Light-driven Hydrogen Evolution from Water by a Tripodal Silane

Department of Chemistry, Savitribai Phule Pune University, Pune 411007, India. Inorg. Chem. , 2017, 56 (21), pp 13286–13292. DOI: 10.1021/acs.inorgc...
0 downloads 14 Views 2MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX-XXX

pubs.acs.org/IC

Light-driven Hydrogen Evolution from Water by a Tripodal Silane Based CoII6L18 Octahedral Cage Mahesh S. Deshmukh,†,∥ Vishwanath S. Mane,§,∥ Avinash S. Kumbhar,*,§ and Ramamoorthy Boomishankar*,†,‡ †

Department of Chemistry and ‡Centre for Energy Science, Indian Institute of Science Education and Research (IISER), Pune, Dr. Homi Bhabha Road, Pune 411008, India § Department of Chemistry, Savitribai Phule Pune University, Pune 411007, India S Supporting Information *

ABSTRACT: The octahedral cage assembly [CoII6L18Cl6(H2O)6]Cl6 has been synthesized in a single-step reaction by using a polypyridyl-functionalized tripodal silane ligand. The electrochemical behavior of the cage in water exhibits the pH dependence of potential as well as catalytic current indicating the possible involvement of proton-coupled electron transfer in H2 evolution. Electrocatalytic hydrogen evolution from an aqueous buffered solution gave a turnover frequency of 16 h−1. Further, this cage assembly has been explored as a photocatalyst (blue light irradiation λ 469 nm) for the evolution of H2 from water in the presence of Ru(bpy)32+ as a photosensitizer and ascorbic acid as a sacrificial electron donor. This catalytic reaction is found to be pseudo first order with a turnover frequency of 20.50 h−1.



medium.32 Interesting examples of pentadentate ligand based on Co complexes have been reported by Alberto,16 Chang,19 Long,23 and Zhao and Wang and their co-workers that show excellent efficacy as PRCs.33 Recently, Lau,25 Thummel,26 and their co-workers have developed a series of Co complexes derived from chelating tetradentate ligands in which the pyridyl donor site offers a square-planar arrangement for the Co(II) ions. The apical positions in all of these Co complexes were taken up by weakly coordinating neutral or anionic donor ligands which were reversibly displaced during the water activation process with the formation of Co−H intermediate species.34 Alternately, some of these complexes were proposed to facilitate H2 evolution by the formation of a dangling protonated pyridyl group during the catalytic cycle.33 Despite the excellent platforms offered by these penta- and tetradentate ligands as H2-evolving Co catalysts, still there are challenges involved in the design of such suitable scaffolds. For example, these geometrically tailored ligands require multistep synthesis and must support the prerequisite of planar and squarepyramidal coordination to the metal ions. Therefore, we looked at a slightly different approach, where a self-assembly procedure can stabilize multinuclear clusters/cages built on Co(II) ions by using simple pyridyl-functionalized multisite ligands. A literature review of such self-assemblies revealed that M6L8 type octahedral cages are ubiquitous when C3-symmetric

INTRODUCTION Hydrogen is perceived as a prominent energy-dense carbon-free fuel to meet the rising demands of global energy and reduce the emission of greenhouse gases from the consumption of fossil fuels.1−4 Sustainable routes to produce hydrogen employs solar energy to split water into its elements (termed as artificial photosynthesis) by two individual half-reactions of water oxidation and proton reduction.5−9 The reductive part of the water splitting yields hydrogen that can be effectively achieved by a photocatalytic system consisting of a photosensitizer, a proton reduction catalyst (PRC), and a sacrificial reductant.10,11 Over the years, polypyridine and cyclometalated complexes of noble and rare-earth metals such as Pt, Rh, Ir, etc. have been shown to be efficient catalysts for the proton reduction process.12−14 However, due to the low abundance of these expensive metal ions, significant efforts have been made to develop homogeneous and heterogeneous PRCs based on nonnoble and earth-abundant transition-metal ions, especially those of Fe, Ni, Co, and Mo.15−20 Among these, cobalt complexes of imine and polypyridyl ligands have shown promising activity as PRCs.21−27 While cobalt imine complexes are excellent electroand photocatalysts, their activity has been largely confined to organic media, as these catalysts show poor solubility in water and show signs of degradation by hydrogenation or hydrolysis.28−31 With the activity of [Co(bpy)3]2+ as a precatalyst for H2 evolution as inspiration, several cobalt pyridyl complexes of penta- and tetradentate ligands have been developed as efficient PRCs in aqueous and buffered water © XXXX American Chemical Society

Received: August 11, 2017

A

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

wrapped with blue LED strips (λ 469 nm). The catalyst concentration was varied from 50 to 250 μM. The evolved hydrogen gas was monitored by recording the increase in headspace pressure every 10 s for 2−3 h. The obtained pressure measurements were converted into number of moles, using the equation PV = nRT. The generated hydrogen gas was identified by gas chromatography (GC) using a 5 Å molecular sieve column (8 ft × 1/8 in.) with argon as the carrier gas and TCD detector. The reaction solution and headspace were purged by nitrogen gas to allow accurate measurement of the amount of H2 produced during the experiment. This experiment was performed in triplicate. The thermodynamic and kinetic stabilities of 1 was measured using turnover number (TON) = [generated H2]/[catalyst] in moles and turnover frequency (TOF) = TON/unit time, respectively. Electrochemical Studies. Electrochemical measurements were carried out with a CH-electrochemical analyzer, Model 1100A, using a glassy-carbon-disk electrode (diameter 3 mm) as the working electrode, platinum wire as the counter electrode, and SCE (saturated calomel electrode) as a reference electrode in 0.1 M (n-Bu4N)PF6 and KNO3 as supporting electrolyte. The acetonitrile solvent and B. R. Buffer (Briton Robinson Buffer) were used as nonaqueous and aqueous solutions, respectively, for electrochemical studies. The bulk electrolysis was performed in 1.0 M aqueous phosphate buffer (pH 7) by using SCE as the working electrode.

tripodal N-donor ligands are employed in the self-assembly reactions.35−37 Herein, as a proof of concept, we describe the synthesis and electro- and photocatalytic H2 evolution activity of the water-soluble cobalt cage [CoII6L18Cl6(H2O)6]Cl6 (1), derived from the rigid tripodal silane ligand MeSi(3Py)3 (L1) featuring 3-pyridyl functionalities. This cage exhibits an octahedral environment for the six Co(II) atoms with coordination of four rigid equatorial N-pyridyl groups and weakly coordinating aqua and chloride ligands at the apex. In addition, the cage 1 is water-soluble (98.8 g/L), due to its high cationic charge. The ligand L1 exhibits a conjugated backbone via the Si atom. Hence, we reasoned that it might facilitate the cage assembly of 1 to accept electrons from the photosensitizer upon light irradiation. The photocatalytic H2 evolution studies of this cage at pH 4.0 in the presence of ascorbic acid (H2A)/ ascorbate (HA−) and [Ru(bpy)3]2+ as a photosensitizer (PS) gave a fairly reasonable turnover number of 43 at a 50 μM concentration of 1.



EXPERIMENTAL SECTION



Materials and Characterization. All manipulations were performed under a dry atmosphere in standard Schlenk glassware. The glassware was oven-dried before use. Water was deionized with the Millipore Synergy system (18.2 MΩ cm resistivities) and placed under vacuum, and the system was refilled with nitrogen. All other chemical reagents were purchased from commercial vendors and used without further purification. The ligand L1 was synthesized by using a previously reported procedure.38 FT-IR spectra and UV−visible spectra were recorded on a PerkinElmer spectrophotometer; DLS analysis was recorded on a Malvern instrument. High-resolution mass spectrometry (HRMS) measurements were performed on a Bruker Impact HD ESI-Q TOF mass spectrometer. The MALDI-TOF/TOF spectra were obtained on an Applied Biosystem spectrometer. Elemental analyses were performed on a Vario-EL cube elemental analyser. Synthesis of Cage 1. To a solution of L1 (10 mg, 36 mmol) in MeOH (1 mL) was added 1 mL of a methanolic solution of CoCl2· 6H2O (6.45 mg, 27 mmol) at room temperature to give a clear pink solution which on slow evaporation at room temperature yielded blue crystals of 1. ESI-MS: m/z 1056.08, {[CoII6L18Cl6(H2O)6]Cl3}3+ + (H 2 O) + K; 999.34, {[Co II 6 L 1 8 Cl 6 (H 2 O) 6 ]Cl 3 } 3+ ; 778.97, [CoII6L18Cl6(H2O)6]Cl2}4+ + K; 520.39, [CoII6L18Cl6(H2O)6]6+ + K. MALDI-TOF: m/z 642.88, 779.30 at pH 2.0 and m/z 642.88, 779.30 at pH 7.0 correspond to the species {[CoII6L18Cl6(H2O)6]Cl}5+ + (H2O) + K and {[CoII6L18Cl6(H2O)6]Cl4}2+. FT-IR data for a powdered sample (cm−1): 2956, 2924, 1605, 1581, 1482, 1432, 1262, 1193, 1135, 1062, 799, 700, 655. Anal. Calcd for C131H134Cl12Co6N24O4Si8: C, 50.55; H, 4.34; N, 10.80. Found: C, 50.57; H, 4.32; N, 10.83. Crystallography. Reflections were collected on a Bruker Smart Apex Duo diffractometer at 100 K using Mo Kα radiation (λ = 0.71073 Å). Structures were refined by full-matrix least squares against F2 using all data (SHELX).38 Crystallographic data for all of these compounds are given in Tables S1 and S2 in the Supporting Information. All nonhydrogen atoms were refined anisotropically if not stated otherwise. Hydrogen atoms were constrained in geometric positions to their parent atoms. The diffuse solvent molecules in cage 1 could not be modeled appropriately. Hence, these were treated as diffuse contributions to the overall scattering and removed by the SQUEEZE/PLATON software for better refinement data. General Procedure for Light-Driven Hydrogen Evolution. The proton reduction studies were performed in a 33 mL roundbottom flask capped with a septum and equipped with a stir bar. The flask that contained 1, 0.5 mM Ru(bpy)32+, and 0.3 M ascorbic acid as electron donor in pH 4 ascorbate buffer solution (5 mL) and connected to a port of a pressure transducer (PXM409-002BAUSBH) was kept in a thermally controlled water bath (25 °C) and was

RESULTS AND DISCUSSION Syntheses. The silane ligand L1 was prepared by our earlier reported procedure in a single step.39 The cage molecule of 1 was obtained by the slow addition of a methanolic solution of cobalt(II) chloride to a stirred solution of L1 in methanol at room temperature (Scheme 1). Blue single crystals of 1 suitable Scheme 1. Preparation of the Cationic Cage Assembly of 1

for X-ray crystallographic analysis were obtained by slow evaporation. The ESI mass spectrum of 1 in water exhibits several peaks centered at m/z 1056.08 for {[CoII6L18Cl6(H2O)6]Cl3}3+ + (H2O) + K, 999.34 for {[CoII6L18Cl6(H2O)6]Cl3}3+, 778.97 for {[CoII6L18Cl6(H2O)6]Cl2}4+ + K, and 520.39 for [CoII6L18Cl6(H2O)6]6+ + K. All of these peaks indicate that the cage remains intact in solution (Figure S1 in the Supporting Information). Further, the stability of the cage at low pH solutions has further been confirmed by MALDI-TOF data. Crystal Structures. The molecular structure of the cage assembly was solved in the orthorhombic space group Pbcn as its solvate adduct of formula 1·3MeOH·H2O (Figure 1 and Figures S2 and S3 in the Supporting Information). The structural analysis reveals that the Co6L18 core has an octahedral geometry in which the six Co(II) atoms are located at the corners and eight equivalent silane ligands act as threeconnected linkers. Each Co(II) atom lies in an octahedral coordination sphere, where the equatorial square plane is occupied by four Npyridyl atoms from four different L1 segments, B

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Electrochemical Behavior. The cyclic voltammogram (CV) of 1 was performed in aqueous solution. It exhibits one quasi-reversible couple at E1/2 = 0.75 V for the CoIII/CoII process and one irreversible event at −0.83 V for the CoII/CoI process (all potentials vs SCE) (Figure 3). Sweeping further

Figure 1. Crystal structure of 1 viewed along (a) the 4-fold axis and (b) the 3-fold axis.

while the axial positions are occupied by one water molecule outside and one chloride ion inside the cage. The Co−N distances range from 2.141(0) to 2.303(0) Å, and the Co−Cl distances range from 2.333(1) to 2.368(1) Å. The average distance between axially located Co centers is 11.32 Å, and that between equatorial Co centers is 8.06 Å. The average distance between the opposite Si atoms situated at the terminus of the 3fold axis is 13.889 Å. There are three crystallographically imposed 4-fold axes passing through two opposite Co(II) centers and four 3-fold axes passing through the opposite pairs of ligand Si−Me groups. The cage adopts a local 432 symmetry (chiral, O) due to the absence of any plane(s) of symmetry. Furthermore, the rigid topology of the ligand (and that of the cage) arises from the absence of any spacers between the Si atom and the pyridyl rings. As a result the cage is very dense, resembling a small nanocluster of metal atoms and has no internal cavity to accommodate guest molecules. However, this arrangement makes this cage very stable in an aqueous environment, as observed by its mass spectrum. The UV−vis spectra of 1 in water exhibit three absorption maxima at 421, 452, and 518 nm, which are attributed to the 4 T1g → 4T1g(P), 4T1g → 4A2g, and 4T1g → 4T2g d−d transitions, respectively, for the octahedral Co(II) complexes.40,41 Further, these peaks did not change even after 48 h (Figure 2), albeit

Figure 3. CV of 1 mM 1 in water at a scan rate of 100 mV/s at 25 °C. Conditions: supporting electrolyte KNO3 (0.1 M); SCE reference electrode; glassy-carbon working electrode; Pt-wire auxiliary electrode.

toward the cathode reveals one irreversible electron wave at −1.5 V assigned to the ligand-centered event. This peak compares well with that of the free ligand L1 (with a slight positive shift), which shows four redox events in acetonitrile (−0.95, −1.33, −1.83, and −2.01 V) (Figure S5 in the Supporting Information). The redox event for CoII/CoI appears at −0.83 V, which is more positive than that of the ligand event at −0.95 V. Thus, the observed irreversible redox event at −0.83 V is clearly attributed to the CoII/CoI process.42 Further, we have calculated the number of electrons involved in the redox events observed in the aqueous system by coulometry using the formula Q = nFm (Q = charge built, n = number of electron, F = Faraday constant, and m = moles of the materials).43 These results show that each metal-centered event is approximately a one-electron process (Figures S6 and S7 in the Supporting Information). We further explored the electrochemical behavior of 1 in buffered aqueous solution between pH 2 and 7.5, the range normally associated with the catalytic proton reduction. In pH 4.0 buffered solution, the CoII/CoI redox process for 1 was observed at E1/2 = −0.62 V (Figure S8 in the Supporting Information). The pH dependence of the redox potentials for 1 in the range 2−7.5 (aqueous Briton Robinson buffer solution) has been determined by using DPV measurements. A Pourbaix diagram indicates the changes in the CoII/CoI redox couples with increasing pH. Initially, it shows pH-independent behavior from pH 1.5 to 2.8. Subsequently, it exhibits a sharp linear dependence from pH 2.8 to 5.6 with a slope approximately equal to −59 mV/pH. Finally, it again shows pH independence from 5.6 to 7.5. However, the CoIII/CoII redox couple was found to be independent of pH (Figure S9 in the Supporting Information). This indicates a proton-coupled electron transfer process (PCET) involved in the complex that could play a vital pathway for the proton reduction reaction. The cathodic scans performed on 1 at different pHs indicate that the onset of the catalytic current is influenced by the variation in pH and shows a linear decline in the corresponding redox potentials (Figure

Figure 2. UV−visible spectra of 1 in water.

with a slight decrease in the intensity of peaks. This might be due to the slow degradation of the cage in low quantities. In addition, measurement of the UV−visible spectra under different pH conditions did not change the aqueous phase absorption profile, indicating the thermodynamic stability of 1 at low pH (Figure S4 in the Supporting Information). C

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry 4). For instance, for a fixed 40 μA catalytic current (j = 0.6 mA/ cm2) the applied potential drops linearly with increasing pH

evolution catalyst. The total charge obtained in the bulk electrolysis was found to increase significantly upon progressively increasing the negative potentials from −1.0 to −1.5 V. The overpotential values reached in each of the electrocatalytic cycles were calculated by using the method described by Evans and co-workers.45 The maximum TOF for the electrocatalytic hydrogen evolution by 1 is 16 h−1 (Figure S13 and Equation S1 in the Supporting Information). The stability of cage 1 was confirmed by recording the ESI mass before and after the bulk electrolysis, which showed no apparent change in the peak patterns (Figure S14 in the Supporting Information). Photocatalytic Activity. These photophysical and electrochemical analyses evidence that 1 can be a molecular catalyst which can drive proton reduction in water. Thus, we set out to probe the suitability of 1 as a PRC in visible-light-driven H2 evolution from water by employing [Ru(bpy)3]Cl2 as a photosensitizer (PS) and ascorbic acid (AA) as a sacrificial electron donor (SED) in aqueous solution. It has earlier been demonstrated that quenching of photoexcited [Ru(bpy)3]2+* by ascorbic acid yields [Ru(bpy)3]+, which drives the PRC to liberate hydrogen from water.10,32 The photocatalytic reaction was monitored in a 33 mL round-bottom flask fitted with a pressure transducer in the presence of an LED light strip (λ 469 nm) as a light source which was wrapped around a temperature-controlled water-jacketed beaker. The pressure developed in the flask due to the evolution of hydrogen was continuously measured by the pressure transducer for a period of up to 2 h. The identity of evolved hydrogen was confirmed by the GC equipped with a thermal conductivity detector (TCD). The liberated hydrogen vs time profile at catalyst concentrations ranging from 50 to 250 μM is shown in Figure 6. The turnover number (TON) after 2 h irradiation of the

Figure 4. Cathodic scan rate of 1 in 0.1 M buffered solution at various pH values. Inset: dependence of the catalytic potential of a catalytic current of 50 μA on the solution pH.

and indicates the involvement of a proton in the initial stages of the electrochemical catalysis by 1 (Figure 4, inset). On the basis of the electrochemical characteristics we propose a probable mechanistic scheme for hydrogen evolution by 1 (Figure S10 in the Supporting Information).44 Furthermore, the electrocatalytic studies of 1 were carried out in its 1.0 M aqueous phosphate buffer solution (pH 7). For this experiment, bulk electrolysis of 1 (0.5 μM) was performed by using a glassy-carbon disk as the working electrode in the potential range of −1.0 to −1.5 V. The charge built on the system was monitored as a function of time during the measurements. For example, when the applied potential was −1.4 V versus SCE, the maximum charge built was found to be 48 mC during the 2 min of electrolysis, which was accompanied by the evolution of gases (Figure 5 and Figure S11 in the Supporting Information). The control experiments under the same potential (−1.4 V) with a catalyst-free solution gave a charge of only 3 mC (Figure S12 in the Supporting Information). This indicates that 1 serves as a hydrogen

Figure 6. Kinetic plots of light-driven H2 evolution by employing 1 at 50, 150, 200, and 250 μM concentrations in the presence of 0.3 M ascorbic acid solution (pH 4) and 0.5 mM [Ru(bpy)3]Cl2 as photosensitizer (PS). Control experiments with 1 mM CoCl2 and 0.5 mM PS have also been plotted. Inset: initial rate of H2 evolution.

reaction mixture containing 50 μM of 1 was found to be 43. If 1 is replaced by CoCl2·6H2O, negligible H2 evolution was observed upon irradiation, even at a relatively higher concentration of 1 mM. Similarly, a much lower amount of H2 evolution was observed when [Ru(bpy)3]Cl2 (0.5 mM) alone was employed in the absence of the catalyst 1. These control experiments rule out any contribution from the simple

Figure 5. Charge buildup over time by the bulk electrolysis of 0.5 μM 1 in 1.0 M aqueous phosphate buffer (pH 7.0). All the data reported have been deducted from the blank. D

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Structural data for 1 and figures and tables as described in the text (PDF)

hydrated Co(II) cation (formed via leaching of 1) or [Ru(bpy)3]Cl2 itself as a catalyst for the light-driven H2 evolution. Further, the absence of cobalt oxide nanoparticles formed during the catalysis was confirmed by DLS analysis (Figure S15 in the Supporting Information). A plot of the initial rate of hydrogen generation vs catalyst concentration indicates the pseudo-first-order origin of the reaction kinetics. The rate constant derived from this plot is found to be 20.50 h−1 and is consistent with the obtained turnover number (TON) at the measured concentration range (Figure 6). A comparison of the photocatalytic activity of 1 with some reported Co-based systems is given in Table S3 in the Supporting Information. To understand the cause of early saturation of the catalytic activity, we performed control experiments. Thus, one of the three components (ascorbic acid, [Ru(bpy)3]2+ as PS, or 1) of the catalytic system was placed (separately or a combination of two) in the reaction flask after cessation of the H2 evolution to see if the H2 evolution could be resumed. However, the addition of any one of the three components, in the same amount as was used in the photocatalytic reaction, resulted in no significant amount of H2 formation, suggesting that decomposition of at least two of the three species occurred during the photocatalytic H2 evolution. A combined addition of 1 and [Ru(bpy)3]2+ as a photosensitizer resumed substantial H2 evolution. However, no significant amount of H2 was produced when ascorbic acid and 1 were added, suggesting the decomposition of [Ru(bpy)3]2+ under these reaction conditions.46 This may be due to the presence of trace amounts of air in the reaction flask that may lead to the decomposition of the catalytic system.47−49 The evolution of H2 in the presence of bulk air was negligible, suggesting that O2 inhibits the evolution of H2 (Figure S16 in the Supporting Information).

Accession Codes

CCDC 1548622 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



Corresponding Authors

*E-mail for A.S.K.: [email protected]. *E-mail for R.B.: [email protected]. ORCID

Avinash S. Kumbhar: 0000-0002-2087-4827 Ramamoorthy Boomishankar: 0000-0002-5933-4414 Author Contributions ∥

M.S.D. and V.S.M. contributed equally.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the SERB, India (EMR/2016/ 000614 and SR/S1/IC-50/2012), and Nanomission Project, DST, India, via Grant No. SR/NM/TP-13/2016 (R.B.). A.S.K. thanks the CSIR, India (01(2861)/16/EMR-II), and UGC (CAS), India (UPE, phase-II). M.S.D. thanks the UGC, India, and V.S.M. thanks S. P. Pune University, India, for the fellowship. We thank Prof. S. B. Ogale for electrochemical measurements and Dr. C. S. Gopinath for GC-TCD measurements.



CONCLUSION In conclusion, we show here for the first time that a selfassembled metal−organic cage based on Co(II) ions can be employed as both electro- and photocatalyst for H2 evolution. The cage assembly of 1 can be synthesized in a one-step reaction and contains six Co(II) ions in octahedral coordination, in which the apical positions are occupied with labile ligands. The photophysical and electrochemical investigations show that the cage can behave as a molecular catalyst. In addition, the pH dependence of the potentials and catalytic currents of 1 indicates the apparent involvement of the protoncoupled electron transfer process in the hydrogen evolution. Electrocatalytic hydrogen evolution experiments with 1 gave a turnover frequency of 16 h−1. Further, 1 has been shown to act as an efficient photocatalyst for the light-driven evolution of H2 from water in the presence of Ru(bpy)32+ as a PS and ascorbic acid as an SED. The obtained turnover number of 43 by application of 50 μM of 1 as a catalyst gave a pseudo-first-order rate constant value of 20.50 h−1. These results promise the utility of a self-assembly approach for designing suitable catalysts for water-splitting reactions. Current efforts are aimed at boosting the efficiency of such self-assembled cages by coupling them with suitable electron-injecting semiconductors.



AUTHOR INFORMATION



REFERENCES

(1) Turner, J. A. Sustainable hydrogen production. Science 2004, 305, 972−974. (2) Lewis, N. S.; Nocera, D. G. Powering the planet: Chemical challenges in solar energy utilization. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 15729−15735. (3) Esswein, A. J.; Nocera, D. G. Hydrogen production by molecular photocatalysis. Chem. Rev. 2007, 107, 4022−4047. (4) Cook, T. R.; Dogutan, D. K.; Reece, S. Y.; Surendranath, Y.; Teets, T. S.; Nocera, D. G. Solar energy supply and storage for the legacy and nonlegacy worlds. Chem. Rev. 2010, 110, 6474−6502. (5) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori, E. A.; Lewis, N. S. Solar water splitting cells. Chem. Rev. 2010, 110, 6446−6473. (6) Karkas, M. D.; Verho, O.; Johnston, E. V.; Åkermark, B. R. Artificial photosynthesis: molecular systems for catalytic water oxidation. Chem. Rev. 2014, 114, 11863−12001. (7) Swierk, J. R.; Mallouk, T. E. Design and development of photoanodes for water-splitting dye-sensitized photoelectrochemical cells. Chem. Soc. Rev. 2013, 42, 2357−2387. (8) Nocera, D. G. The artificial leaf. Acc. Chem. Res. 2012, 45, 767− 776. (9) Tran, P. D.; Artero, V.; Fontecave, M. Water electrolysis and photoelectrolysis on electrodes engineered using biological and bioinspired molecular systems. Energy Environ. Sci. 2010, 3, 727−747. (10) Creutz, C.; Sutin, N. Photogeneration and reactions of cobalt (I) complexes. Coord. Chem. Rev. 1985, 64, 321−341. (11) Alstrum-Acevedo, J. H.; Brennaman, M. K.; Meyer, T. J. Chemical approaches to artificial photosynthesis. Inorg. Chem. 2005, 44, 6802−6827.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b02074. E

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (12) Du, P.; Schneider, J.; Jarosz, P.; Zhang, J.; Brennessel, W. W.; Eisenberg, R. Photoinduced electron transfer in platinum (II) terpyridyl acetylide chromophores: Reductive and oxidative quenching and hydrogen production. J. Phys. Chem. B 2007, 111, 6887−6894. (13) Cline, E. D.; Adamson, S. E.; Bernhard, S. Homogeneous catalytic system for photoinduced hydrogen production utilizing iridium and rhodium complexes. Inorg. Chem. 2008, 47, 10378−10388. (14) Curtin, P. N.; Tinker, L. L.; Burgess, C. M.; Cline, E. D.; Bernhard, S. Structure-Activity Correlations Among Iridium (III) Photosensitizers in a Robust Water-Reducing System. Inorg. Chem. 2009, 48, 10498−10506. (15) Du, P.; Eisenberg, R. Catalysts made of earth-abundant elements (Co, Ni, Fe) for water splitting: recent progress and future challenges. Energy Environ. Sci. 2012, 5, 6012−6021. (16) Bachmann, C.; Guttentag, M.; Spingler, B.; Alberto, R. 3d element complexes of pentadentate bipyridine-pyridine-based ligand scaffolds: structures and photocatalytic activities. Inorg. Chem. 2013, 52, 6055−6061. (17) Eckenhoff, W. T.; Brennessel, W. W.; Eisenberg, R. LightDriven Hydrogen Production from Aqueous Protons using Molybdenum Catalysts. Inorg. Chem. 2014, 53, 9860−9869. (18) Han, Z.; Shen, L.; Brennessel, W. W.; Holland, P. L.; Eisenberg, R. Nickel pyridinethiolate complexes as catalysts for the light-driven production of hydrogen from aqueous solutions in noble-metal-free systems. J. Am. Chem. Soc. 2013, 135, 14659−14669. (19) Thoi, V. S.; Sun, Y.; Long, J. R.; Chang, C. J. Complexes of earth-abundant metals for catalytic electrochemical hydrogen generation under aqueous conditions. Chem. Soc. Rev. 2013, 42, 2388− 2400. (20) DuBois, D. L. Development of molecular electrocatalysts for energy storage. Inorg. Chem. 2014, 53, 3935−3960. (21) Artero, V.; Chavarot-Kerlidou, M.; Fontecave, M. Splitting water with cobalt. Angew. Chem., Int. Ed. 2011, 50, 7238−7266. (22) Singh, W. M.; Baine, T.; Kudo, S.; Tian, S.; Ma, X. A. N.; Zhou, H.; DeYonker, N. J.; Pham, T. C.; Bollinger, J. C.; Baker, D. L. Electrocatalytic and photocatalytic hydrogen production in aqueous solution by a molecular cobalt complex. Angew. Chem., Int. Ed. 2012, 51, 5941−5944. (23) Khnayzer, R.; Thoi, V.; Nippe, M.; King, A.; Jurss, J.; El Roz, K.; Long, J.; Chang, C.; Castellano, F. Towards a comprehensive understanding of visible-light photogeneration of hydrogen from water using cobalt (II) polypyridyl catalysts. Energy Environ. Sci. 2014, 7, 1477−1488. (24) Call, A.; Codolà, Z.; Acuña-Parés, F.; Lloret-Fillol, J. Photo-and Electrocatalytic H2 Production by New First-Row Transition-Metal Complexes Based on an Aminopyridine Pentadentate Ligand. Chem. Eur. J. 2014, 20, 6171−6183. (25) Leung, C.-F.; Ng, S.-M.; Ko, C.-C.; Man, W.-L.; Wu, J.; Chen, L.; Lau, T.-C. A cobalt (II) quaterpyridine complex as a visible lightdriven catalyst for both water oxidation and reduction. Energy Environ. Sci. 2012, 5, 7903−7907. (26) Tong, L.; Zong, R.; Thummel, R. P. Visible light-driven hydrogen evolution from water catalyzed by a molecular cobalt complex. J. Am. Chem. Soc. 2014, 136, 4881−4884. (27) Du, P.; Schneider, J.; Luo, G.; Brennessel, W. W.; Eisenberg, R. Visible light-driven hydrogen production from aqueous protons catalyzed by molecular cobaloxime catalysts. Inorg. Chem. 2009, 48, 4952−4962. (28) Losse, S.; Vos, J. G.; Rau, S. Catalytic hydrogen production at cobalt centres. Coord. Chem. Rev. 2010, 254, 2492−2504. (29) Hawecker, J.; Lehn, J. M.; Ziessel, R. Efficient homogeneous photochemical hydrogen generation and water reduction mediated by cobaloxime or macrocyclic cobalt complexes. New J. Chem. 1983, 7, 271−277. (30) Jacques, P.-A.; Artero, V.; Pécaut, J.; Fontecave, M. Cobalt and nickel diimine-dioxime complexes as molecular electrocatalysts for hydrogen evolution with low overvoltages. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 20627−20632.

(31) McCrory, C. C.; Uyeda, C.; Peters, J. C. Electrocatalytic hydrogen evolution in acidic water with molecular cobalt tetraazamacrocycles. J. Am. Chem. Soc. 2012, 134, 3164−3170. (32) Creutz, C.; Sutin, N. Photogeneration and reactions of cobalt(I) complexes. Coord. Chem. Rev. 1985, 64, 321−341. (33) Xie, J.; Zhou, Q.; Li, C.; Wang, W.; Hou, Y.; Zhang, B.; Wang, X. An unexpected role of the monodentate ligand in photocatalytic hydrogen production of the pentadentate ligand-based cobalt complexes. Chem. Commun. 2014, 50, 6520−6522. (34) Tong, L.; Kopecky, A.; Zong, R.; Gagnon, K. J.; Ahlquist, M. S.; Thummel, R. P. Light-Driven Proton Reduction in Aqueous Medium Catalyzed by a Family of Cobalt Complexes with Tetradentate Polypyridine-Type Ligands. Inorg. Chem. 2015, 54, 7873−7884. (35) Li, X.-J.; Jiang, F.-L.; Wu, M.-Y.; Zhang, S.-Q.; Zhou, Y.-F.; Hong, M.-C. Self-assembly of discrete M6L8 coordination cages based on a conformationally flexible tripodal phosphoric triamide ligand. Inorg. Chem. 2012, 51, 4116−4122. (36) Yoneya, M.; Yamaguchi, T.; Sato, S.; Fujita, M. Simulation of metal−ligand self-assembly into spherical complex M6L8. J. Am. Chem. Soc. 2012, 134, 14401−14407. (37) Moon, D.; Kang, S.; Park, J.; Lee, K.; John, R. P.; Won, H.; Seong, G. H.; Kim, Y. S.; Kim, G. H.; Rhee, H. Face-Driven CornerLinked Octahedral Nanocages: M6L8 Cages Formed by C3-Symmetric Triangular Facial Ligands Linked via C4-Symmetric Square Tetratopic PdII Ions at Truncated Octahedron Corners. J. Am. Chem. Soc. 2006, 128, 3530−3531. (38) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (39) Deshmukh, M. S.; Yadav, A.; Pant, R.; Boomishankar, R. Thermochromic and mechanochromic luminescence umpolung in isostructural metal−organic frameworks based on Cu6I6 clusters. Inorg. Chem. 2015, 54, 1337−1345. (40) Aizawa, S.-i.; Funahashi, S. Octahedral-Tetrahedral Equilibrium and Solvent Exchange of Cobalt(II) Ions in Primary Alkylamines. Inorg. Chem. 2002, 41, 4555−4549. (41) Pandey, S.; Das, P. P.; Singh, A. K.; Mukherjee, R. Cobalt(II), nickel(II) and copper(II) complexes of a hexadentate pyridine amide ligand. Effect of donor atom (ether vs. thioether) on coordination geometry, spin-state of cobalt and MIII−MII redox potential. Dalton Trans. 2011, 40, 10758−10768. (42) Silva, T. F. S.; Martins, L. M. D. R. S.; Guedes da Silva, M. F. C.; Fernandes, A. R.; Silva, A.; Borralho, P. M.; Santos, S.; Rodrigues, C. M. P.; Pombeiro, A. J. L. Cobalt complexes bearing scorpionate ligands: synthesis, characterization, cytotoxicity and DNA cleavage. Dalton Trans. 2012, 41, 12888−12897. (43) Feldman, B. J.; Gheller, S. F.; Bailey, G. F.; Newton, W. E.; Schultz, F. A. Electrochemical cells for voltammetry, coulometry, and protein activity assays of small-volume biological samples. Anal. Biochem. 1990, 185, 170−175. (44) Kochubey, D.; Kaichev, V.; Saraev, A.; Tomyn, S.; Belov, A. Y. Combined X-ray Absorption Near-Edge Structure and X-ray Photoelectron Study of the Electrocatalytically Active Cobalt(I) Cage Complexes and the Clathrochelate Cobalt(II)- and Cobalt(III)Containing Precursors and Analogs. J. Phys. Chem. C 2013, 117, 2753−2759. (45) Felton, G. A. N.; Glass, R. S.; Lichtenberger, D. L.; Evans, D. H. Iron-only hydrogenase mimics. Thermodynamic aspects of the use of electrochemistry to evaluate catalytic efficiency for hydrogen generation. Inorg. Chem. 2006, 45, 9181−9184. (46) Khnayzer, R. S.; Thoi, V. S.; Nippe, M.; King, A. E.; Jurss, J. W.; El Roz, K. A.; Long, J. R.; Chang, C. J.; Castellano, F. N. Towards a comprehensive understanding of visible-light photogeneration of hydrogen from water using cobalt(II) polypyridyl catalysts. Energy Environ. Sci. 2014, 7, 1477−1488. (47) Mondal, B.; Dey, A. Development of air-stable hydrogen evolution catalysts. Chem. Commun. 2017, 53, 7707−7715. (48) Dey, S.; Rana, A.; Crouthers, D.; Mondal, B.; Das, P. K.; Darensbourg, M. Y.; Dey, A. Electrocatalytic O2 Reduction by [Fe-Fe]F

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Hydrogenase Active Site Models. J. Am. Chem. Soc. 2014, 136, 8847− 8850. (49) Swanson, K. D.; Ratzloff, M. W.; Mulder, D. W.; Artz, J. H.; Ghose, S.; Hoffman, A.; White, S.; Zadvornyy, O. A.; Broderick, J. B.; Bothner, B.; King, P. W.; Peters, J. W. [FeFe]-Hydrogenase Oxygen Inactivation Is Initiated at the H Cluster 2Fe Subcluster. J. Am. Chem. Soc. 2015, 137, 1809−1816.

G

DOI: 10.1021/acs.inorgchem.7b02074 Inorg. Chem. XXXX, XXX, XXX−XXX