetic isotope effect of 3.0 ± 0.1 was found in cyclohexane oxidation at 23 °C, suggesting C–H bond cleavage as the rate-limiting step. Competition experiments between C–H and water oxidation show that C–H oxidation of sodium 4-ethylbenzene sulfonate is favored by 4 orders of magnitude. In operando time-resolved dynamic light scattering and kinetic analysis exclude the involvement of metal oxide nanoparticles and support our previously suggested homogeneous pathway.
Co-reporter:Stafford W. Sheehan, Heeso Noh, Gary W. Brudvig, Hui Cao, and Charles A. Schmuttenmaer
The Journal of Physical Chemistry C 2013 Volume 117(Issue 2) pp:927-934
Publication Date(Web):December 17, 2012
DOI:10.1021/jp311881k
We report the synthesis of core–shell–shell Au@SiO2@TiO2 nanostructures and demonstrate near-field plasmonic enhancement of dye-sensitized solar cells (DSSCs) incorporating them. Isolated nanoparticles as well as nanostructured plasmonic aggregates with broadband light absorption throughout the visible light region are developed. Comparisons to theoretical calculations are performed for the nanoparticles to provide further insight into their structure. We show that Au@SiO2@TiO2 nanoparticles provide efficiency enhancements greater than that of Au@SiO2 plasmonic nanoparticles and vary the distance between the molecular chromophore and the gold NP surface to demonstrate that this arises from near-field plasmonic effects. Finally, enhancement of dye absorption in DSSCs using a coupled plasmonic system is shown for the first time and results in a broadband enhancement of quantum efficiency.
Co-reporter:Lauren A. Martini, Gary F. Moore, Rebecca L. Milot, Lawrence Z. Cai, Stafford W. Sheehan, Charles A. Schmuttenmaer, Gary W. Brudvig, and Robert H. Crabtree
The Journal of Physical Chemistry C 2013 Volume 117(Issue 28) pp:14526-14533
Publication Date(Web):June 11, 2013
DOI:10.1021/jp4053456
Efforts to improve the ease of assembly and robustness of photoanodes for light-driven water oxidation have led to the development of a modular assembly method for anchoring high-potential zinc porphyrins to TiO2 via coordination to surface-bound pyridine linkers. It is essential that the anchoring groups provide strong electronic coupling between the molecular dye and metal oxide surface for optimal electron injection and that they are robust under the operating conditions of the system. Here, four linker molecules functionalized with either carboxylate, phosphonate, acetylacetonate, or hydroxamate anchoring groups are compared for their relative water stability on TiO2. We also report the relative electron injection efficiencies, as measured by terahertz spectroscopy, for high-potential zinc porphyrins coordinated to TiO2 via pyridyl linkers with the series of anchoring groups.
Co-reporter:Rebecca L. Milot ; Gary F. Moore ; Robert H. Crabtree ; Gary W. Brudvig ;Charles A. Schmuttenmaer
The Journal of Physical Chemistry C 2013 Volume 117(Issue 42) pp:21662-21670
Publication Date(Web):September 24, 2013
DOI:10.1021/jp406734t
The photoexcited electron injection dynamics of free-base and metallo-derivatives of tris(pentafluorophenyl)porphyrins bound to TiO2 and SnO2 nanoparticle surfaces have been investigated using time-resolved terahertz spectroscopy (TRTS). The metallo-derivatives include Zn(II), Cu(II), Ni(II), and Pd(II). For the TiO2–porphyrin assemblies, electron injection from the photoexcited dye to the semiconductor occurs only when using the zinc derivative as the sensitizer because it is the only dye studied in this report with long-lived excited states higher in energy than the TiO2 conduction band edge. All of the dyes, however, have excited-state energies above the SnO2 conduction band edge, and the electron injection rates vary widely from 0.4 to 200 ps depending on the sensitizer. For the SnO2–porphyrin assemblies, electron injection is strongly influenced by competition with alternate deactivation routes that are accessible following Soret band excitation. These results offer thermodynamic and kinetic considerations for designing improved high-potential porphyrin photoanodes with applications to solar-powered water oxidation.
Co-reporter:Robert C. Snoeberger ; III; Karin J. Young ; Jiji Tang ; Laura J. Allen ; Robert H. Crabtree ; Gary W. Brudvig ; Philip Coppens ; Victor; S. Batista ;Jason B. Benedict
Journal of the American Chemical Society 2012 Volume 134(Issue 21) pp:8911-8917
Publication Date(Web):May 1, 2012
DOI:10.1021/ja301238t
Interfacial electron transfer (IET) between a chromophore and a semiconductor nanoparticle is one of the key processes in a dye-sensitized solar cell. Theoretical simulations of the electron transfer in polyoxotitanate nanoclusters Ti17O24(OPri)20 (Ti17) functionalized with four p-nitrophenyl acetylacetone (NPA-H) adsorbates, of which the atomic structure has been fully established by X-ray diffraction measurements, are presented. Complementary experimental information showing IET has been obtained by EPR spectroscopy. Evolution of the time-dependent photoexcited electron during the initial 5 fs after instantaneous excitation to the NPA LUMO + 1 has been evaluated. Evidence for delocalization of the excitation over multiple chromophores after excitation to the NPA LUMO + 2 state on a 15 fs time scale is also obtained. While chromophores are generally considered electronically isolated with respect to neighboring sensitizers, our calculations show that this is not necessarily the case. The present work is the most comprehensive study to date of a sensitized semiconductor nanoparticle in which the structure of the surface and the mode of molecular adsorption are precisely defined.
Co-reporter:Karin J. Young, Lauren A. Martini, Rebecca L. Milot, Robert C. Snoeberger III, Victor S. Batista, Charles A. Schmuttenmaer, Robert H. Crabtree, Gary W. Brudvig
Coordination Chemistry Reviews 2012 Volume 256(21–22) pp:2503-2520
Publication Date(Web):November 2012
DOI:10.1016/j.ccr.2012.03.031
Light-driven water oxidation is an essential step for conversion of sunlight into storable chemical fuels. Fujishima and Honda reported the first example of photoelectrochemical water oxidation in 1972. In their system, TiO2 was irradiated with ultraviolet light, producing oxygen at the anode and hydrogen at a platinum cathode. Inspired by this system, more recent work has focused on functionalizing nanoporous TiO2 or other semiconductor surfaces with molecular adsorbates, including chromophores and catalysts that absorb visible light and generate electricity (i.e., dye-sensitized solar cells) or trigger water oxidation at low overpotentials (i.e., photocatalytic cells). The physics involved in harnessing multiple photochemical events for multi-electron reactions, as required in the four-electron water-oxidation process, has been the subject of much experimental and computational study. In spite of significant advances with regard to individual components, the development of highly efficient photocatalytic cells for solar water splitting remains an outstanding challenge. This article reviews recent progress in the field with emphasis on water-oxidation photoanodes inspired by the design of functionalized thin-film semiconductors of typical dye-sensitized solar cells.Graphical abstract.Highlights► Anodes for light-driven water oxidation. ► Design includes semiconductor, light-harvesting molecule, and catalyst. ► Integration of components is greatest challenge.
Co-reporter:Oana R. Luca, James D. Blakemore, Steven J. Konezny, Jeremy M. Praetorius, Timothy J. Schmeier, Glendon B. Hunsinger, Victor S. Batista, Gary W. Brudvig, Nilay Hazari, and Robert H. Crabtree
Inorganic Chemistry 2012 Volume 51(Issue 16) pp:8704-8709
Publication Date(Web):July 31, 2012
DOI:10.1021/ic300009a
Nonplatinum metals are needed to perform cost-effective water reduction electrocatalysis to enable technological implementation of a proposed hydrogen economy. We describe electrocatalytic proton reduction and H2 production by two organometallic nickel complexes with tridentate pincer ligands. The kinetics of H2 production from voltammetry is consistent with an overall third order rate law: the reaction is second order in acid and first order in catalyst. Hydrogen production with 90–95% Faradaic yields was confirmed by gas analysis, and UV–vis spectroscopy suggests that the ligand remains bound to the catalyst over the course of the reaction. A computational study provides mechanistic insights into the proposed catalytic cycle. Furthermore, two proposed intermediates in the proton reduction cycle were isolated in a representative system and show a catalytic response akin to the parent compound.
Co-reporter:James D. Blakemore, Nathan D. Schley, Maxwell N. Kushner-Lenhoff, Andrew M. Winter, Francis D’Souza, Robert H. Crabtree, and Gary W. Brudvig
Inorganic Chemistry 2012 Volume 51(Issue 14) pp:7749-7763
Publication Date(Web):June 22, 2012
DOI:10.1021/ic300764f
Electrodeposition of iridium oxide layers from soluble precursors provides a route to active thin-layer electrocatalysts for use on water-oxidizing anodes. Certain organometallic half-sandwich aqua complexes of iridium form stable and highly active oxide films upon electrochemical oxidation in aqueous solution. The catalyst films appear as blue layers on the anode when sufficiently thick, and most closely resemble hydrous iridium(III,IV) oxide by voltammetry. The deposition rate and cyclic voltammetric response of the electrodeposited material depend on whether the precursor complex contains a pentamethylcyclopentadieneyl (Cp*) or cyclopentadienyl ligand (Cp), and do not match, in either case, iridium oxide anodes prepared from non-organometallic precursors. Here, we survey our organometallic precursors, iridium hydroxide, and pre-formed iridium oxide nanoparticles. From electrochemical quartz crystal nanobalance (EQCN) studies, we find differences in the rate of electrodeposition of catalyst layers from the two half-sandwich precursors; however, the resulting layers operate as water-oxidizing anodes with indistinguishable overpotentials and H/D isotope effects. Furthermore, using the mass data collected by EQCN and not otherwise available, we show that the electrodeposited materials are excellent catalysts for the water-oxidation reaction, showing maximum turnover frequencies greater than 0.5 mol O2 (mol iridium)−1 s–1 and quantitative conversion of current to product dioxygen. Importantly, these anodes maintain their high activity and robustness at very low iridium loadings. Our organometallic precursors contrast with pre-formed iridium oxide nanoparticles, which form an unstable electrodeposited material that is not stably adherent to the anode surface at even moderately oxidizing potentials.
Co-reporter:Alexander R. Parent, Timothy P. Brewster, Wendy De Wolf, Robert H. Crabtree, and Gary W. Brudvig
Inorganic Chemistry 2012 Volume 51(Issue 11) pp:6147-6152
Publication Date(Web):May 15, 2012
DOI:10.1021/ic300154x
Sodium periodate was characterized as a primary chemical oxidant for the catalytic evolution of oxygen at neutral pH using a variety of water-oxidation catalysts. The visible spectra of solutions formed from Cp*Ir(bpy)SO4 during oxygen-evolution catalysis were measured. NMR spectroscopy suggests that the catalyst remains molecular after several turnovers with sodium periodate. Two of our [Cp*Ir(bis-NHC)][PF6]2 complexes, along with other literature catalysts, such as the manganese terpyridyl dimer, Hill’s cobalt polyoxometallate, and Meyer’s blue dimer, were also tested for activity. Sodium periodate was found to function only for water-oxidation catalysts with low overpotentials. This specificity is attributed to the relatively low oxidizing capability of sodium periodate solutions relative to solutions of other common primary oxidants. Studying oxygen-evolution catalysis by using sodium periodate as a primary oxidant may, therefore, provide preliminary evidence that a given catalyst has a low overpotential.
Co-reporter:Yunlong Gao ; Robert H. Crabtree
Inorganic Chemistry 2012 Volume 51(Issue 7) pp:4043-4050
Publication Date(Web):March 21, 2012
DOI:10.1021/ic2021897
The tetranuclear manganese complex [MnIV4O5(terpy)4(H2O)2](ClO4)6 (1; terpy = 2,2′:6′,2″-terpyridine) gives catalytic water oxidation in aqueous solution, as determined by electrochemistry and GC-MS. Complex 1 also exhibits catalytic water oxidation when adsorbed on kaolin clay, with CeIV as the primary oxidant. The redox intermediates of complex 1 adsorbed on kaolin clay upon addition of CeIV have been characterized by using diffuse reflectance UV/visible and EPR spectroscopy. One of the products in the reaction on kaolin clay is MnIII, as determined by parallel-mode EPR spectroscopic studies. When 1 is oxidized in aqueous solution with CeIV, the reaction intermediates are unstable and decompose to form MnII, detected by EPR spectroscopy, and MnO2. DFT calculations show that the oxygen in the mono-μ-oxo bridge, rather than MnIV, is oxidized after an electron is removed from the Mn(IV,IV,IV,IV) tetramer. On the basis of the calculations, the formation of O2 is proposed to occur by reaction of water with an electrophilic manganese-bound oxyl radical species, •O–Mn2IV/IV, produced during the oxidation of the tetramer. This study demonstrates that [MnIV4O5(terpy)4(H2O)2](ClO4)6 may be relevant for understanding the role of the Mn tetramer in photosystem II.
Co-reporter:James D. Blakemore, Jonathan F. Hull, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2012 vol. 41(Issue 25) pp:7681-7688
Publication Date(Web):17 Apr 2012
DOI:10.1039/C2DT30371J
The speciation behavior of a water-soluble manganese(III) tetrasulfonated phthalocyanine complex was investigated with UV-visible and electron paramagnetic resonance (EPR) spectroscopies, as well as cyclic voltammetry. Parallel-mode EPR (in dimethylformamide:pyridine solvent mix) reveals a six-line hyperfine signal, centered at a g-value of 8.8, for the manganese(III) monomer, characteristic of the d4S = 2 system. The color of an aqueous solution containing the complex is dependent upon the pH of the solution; the phthalocyanine complex can exist as a water-bound monomer, a hydroxide-bound monomer, or an oxo-bridged dimer. Addition of coordinating bases such as borate or pyridine changes the speciation behavior by coordinating the manganese center. From the UV-visible spectra, complete speciation diagrams are plotted by global analysis of the pH-dependent UV-visible spectra, and a complete set of pKa values is obtained by fitting the data to a standard pKa model. Electrochemical studies reveal a pH-independent quasi-reversible oxidation event for the monomeric species, which likely involves oxidation of the organic ligand to the radical cation species. Adsorption of the phthalocyanine complex on the carbon working electrode was sometimes observed. The pKa values and electrochemistry data are discussed in the context of the development of mononuclear water-oxidation catalysts.
Co-reporter:Oana R. Luca, Steven J. Konezny, James D. Blakemore, Dominic M. Colosi, Shubhro Saha, Gary W. Brudvig, Victor S. Batista and Robert H. Crabtree
New Journal of Chemistry 2012 vol. 36(Issue 5) pp:1149-1152
Publication Date(Web):19 Mar 2012
DOI:10.1039/C2NJ20912H
A NiII complex with a redox-active pincer ligand reduces protons at a low overpotential in aqueous acidic conditions. A combined experimental and computational study provides mechanistic insights into a putative catalytic cycle.
Co-reporter:Jonathan Graeupner, Timothy P. Brewster, James D. Blakemore, Nathan D. Schley, Julianne M. Thomsen, Gary W. Brudvig, Nilay Hazari, and Robert H. Crabtree
Organometallics 2012 Volume 31(Issue 20) pp:7158-7164
Publication Date(Web):October 3, 2012
DOI:10.1021/om300696t
Cp*IrIII and CpIrIII complexes have attracted interest as catalysts for oxidative transformations, and highly oxidizing iridium species are postulated as key intermediates in both catalytic water and C–H bond oxidation. Strongly electron-donating ligand sets have been shown to stabilize IrIV complexes. We describe the synthesis and reactivity of complexes containing the CpIr(biphenyl-2,2′-diyl) moiety stabilized by a series of strong donor carbon-based ligands. The oxidation chemistry of these complexes has been characterized electrochemically, and a singly oxidized IrIV species has been observed by X-band EPR for the complex CpIr(biph)(p-CNCH2SO2C6H4CH3).
Co-reporter:Gary F. Moore ; Steven J. Konezny ; Hee-eun Song ; Rebecca L. Milot ; James D. Blakemore ; Minjoo L. Lee ; Victor S. Batista ; Charles A. Schmuttenmaer ; Robert H. Crabtree
The Journal of Physical Chemistry C 2012 Volume 116(Issue 7) pp:4892-4902
Publication Date(Web):January 23, 2012
DOI:10.1021/jp210096m
We report a selection of high-potential porphyrin photoanodes (HPPPs) for use in photoelectrochemical cells (PECs). The anodes consist of bispentafluorophenyl free-base and metallo-porphyrin sensitizers bearing anchoring groups for attachment to metal-oxide surfaces including TiO2 and SnO2 nanoparticles. The term “high potential” refers to the relatively large and positive value of the electrochemical reduction potential for the bispentafluorophenyl porphyrin radical cation (P•+ + e– → P) as compared with more conventional nonfluorinated analogues. Photoelectrochemical measurements demonstrate the sensitizers used in these HPPPs extend the absorption of the bare anode well into the visible region. Terahertz spectroscopic studies show the photoexcited dyes are capable of injecting electrons into the conduction band of an underlying metal-oxide with appropriate energetics. The reduction potentials of the resulting photogenerated porphyrin radical cations are relatively high (ranging from ∼1.35 to 1.65 V vs NHE depending on the sensitizer). This is demonstrated by the ability of dye-sensitized solar cells, containing our HPPPs, to use the Br3–/Br– redox couple as a regenerative electron mediator with superior performance in comparison to results obtained using the lower-potential I3–/I– relay. Computational modeling of the structures and equivalent circuits assists in a molecular-based understanding of these systems. Further, the oxidation power of the porphyrin radical cations generated in these bioinspired constructs is similar to that found in the reaction centers of their natural counterpart (photosystem II); thus, HPPPs are promising as components in artificial systems for photochemical water spitting applications.
Co-reporter:Gary F. Moore, James D. Blakemore, Rebecca L. Milot, Jonathan F. Hull, Hee-eun Song, Lawrence Cai, Charles A. Schmuttenmaer, Robert H. Crabtree and Gary W. Brudvig
Energy & Environmental Science 2011 vol. 4(Issue 7) pp:2389-2392
Publication Date(Web):11 May 2011
DOI:10.1039/C1EE01037A
A high-potential porphyrin is codeposited on TiO2 nanoparticles together with our Cp*–iridium water-oxidation catalyst to give a photoanode for a water-splitting cell. The photoanode optically resembles the porphyrin yet electrochemically responds like the Ir catalyst when it is immersed in aqueous solutions. Photoelectrochemical data show that illumination of the codeposited anode in water results in a marked enhancement and stability of the photocurrent, providing evidence for light-induced activation of the catalyst.
Co-reporter:Nathan D. Schley ; James D. Blakemore ; Navaneetha K. Subbaiyan ; Christopher D. Incarvito ; Francis D’Souza ; Robert H. Crabtree
Journal of the American Chemical Society 2011 Volume 133(Issue 27) pp:10473-10481
Publication Date(Web):June 15, 2011
DOI:10.1021/ja2004522
Molecular water-oxidation catalysts can deactivate by side reactions or decompose to secondary materials over time due to the harsh, oxidizing conditions required to drive oxygen evolution. Distinguishing electrode surface-bound heterogeneous catalysts (such as iridium oxide) from homogeneous molecular catalysts is often difficult. Using an electrochemical quartz crystal nanobalance (EQCN), we report a method for probing electrodeposition of metal oxide materials from molecular precursors. Using the previously reported [Cp*Ir(H2O)3]2+ complex, we monitor deposition of a heterogeneous water oxidation catalyst by measuring the electrode mass in real time with piezoelectric gravimetry. Conversely, we do not observe deposition for homogeneous catalysts, such as the water-soluble complex Cp*Ir(pyr-CMe2O)X reported in this work. Rotating ring-disk electrode electrochemistry and Clark-type electrode studies show that this complex is a catalyst for water oxidation with oxygen produced as the product. For the heterogeneous, surface-attached material generated from [Cp*Ir(H2O)3]2+, we can estimate the percentage of electroactive metal centers in the surface layer. We monitor electrode composition dynamically during catalytic turnover, providing new information on catalytic performance. Together, these data suggest that EQCN can directly probe the homogeneity of molecular water-oxidation catalysts over short times.
Co-reporter:Gözde Ulas
Journal of the American Chemical Society 2011 Volume 133(Issue 34) pp:13260-13263
Publication Date(Web):August 2, 2011
DOI:10.1021/ja2049226
A negatively charged region on the surface of photosystem II (PSII) near QA has been identified as a docking site for cationic exogenous electron acceptors. Oxygen evolution activity, which is inhibited in the presence of the herbicide 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), is recovered by adding CoIII complexes. Thus, a new electron-transfer pathway is created with CoIII as the new terminal electron acceptor from QA–. This binding site is saturated at ∼2.5 mM [CoIII], which is consistent with the existence of low-affinity interactions with a solvent-exposed surface. This is the first example of a higher plant PSII in which the electron-transfer pathway has been redirected from the normal membrane-associated quinone electron acceptors to water-soluble electron acceptors. The proposed CoIII binding site may enable efficient collection of electrons generated from photochemical water oxidation by PSII immobilized on an electrode surface.
Co-reporter:Iain L. McConnell ; Vladimir M. Grigoryants ; Charles P. Scholes ; William K. Myers ; Ping-Yu Chen ; James W. Whittaker
Journal of the American Chemical Society 2011 Volume 134(Issue 3) pp:1504-1512
Publication Date(Web):December 5, 2011
DOI:10.1021/ja203465y
The synthesis of efficient water-oxidation catalysts demands insight into the only known, naturally occurring water-oxidation catalyst, the oxygen-evolving complex (OEC) of photosystem II (PSII). Understanding the water oxidation mechanism requires knowledge of where and when substrate water binds to the OEC. Mn catalase in its Mn(III)–Mn(IV) state is a protein model of the OEC’s S2 state. From 17O-labeled water exchanged into the di-μ-oxo di-Mn(III,IV) coordination sphere of Mn catalase, CW Q-band ENDOR spectroscopy revealed two distinctly different 17O signals incorporated in distinctly different time regimes. First, a signal appearing after 2 h of 17O exchange was detected with a 13.0 MHz hyperfine coupling. From similarity in the time scale of isotope incorporation and in the 17O μ-oxo hyperfine coupling of the di-μ-oxo di-Mn(III,IV) bipyridine model (Usov, O. M.; Grigoryants, V. M.; Tagore, R.; Brudvig, G. W.; Scholes, C. P. J. Am. Chem. Soc. 2007,, 129, 11886−11887), this signal was assigned to μ-oxo oxygen. EPR line broadening was obvious from this 17O μ-oxo species. Earlier exchange proceeded on the minute or faster time scale into a non-μ-oxo position, from which 17O ENDOR showed a smaller 3.8 MHz hyperfine coupling and possible quadrupole splittings, indicating a terminal water of Mn(III). Exchangeable proton/deuteron hyperfine couplings, consistent with terminal water ligation to Mn(III), also appeared. Q-band CW ENDOR from the S2 state of the OEC was obtained following multihour 17O exchange, which showed a 17O hyperfine signal with a 11 MHz hyperfine coupling, tentatively assigned as μ-oxo-17O by resemblance to the μ-oxo signals from Mn catalase and the di-μ-oxo di-Mn(III,IV) bipyridine model.
Co-reporter:James D. Blakemore, Nathan D. Schley, Gerard W. Olack, Christopher D. Incarvito, Gary W. Brudvig and Robert H. Crabtree
Chemical Science 2011 vol. 2(Issue 1) pp:94-98
Publication Date(Web):28 Oct 2010
DOI:10.1039/C0SC00418A
Artificial photosynthesis, modeled on natural light-driven oxidation of water in Photosystem II, holds promise as a sustainable source of reducing equivalents for producing fuels. Few robust water-oxidation catalysts capable of mediating this difficult four-electron, four-proton reaction have yet been described. We report a new method for generating an amorphous electrodeposited material, principally consisting of iridium and oxygen, which is a robust and long-lived catalyst for water oxidation, when driven electrochemically. The catalyst material is generated by a simple anodic deposition from Cp*Ir aqua or hydroxo complexes in aqueous solution. This work suggests that organometallic precursors may be useful in electrodeposition of inorganic heterogeneous catalysts.
Co-reporter:Alexander R. Parent, James D. Blakemore, Gary W. Brudvig and Robert H. Crabtree
Chemical Communications 2011 vol. 47(Issue 42) pp:11745-11747
Publication Date(Web):28 Sep 2011
DOI:10.1039/C1CC15501F
The catalytic water-oxidation activity of Wilkinson's iridium acetate trimer (1) has been characterized electrochemically and by using chemical oxidants. We show that 1 can function as an operationally homogeneous water-oxidation catalyst when driven with sodium periodate as a primary oxidant, but rapidly decomposes using Ce(IV) as a primary oxidant.
Co-reporter:Ravi Pokhrel, Iain L. McConnell, and Gary W. Brudvig
Biochemistry 2011 Volume 50(Issue 14) pp:
Publication Date(Web):March 2, 2011
DOI:10.1021/bi2000388
Chloride-dependent α-amylases, angiotensin-converting enzyme (ACE), and photosystem II (PSII) are activated by bound chloride. Chloride-binding sites in these enzymes contain a positively charged Arg or Lys residue crucial for chloride binding. In α-amylases and ACE, removal of chloride from the binding site triggers formation of a salt bridge between the positively charged Arg or Lys residue involved in chloride binding and a nearby carboxylate residue. The mechanism for chloride activation in ACE and chloride-dependent α-amylases is 2-fold: (i) correctly positioning catalytic residues or other residues involved in stabilizing the enzyme−substrate complex and (ii) fine-tuning of the pKa of a catalytic residue. By using examples of how chloride activates α-amylases and ACE, we can gain insight into the potential mechanisms by which chloride functions in PSII. Recent structural evidence from cyanobacterial PSII indicates that there is at least one chloride-binding site in the vicinity of the oxygen-evolving complex (OEC). Here we propose that, in the absence of chloride, a salt bridge between D2:K317 and D1:D61 (and/or D1:E333) is formed. This can cause a conformational shift of D1:D61 and lower the pKa of this residue, making it an inefficient proton acceptor during the S-state cycle. Movement of the D1:E333 ligand and the adjacent D1:H332 ligand due to chloride removal could also explain the observed change in the magnetic properties of the manganese cluster in the OEC upon chloride depletion.
Co-reporter:Ivan Rivalta, Muhamed Amin, Sandra Luber, Serguei Vassiliev, Ravi Pokhrel, Yasufumi Umena, Keisuke Kawakami, Jian-Ren Shen, Nobuo Kamiya, Doug Bruce, Gary W. Brudvig, M. R. Gunner, and Victor S. Batista
Biochemistry 2011 Volume 50(Issue 29) pp:
Publication Date(Web):June 16, 2011
DOI:10.1021/bi200685w
Chloride binding in photosystem II (PSII) is essential for photosynthetic water oxidation. However, the functional roles of chloride and possible binding sites, during oxygen evolution, remain controversial. This paper examines the functions of chloride based on its binding site revealed in the X-ray crystal structure of PSII at 1.9 Å resolution. We find that chloride depletion induces formation of a salt bridge between D2-K317 and D1-D61 that could suppress the transfer of protons to the lumen.
Co-reporter:Steven J. Konezny, Christiaan Richter, Robert C. Snoeberger III, Alexander R. Parent, Gary W. Brudvig, Charles A. Schmuttenmaer, and Victor S. Batista
The Journal of Physical Chemistry Letters 2011 Volume 2(Issue 15) pp:1931-1936
Publication Date(Web):July 14, 2011
DOI:10.1021/jz200853v
The electronic mechanisms responsible for dark conductivity in nanoporous TiO2 thin films remain only partially understood, although they control the efficiency of charge transport in a wide range of technological applications. Measurements in the 78–335 K temperature range show DC conductivity values spanning over 4 orders of magnitude, with a high-temperature Arrhenius dependence that gradually changes into a temperature-independent plateau at low temperatures. We show evidence that a fluctuation-induced tunneling conductivity (FITC) mechanism is fully consistent with the experimental data. Quantitative agreement is demonstrated for the entire temperature range (T = 78–335 K) with a FITC model parametrized according to atomistic models of nanoporous TiO2 and the characterization of the films by X-ray diffraction and scanning electron microscopy measurements. These findings suggest that dark DC conductivity in nanoporous TiO2 films depends strongly on the properties of the junctions linking the constituent nanoparticles.Keywords: charge transport; chemical sensors; DC conductivity; nanoporous TiO2; photocatalytic cells; solar cells; tunneling;
Co-reporter:Timothy P. Brewster, James D. Blakemore, Nathan D. Schley, Christopher D. Incarvito, Nilay Hazari, Gary W. Brudvig, and Robert H. Crabtree
Organometallics 2011 Volume 30(Issue 5) pp:965-973
Publication Date(Web):February 8, 2011
DOI:10.1021/om101016s
The Ir precatalyst (3) contains both a Cp* and a κ2C2,C2′-1,3-diphenylimidazol-2-ylidene ligand, a C−C chelate, where one C donor is the NHC and the other is a cyclometalated N-phenyl wingtip group. The structure of 3 was confirmed by X-ray crystallography. Like our other recently described Cp*Ir catalysts, this compound is a precursor to a catalyst that can oxidize water to dioxygen. Electrochemical characterization of the new compound shows that it has a stable iridium(IV) oxidation state, [Cp*IrIV(NHC)Cl]+, in contrast with the unstable Ir(IV) state seen in our previous cyclometalated [Cp*IrIII(2-pyridyl-2′-phenyl)Cl] catalyst. The new iridium(IV) species has been characterized by EPR spectroscopy and has a rhombic symmetry, a consequence of the ligand environment. These results both support previous studies which suggest that Cp*Ir catalysts can be advanced through the relevant catalytic cycle(s) in one-electron steps and help clarify the electrochemical behavior of this class of water-oxidation catalysts.
Co-reporter:William R. McNamara, Rebecca L. Milot, Hee-eun Song, Robert C. Snoeberger III, Victor S. Batista, Charles A. Schmuttenmaer, Gary W. Brudvig and Robert H. Crabtree
Energy & Environmental Science 2010 vol. 3(Issue 7) pp:917-923
Publication Date(Web):26 Apr 2010
DOI:10.1039/C001065K
A novel class of derivatized hydroxamic acid linkages for robust sensitization of TiO2 nanoparticles (NPs) under various aqueous conditions is described. The stability of linkages bound to metal oxides under various conditions is important in developing photocatalytic cells which incorporate transition metal complexes for solar energy conversion. In order to compare the standard carboxylate anchor to hydroxamates, two organic dyes differing only in anchoring groups were synthesized and attached to TiO2 NPs. At acidic, basic, and close to neutral pH, hydroxamic acid linkages resist detachment compared to the labile carboxylic acids. THz spectroscopy was used to compare ultrafast interfacial electron transfer (IET) into the conduction band of TiO2 for both linkages and found similar IET characteristics. Observable electron injection and stronger binding suggest that hydroxamates are a suitable class of anchors for designing water stable molecules for functionalizing TiO2.
Co-reporter:Jonathan F. Hull ; David Balcells ; Effiette L. O. Sauer ; Christophe Raynaud ; Gary W. Brudvig ; Robert H. Crabtree ;Odile Eisenstein
Journal of the American Chemical Society 2010 Volume 132(Issue 22) pp:7605-7616
Publication Date(Web):May 18, 2010
DOI:10.1021/ja908744w
We describe competitive C−H bond activation chemistry of two types, desaturation and hydroxylation, using synthetic manganese catalysts with several substrates. 9,10-Dihydrophenanthrene (DHP) gives the highest desaturation activity, the final products being phenanthrene (P1) and phenanthrene 9,10-oxide (P3), the latter being thought to arise from epoxidation of some of the phenanthrene. The hydroxylase pathway also occurs as suggested by the presence of the dione product, phenanthrene-9,10-dione (P2), thought to arise from further oxidation of hydroxylation intermediate 9-hydroxy-9,10-dihydrophenanthrene. The experimental work together with the density functional theory (DFT) calculations shows that the postulated Mn oxo active species, [Mn(O)(tpp)(Cl)] (tpp = tetraphenylporphyrin), can promote the oxidation of dihydrophenanthrene by either desaturation or hydroxylation pathways. The calculations show that these two competing reactions have a common initial step, radical H abstraction from one of the DHP sp3 C−H bonds. The resulting Mn hydroxo intermediate is capable of promoting not only OH rebound (hydroxylation) but also a second H abstraction adjacent to the first (desaturation). Like the active MnV═O species, this MnIV−OH species also has radical character on oxygen and can thus give H abstraction. Both steps have very low and therefore very similar energy barriers, leading to a product mixture. Since the radical character of the catalyst is located on the oxygen p orbital perpendicular to the MnIV−OH plane, the orientation of the organic radical with respect to this plane determines which reaction, desaturation or hydroxylation, will occur. Stereoelectronic factors such as the rotational orientation of the OH group in the enzyme active site are thus likely to constitute the switch between hydroxylase and desaturase behavior.
Co-reporter:James D. Blakemore ; Nathan D. Schley ; David Balcells ; Jonathan F. Hull ; Gerard W. Olack ; Christopher D. Incarvito ; Odile Eisenstein ; Gary W. Brudvig ;Robert H. Crabtree
Journal of the American Chemical Society 2010 Volume 132(Issue 45) pp:16017-16029
Publication Date(Web):October 21, 2010
DOI:10.1021/ja104775j
Iridium half-sandwich complexes of the types Cp*Ir(N−C)X, [Cp*Ir(N−N)X]X, and [CpIr(N−N)X]X are catalyst precursors for the homogeneous oxidation of water to dioxygen. Kinetic studies with cerium(IV) ammonium nitrate as primary oxidant show that oxygen evolution is rapid and continues over many hours. In addition, [Cp*Ir(H2O)3]SO4 and [(Cp*Ir)2(μ-OH)3]OH can show even higher turnover frequencies (up to 20 min−1 at pH 0.89). Aqueous electrochemical studies on the cationic complexes having chelate ligands show catalytic oxidation at pH > 7; conversely, at low pH, there are no oxidation waves up to 1.5 V vs NHE for the complexes. H218O isotope incorporation studies demonstrate that water is the source of oxygen atoms during cerium(IV)-driven catalysis. DFT calculations and kinetic experiments, including kinetic-isotope-effect studies, suggest a mechanism for homogeneous iridium-catalyzed water oxidation and contribute to the determination of the rate-determining step. The kinetic experiments also help distinguish the active homogeneous catalyst from heterogeneous nanoparticulate iridium dioxide.
Co-reporter:Clyde W. Cady, Katherine E. Shinopoulos, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2010 vol. 39(Issue 16) pp:3985-3989
Publication Date(Web):17 Mar 2010
DOI:10.1039/B922087A
Photosynthetic water oxidation occurs naturally at a tetranuclear manganese center in the photosystem II protein complex. Synthetically mimicking this tetramanganese center, known as the oxygen-evolving complex (OEC), has been an ongoing challenge of bioinorganic chemistry. Most past efforts have centered on water-oxidation catalysis using chemical oxidants. However, solar energy applications have drawn attention to electrochemical methods. In this paper, we examine the electrochemical behavior of the biomimetic water-oxidation catalyst [(H2O)(terpy)Mn(μ-O)2Mn(terpy)(H2O)](NO3)3 [terpy = 2,2′:6′,2′′-terpyridine] (1) in water under a variety of pH and buffered conditions and in the presence of acetate that binds to 1 in place of one of the terminal water ligands. These experiments show that 1 not only exhibits proton-coupled electron-transfer reactivity analogous to the OEC, but also may be capable of electrochemical oxidation of water to oxygen.
Co-reporter:Gözde Ulas and Gary W. Brudvig
Biochemistry 2010 Volume 49(Issue 37) pp:
Publication Date(Web):August 13, 2010
DOI:10.1021/bi101027a
Photosystem II (PSII) is the only enzyme in nature that can catalyze the challenging catalytic photooxidation of H2O into four protons, four electrons, and O2. Slowing down turnover of the O2-evolving complex (OEC) is a plausible approach to gain mechanistic information on the reaction. However, modulating the kinetics of the reaction without perturbing the active site is a challenge. In this study, it is shown that the steady-state activity of cyanobacterial PSII is inhibited by small zwitterions, such as glycine betaine and β-alanine. We show that the binding of zwitterions is nondenaturing, is highly reversible, and results in the decrease of the rate of catalytic turnover by ∼50% in the presence of excess zwitterion. Control measurements of photoinduced electron transfer in O2-inactive PSII show that the inhibition by zwitterions is the result of a specific decrease in the rate of catalytic turnover of the OEC. Recovery of activity upon addition of an exogenous proton carrier (HCO3−) provides evidence that proton-transfer pathways, thought to be essential for the relay of protons from the OEC to the lumen, are affected. Interestingly, no inhibition is observed for spinach PSII, suggesting that zwitterions act specifically by binding to the extrinsic proteins on the lumenal side of PSII, which differ significantly between plants and cyanobacteria, to slow proton transfer on the electron donor side of PSII.
Co-reporter:Gonghu Li, Eduardo M. Sproviero, William R. McNamara, Robert C. Snoeberger III, Robert H. Crabtree, Gary W. Brudvig, and Victor S. Batista
The Journal of Physical Chemistry B 2010 Volume 114(Issue 45) pp:14214-14222
Publication Date(Web):November 19, 2009
DOI:10.1021/jp908925z
Several polynuclear transition-metal complexes, including our own dinuclear di-μ-oxo manganese compound [H2O(terpy)MnIII(μ-O)2MnIV(terpy)H2O](NO3)3 (1, terpy = 2,2′:6′,2′′-terpyridine), have been reported to be homogeneous catalysts for water oxidation. This paper reports the covalent attachment of 1 onto nanoparticulate TiO2 surfaces using a robust chromophoric linker L. L, a phenylterpy ligand attached to a 3-phenyl-acetylacetonate anchoring moiety via an amide bond, absorbs visible light and leads to photoinduced interfacial electron transfer into the TiO2 conduction band. We characterize the electronic and structural properties of the 1−L−TiO2 assemblies by using a combination of methods, including computational modeling and UV−visible, IR, and EPR spectroscopies. We show that the Mn(III,IV) state of 1 can be reversibly advanced to the Mn(IV,IV) state by visible-light photoexcitation of 1−L−TiO2 nanoparticles (NPs) and recombines back to the Mn(III,IV) state in the dark, in the absence of electron scavengers. Our findings also indicate that a high degree of crystallinity of the TiO2 NPs is essential for promoting photooxidation of the adsorbates by photoinduced charge separation when the TiO2 NPs serve as electron acceptors in artificial photosynthetic assemblies. The reported results are particularly relevant to the development of photocatalytic devices for oxidation chemistry based on inexpensive materials (e.g., TiO2 and Mn complexes) that are robust under aqueous and oxidative conditions.
Co-reporter:Gonghu Li, Eduardo M. Sproviero, Robert C. Snoeberger III, Nobuhito Iguchi, James D. Blakemore, Robert H. Crabtree, Gary W. Brudvig and Victor S. Batista
Energy & Environmental Science 2009 vol. 2(Issue 2) pp:230-238
Publication Date(Web):12 Jan 2009
DOI:10.1039/B818708H
Inexpensive water oxidation catalysts are needed to develop photocatalytic solar cells that mimic photosynthesis and produce fuel from sunlight and water. This paper reports the successful attachment of a dinuclear di-µ-oxo manganese water oxidation catalyst [H2O(terpy)MnIII(µ-O)2 MnIV(terpy)H2O](NO3)3 (1, terpy = 2,2′:6′2″-terpyridine) onto TiO2 nanoparticles (NPs) via direct adsorption, or in situ synthesis. The resulting surface complexes are characterized by EPR and UV-visible spectroscopy, electrochemical measurements and computational modeling. We conclude that the mixed-valence (III,IV) state of 1 attaches to near-amorphous TiO2 NPs by substituting one of its water ligands by the TiO2 NP, as suggested by low-temperature (7 K) EPR data. In contrast, the analogous attachment onto well-crystallized TiO2 NPs leads to dimerization of 1 forming Mn(IV) tetramers on the TiO2 surface as suggested by EPR spectroscopy and electrochemical studies.
Co-reporter:William R. McNamara, Robert C. Snoeberger III, Gonghu Li, Christiaan Richter, Laura J. Allen, Rebecca L. Milot, Charles A. Schmuttenmaer, Robert H. Crabtree, Gary W. Brudvig and Victor S. Batista
Energy & Environmental Science 2009 vol. 2(Issue 11) pp:1173-1175
Publication Date(Web):07 Aug 2009
DOI:10.1039/B910241H
A graphical abstract is available for this content
Co-reporter:Jonathan F. Hull ; David Balcells ; James D. Blakemore ; Christopher D. Incarvito ; Odile Eisenstein ; Gary W. Brudvig ;Robert H. Crabtree
Journal of the American Chemical Society 2009 Volume 131(Issue 25) pp:8730-8731
Publication Date(Web):June 4, 2009
DOI:10.1021/ja901270f
A series of Cp*Ir catalysts are the most active known by over an order of magnitude for water oxidation with Ce(IV). DFT calculations support a Cp*Ir═O complex as an active species.
Co-reporter:Jonathan F. Hull, Effiette L. O. Sauer, Christopher D. Incarvito, J. W. Faller, Gary W. Brudvig and Robert H. Crabtree
Inorganic Chemistry 2009 Volume 48(Issue 2) pp:488-495
Publication Date(Web):December 18, 2008
DOI:10.1021/ic8013464
Selective epoxidation of alkenes is possible with a new manganese porphyrin catalyst, CPMR, that uses hydrogen bonding between the carboxylic acid on the substrate molecule and a Kemp’s triacid unit. For two out of three olefin substrates employed, molecular recognition prevents the unselective oxidation of C−H bonds, and directs oxidation to the olefin moiety, giving only epoxide products. Weak diastereoselectivity is observed in the epoxide products, suggesting that molecular recognition affects the orientation of the catalyst-bound substrate. The previously reported manganese terpyridine complex CTMR is shown to be a superior epoxidation catalyst to the porphyrin catalyst CPMR. Good conversion of 2-cyclopentene acetic acid (substrate S2) with CPMR is consistent with molecular modeling, which indicates a particularly good substrate/catalyst match. Evidence suggests that hydrogen bonding between the substrate and the catalyst is critical in this system.
Co-reporter:Gonghu Li, Christiaan P. Richter, Rebecca L. Milot, Lawrence Cai, Charles A. Schmuttenmaer, Robert H. Crabtree, Gary W. Brudvig and Victor S. Batista
Dalton Transactions 2009 (Issue 45) pp:10078-10085
Publication Date(Web):02 Sep 2009
DOI:10.1039/B908686B
A synergistic effect between anatase and rutile TiO2 is known, in which the addition of rutile can remarkably enhance the photocatalytic activity of anatase in the degradation of organic contaminants. In this study, mixed-phase TiO2 nanocomposites consisting of anatase and rutile nanoparticles (NPs) were prepared for use as photoanodes in dye-sensitized solar cells (DSSCs) and were characterized by using UV-vis spectroscopy, powder X-ray diffraction and scanning electron microscopy. The addition of 10–15% rutile significantly improved light harvesting and the overall solar conversion efficiency of anatase NPs in DSSCs. The underlying mechanism for the synergistic effect in DSSCs is now explored by using time-resolved terahertz spectroscopy. It is clearly demonstrated that photo-excited electrons injected into the rutile NPs can migrate to the conduction band of anatase NPs, enhancing the photocurrent and efficiency. Interfacial electron transfer from rutile to anatase, similar to that in heterogeneous photocatalysis, is proposed to account for the synergistic effect in DSSCs. Our results further suggest that the synergistic effect can be used to explain the beneficial effect of TiCl4 treatment on DSSC efficiency.
Co-reporter:Siddhartha Das, Gary W. Brudvig, Robert H. Crabtree
Inorganica Chimica Acta 2009 Volume 362(Issue 4) pp:1229-1233
Publication Date(Web):2 March 2009
DOI:10.1016/j.ica.2008.06.009
We report a rapid method for assembling our di-μ-oxo dimanganese catalyst, verified by ESI-MS and EPR, assessing its water oxidation activity by a Clark electrode O2-assay study and its regioselective C–H activation activity by product analysis in catalytic runs.A rapid method is described for assembling a CH activation catalyst that uses molecular recognition to attain high selectivity.
Co-reporter:Yunlong Gao, Katherine E. Shinopoulos, Cara A. Tracewell, A. Ligia Focsan, Gary W. Brudvig and Lowell D. Kispert
The Journal of Physical Chemistry B 2009 Volume 113(Issue 29) pp:9901-9908
Publication Date(Web):June 24, 2009
DOI:10.1021/jp8075832
β-Carotene radicals produced in the hexagonal pores of the molecular sieve Cu(II)−MCM-41 were studied by ENDOR and visible/near-IR spectroscopies. ENDOR studies showed that neutral radicals of β-carotene were produced in humid air under ambient fluorescent light. The maximum absorption wavelengths of the neutral radicals were measured and were additionally predicted by using time-dependent density functional theory (TD-DFT) calculations. An absorption peak at 750 nm, assigned to the neutral radical with a proton loss from the 4(4′) position of the β-carotene radical cation in Cu(II)−MCM-41, was also observed in photosystem II (PS II) samples using near-IR spectroscopy after illumination at 20 K. This peak was previously unassigned in PS II samples. The intensity of the absorption peak at 750 nm relative to the absorption of chlorophyll radical cations and β-carotene radical cations increased with increasing pH of the PS II sample, providing further evidence that the absorption peak is due to the deprotonation of the β-carotene radical cation. Based on a consideration of possible proton acceptors that are adjacent to β-carotene molecules in photosystem II, as modeled in the X-ray crystal structure of Guskov et al. Nat. Struct. Mol. Biol. 2009, 16, 334−342, an electron-transfer pathway from a β-carotene molecule with an adjacent proton acceptor to P680•+ is proposed.
Co-reporter:Ranitendranath Tagore ; Robert H. Crabtree
Inorganic Chemistry 2008 Volume 47(Issue 6) pp:1815-1823
Publication Date(Web):March 10, 2008
DOI:10.1021/ic062218d
[Mn2III/IV(μ-O)2(terpy)2(OH2)2](NO3)3 (1, where terpy = 2,2′:6′2′′-terpyridine) acts as a water-oxidation catalyst with HSO5− as the primary oxidant in aqueous solution and, thus, provides a model system for the oxygen-evolving complex of photosystem II (Limburg, J.; et al. J. Am. Chem. Soc. 2001, 123, 423–430). The majority of the starting [Mn2III/IV(μ-O)2]3+ complex is converted to the[Mn2IV/IV(μ-O)2]4+ form (2) during this reaction (Chen, H.; et al. Inorg. Chem. 2007, 46, 34–43). Here, we have used stopped-flow UV–visible spectroscopy to monitor UV–visible absorbance changes accompanying the conversion of 1 to 2 by HSO5−. With excess HSO5−, the rate of absorbance change was found to be first-order in [1] and nearly zero-order in [HSO5−]. At relatively low [HSO5−], the change of absorbance with time is distinctly biphasic. The observed concentration dependences are interpreted in terms of a model involving the two-electron oxidation of 1 by HSO5−, followed by the rapid reaction of the two-electron-oxidized intermediate with another molecule of 1 to give two molecules of 2. In order to rationalize biphasic behavior at low [HSO5−], we propose a difference in reactivity of the [Mn2III/IV(μ-O)2]3+ complex upon binding of HSO5− to the MnIII site as compared to the reactivity upon binding HSO5− to the MnIV site. The kinetic distinctness of the MnIII and MnIV sites allows us to estimate upper limits for the rates of intramolecular electron transfer and terminal ligand exchange between these sites. The proposed mechanism leads to insights on the optimization of 1 as a water-oxidation catalyst. The rates of terminal ligand exchange and electron transfer between oxo-bridged Mn atoms in the oxygen-evolving complex of photosystem II are discussed in light of these results.
Co-reporter:James P. McEvoy and Gary W. Brudvig
Biochemistry 2008 Volume 47(Issue 50) pp:13394-13403
Publication Date(Web):November 17, 2008
DOI:10.1021/bi8013888
Photosystem II (PSII) contains a non-heme ferrous ion, located on the stromal side of the protein in close proximity to quinones A and B (QA and QB). We used EPR spectroscopy to examine the temperature-dependent redox reactions of the iron-quinone site, using it as a probe of potentially physiologically relevant proton-coupled electron-transfer (PCET) reactions. Complete chemical oxidation of the non-heme iron at ambient temperatures was followed by cryogenic photoreduction, producing a temperature-dependent yield of Fe2+QA (or Fe3+QA−)···Chl+/Car+/YD• charge separations. These charge separations were subsequently observed to partially recombine in the dark at cryogenic temperatures. We observed no double photochemical charge separations upon illumination at temperatures ≤30 K, demonstrating that QA and Fe3+ together act as a single electron-accepting moiety at very low temperatures. Our results indicate the existence of two populations of the iron-quinone site in PSII, one whose Fe3+ signal is abolished by illumination at liquid helium temperatures and one whose Fe3+ signal is abolished by illumination only above 75 K. The observation of non-heme iron photoreduction at cryogenic temperatures (possibly at liquid helium temperatures and certainly above 75 K) implies the existence of a low reorganization energy proton-transfer (ET) pathway within the protein to the non-heme iron environment, of possible relevance to the PCET reactions of QB and/or the non-heme iron itself. Furthermore, we observed the partial reoxidation of the non-heme iron by charge recombination with previously oxidized chlorophyll, carotenoid, and YD within PSII. This electron transfer might be important in the photoprotective transfer of oxidative power away from P680+ and the oxygen-evolving complex in stressed PSII centers.
Co-reporter:Cara A. Tracewell and Gary W. Brudvig
Biochemistry 2008 Volume 47(Issue 44) pp:
Publication Date(Web):October 14, 2008
DOI:10.1021/bi801461d
Photosystem II (PS II) is unique among photosynthetic reaction centers in having secondary electron donors that compete with the primary electron donors for reduction of P680+. We have characterized the photooxidation and dark decay of the redox-active accessory chlorophylls (Chl) and β-carotenes (Car) in oxygen-evolving PS II core complexes by near-IR absorbance and EPR spectroscopies at cryogenic temperatures. In contrast to previous results for Mn-depleted PS II, multiple near-IR absorption bands are resolved in the light-minus-dark difference spectra of oxygen-evolving PS II core complexes including two fast-decaying bands at 793 and 814 nm and three slow-decaying bands at 810, 825, and 840 nm. We assign these bands to chlorophyll cation radicals (Chl+). The fast-decaying bands observed after illumination at 20 K could be generated again by reilluminating the sample. Quantization by EPR gives a yield of 0.85 radicals per PS II, and the yield of oxidized cytochrome b559 by optical difference spectroscopy is 0.15 per PS II. Potential locations of Chl+ and Car+ species, and the pathways of secondary electron transfer based on the rates of their formation and decay, are discussed. This is the first evidence that Chls in the light-harvesting proteins CP43 and CP47 are oxidized by P680+ and may have a role in Chl fluorescence quenching. We also suggest that a possible role for negatively charged lipids (phosphatidyldiacylglycerol and sulfoquinovosyldiacylglycerol identified in the PS II structure) could be to decrease the redox potential of specific Chl and Car cofactors. These results provide new insight into the alternate electron-donation pathways to P680+.
Co-reporter:Gözde Ulas, Gerard Olack and Gary W. Brudvig
Biochemistry 2008 Volume 47(Issue 10) pp:
Publication Date(Web):February 15, 2008
DOI:10.1021/bi8000424
The oxidation of water to molecular oxygen by photosystem II (PSII) is inhibited in bicarbonate-depleted media. One contribution to the inhibition is the binding of bicarbonate to the non-heme iron, which is required for efficient electron transfer on the electron-acceptor side of PSII. There are also proposals that bicarbonate is required for formation of O2 by the manganese-containing O2-evolving complex (OEC). Previous work indicates that a bicarbonate ion does not bind reversibly close to the OEC, but it remains possible that bicarbonate is bound sufficiently tightly to the OEC that it cannot readily exchange with bicarbonate in solution. In this study, we have used NH2OH to destroy the OEC, which would release any tightly bound bicarbonate ions from the active site, and mass spectrometry to detect any released bicarbonate as CO2. The amount of CO2 per PSII released by the NH2OH treatment is observed to be comparable to the background level, although N2O, a product of the reaction of NH2OH with the OEC, is detected in good yield. These results strongly argue against tightly bound bicarbonate ions in the OEC.
Co-reporter:Ranitendranath Tagore, Hongyu Chen, Hong Zhang, Robert H. Crabtree, Gary W. Brudvig
Inorganica Chimica Acta 2007 Volume 360(Issue 9) pp:2983-2989
Publication Date(Web):10 June 2007
DOI:10.1016/j.ica.2007.02.020
O2 evolution was observed upon mixing aqueous [(terpy)(H2O)Mn(O)2Mn(H2O)(terpy)](NO3)3 (1, terpy = 2,2′:6′,6″-terpyridine) with aqueous solutions of Ce4+. However, when the solution of 1 was incubated at pH 1 (by dissolving in dilute HNO3) before mixing with Ce4+, very small amounts of O2 were observed. This observation of acid-induced deactivation suggests an explanation, both for the previously reported lack of O2 evolution from aqueous solutions of 1 with Ce4+ as oxidant, and the present observation of low amounts of O2 production with the very acidic Ce4+ reagent. Evidence is provided for water being the source of evolved O2, and for the requirement of a high valent multinuclear Mn species for O2 evolution. We test the possibility of complications in the use of ceric ammonium nitrate (CAN) in oxidation chemistry due to the presence of the oxidizable NH4+ ion.The title compound 1 is shown to oxidize water in the presence of Ce4+ as the primary electron acceptor. Evidence is provided for higher oxidation states of 1 as the water-oxidizing species, and low-valent mononuclear manganese species formed from 1 under acidic conditions are shown to be inactive. The potential for catalytic water oxidation by 1 in the presence of electron acceptor oxidants is discussed.
Co-reporter:Siddhartha Das;Christopher D. Incarvito;Robert H. Crabtree
Science 2006 Vol 312(5782) pp:1941-1943
Publication Date(Web):30 Jun 2006
DOI:10.1126/science.1127899
Abstract
Although enzymes often incorporate molecular recognition elements to orient substrates selectively, such strategies are rarely achieved by synthetic catalysts. We combined molecular recognition through hydrogen bonding with C-H activation to obtain high-turnover catalytic regioselective functionalization of sp3 C-H bonds remote from the –COOH recognition group. The catalyst contains a Mn(μ-O)2Mn reactive center and a ligand based on Kemp's triacid that directs a –COOH group to anchor the carboxylic acid group of the substrate and thus modify the usual selectivity for oxidation. Control experiments supported the role of hydrogen bonding in orienting the substrate to achieve high selectivity.
Co-reporter:Clyde W. Cady, Christopher Incarvito, Gary W. Brudvig, Robert H. Crabtree
Inorganica Chimica Acta 2006 Volume 359(Issue 8) pp:2509-2512
Publication Date(Web):15 May 2006
DOI:10.1016/j.ica.2006.02.005
The new hexadentate, bis-pincer ligand, (dipyCH2)MeNCH2CH2NMe(CH2dipy) (dipy = 2,2′-dipyridyl-6-yl) forms a crystallographically characterized Mn(II) complex in which each half of the ligand binds a separate Mn(OAc)2 unit. The structure consists of a distorted N3Mn(η2-OAc)(η1-OAc) core with six normal coordinate bonds and a long (2.85 Å) secondary bond to a seventh ligand atom, an oxygen of the η1-acetate. In addition to demonstrating an interesting coordination mode, the structure also mimics a predicted transition state in the associative ligand exchange of octahedral Mn(II) complexes.The crystal structure of a novel pseudo-seven-coordinate Mn(II) complex containing a long secondary bond to the seventh ligand atom demonstrates an interesting coordination mode, and mimics a previously predicted transition state in the associative ligand exchange of octahedral Mn(II) complexes.
Co-reporter:James P. McEvoy, Jose A. Gascon, Victor S. Batista and Gary W. Brudvig
Photochemical & Photobiological Sciences 2005 vol. 4(Issue 12) pp:940-949
Publication Date(Web):04 Oct 2005
DOI:10.1039/B506755C
Oxygenic photosynthesis, which provides the biosphere with most of its chemical energy, uses water as its source of electrons. Water is photochemically oxidized by the protein complex photosystem II (PSII), which is found, along with other proteins of the photosynthetic light reactions, in the thylakoid membranes of cyanobacteria and of green plant chloroplasts. Water splitting is catalyzed by the oxygen-evolving complex (OEC) of PSII, producing dioxygen gas, protons and electrons. O2 is released into the atmosphere, sustaining all aerobic life on earth; product protons are released into the thylakoid lumen, augmenting a proton concentration gradient across the membrane; and photo-energized electrons pass to the rest of the electron-transfer pathway. The OEC contains four manganese ions, one calcium ion and (almost certainly) a chloride ion, but its precise structure and catalytic mechanism remain unclear. In this paper, we develop a chemically complete structure of the OEC and its environment by using molecular mechanics calculations to extend and slightly adjust the recently-obtained X-ray crystallographic model [K. N. Ferreira, T. M. Iverson, K. Maghlaoui, J. Barber and S. Iwata, Science, 2004, 303, 1831–1838]. We discuss our mechanistic hypothesis [J. P. McEvoy and G. W. Brudvig, Phys. Chem. Chem. Phys., 2004, 6, 4754–4763] with reference to this structure and to some important recent experimental results.
Co-reporter:James P. McEvoy and Gary W. Brudvig
Physical Chemistry Chemical Physics 2004 vol. 6(Issue 20) pp:4754-4763
Publication Date(Web):19 Jul 2004
DOI:10.1039/B407500E
The recently-published 3.5 Å resolution X-ray crystal structure of a cyanobacterial photosystem II (PDB entry 1S5L) provides a detailed architecture of the oxygen-evolving complex (OEC) and the surrounding amino-acids [K. N. Ferreira, T. M. Iverson, K. Maghlaoui, J. Barber and S. Iwata, Science, 2004, 203, 1831–1838]. The revealed geometry of the OEC lends weight to certain hypothesized mechanisms for water-splitting, including the one propounded by this group, in which a calcium-bound water acts as a nucleophile to attack the oxygen of a MnVO group in the crucial O–O bond-forming step [J. S. Vrettos, J. Limburg and G. W. Brudvig, Biochim. Biophys. Acta, 2001, 1503, 229–245]. Here we re-examine this mechanism in the light of the new crystallographic information and make detailed suggestions concerning the mechanistic functions (especially the redox and proton-transfer roles) of calcium, chloride and certain amino-acid residues in and around the OEC. In particular, we propose an important role for an arginine residue, CP43–Arg357, in abstracting protons from a substrate water molecule during the water-splitting reaction.
Co-reporter:Olaf Kievit, Gary W Brudvig
Journal of Electroanalytical Chemistry 2001 Volume 497(1–2) pp:139-149
Publication Date(Web):2 February 2001
DOI:10.1016/S0022-0728(00)00467-8
Direct electrochemistry has been developed into a powerful tool for the analysis of soluble proteins and enzymes over the past two decades. In the present study, we have expanded the range of proteins to include a solubilized membrane protein, photosystem I. Taking into account the difficulties caused by the presence of detergent, we have shown it to be possible to obtain an electrochemical signal of the primary electron donor, a chlorophyll dimer called P700, of photosystem I. A reduction potential of vs. SHE with a peak separation of was found, in good agreement with literature values.
Co-reporter:Carole Baffert, Hongyu Chen, Robert H. Crabtree, Gary W. Brudvig, Marie-Noëlle Collomb
Journal of Electroanalytical Chemistry 2001 Volume 506(Issue 2) pp:99-105
Publication Date(Web):29 June 2001
DOI:10.1016/S0022-0728(01)00488-0
The electrochemical behavior of [MnIII(terpy)(N3)3] (1) (terpy=2,2′:6′,2″-terpyridine), a structural model of the azide complex of the manganese superoxide dismutase (MnSOD), has been investigated in acetonitrile (CH3CN) solution. In CH3CN containing either 0.1 M tetra-n-butylammonium perchlorate (Bu4NClO4) or tetraethylammonium trifluoroacetate (Et4NCF3CO2) as supporting electrolytes, the cyclic voltammogram of 1 exhibits one quasi-reversible reduction wave at E1/2=−0.170 V versus Ag ∣ 10 mM Ag+ and one quasi-reversible oxidation wave at E1/2=+0.675 V. These are both one-electron waves, corresponding to the Mn(III)/Mn(II) and Mn(III)/Mn(IV) redox couples respectively. To evaluate the stability of the oxidized and reduced species of 1, exhaustive electrolyses have been carried out. Controlled-potential reductions at −0.35 V of solutions of 1 in CH3CN containing 0.1 M Bu4NClO4 or 0.1 M Et4NCF3CO2 lead to the quantitative conversion of 1 into the bridging N3− dimanganese(II) complex, [(N3)(terpy)MnII(μ-N3)2MnII(terpy)(N3)] (2). This transformation is chemically reversible by an oxidation process. Controlled-potential oxidation at 0.8 V of a solution of 1 in CH3CN+0.1 M Bu4NClO4 produces a new mononuclear Mn(IV) complex characterized by electron paramagnetic resonance spectroscopy, which is stable only at or below −15°C. If this oxidation is conducted in CH3CN+0.1 M Et4NCF3CO2, the stable dimanganese(IV) di-μ-oxo complex [(CF3CO2)(terpy)MnIV(μ-O)2MnIV(terpy)(CF3CO2)]2+ (3) is formed quantitatively owing to the presence of an excess of the coordinating CF3CO2− anions and residual H2O in the CH3CN solution.
Co-reporter:Michael J. Reifler;Veronika A. Szalai;Celeste N. Peterson
Journal of Molecular Recognition 2001 Volume 14(Issue 3) pp:157-165
Publication Date(Web):11 APR 2001
DOI:10.1002/jmr.529
The QB quinone-binding site of photosystem II is an important target for herbicides. Two major classes of herbicides are based on s-triazine and phenylurea moieties. A small library of triazine and phenylurea compounds has been synthesized which have tail-like substituents in order to test the effects of charge, hydrophobicity and size of the tail on binding properties. It is found that a tail can be attached to one of the alkylamino groups of triazine-type herbicides or to the para position of phenylurea-type herbicides without loss of binding, provided that the tail is hydrophobic. This indicates that the herbicides must be oriented in the QB site such that these positions point toward the natural isoprenyl tail-binding pocket that extends out of the QB site. In turn, the requirement that the tail must extend out of the QB site constrains the size of the other herbicide substituents in the pocket. This is in agreement with the presumed orientation and fit of ligands in the QB site. When longer hydrophobic tails are used, the binding penalty that occurs upon adding a charged substituent at the distal end is reduced. This allows the use of a series of tail substituents possessing a distal charge as an approximate molecular ruler to measure the distance from the QB site to the aqueous phase. Even a 10-carbon alkyl chain still shows a 4-fold effect from the presence or absence of a distal charge. Such a chain does not appear to be long enough to extend from the bulk aqueous phase to the QB site because binding is completely lost when a large hydrophilic domain (PEG4000) is attached to the distal end. Longer tails are effective only if they are sufficiently hydrophobic. An effort was made to use tailed herbicides for affinity binding of photosystem II. It was found that hydrophobic linkers promote nonspecific binding, but careful choice of solvent conditions, such as the use of excess nonionic detergent well above its critical micelle concentration, might obviate this problem during affinity-binding applications. Copyright © 2001 John Wiley & Sons, Ltd.
Abbreviations used:- BOC
t-butoxycarbonyl
- bRC
bacterial photosynthetic reaction center
- DCC
N,N′-dicyclohexylcarbodiimide
- DCMU
3-(3,4-dichlorophenyl)-1,1-dimethylurea
- NHS
N-hydroxysuccinimide
- PEG
polyethylene glycol
- PSII
photosystem II
- QA
tightly bound quinone electron acceptor in PSII
- QB
exchangeable quinone electron acceptor in PSII
- THF
tetrahydrofuran
- TLC
thin-layer chromatography.
Co-reporter:Katherine E. Shinopoulos, Gary W. Brudvig
Biochimica et Biophysica Acta (BBA) - Bioenergetics (January 2012) Volume 1817(Issue 1) pp:66-75
Publication Date(Web):January 2012
DOI:10.1016/j.bbabio.2011.08.002
Co-reporter:Alec C. Durrell, Gonghu Li, Matthieu Koepf, Karin J. Young, Christian F.A. Negre, Laura J. Allen, William R. McNamara, Hee-eun Song, Victor S. Batista, Robert H. Crabtree, Gary W. Brudvig
Journal of Catalysis (February 2014) Volume 310() pp:37-44
Publication Date(Web):1 February 2014
DOI:10.1016/j.jcat.2013.07.001
•H2DCF is an effective fluorescence sensor to evaluate anode performance in photocatalytic cells.•We show that increased photocurrent does not necessarily correlate with increased catalysis.•Redox-active Mn photoanodes outperform redox-inactive Zn anodes.A manganese complex covalently attached to a TiO2 electrode via a light-absorbing organic linker (L) was used in the photooxidation of 2′,7′-dihydrodichlorofluorescein (H2DCF). Significant and sustained photocurrent was observed upon visible-light illumination of the fully assembled anode in the presence of the substrate. The two-electron, two-proton oxidation of H2DCF yields the fluorescent compound, 2′,7′-dichlorofluorescein (DCF). Our studies suggest that the MnII–L–TiO2 architecture is an effective photoanode for multielectron chemistry, as production of DCF under visible-light illumination exceeds yields observed for bare TiO2 as well as ZnII–L–TiO2 anodes. The turn-on fluorescent behavior of H2DCF upon oxidation makes it an excellent substrate for the study of new photoanodes. The high fluorescence quantum yield of DCF allows for nanomolar sensitivity and real-time monitoring of substrate oxidation.Graphical abstractDownload high-res image (66KB)Download full-size image
Co-reporter:Bradley J. Brennan, Jeffrey Chen, Benjamin Rudshteyn, Subhajyoti Chaudhuri, Brandon Q. Mercado, Victor S. Batista, Robert H. Crabtree and Gary W. Brudvig
Chemical Communications 2016 - vol. 52(Issue 14) pp:NaN2975-2975
Publication Date(Web):2016/01/12
DOI:10.1039/C5CC09857B
Hydroxamate binding modes and protonation states have yet to be conclusively determined. Molecular titanium(IV) phenylhydroxamate complexes were synthesized as structural and spectroscopic models, and compared to functionalized TiO2 nanoparticles. In a combined experimental–theoretical study, we find that the predominant binding form is monodeprotonated, with evidence for the chelate mode.
Co-reporter:Yuta Tsubonouchi, Shu Lin, Alexander R. Parent, Gary W. Brudvig and Ken Sakai
Chemical Communications 2016 - vol. 52(Issue 51) pp:NaN8021-8021
Publication Date(Web):2016/05/31
DOI:10.1039/C6CC02816K
A μ-oxido-bridged triruthenium complex (RuT2+), formed by air-oxidation of a previously reported monoruthenium water oxidation catalyst (WOC), serves as an efficient photochemical WOC with the turnover frequency (TOF) and turnover number (TON) 0.90 s−1 and 610, respectively. The crystal structures of RuT2+ and its one-electron oxidized RuT3+ are also reported.
Co-reporter:Alexander R. Parent, James D. Blakemore, Gary W. Brudvig and Robert H. Crabtree
Chemical Communications 2011 - vol. 47(Issue 42) pp:NaN11747-11747
Publication Date(Web):2011/09/28
DOI:10.1039/C1CC15501F
The catalytic water-oxidation activity of Wilkinson's iridium acetate trimer (1) has been characterized electrochemically and by using chemical oxidants. We show that 1 can function as an operationally homogeneous water-oxidation catalyst when driven with sodium periodate as a primary oxidant, but rapidly decomposes using Ce(IV) as a primary oxidant.
Co-reporter:James D. Blakemore, Nathan D. Schley, Gerard W. Olack, Christopher D. Incarvito, Gary W. Brudvig and Robert H. Crabtree
Chemical Science (2010-Present) 2011 - vol. 2(Issue 1) pp:NaN98-98
Publication Date(Web):2010/10/28
DOI:10.1039/C0SC00418A
Artificial photosynthesis, modeled on natural light-driven oxidation of water in Photosystem II, holds promise as a sustainable source of reducing equivalents for producing fuels. Few robust water-oxidation catalysts capable of mediating this difficult four-electron, four-proton reaction have yet been described. We report a new method for generating an amorphous electrodeposited material, principally consisting of iridium and oxygen, which is a robust and long-lived catalyst for water oxidation, when driven electrochemically. The catalyst material is generated by a simple anodic deposition from Cp*Ir aqua or hydroxo complexes in aqueous solution. This work suggests that organometallic precursors may be useful in electrodeposition of inorganic heterogeneous catalysts.
Co-reporter:Alexander R. Parent, Robert H. Crabtree and Gary W. Brudvig
Chemical Society Reviews 2013 - vol. 42(Issue 6) pp:NaN2252-2252
Publication Date(Web):2012/09/13
DOI:10.1039/C2CS35225G
In this tutorial review, we compare chemical oxidants for driving water-oxidation catalysts, focusing on the advantages and disadvantages of each oxidant.
Co-reporter:Julianne M. Thomsen, Daria L. Huang, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2015 - vol. 44(Issue 28) pp:NaN12472-12472
Publication Date(Web):2015/05/05
DOI:10.1039/C5DT00863H
Organometallic Ir precatalysts have been found to yield homogeneous Ir-based water-oxidation catalysts (WOCs) with very high activity. The Cp*Ir catalyst series can operate under a variety of regimes: it can either act as a homogeneous or a heterogeneous catalyst; it can be driven by chemical, photochemical, or electrochemical methods; and the molecular catalyst can either act in solution or supported as a molecular unit on a variety of solid oxides. In addition to optimizing the various reaction conditions, work has continued to elucidate the catalyst activation mechanism and identify water-oxidation intermediates. This Perspective will describe the development of the Cp*Ir series, their many forms as WOCs, and their ongoing characterization.
Co-reporter:Wojciech T. Osowiecki, Stafford W. Sheehan, Karin J. Young, Alec C. Durrell, Brandon Q. Mercado and Gary W. Brudvig
Dalton Transactions 2015 - vol. 44(Issue 38) pp:NaN16881-16881
Publication Date(Web):2015/09/02
DOI:10.1039/C5DT02390D
Splitting water into hydrogen and oxygen is one of the most promising ways of storing energy from intermittent, renewable sources in the future. Toward this goal, development of inexpensive, stable, and non-toxic catalysts for water oxidation is crucial. We report that the electrodeposition of manganese oxide in the presence of sodium dodecyl sulfate (SDS) produces a material that is highly active for electrocatalytic water oxidation at pH near 7 and remains stable for over 24 hours of sustained electrolysis. Clark electrode measurements demonstrate more than 95% Faradaic efficiency for oxygen evolution after an initial charging period. We found that catalytic performance was optimized in films prepared by electrodeposition using a precursor solution containing moderate concentration of substrates, namely 25 mM Mn2+ and 25 mM SDS. Microstructure and elemental analyses revealed that the deposited material, a mixed-phase manganese oxide, is structurally similar to materials used for electrochemical capacitors and batteries, drawing a parallel between highly studied cathode materials for rechargeable batteries and heterogeneous catalysts for water oxidation.
Co-reporter:Maxwell N. Kushner-Lenhoff, James D. Blakemore, Nathan D. Schley, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2013 - vol. 42(Issue 10) pp:NaN3622-3622
Publication Date(Web):2012/12/13
DOI:10.1039/C2DT32326E
A thin layer of an amorphous, mixed-valence iridium oxide (electrodeposited from an organometallic precursor, [Cp*Ir(H2O)3]2+) is a heterogeneous catalyst among the most active and stable currently available for electrochemical water oxidation. We show that buffers can improve the oxygen-evolution activity of such thin-layer catalysts near neutral pH, but that buffer identity and concentration, as well as the solution pH, remain key determinants of long-term electrocatalyst activity and stability; for example, phosphate buffer can reduce the overpotential by up to 173 mV.
Co-reporter:James D. Blakemore, Jonathan F. Hull, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2012 - vol. 41(Issue 25) pp:NaN7688-7688
Publication Date(Web):2012/04/17
DOI:10.1039/C2DT30371J
The speciation behavior of a water-soluble manganese(III) tetrasulfonated phthalocyanine complex was investigated with UV-visible and electron paramagnetic resonance (EPR) spectroscopies, as well as cyclic voltammetry. Parallel-mode EPR (in dimethylformamide:pyridine solvent mix) reveals a six-line hyperfine signal, centered at a g-value of 8.8, for the manganese(III) monomer, characteristic of the d4S = 2 system. The color of an aqueous solution containing the complex is dependent upon the pH of the solution; the phthalocyanine complex can exist as a water-bound monomer, a hydroxide-bound monomer, or an oxo-bridged dimer. Addition of coordinating bases such as borate or pyridine changes the speciation behavior by coordinating the manganese center. From the UV-visible spectra, complete speciation diagrams are plotted by global analysis of the pH-dependent UV-visible spectra, and a complete set of pKa values is obtained by fitting the data to a standard pKa model. Electrochemical studies reveal a pH-independent quasi-reversible oxidation event for the monomeric species, which likely involves oxidation of the organic ligand to the radical cation species. Adsorption of the phthalocyanine complex on the carbon working electrode was sometimes observed. The pKa values and electrochemistry data are discussed in the context of the development of mononuclear water-oxidation catalysts.
Co-reporter:Clyde W. Cady, Katherine E. Shinopoulos, Robert H. Crabtree and Gary W. Brudvig
Dalton Transactions 2010 - vol. 39(Issue 16) pp:NaN3989-3989
Publication Date(Web):2010/03/17
DOI:10.1039/B922087A
Photosynthetic water oxidation occurs naturally at a tetranuclear manganese center in the photosystem II protein complex. Synthetically mimicking this tetramanganese center, known as the oxygen-evolving complex (OEC), has been an ongoing challenge of bioinorganic chemistry. Most past efforts have centered on water-oxidation catalysis using chemical oxidants. However, solar energy applications have drawn attention to electrochemical methods. In this paper, we examine the electrochemical behavior of the biomimetic water-oxidation catalyst [(H2O)(terpy)Mn(μ-O)2Mn(terpy)(H2O)](NO3)3 [terpy = 2,2′:6′,2′′-terpyridine] (1) in water under a variety of pH and buffered conditions and in the presence of acetate that binds to 1 in place of one of the terminal water ligands. These experiments show that 1 not only exhibits proton-coupled electron-transfer reactivity analogous to the OEC, but also may be capable of electrochemical oxidation of water to oxygen.
Co-reporter:Gonghu Li, Christiaan P. Richter, Rebecca L. Milot, Lawrence Cai, Charles A. Schmuttenmaer, Robert H. Crabtree, Gary W. Brudvig and Victor S. Batista
Dalton Transactions 2009(Issue 45) pp:NaN10085-10085
Publication Date(Web):2009/09/02
DOI:10.1039/B908686B
A synergistic effect between anatase and rutile TiO2 is known, in which the addition of rutile can remarkably enhance the photocatalytic activity of anatase in the degradation of organic contaminants. In this study, mixed-phase TiO2 nanocomposites consisting of anatase and rutile nanoparticles (NPs) were prepared for use as photoanodes in dye-sensitized solar cells (DSSCs) and were characterized by using UV-vis spectroscopy, powder X-ray diffraction and scanning electron microscopy. The addition of 10–15% rutile significantly improved light harvesting and the overall solar conversion efficiency of anatase NPs in DSSCs. The underlying mechanism for the synergistic effect in DSSCs is now explored by using time-resolved terahertz spectroscopy. It is clearly demonstrated that photo-excited electrons injected into the rutile NPs can migrate to the conduction band of anatase NPs, enhancing the photocurrent and efficiency. Interfacial electron transfer from rutile to anatase, similar to that in heterogeneous photocatalysis, is proposed to account for the synergistic effect in DSSCs. Our results further suggest that the synergistic effect can be used to explain the beneficial effect of TiCl4 treatment on DSSC efficiency.
Co-reporter:Jianbing Jiang, John R. Swierk, Svante Hedström, Adam J. Matula, Robert H. Crabtree, Victor S. Batista, Charles A. Schmuttenmaer and Gary W. Brudvig
Physical Chemistry Chemical Physics 2016 - vol. 18(Issue 28) pp:NaN18682-18682
Publication Date(Web):2016/06/30
DOI:10.1039/C6CP04377A
Interfacial electron transfer dynamics of a series of photosensitizers bound to TiO2via linkers of varying conjugation strength are explored by spectroscopic and computational techniques. Injection and recombination depend on the extent of conjugation in the linker, where the LUMO delocalization determines the injection dynamics but both the HOMO and HOMO−1 are involved in recombination.
Co-reporter:Bradley J. Brennan, Alec C. Durrell, Matthieu Koepf, Robert H. Crabtree and Gary W. Brudvig
Physical Chemistry Chemical Physics 2015 - vol. 17(Issue 19) pp:NaN12734-12734
Publication Date(Web):2015/04/16
DOI:10.1039/C5CP01683E
Current molecular water-oxidation photoelectrocatalytic cells have substantial kinetic limitations under normal solar photon flux where electron–hole recombination processes may outcompete charge buildup on the catalytic centers. One method of overcoming these limitations is to design a system where multiple light-harvesting dyes work cooperatively with a single catalyst. We report a porphyrin monomer/dyad array for analysis of lateral hole transfer on a SnO2 surface consisting of a free-base porphyrin that functions to absorb light and initiate charge injection into the conduction band of SnO2, which leaves a positive charge on the organic moiety, and a free-base porphyrin/Zn-porphyrin dyad molecule that functions as a thermodynamic trap for the photoinduced holes. By using transient absorption spectroscopy, we have determined that the holes on the surface-bound free-base porphyrins are highly mobile via electron self-exchange between close-packed neighbors. The lateral charge-transfer processes were modelled by treating the system statistically with a random-walk method that utilizes experimentally derived kinetic parameters. The results of the modelling indicate that each self-exchange (hop) occurs within 25 ns and that the holes are efficiently transferred to the Zn-porphyrin. This hole-harvesting scheme provides a framework for enhancing the efficiency of multielectron photoelectrocatalytic reactions such as the four-electron oxidation of water.
Co-reporter:Ravi Pokhrel and Gary W. Brudvig
Physical Chemistry Chemical Physics 2014 - vol. 16(Issue 24) pp:
Publication Date(Web):
DOI:10.1039/C4CP00493K
Co-reporter:C. Koenigsmann, T. S. Ripolles, B. J. Brennan, C. F. A. Negre, M. Koepf, A. C. Durrell, R. L. Milot, J. A. Torre, R. H. Crabtree, V. S. Batista, G. W. Brudvig, J. Bisquert and C. A. Schmuttenmaer
Physical Chemistry Chemical Physics 2014 - vol. 16(Issue 31) pp:NaN16641-16641
Publication Date(Web):2014/07/04
DOI:10.1039/C4CP02405B
An efficient synthetic protocol to functionalize the cyanoacrylic acid anchoring group of commercially available MK-2 dye with a highly water-stable hydroxamate anchoring group is described. Extensive characterization of this hydroxamate-modified dye (MK-2HA) reveals that the modification does not affect its favorable optoelectronic properties. Dye-sensitized solar cells (DSSCs) prepared with the MK-2HA dye attain improved efficiency (6.9%), relative to analogously prepared devices with commercial MK-2 and N719 dyes. The hydroxamate anchoring group also contributes to significantly increased water stability, with a decrease in the rate constant for dye desorption of MK-2HA relative to MK-2 in the presence of water by as much as 37.5%. In addition, the hydroxamate-anchored dye undergoes essentially no loss in DSSC efficiency and the external quantum efficiency improves when up to 20% water is purposefully added to the electrolyte. In contrast, devices prepared with the commercial dye suffer a 50% decline in efficiency under identical conditions, with a concomitant decrease in external quantum efficiency. Collectively, our results indicate that covalent functionalization of organic dyes with hydroxamate anchoring groups is a simple and efficient approach to improving the water stability of the dye–semiconductor interface and overall device durability.
Co-reporter:Jonathan Graeupner ; Ulrich Hintermair ; Daria L. Huang ; Julianne M. Thomsen ; Mike Takase ; Jesús Campos ; Sara M. Hashmi ; Menachem Elimelech ; Gary W. Brudvig ;Robert H. Crabtree
Organometallics () pp:
Publication Date(Web):September 24, 2013
DOI:10.1021/om400658a
A series of Cp*IrIII dimers have been synthesized to elucidate the mechanistic viability of radical oxo-coupling pathways in iridium-catalyzed O2 evolution. The oxidative stability of the precursors toward nanoparticle formation and their oxygen evolution activity have been investigated and compared to suitable monomeric analogues. We found that precursors bearing monodentate NHC ligands degraded to form nanoparticles (NPs), and accordingly their O2 evolution rates were not significantly influenced by their nuclearity or distance between the two metals in the dimeric precursors. A doubly chelating bis-pyridine–pyrazolide ligand provided an oxidation-resistant ligand framework that allowed a more meaningful comparison of catalytic performance of dimers with their corresponding monomers. With sodium periodate (NaIO4) as the oxidant, the dimers provided significantly lower O2 evolution rates per [Ir] than the monomer, suggesting a negative interaction instead of cooperativity in the catalytic cycle. Electrochemical analysis of the dimers further substantiates the notion that no radical oxyl-coupling pathways are accessible. We thus conclude that the alternative path, nucleophilic attack of water on high-valent Ir-oxo species, may be the preferred mechanistic pathway of water oxidation with these catalysts, and bimolecular oxo-coupling is not a valid mechanistic alternative as in the related ruthenium chemistry, at least in the present system.