“This document is the Accepted Manuscript version of a Published Work that appeared in final form in Mater. Today, copyright ©2018 Elsevier B.V.after peer review and technical editing by the publisher. To access the final edited and published work see: https://www.sciencedirect.com/science/article/pii/S2468606918301060 Light-driven water oxidation using hybrid photosensitizer-decorated Co 3O4 nanoparticles Jonathan De Tovara, Nuria Romeroa, Sergey Denisovb, Roger Bofilla, Carolina GimbertSuriñachc, Diana Ciuculescu-Pradinesd, Samuel Drouetd, Antoni Llobeta,c, Pierre Lecantee, Vincent Colliered, Zoraida Freixaf, Nathan McClenaghanb, Catherine Amiensd, Jordi GarcíaAntóna,*, Karine Philippotd,*, and Xavier Salaa,* a Departament de Química, Facultat de Ciències Universitat Autònoma de Barcelona 08193 Bellaterra, Catalonia (Spain) b Institut des Sciences Moléculaires, Université Bordeaux / CNRS 351 Cours de la Libération 33405 Talence Cedex (France) c Institut Català d’Investigació Química (ICIQ) Barcelona Institute of Science and Technology (BIST) Av. Països Catalans 16 43007 Tarragona, Catalonia (Spain) d LCC-CNRS, Université de Toulouse, CNRS, UPS 205, route de Narbonne F-31077 Toulouse (France) E-mail: e CNRS, CEMES (Centre d'Elaboration de Matériaux et d'Etudes Structurales) 29 rue J. Marvig F-31055 Toulouse (France) f Department of Applied Chemistry, Faculty of Chemistry University of the Basque Country (UPV-EHU) 20080 San Sebastián (Spain) & IKERBASQUE, Basque Foundation for Science Bilbao (Spain) * Corresponding authors: E-mail adresses: Jordi.GarciaAnton@uab.es (J. García-Antón), Xavier.Sala@uab.cat (X. Sala), Karine.Philippot@lcc-toulouse.fr (K. Philippot) 1 Keywords: Co3O4 photocatalysis nanoparticles, water oxidation, 2 dyad systems, electrocatalysis, ABSTRACT Cobalt nanoparticles (NPs) have been prepared by hydrogenation of the organometallic complex [Co(3C8H13)(4-C8H12)] in 1-heptanol in the absence of any other stabilizer and then transformed to Co3O4 NPs using mild oxidative reaction conditions. After deposition onto glassy carbon rotating disk electrodes, the electrocatalytic performance of the Co3O4 NPs in water oxidation has been tested in 1M NaOH. The activity has been benchmarked with that of state-of-the-art Co3O4 NPs through electrochemically-active surface area (ECSA) and specific current density measurements. Furthermore, the covalent grafting of photosensitive polypyridyl-based RuII complexes onto the surface of Co3O4 NPs afforded hybrid nanostructured materials able to photo-oxidize water into O2, while steady-state and time-resolved spectroscopic measurements gave some further insight into kinetics and pertinent reaction steps following excitation. These first-row transition metal oxide hybrid nanocatalysts display better catalytic performance than simple mixtures of non-grafted photosensitizers and Co3O4 NPs, thus evidencing the advantage of the direct coupling between the two entities for the photo-induced water oxidation reaction. elements[12,13] have been successfully employed in photocatalytic WO. Transition from homogeneous to colloidal/heterogeneous species takes place recurrently during WO catalysis[ 14 , 15 , 16 ], particularly when first row transition metals and/or easily oxidizable ligands are employed[17]. Despite the high activity often shown by these in situ formed species, their size, composition and reactivity are poorly controlled. Therefore, the ex situ synthesis of metal nanoparticle (NP) catalysts with wellcontrolled size and surface properties might allow a better tuning of their catalytic performance. In this regard, the WO photocatalytic performance of colloidal dye (PS)-decorated IrOx suspensions has been reported by Mallouk and co-workers[ 7,8,18]. In these hybrid dyads, the PS serves both as NP stabilizer and light-harvester, thus helping to overcome the thermodynamically unfavorable barrier associated with water oxidation. Also, Frei and collaborators have reported that the use of electron conducting p-oligo(phenylenevinylene) molecular wires covalently bonded to the Co3O4 core of Co3O4/SiO2 core/shell NPs boost hole injection from oxidized PS molecules either present in solution[ 19,20] or electrostatically adsorbed onto the SiO2 shell[21] to the Co3O4 core. However, to our knowledge, no examples of WOCs are known so far in which the PS is covalently bound to abundant first row transition metal oxide NPs, which would reduce the global cost of the water splitting process. Recent literature data thus underline the interest of a rational design of hybrid PS-NP catalysts for the WO reaction. Within this context, the organometallic approach, which is recognized to be highly efficient to prepare NPs with high control over the size, composition and surface properties[22], may offer new opportunities in the synthesis of hybrid PS-NP materials with covalent bonding between the PS and the NP surface. In this way, well-controlled hybrid catalytic materials would be accessible. Eventually, they would permit studies of the influence of the direct link between the PS and the NP surface on their reactivity, adjusting the dyads to reach better catalytic performances. With this idea in mind, we decided to investigate the grafting of polypyridyl- 1. Introduction Non-renewable fossil fuels are still nowadays the main energy source used by mankind. However, their fast depletion due to the constant increase of the global energy consumption and their relationship with worrying levels of greenhouse gases and climate change make the development of less polluting and renewable energy sources a central topic for the scientific community. In this context, the development of sunlight-driven water-splitting procedures is gaining momentum[ 1 ]. The water splitting reaction, in which both oxygen and hydrogen gas are generated in the anode and cathode, respectively, represents an attractive method for obtaining energy in the form of the highly energetic H–H chemical bond[2], as long as the energy used to produce the hydrogen gas is renewable. Nevertheless, the anodic oxidation of water into O2 is a thermodynamically uphill, mechanistically complex and kinetically slow process, in which four electrons have to be removed from two water molecules and an O=O double bond has to be formed[3]. With the aim of improving the kinetics of this half reaction and inspired by Nature’s photosystem II, which is responsible for oxygen formation during photosynthesis, a range of photocatalytic systems has been developed during the past years[4,5]. Thus, several homogeneous and heterogeneous water oxidation catalysts (WOCs) have been tested in the presence of a photosensitizer (PS), a chromophore able to harvest photons and harness their energy to transfer electrons[ 6 ]. However, the catalytic performance of these systems is often limited by the insufficient rate of electron transfer from the WOC to the PS and the undesired back-electron transfer phenomena between the PS and the WOC[6,7]. In this context, covalently-bound molecular PS-WOC dyad systems have proven to be more efficient in photocatalytic WO since the rate of electron transfer from the WOC to the PS is significantly faster[8,9]. Thus, molecular PS-WOC dyad systems based on second and third row transition elements[ 9,10,11] and abundant first row 3 RuII complexes acting as the PS at the surface of Co3O4 NPs prepared by organometallic chemistry as a strategy to obtain advanced WO photocatalysts. Therefore, herein we report the synthesis and full characterization of novel hybrid PS-NP materials obtained by covalently anchoring light-harvesting [Ru(bpy)3]2+ derived-complexes (bpy = 2,2’bipyridine) to preformed abundant first row cobalt oxide NPs (Co3O4 NPs), and their use as catalysts in the photoinduced oxidation of water. Their catalytic performance is compared with that of colloidal mixtures of non-bonded PS and Co3O4 NPs, and Co3O4 NPs alone. Finally, flash photolysis/transient absorption methods as well as detailed photochemical analyses are used to gain further insight into the specific processes and rates associated with each component of the catalytic system and thoroughly discussed. controlled surface chemistry, the reducing agent and metal precursor should be carefully chosen. One of the best precursors in this regard is the organometallic complex [Co(3-C8H13)(4-C8H12)], which upon hydrogenation releases only cobalt atoms and cyclooctane, a non-coordinating molecule[ 24 ]. This synthetic strategy avoids competition between the stabilizing agent introduced and reaction by-products observed in other cases. The NPs produced by this way display high reactivity towards air[ 25 , 26 ], even at low temperature, which is a prerequisite to keep the morphological parameters unchanged during the oxidation process and reach Co3O4 NPs of controlled average size and size distribution. To facilitate the eventual grafting of the polypyridyl RuII complexes on the surface of the NPs, the use of strongly coordinating stabilizers like carboxylic acids should be avoided. For this reason and based on our former work on Ru NPs, 1-heptanol was chosen as solvent and stabilizer for the preparation of Co NPs[27]. Briefly, the synthesis of Co NPs was performed by decomposition of a 1-heptanol solution of the organometallic precursor [Co(3-C8H13)(4-C8H12)] under 3 bar of H2 at room temperature. TEM (Transmission Electron Microscopy) analysis carried out from the so-obtained crude dark-brown colloidal solution revealed the formation of a monodisperse population of spherical NPs of mean size 3.0 ± 0.1 nm (Fig. 1). These particles are welldispersed on the TEM grid and display a narrow size distribution with a standard deviation below 5% of the mean size. 2. Results and Discussion 2.1. Synthesis and characterization of the nanomaterials Since the direct synthesis of ultrafine Co3O4 NPs is a challenging process[23], we designed a threestep strategy to obtain the target PS-Co3O4 NPs hybrids: 1) synthesis of Co NPs with an average size of a few nanometers and narrow size distribution; 2) oxidation of these NPs into the target Co3O4 nanomaterial while keeping their morphological features unchanged, and 3) grafting of RuIIpolypyridyl complexes with anchoring groups on the surface of the Co3O4 NPs. A literature survey revealed that to reach Co NPs with an average size of a few nanometers, narrow size distribution and 4 Fig. 1. From left to right, synthesis and TEM images of Co, Co3O4 and PS-Co3O4 NPs (PS in the example shown is PS1). Bottom left: Schematic drawing of [Ru(bpy)3]2+ (PS0) and modified [Ru(bpy)3]2+ complexes with 2 or 4 phosphonic acid coordinating pending groups (PS1 and PS2, respectively) used as photosensitizers. heptanol, indicating the absence of extended aggregation during the oxidation process. The Recovery of the NPs from the 1-heptanol average size of the final NPs was estimated to be 3.0 colloidal solution was performed by application of a ± 0.2 nm (Fig. 1). Taking into account the change in magnet on the reactor walls (magnetic filtration). material density (8.9 g·cm-3 for bulk Co compared to Then, successive washings with pentane followed by 6.11 g·cm-3 for bulk Co3O4) one could expect the drying of the obtained solid under vacuum afforded volume of the NP to double upon oxidation, leading Co NPs under the form of a black fine powder, which to an expected diameter value of around 3.9 nm. This was used for further characterizations. WAXS apparent absence of volume change was already (Wide-Angle X-Ray Scattering) analysis of the reported for the study of the mild oxidation of Co sample confirmed the presence of Co NPs in the α NPs embedded in a polymer matrix[25]. This can be (hcp) and ε (metastable cubic) crystalline structures related to the lower density of the Co oxide NPs, as evidenced by the good match observed between which makes the determination of their exact experimental data and the combination of reference diameter from TEM images less precise and may patterns (PDF 01-080-6668 and PDF 04-017-5578, induce its systematic underestimation. Oxidation of respectively, Fig. 2) and with a coherence length of the CoNPs into Co3O4 could be confirmed by ca. 2.5 nm (Fig. S1). The Co content in the sample WAXS, XPS (X-Ray Photoelectron Spectroscopy) determined by ICP-OES (35.35%) suggests an analyses and by IR spectroscopy. empirical formula (1-heptanol)0.9Co1 (see Supporting Information) which points to more than WAXS measurements evidenced the presence of one 1-heptanol molecule per surface Co atom, nanoparticles displaying a coherence length of ca. suggesting the formation of interacting multilayers 3.0 nm (Fig. S1) and a crystalline structure around the NPs. corresponding to the Co3O4 phase as shown by the good match observed between experimental data and the reference pattern (PDF 04-005-4386) (Fig. 2). The perfect agreement between the coherence length and the mean size determined by TEM points at first sight towards well crystallized nanoparticles. This is surprising given the mild conditions used during the oxidation of the Co NPs and would rather suggest that the diameter determined from the analysis of the TEM images is underestimated. In XPS (Fig. S2), the Co3O4 NPs show two main peaks at ca. 780.0 and 795.5 eV, corresponding to the Co 2p3/2 and Co 2p1/2 components, respectively, accompanied by two broad satellite peaks with very low intensity at higher energies (ca. 788 and 805 eV). According to the literature data, the presence of satellite peaks are an indication of the presence of unpaired electrons in the sample, i.e. the presence of CoII (d7) atoms[28], while both the positions and the intensities of all 4 types of bands observed in the XPS spectrum clearly match the data already reported for the mixed CoII-CoIII oxide Co3O4[28,29]. IR spectroscopy revealed the presence of typical Fig. 2. WAXS analysis of Co NPs (top) and Co3O4 Co-O stretching bands centered at 666 and 575 cm-1 NPs (bottom) in comparison with Co  / Co  and [30] within the Co3O4 structure (green line, Fig. 3) Co3O4 phase diagrams for Co NPs and Co3O4 NPs, confirming the crystallinity and structure of the respectively. nanomaterial. Poorly defined absorptions in the 1600-1400 cm-1 region could correspond to adsorbed Complete conversion of the Co NPs previously water and carbonate molecules. All together these described into Co3O4 NPs was achieved in soft results show the transformation of preformed Co reaction conditions by simple exposure of the NPs into crystalline Co3O4 NPs while preventing particles in the solid state under ambient air for 6 aggregation and coalescence of the NPs thanks to the days. The NPs are well dispersed on the TEM images use of soft reaction conditions. recorded after dissolution of the powder in 1- 5 the 1465-1394 cm-1 region, indicative of the presence of the bipyridine backbone of the PS. No free P=O nor P-OH bands are visible, but there is a band at 1061 cm-1 that can be attributed to -P(O-Co) functionalization[35] (by comparison with the 1016 cm-1 value found for -P(O-Zn)3), thus supporting the grafting of PS1 to the surface of Co3O4 NPs and the use of all possible anchoring points. Additionally, neither -CH3 nor -CH2- stretching bands are clearly observed at ca. 2900 cm-1 (such as those seen for Co3O4 NPs (red spectrum) in Fig. 3), indicating that a partial replacement of the 1-heptanol molecules by the PS1 molecules at the NP surface may have taken place and/or that part of the 1-heptanol initially present could have been released from the NPs surface during dialysis. Therefore, we cannot exclude the possible presence of residual 1-heptanol molecules on the surface of the hybrid NPs. TEM analysis from the aqueous colloidal dispersion of the hybrid material revealed the presence of nanoobjects with homogeneous shape (spherical) that display a mean size of 3.1 ± 0.3 nm (Fig. 1). The results show that the initial size and morphology of the Co3O4 NPs are maintained after anchoring the PS1 at their surface. The presence of Ru only in the vicinity of the NPs was evidenced through STEMEDX analysis, indicating a successful purification process (Fig. S4). The PS2-Co3O4 NP hybrid material was characterized accordingly, with very similar results to the PS1-Co3O4 hybrid material (Fig. S3, Fig. S5 and Fig. S6). Based on ICP-OES analysis, incorporation of 19 PS1 and 32 PS2 complexes per Co3O4 NP is estimated (see Supporting Information for further details). The higher grafting density obtained in the case of the PS2-Co3O4 hybrid could be related to the statistically more favored interaction between the surface and the chelating biphosphonate bipyridine ligands. To sum-up this section, easily obtained Co NPs could be transformed into Co3O4 NPs in mild reaction conditions while preserving their morphology. These results highlight: 1) the efficiency of the synthetic approach, where 1heptanol acts as both solvent and stabilizer for the preparation of small and size-controlled Co NPs; 2) the possibility to access small Co3O4 NPs by a soft oxidative treatment with a preserved morphology, and 3) the possibility to graft RuII-polypyridyl complexes at the Co3O4 NPs surface, despite the coverage of the surface by 1-heptanol molecules, to get hybrid nanostructured materials with multiple functionalities, namely PS and catalyst. The catalytic properties of the obtained Co3O4 NPs and hybrid PSCo3O4 NPs materials are described hereafter. 2.2. Electrocatalytic behavior in water oxidation catalysis The electrocatalytic performance of the prepared ultrafine Co3O4 NPs was studied in 1M NaOH solution after their deposition onto a glassy carbon Fig. 3. Overlay of ATR-IR spectra of Co3O4 NPs (red), free PS1 (black) and PS1-Co3O4 NPs hybrid material (blue) and of IR spectrum of a KBr pellet of Co3O4 NPs (green). The transmittance of each sample has been shifted along the y-axis for comparison purposes. The positions of the bands described along the text have been added as dotted lines. The next step was the building of the hybrid nanomaterials (PS-Co3O4 NPs) by grafting RuIIpolypyridyl complexes at the surface of the Co3O4 NPs. For this purpose, the [Ru(bpy)3]2+ complexes bearing phosphonic acid pending groups (PS1 with two and PS2 with four) were prepared following literature procedures[31,32] (Fig. 1). The phosphonic acid anchoring group is well known to efficiently interact with metal oxide surfaces[ 33 ]. The PS grafting was performed by mixing a 1-heptanol colloidal dispersion of Co3O4 NPs with a methanol solution of the chosen complex ([RuII complex]/[Co3O4 NPs] = 0.24) and leaving the obtained reaction mixture under vigorous stirring in the dark for 4 days. The nanohybrid materials were recovered by centrifugation and further purified from unreacted Ru complexes by dialysis against deionized water. This procedure was applied for PS1 and PS2 complexes giving rise to the PS1-Co3O4 and PS2-Co3O4 hybrids, which could be recovered as black powders after evaporation of water under vacuum. From ICP-OES analysis, it can be inferred that these nanomaterials show (1heptanol)0.80(PS1)0.09-Co3O4 and (1heptanol)0.97(PS2)0.15-Co3O4 empirical formulas (see Supporting Information). The attachment of PS1 and PS2 complexes at the surface of the Co3O4 NPs was attested by ATR-IR spectroscopy (Fig. 3 and Fig. S3, respectively). As shown in Fig. 3 (black spectrum) the PS1 complex shows two absorption bands at 1149 and 928 cm-1 corresponding to the free P=O and P-OH units, respectively[34,35]. The PS1-Co3O4 hybrid nanomaterial (Fig. 3, blue spectrum) shows bands in 6 rotating disk electrode (GC-RDE, Fig. S7). The electrode was prepared by depositing three 5 L drops from a THF dispersion (1 mg of Co3O4 NPs in 250 L of THF) of the NPs onto the glassy carbon disk of the RDE. Fig. S7a shows the rotating disk voltammetry (RDV) of the as-deposited Co3O4 NPs, where a steep increase in intensity above 0.7 V vs. NHE is visible. This increase in intensity is attributed to the oxidation of water into molecular oxygen[36,37,38], which in our case happens at an onset overpotential () of ca. 0.29 V. The electrocatalytic performance of our Co3O4 NPs deposited onto GC-RDE was compared with that of other electrocatalysts following the benchmarking methodology reported by Jaramillo et al.[36,37,38]. Thus, the electrochemically-active surface area (ECSA) of the Co3O4 modified GCRDE was estimated to be 0.175 cm2 from the electrochemical double-layer capacitance (Cdl) by measuring the non-Faradaic capacitive current associated with double-layer charging from the scanrate dependence of cyclic voltammograms (CVs)[ 39 , 40 ] over a 0.145-0.245 V vs. NHE potential range (Fig. S9). The roughness factor (RF) was calculated by dividing the estimated ECSA by the geometric area of the electrode, and these factors as well as those corresponding to the electrocatalytic activity of our nanocatalyst are compared with those of the state-of-the-art Co3O4 NPs in the same electrolyte in Table 1. However, it is important to note that the ECSA serves only as an approximate guide for the determination of the RF, since the accuracy of the data lies normally within an order of magnitude[37]. Therefore, comparison with literature data can only be analyzed in terms of general trends. As depicted in Table 1, the Co3O4 modified GC-RDE catalyst reported herein (entry 1) shows an onset overpotential (ηonset) of ca. 0.29 V vs. NHE, which lies close to the reported values for Co3O4 NPs in graphene[41,42] (entries 3 and 6) and single walled carbon nanotubes (SWCNTs)[ 43 ] (entry 7). To achieve a current density of 10 mA·cm2 , approximately the current density expected for a 10% efficient solar-to-fuel conversion device[37, 44 , 45 , 46 ], a η10mA/cm2 of 0.486 V is required (entry 1, this work). This value is close to the η10mA/cm2 reported for Co3O4 NPs of similar RF by Jaramillo et al.[38] (entry 4) but higher than the η10mA/cm2 values published for catalytic systems with nearly two orders of magnitude higher RF[47,48] (entries 2 and 5). Even so, it falls within the reported area of interest for catalyst benchmarking[36,37,38]. Most interestingly, when the current at an η of 0.35 V (jg) is normalized by the ECSA[ 49 , 50 ], a normalized current density (js) of 1.04 mA cm-2 is obtained. This value is significantly higher than that reported by Jaramillo et al. for their Co3O4-based nanocatalyst of similar RF and η10mA/cm2 (entry 4) and also for all other nanostructured metal oxide electrocatalysts deposited onto GC-RDE[38]. Thus, the Co3O4 modified GC-RDE reported herein is very active at low η (0.35 V). The system becomes less competitive at higher overpotentials due to the relatively high slope of the Tafel plot above η = 0.35 V (approx. 100 mV·dec-1, Fig. S7b). Table 1. Benchmarking parameters vs. NHE for Co3O4 NPs in 1 M NaOH and comparison with state-of-the-art data in the same electrolyte. Entry Catalyst ECSA /cm2 RF ηonset/V η10mA/cm2, η10mA/cm2, t =0/V t=1h/V jg,η=0.35 V /mA·cm-2 js,η=0.35 V /mA·cm-2  (%) Ref. 1 3.0 nm Co3O4 NPsa 0.175 2.5 0.29 0.486h 0.480h 2.6 1.04 95g this work 2 10 nm Co3O4 NPsb 429 429 --- 0.290 --- ca. 25 ca. 0.06 95 47 3 ca. 10 nm Co3O4 NPsc --- --- 0.24 0.313 0.313 ca. 30 --- --- 41 4 ca. 70 nm Co3O4 NPsd 1.52 7.8 --- 0.50 0.51 0.06 0.039 --- 38 5 50 nm Co3O4 nanocubese 6.44 91.2 --- 0.28 --- ca. 70 ca. 0.77 --- 48 6 4-8 nm Co3O4 NPsf --- --- ca. 0.27 0.31 --- ca. 35 --- --- 42 7 6 nm Co3O4 NPsg --- --- ca. 0.28 0.593 --- ca. 1.8 --- --- 43 a Using a GC-RDE; b deposited on Ni foam; c in graphene nanocomposite; d deposited onto GC-RDE; e In N-doped graphene nanocomposite; f in N-doped graphene deposited onto Ni foam; g in SWCNTs; h deposited onto an FTO electrode. FTO electrode was loaded with 15 L of a dispersion of Co3O4 NPs (2 mg in 500 L of 1-heptanol) by spin-coating, and afterwards a poly(methyl Both stability and Faradaic efficiency () are key parameters for a catalyst to be suitable for practical applications in WO catalysis. To analyze them, an 7 *PSx+ + S2O82-  PS(x+1)+ + SO42- + SO4-. (Equation 2) methacrylate) (PMMA) layer was added as “gluing” material[ 51 ]. The generated FTO/Co3O4NPsPMMA electrode showed onset  of ca. 0.29 V (see Fig. S10b), identical to that of the Co3O4 NPs on GCRDE (see Table 1), highlighting the negligible effect of the support used. The FTO/Co3O4NPs-PMMA electrode was then held at a constant current density of 10 mA·cm-2 in a current-controlled experiment for 1h in 1M NaOH. As shown in Fig. S10a, the system showed a stable operating potential, changing negligibly from 10mA/cm2, t=0 h = 0.486 V to 10mA/cm2, t=1 h = 0.480 V during 1 h (entry 1, Table 1). Comparison of the CV polarization curves measured before and after this current-controlled experiment shows a slight increase in the observed current density after catalytic turnover (Fig. S10b). This global increase in current density is indicative of a certain activation of the Co ions in the Co3O4 NPs. An increase in the Co+3/+4/Co+2 population ratio, as suggested by Frei and co-workers[52], and / or the elimination of 1-heptanol molecules present at the surface of the NPs increasing the number of exposed active sites are plausible reasons for this behavior. XPS analysis of the resulting electrode after electrolysis shows the intact composition of the NPs as Co3O4 (Fig. S11), thus pointing to the removal of 1-heptanol molecules from their surface under catalytic conditions as the origin of the observed activation process. Furthermore, the activation of the NPs does not significantly affect the Tafel slopes (Fig. S10c), which are very similar before and after catalytic turnover. In addition, a Faradaic efficiency of 95% was determined by quantifying the amount of O2 generated during a bulk electrolysis (0.886 V vs. NHE corresponding to an initial current density of 10 mA·cm-2) using an O2-probe and dividing it by the theoretical O2 amount calculated from the total charge passed through the system (Fig. S12), thus confirming the production of O2 as the only reaction taking place. PSx+ + SO4-.  PS(x+1)+ + SO42- (Equation 3) The influence of PS concentration on the turnover number (TON) has been studied with PS1 in order to determine the best [PS]/[Co3O4(heptanol)2.8 units] ratio to be used (Fig. S16). The curve (TON vs. PS1 equiv) obtained reaches a plateau at ca. 6.0 equiv of PS. Consequently, a PS:Co3O4(heptanol)2.8 units ratio of 6.0:1.0 was chosen for the following photochemical WO studies. As reported in Table 2 (entries 1-3), all the tested PS (Fig. 1) could oxidize the Co3O4 NPs under catalytic conditions. These results are consistent with the electrochemical analysis of the Co3O4 NPs shown above, since their onset potential at pH 5.6 (ca. 1.1 V vs. NHE, Fig. S17) is lower than that of the RuIII/RuII redox couple for all PS tested (1.2-1.3 V vs. NHE, Fig. S18). TON and TOF (turnover frequency, min-1) values were determined from the estimated total number of NPs and of PS molecules present on each sample (as presented in the Supporting Information, these values are underestimated; yet relative values can be obtained). Thus, the TON per NP obtained for the catalytic mixtures made of Co3O4 NPs plus PS0, PS1 or PS2 are all similar and within the range 453-604 (entries 1, 2, 3 in Table 2). Concerning the TOF per NP, it is reduced by approximately half when doubling the number of phosphonate groups present (67.9 vs. 36.2 min-1 for PS1 and PS2, respectively; entry 2 vs. entry 3). The TOF per PS is also reduced by a similar amount (0.053 vs. 0.028 min-1 for PS1 and PS2, respectively). One potential reason for the reduced TOF values for the PS2 case could be that the presence of four phosphonate binding groups could allow the simultaneous binding of a single PS2 unit onto two Co3O4 NPs, thus favoring their aggregation, in contrast to the PS1 system. This hypothesis is confirmed when comparing the TEM images of the as-synthesized hybrid PS1-Co3O4 and PS2-Co3O4 systems (Fig. 1 and Fig. S5, respectively), where higher aggregation and lower dispersion of the NPs are obtained for PS2-Co3O4. At this point, it is also worth mentioning that the limiting factors that may stop the O2 evolution in such photocatalytic systems are usually the degradation of the PS and the pH decrease due to the release of protons during the reaction[53,54,55]. However, it is interesting to note that O2 evolution was not resumed under our conditions after the addition of an extra aliquot of PS. Thus, some kind of inhibition of the whole photocatalytic system and not only PS degradation takes place. This inhibition could be caused by the progressive increase in the ionic strength of the medium (sulfate ions are produced during catalysis), as this can negatively affect the performance of the NPs due to a reduction of the quenching efficiency of the photoexcited PS (*PSx+) by peroxodisulfate 2.3. Photochemical water oxidation catalysis The efficiency of the Co3O4 NPs as a photocatalyst for WO was first evaluated in Na2SiF6NaHCO3 (0.02-0.04 M, pH 5.60) in the presence of PS0 ([Ru(bpy)3]2+) and its phosphonate derivatives PS1 or PS2 as photosensitizers (Fig. 1), using sodium peroxodisulfate as the sacrificial electronacceptor (SEA) and liquid phase Hansatech-type microsensors for measuring the evolved oxygen (Fig. S13, Fig. S14 and Fig. S15). It is noteworthy that the semiconducting Co3O4 NPs alone do not behave as photocatalysts for this reaction. The photoactive species responsible for the activation of the Co3O4 NPs in WO is PS(x+1)+, which is generated after a three-step process as follows: PSx+ + h  *PSx+ (Equation 1) 8 (Equation 2)[56]. In addition, another phenomenon that typically explains the reduced performance of nanocatalysts with time in the presence of the NaSiF6-NaHCO3 buffer is the hydrolysis of the buffer to generate SiO2 particles, which can provoke the adsorption of the cationic PS molecules onto their surface, thus competing with the catalytic NPs and reducing the global catalytic performance[18]. This last hypothesis has been confirmed by HREM (High Resolution TEM) and STEM (Scanning TEM) analyses of the recovered nanocatalyst after photocatalytic turnover in the presence of PS1, in which Co3O4 NP aggregates of ca. 50 nm attached to a bigger Si-containing aggregate are observed (Fig. S19). Table 2. TON and TOF (min-1) per NP and per PS obtained as a function of PS nature in photochemical WO measurements with single Co3O4 NPs and hybrid PS-Co3O4 NPs at pH 5.6. Entry System PS:Co3O4 ratio TON (O2/NP) TOF min-1 (O2/NP) TON (O2/PS) TOF min-1 (O2/PS) 1 Co3O4 + PS0 6.0:1.0 453 49.4 0.35 0.038 2 Co3O4 + PS1 6.0:1.0 566 67.9 0.44 0.053 3 Co3O4 + PS2 6.0:1.0 604 36.2 0.47 0.028 4 Co3O4 + PS1 0.09:1.0 <1 - < 0.1 - 5 Co3O4 + PS2 0.15:1.0 <1 - < 0.1 - 6 PS1-Co3O4 0.09:1.0 5.4 0.90 0.28 0.046 7 PS2-Co3O4 0.15:1.0 82.0 2.05 2.53 0.063 (9.66·10-4 M) in the presence of PS1 (1.23·10-4 M) and Na2S2O8 (7.85·10-2 M). Khaki line: Co3O4 NPs (9.95·10-4 M) in the presence of PS2 (1.05·10-4 M) and Na2S2O8 (8.08·10-2 M). Blue line: PS1-Co3O4 NPs (1.02·10-3 M) in the presence of Na2S2O8 (7.32·10-2 M). Purple line: PS2-Co3O4 NPs (1.09·103 M) in the presence of Na2S2O8 (7.87·10-2 M). All measurements were performed in Na2SiF6-NaHCO3 (0.02-0.04 M, pH 5.60) buffer solution. Irradiation provided by a Xe lamp equipped with a 400 nm cutoff filter and calibrated to 1 sun (100 mW cm-2). T = 25°C. As shown in Table 2, the oxygen evolved by both unbound systems is almost negligible (entries 4 and 5 and Fig. 4). This can be attributed to the kinetic prevalence of the deactivation processes described above (PS degradation, NP aggregation) competing with oxygen evolution when very low concentrations of unbound PS are used. Conversely, the same PS / Co3O4 ratio is rather more active when bound PS-Co3O4 hybrid systems are employed (entries 6 and 7). These results could be expected in part if taking into account that in PS-NP dyad systems the electron transfer between both entities is obviously not limited by diffusion, as for the unbound systems. Furthermore, comparison of entries 6 and 7 in Table 2 shows the rather superior activity of PS2-Co3O4 (TON and TOF per NP of 82 and 2.05 min-1, respectively) versus PS1-Co3O4 (TON and TOF per NP of 5.4 and 0.90 min-1, respectively). Thus, the superior PS surface functionalization in PS2-Co3O4 Concerning the PS1-Co3O4 and PS2-Co3O4 hybrid materials, the number of PS per Co3O4 NP unit were estimated to be 0.09 and 0.15, respectively (see Supporting Information). Thus, in order to compare the photocatalytic performance of these hybrid dyads with that of the corresponding unbound systems, similar PS / Co3O4 ratios were applied under catalytic conditions for the two control experiments (Fig. 4). Note that for experiments in which little oxygen is evolved, we observe a decrease of the signal inside the chamber when exposed to light due to the reaction of the singlet state of PS with residual oxygen traces[57]. Fig. 4. Photocatalytic oxygen production by different Co3O4-PS systems. Green line: Co3O4 NPs 9 (32 PS molecules per NP vs. the 19 molecules present in PS1-Co3O4) enhances the kinetics of oxygen evolution and better stabilizes the catalytic system, increasing its durability under photocatalytic conditions (purple line, Fig. 4). When the kinetics of oxygen evolution are normalized by the PS concentration, TOF values of both hybrid systems get closer (rightmost column in Table 2, entries 6 and 7), thus confirming the relationship between the rate of photocatalytic turnover and the degree of functionalization of the NPs surface. In terms of stability, the weakness of P-O-M bonds has been extensively identified as a main deactivation pathway of grafted molecular complexes and dyesensitized systems when employed as catalysts for the oxidation of water[34,35,58]. Thus, the higher number of anchoring groups present in PS2 (4 phosphonate groups vs. the 2 present in PS1) can also contribute to the superior longevity of the PS2Co3O4 hybrid system. To further study the fate of the hybrid dyad nanocatalysts under photocatalytic WO conditions, the crude reaction mixture after a 1h photocatalytic test with PS1-Co3O4 was dialyzed against 2 L of deionized water for 4 days followed by centrifugation and air-drying. Some partial aggregation was observed by TEM analysis (Fig. S20), although less intense than for the non-anchored system (Fig. S19), thus confirming the above proposed protective role of the attached PS1 molecules against aggregation. Similar studies with Co3O4 NPs stabilized by phosphonate-derived ligands have also shown the preservation of the NP size along light driven WO catalysis[ 59]. On the other hand, IR spectroscopy showed the loss of PS1 (Fig. S21), since the intensity of the bpy bands at 1400-1500 cm-1 and the -P(O-Co) bands at ca. 1060 cm-1 decreases after photocatalysis. Also, ICP-OES analyses evidence the decrease in the [Ru] / [Co] ratio from 0.03 to 0.003 after catalysis. This loss in PS can be not only due to its decomposition, which is kinetically competitive with the oxidation of water[55], but also to its decoordination from the surface of the Co3O4 NPs, as commonly observed in related systems[34,35,58]. In summary, the results shown in this section highlight the benefits of the dyad approach where the direct connection between the Co3O4 nanocatalyst and the PS facilitates electron-transfer and stabilizes the system against aggregation under turnover conditions due to the protective effect of the PS at the surface of the Co3O4 NPs. However, they also emphasize the relative instability of the -P(O-Co) bonds under turnover conditions and the need of further research for developing more stable systems with higher durability. In order to rationalize the kinetics of light induced oxygen evolution described above, and the precise role and contribution of each component in the catalytic systems, a series of steady-state and time-resolved measurements were undertaken. Indeed, photo-oxidation of different PS upon addition of the electron acceptor peroxodisulfate, according to Equation 2, was investigated by steadystate MLCT luminescence quenching and emission lifetime changes of the inherently luminescent PS in Na2SiF6-NaHCO3 0.02-0.04 M, pH 5.60 (Fig. S22, see Supporting Information for experimental details). Information on the nature of the quenching processes (i.e. static vs dynamic) was sought via Stern-Volmer (SV) plots, I0/I vs. [S2O82-] (where I0 and I are the emission intensity of the excited PS (*PS) in the absence and presence of the quencher, respectively) for all PS (alone and attached to Co3O4 NPs), and are shown in Fig. S23. The SV plots for the PS1-Co3O4 and PS2-Co3O4 hybrid systems show a linear trend, while for the free PS the SV plots significantly deviate from linearity. These results are similar to those previously reported by Musaev[60] and Bard[61] for [Ru(bpy)3]2+ (PS0 in our work) and the same quencher in a different electrolyte. They showed that the SV plot is described by a model that takes into account the formation of ground-state ion pairs between the emitter and the quencher. In this case, two different quenching processes can occur: collisional or dynamic quenching (bimolecular pathway) or a static or complex formation quenching (unimolecular pathway)[62,63]. Dynamic quenching occurs when the excited photosensitizer collides with the quencher, following the conventional SternVolmer behavior. However, in some cases the photosensitizer can initially form a stable ion-pair complex with the quencher, followed by a photoexcitation of the whole system (Scheme S1). The model that considers both quenching processes at the same time gives rise to the following equation for quenching emission[61]: = ( )( ( ) )( ( ) (Equation 4) ) where Keq is the equilibrium constant of ion pair formation, and the unquenched unimolecular decay time and the bimolecular quenching constant of the excited emitter in the free form are, respectively, τ0 and kq, while in ion pair state are τ’ and kq’. Under conditions of negligible ion-pair formation, Keq[S2O82-]<<1, that is, under dynamic (or bimolecular) quenching, Equation 4 is simplified to the Stern-Volmer equation: = = = 1 + 𝑘 𝜏 [𝑆 𝑂 ] (Equation 5) while for Keq[S2O82-]>>1, that is, under static (or unimolecular) quenching 2.4. Photophysical studies 10 = [ ] PS there is a deviation from linearity of the SV plots at [S2O82-]<15 mM, which indicates the formation of a ground-state ion pair between PS and S2O82-. Assuming that kq and kq’ are similar[61], these SV plots can be fitted by Equation 4, and the results of all fittings are listed in Table 3. Also, the SV plots of PS0, at [S2O82-] > 50 mM show a change in the curvature, with an extrapolated intercept at the x axis origin away from (0,1) (Fig. S23), which suggests that adsorption of the cationic PS to negatively charged silica particles (originated from Na2SiF6 hydrolysis)[18] or dynamic quenching processes are taking place as a result of the increased ionic strength[61,64,65]. (Equation 6) and when 𝑘 𝜏 [𝑆 𝑂 ] in the denominator is large compared to 1, then = + 𝑘 𝜏 [𝑆 𝑂 ] (Equation 7) As shown in Fig. S23, data for PS1-Co3O4 and PS2-Co3O4 fit well into Equation 5, suggesting that the equilibrium constant corresponding to the formation of an S2O82-/PS-Co3O4 ion pair is negligible and most quenchers are not in the ion-pair ground-state. Thus, the quenching proceeds through a bimolecular pathway. On the contrary, for all free 11 Table 3. Results of kinetic analysis of the {PS*,S2O82-} and {Co3O4-PS*,S2O82-} systems at pH 5.6. En try System ΦO2a ΦArb Ar0,c ns O20, d ns NPs0 ,e ns ’,f ns 0/’g kq (kq’),h mol-1·L· s-1 Keq,i mol1 ·L kET,j s-1 kq[S2O82]k, s-1 kq[S2O82]l, s-1 1 PS0 no buffer 0.028 0.042 550 366 N/A 108 5.09 3.9·108 6.93·102 7.4·106 1.17·108 7.41·106 2 PS0 0.021 0.032 558 364 550 55 10.22 4.39·108 2.46·102 1.65·107 1.32·108 8.34·106 5 PS1 0.026 0.031 501 420 512 109 4.59 3.17·107 1.29·102 7.17·106 9.52·106 6.02·105 6 PS2 0.023 0.029 429 340 425 312 1.37 7.50·107 2.81·101 8.72·105 2.25·107 1.43·106 7 PS1Co3O4 0.025 0.030 497 414 497 - - 5.05·107 - - PS2Co3O4 0.022 8 9.60·105 1.52·107 0.030 412 327 412 - - a 1.94·107 - - 5.82·106 3.69·105 Quantum yield in air-saturated buffer solutions in the absence of quencher; b quantum yield in argon-saturated buffer solutions in the absence of quencher; c from lifetime measurements in the absence of S2O82- in an argonsaturated buffer solution; d from lifetime measurements in the absence of S2O82- in an air-saturated buffer solution; e from lifetime measurements in the absence of S2O82- in an argon-saturated buffer solution in the presence of Co3O4 NPs; f calculated from intercept of SV plots (τ0/ τ’) and τ0; g intercept of SV plots at variable [S2O82-]; h from slope of SV plots at variable [S2O82-]; i computational best fit to Equation 4 (see text); j Calculated from Equation 8 (see text); k Calculated for 0.3 M S2O82-; l Calculated for 0.019 M S2O82-. Adding Co3O4 NPs to the buffered solutions containing the different PS and measuring Data gathered in Table 3 shows that the quantum luminescence lifetimes in the presence of Ar (τNPs0) yield Φ (the ratio between emitted photons and allowed testing of significance of charge transfer absorbed photons) in the presence of oxygen (ΦO2) is between *PS and Co3O4 NPs (processes 3 and 7 in lower than in the presence of Ar (ΦAr) because of the Scheme 1). As shown in Table 3, no significant luminescence quenching effect of oxygen in the differences are observed between τAr0 and τNPs0. former case. Accordingly, lifetimes in the presence Thus, we can assume that no direct interactions occur of O2 (τO20) are shorter than in the presence of Ar between the *PS and free Co3O4 NPs. (τAr0). Moreover, the similarity of ΦAr in the free and hybrid systems (entries 5,6 vs. 7,8) indicates that no direct electron transfer process exists between the excited PS and the Co3O4 NPs in the hybrid systems. 12 Scheme 1. Scheme of pertinent processes leading to photoinduced water oxidation using prototype dye-sensitized cobalt oxide NPs in a photoelectrochemical system along with main energy levels and relevant desired (in green) and undesired (in red) electron transfer processes. (1) Dye photoexcitation; (2) electron injection/electron acceptor reduction; (3) hole injection/catalyst oxidation; (4) water oxidation; (5) dye (radiative or nonradiative) deexcitation; (6) electron-hole recombination to the oxidized dye; (7) oxidative dye excited-state quenching by the catalyst; (8) electron-hole recombination to the oxidized catalyst. Note that for the sake of simplicity the position of the different species along the y axis does not necessarily correspond to their energy level. behavior at pH 5.6 is observed at pH 8.4 for PS0 since PS0 is not affected by pH. On the other hand, the intrinsic radiative and nonradiative rate constants of {PS···S2O82-}* ion pairs Transient absorption spectroscopy was also used should be similar to those of the *PS due to the weak to study the difference between the free PS1 and the (a few kilocalories per mole) electrostatic interaction PS1-Co3O4 hybrid systems. No significant between PS and S2O82-[60]. Then, the photoinduced differences were observed between the free PS and unimolecular electron transfer (ET) rate (kET) can be that attached to Co3O4 NPs (Fig. S26), and in both cases the ground-state bleaching of PS1 at 452 nm estimated according to Equation 8, where  is the and the formation of the PS1* excited state near 360 observed lifetime and 0 is the unquenched lifetime. nm could be observed (Fig. S27). Thus, direct charge 𝑘 = − (Equation 8) transfer from the Co3O4 NPs to PS1* is not observed 2in any system, as otherwise deduced from the At 0.3 M concentration of S2O8 , the bimolecular 2comparison between τAr0 and τNPs0 (Table 3), ET rates kq[S2O8 ] are faster than the unimolecular meaning that PS1* remains excited until its reaction ET rates (kET) for all systems at pH 5.6 (Table 3), with the sacrificial electron acceptor S2O82- in suggesting that at this concentration S2O82agreement with luminescence analyses, and in deactivates the unimolecular quenching process accordance with Scheme 1. since it inhibits the formation of the ground-state {PS···S2O82-} ion pairs[60]. However, under catalytic conditions (0.019 M S2O82-, last column in 3. Conclusions Table 3) the bimolecular quenching is basically In conclusion, we have demonstrated that 1favored for PS2 (kq[S2O82-]  kET), whereas the heptanol serves as both an effective solvent and unimolecular quenching is preferential for PS0 and stabilizing agent for the synthesis of ultra-small Co PS1 (kET  kq[S2O82-]). NPs (ca. 3 nm), preventing their aggregation. The Co For all systems, their photo-oxidation by S2O82NPs have been oxidized into Co3O4 NPs by airhas also been studied at pH 8.4, in which PS1 and exposure in mild conditions preserving their PS2 are deprotonated (Table S1, Fig. S24 and Fig. morphology and dispersion. When deposited at the S25). Under these conditions we can see no surface of a GC-RDE electrode and in 1M NaOH, deviations from linearity for the Stern-Volmer plots, these Co3O4 NPs electrocatalytically oxidize water since the unimolecular ET mechanism is clearly with an onset  of ca. 0.29 V and η10mA/cm2 of 0.486 disfavored now due to the repulsion that appears V, showing ECSA normalized current densities (js) between the negatively charged PS and S2O82- at of 1.04 mA cm-2 at  = 0.35 V, a value that fairly basic pH. On the other hand, the same Stern-Volmer outperforms that of all benchmarked nanostructured metal oxide electrocatalysts deposited onto GC- 13 RDE. Despite stable and showing 95% Faradaic efficiency, the system is less competitive at higher current densities due to its Tafel slope of ca. 100 mV·dec-1 at  > 0.35 V. RuII photosensitizers displaying phosphonic acid pending groups (PS1 with two and PS2 with four) were attached to the surface of Co3O4 NPs, yielding PS-Co3O4 hybrid systems with a different degree of surface functionalization, namely incorporation of ca. 19 PS1 and 32 PS2 complexes per Co3O4 NP. The capacity of these dyad systems to photo-oxidize water into oxygen using visible light and S2O82- as sacrificial electron acceptor at pH 5.6 was evaluated and the results compared with those of unbound systems of the same components and concentrations. The benefits of the dyad approach arise when observing the inactivity of the unbound Co3O4/PS systems with regards to the significant TON and TOF values per NP (5.4 / 0.90 min-1 and 82 / 2.05 min-1) obtained for PS1-Co3O4 and PS2-Co3O4, respectively. The better catalytic performance of the latter over the former was attributed to the higher surface functionalization of PS2-Co3O4, which enhances the kinetics of WO and protects better the catalytic entity under working conditions against aggregation. These data stress the important role of the direct connection between the PS and the nanocatalyst by; 1) favoring their efficient electronic communication that allows being kineticallycompetitive with the typical side deactivation processes of light-driven WO and 2) minimizing catalyst aggregation under turnover conditions thanks to the protective / stabilizing effect of the PS coordinated at the surface of the Co3O4 NPs. Photophysical measurements show the determinant roles of the sacrificial electron acceptor, diffusion and ground state acceptor-PS complexes, and lack of direct electron transfer between NPs and *PS. In summary, this work opens the way towards precisely defined first row PS-NP hybrid dyads by means of the so-called organometallic approach as synthetic methodology, which provides wellcontrolled and fully characterized cobalt, cobalt oxide and RuPS-Co3O4 nanomaterials. The latter species have proven capable to photo-oxidize water into dioxygen at pH 5.6, being fairly superior to their unbound Co3O4 NPs / PS counterparts under identical conditions. Therefore, the fine tuning of this system through the length and nature of the PSCo3O4 NPs connection is expected to lead to a better understanding of the key parameters governing the catalytic process and is already under way in our laboratories. nanoparticles were carried out using standard Schlenk tubes, Fisher-Porter glassware and vacuum line techniques or in a glove-box (Braun) under an argon atmosphere. Reagents and solvents were degassed before use via a multi-cycle freeze-pumpthaw process. The (cyclooctadienyl)(1,5cyclooctadiene)cobalt(I) complex, [Co(3-C8H13)( 4-C8H12)], was purchased from NanomepsToulouse. 1-Heptanol, sodium persulfate, sodium hydroxide, sodium hexafluorosilicate, and sodium bicarbonate were acquired from Sigma-Aldrich. Hydrogen and argon were purchased from Alphagaz. 1-Heptanol was dried over activated molecular sieves (4 Å) prior to use and other reagents were employed as received unless otherwise specified. Solvents (THF, pentane, dichloromethane, diethyl ether) were purified before use by filtration on adequate alumina columns in a purification apparatus (MBraun) and handled under argon atmosphere. 4.2. Synthesis protocols Photosensitizers: The photosensitizers used in this work (see Fig. 1) were prepared according to literature data[31,32] and obtained with Cl- as counterion. Co nanoparticles: [Co(3-C8H13)(4-C8H12)] (120 mg, 0.43 mmol) as cobalt source and anhydrous 1-heptanol (20 mL) as both solvent and stabilizer were mixed into a Fisher-Porter reactor under an argon atmosphere inside a glove-box, leading to a brownish solution. Then, the Fisher-Porter reactor was pressurized with 3 bar of H2 and the reaction mixture was kept under vigorous stirring overnight, after which a dark colloidal dispersion was obtained. Excess H2 was eliminated under vacuum. A TEM grid was prepared under argon for TEM analysis of the crude colloidal solution. The application of a magnet on the reactor walls allowed to attract the Co NPs as a solid and then to isolate them from 1heptanol, which was then removed via cannula. The Co NPs were then washed with degassed anhydrous pentane (4 x 20 mL) and dried under vacuum. ICPOES (wt%): Co (35.35%). Co3O4 nanoparticles: Co3O4 NPs were prepared by treatment of isolated Co NPs under ambient air at room temperature during 6 days. Estimated Co content: 31.33%. PS-Co3O4 NP hybrids: A solution of PS1 (11 mg, 0.014 mmol, 0.24 eq) or PS2 (13 mg, 0.014 mmol, 0.24 eq) in methanol (0.4 mL) was added to a colloidal dispersion of Co3O4 NPs (33 mg, 0.058 mmol of Co3O4(heptanol)2.8 units) in 1-heptanol (0.6 mL). The reaction mixture was kept under vigorous stirring for 4 days in the dark. Then, precipitation of the crude product was achieved by adding isopropanol (1.5 mL) and diethyl ether (10 mL) and centrifuging at 1000 rpm for 10 min. The obtained 4. Experimental Section 4.1. General All procedures concerning the synthesis and preparation of samples for characterization of Co 14 crude product and water (1 mL) were introduced in a cellulose membrane bag for dialysis against deionized water (2 L). Dialysis was pursued until the external dialysis solution remained colorless. Then, centrifugation allowed to recover a solid, which was washed 3 times with a mixture of diethyl ether / isopropanol (8:2, v/v) to remove water and 3 times again with diethyl ether before drying under vacuum. The Ru/Co ratio of the obtained PS-Co3O4 NPs hybrids was calculated from ICP-OES measurements. PS-Co3O4 NPs hybrids were stored in the dark. ICP-OES (wt%): PS1-Co3O4, Ru (2.23%), Co (43.60%); PS2-Co3O4, Ru (3.12%), Co (35.22%). Recovered: 19 mg for PS1-Co3O4; 18 mg for PS2-Co3O4. 10 min. After that, the FTO/Co3O4 NPs electrode was dipped into a 0.5% wt PMMA dichloromethane solution for a few seconds (<10 sec) and air-dried. PMMA Coating. A poly(methyl methacrylate) (PMMA) coating was formed onto the FTO electrode supporting Co3O4 NPs by simply dipping this electrode in dichloromethane (DCM) with 0.5 % wt concentration of PMMA. After soaking the electrode in the PMMA solution for a few seconds (< 10 sec), the electrode was air-dried. Acknowledgements J. De T. acknowledges the Universitat Autònoma de Barcelona for a PIF doctoral grant. Financial supports were provided by MINECO / FEDER (CTQ2015-64261-R and CTQ2015-65268-C2-1-P), IDEX UNITI Emergence (UFTMIP: 2015-209-CIFD-DRD-127185) and CNRS. GDRI HC3A CNRS action (Catalunya / Midi-Pyrénées) and CTP regional action (Catalunya / CTP2013-00016, MidiPyrénées / n°13053026, Basque Country/CTP2013R03 and Région Aquitaine / n°13010761) are gratefully acknowledged for exchange funding between the partners. J.G.-A. acknowledges the Serra Húnter Program. 4.3. Preparation of electrodes RDE/Co3O4 and GC/Co3O4 NPs: 15 L of a dispersion of Co3O4 NPs (1 mg) in THF (250 L) were deposited onto the 0.07 cm2 glassy carbon (GC) disk from the rotatory disk electrode (RDE) or onto a GC electrode and let it dry under air. FTO/Co3O4 NPs-PMMA: A dispersion of Co3O4 NPs (2 mg) in 1-heptanol (500 μL) was prepared. Then, the NPs were deposited by spin-coating 15 μL of this dispersion onto an FTO electrode followed by evaporation of the solvent in a furnace at 100ºC for 15 References [1] S. Berardi, S. Drouet, L. Francàs, C. Gimbert-Suriñach, M. Guttentag, C. Richmond, T. Stoll, A. Llobet, Chem. Soc. Rev. 43 (2014) 7501-7519. [2] R. Bofill, J. García-Antón, L. Escriche, X. Sala, A. Llobet, Water Oxidation, in: J. Reedijk, K. Poeppelmeier (Eds.), Comprehensive Inorganic Chemistry, II, Elsevier, 2013, Vol 8, pp. 505-523. [3] R. Bofill, J. García-Antón, L. Escriche, X. Sala, J. Photochem. Photobiol. B: Biol. 152 (2015) 71-81. [4] N.S. Lewis, Science 351 (2016) aad19201. [5] M.D. Kärkäs, O. Verho, E.V. Johnston, B. Akermark, Chem. Rev. 114 (2014) 11863-12001. [6] B. Limburg, E. Bouwman, S. Bonnet, ACS Catal. 6 ( 2016) 5273–5284. [7] W.J. Youngblood, S.-H.A. Lee, Y. Kobayashi, E.A. Hernandez-Pagan, P.G. Hoertz, T.A. Moore, A.L. Moore, D. Gust, T.E. Mallouk, J. Am. Chem. Soc. 131( 2009) 926-927. [8] J. Youngblood, S.H. Lee, K. Maeda, T.E. Mallouk, Acc. Chem. Res. 42 (2009) 1966-1973. [9] D.L. Ashford, M.K. Gish, A.K. Vannucci, M.K. Brennaman, J.L. Templeton, J.M. Papanikolas, T.J. Meyer, Chem. Rev. 115 (2015) 13006-13049. [10] N. Kaveevivitchai, R. Chitta, R. Zong, M. El Ojaimi, R.P. Thummel, J. Am. Chem. Soc. 134 (2012) 1072110724. [11] H. Li, F. Li, B. Zhang, X. Zhou, F. Yu, L. Sun, J. Am. Chem. Soc. 137 (2015) 4332-4335. [12] A.M. Lopez, M. Natali, E. Pizzolato, C. Chiorboli, M. Bonchio, A. Sartorel, F. Scandola, Phys.Chem. Chem. Phys. 16 (2014) 12000-12007. [13] E.A. Karlsson, B.-L. Lee, R.-Z. Liao, T. Akermark, M.D. Kärkäs, V.S. Becerril, P.E.M. Siegbahn, X. Zou, M. Abrahamsson, B. Akermark, ChemPlusChem 79 (2014) 936-950. [14] R.K. Hocking, R. Brimblecombe, L.-Y. Chang, A. Singh, M.H. Cheah, C. Glover, W.H. Casey, L. Spiccia, Nat. Chem. 3 (2011) 461-466. [15] M.W. Kanan, D.G. Nocera, Science 321 (2008) 1072-1075. [16] H.-Y. Wang, J. Liu, J. Zhu, S. Styring, S. Ott, A. Thapper, Phys. Chem. Chem. Phys. 16 (2014) 3661-3669. [17] P. Garrido-Barrios, C. Gimbert-Suriñach, R. Matheu, X. Sala, A. Llobet, Chem. Soc. Rev. 46 (2017) 60886098. [18] P.G. Hoertz, Y.-I. Kim, W.J. Youngblood, T.E. Mallouk, J. Phys. Chem. B 111 (2007) 6845-6856. [19] H.S. Soo, A. Agiral, A. Bachmeier, H. Frei, J. Am. Chem. Soc. 134 (2012) 17104-17116. [20] A. Agiral, H.S. Soo, H. Frei, Chem. Mater. 25 (2013) 2264 -2273. [21] E. Edri, J.K. Cooper, I.D. Sharp, D.M. Guldi, H. Frei, J. Am. Chem. Soc. 139 (2017) 5458-5466. [22] C. Amiens, B. Chaudret, D. Ciuculescu-Pradines, V. Collière, K. Fajerwerg, P. Fau, M. Kahn, A. Maisonnat, K. Philippot, New J. Chem. 37 (2013) 3374-3401. [23] M. Salavati-Niasari, A. Khansari, CR. Chimie 17 (2014) 352-358. [24] J. Osuna, D. de Caro, C. Amiens, B. Chaudret, E. Snoeck, M. Respaud, J.-M. Broto, A. Fert, J. Phys. Chem. 100 (1996) 14571-14574. [25] M. Verelst, T.O. Ely, C. Amiens, E. Snoeck, P. Lecante, A. Mosset, M. Respaud, J.-M. Broto, B. Chaudret, Chem. Mater. 11 (1999) 2702-2708. [26] C. Amiens, Faraday Discuss. 125 (2004) 293-309. [27] K. Pelzer, K. Philippot, B. Chaudret, Z. Phys. Chem. 217 (2003) 1539-1547. [28] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.St.C. Smart, Appl. Surf. Sci. 257 (2011) 2717-2730. [29] J. Yang, H. Lu, W.N. Martens, R.L. Frost, J. Phys. Chem. C 114 (2010) 111-119. [30] J. Jiang, L. Li, Materials Letters 61 (2007) 4894–4896. [31] M.R. Norris, J.J. Concepcion, C.R.K. Glasson, Z. Fang, A.M. Lapides, D.L. Ashford, J.L. Templeton, T.J. Meyer, Inorg. Chem. 52 (2013) 12492-12501. [32] P. Jansa, O. Baszczyňski, E. Procházková, M. Dračínský, Z. Janeba, Green Chem. 14 (2012) 2282-2288. [33] S.A. Paniagua, A.J. Giordano, O’Neil L. Smith, S. Barlow, H. Li, N.R. Armstrong, J.E. Pemberton, J.-L. Brédas, D. Ginger, S.R. Marder, Chem. Rev. 12 (2016) 7117-7158. 16 [34] G. Guerrero, P.H. Mutin, A. Vioux, Chem. Mater. 13 (2001) 4367-4373. [35] G. Guerrero, J.G. Alauzun, M. Granier, D. Laurencin, P.H. Mutin, Dalton Trans. 42 (2013) 12569-12585. [36] C.C.L McCrory, S. Jung, I.M. Ferrer, S.M. Chatman, J.C. Peters, T.F. Jaramillo, J. Am. Chem. Soc. 137 (2015) 4347-4357. [37] C.C.L. McCrory, S. Jung, J.C. Peters, T.F. Jaramillo, J. Am. Chem. Soc. 135 (2013) 16977-16987. [38] S. Jung, C.C.L. McCrory, I.M. Ferrer, J.C. Peters, T.F. Jaramillo, J. Mater. Chem. A 4 (2016) 3068-3076. [39] S. Trasatti, O.A. Petrii, Pure Appl. Chem. 63 (1991) 711-734. [40] J.D. Benck, Z. Chen, L.Y. Kuritzky, A.J. Forman, T.F. Jaramillo, ACS Catal. 2 (2012) 1916-1923. [41] Y. Zhao, S. Chen, B. Sun, D. Su, X. Huang, H. Liu, Y. Yan, K. Sun, G. Wang, Sci. Rep. 5 (2015) 7629, DOI: 10.1038/srep07629. [42] Y. Liang, Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier, H. Dai, Nat. Mater. 10 (2011) 780-786. [43] J. Wu, Y. Xue, X. Yan, W. Yan, Q. Cheng, Y. Xie, Nano Res. 5 (2012) 521-530. [44] M.G. Walter, E.L. Warren, J.R. McKone, S.W. Boettcher, Q. Mi, E.A. Santori, N.S. Lewis, Chem. Rev. 110 (2010) 6446-6473. [45] M.F. Weber, M.J. Dignam, J. Electrochem. Soc. 131 (1984) 1258-1265. [46] Y. Gorlin, T.F. Jaramillo, J. Am. Chem. Soc. 132 (2010) 13612-13614. [47] N.H. Chou, P.N. Ross, A.T. Bell, T.D. Tilley, ChemSusChem 4 (2011) 1566-1569. [48] S.K. Singh, V.M. Dhavale, S. Kurungot, ACS Appl. Mater. Interfaces 7 (2015) 442-451. [49] H.A. Gasteiger, S.S. Kocha, B. Sompalli, F.T. Wagner, Appl. Catal. B 56 (2005) 9-35. [50] J. Suntivich, K.J. May, H.A. Gasteiger, J.B. Goodenough, Y. Shao-Horn, Science 334 (2011) 1383-1385. [51] K.-R. Wee, M.K. Brennaman, L. Alibabaei, B.H. Farnum, B. Sherman, A.M. Lapides, T.J. Meyer, J. Am. Chem. Soc. 136 (2014) 13514-13517. [52] M. Zhang, M. de Respinis, H. Frei, Nat. Chem. 6 (2014) 362-367. [53] H.-C. Chen, D.G.H. Hetterscheid, R.M. Williams, J.I. van der Vlugt, J.N.H. Reek, A.M. Brouwer, Energy Environ. Sci. 8 (2015) 975-982. [54] P. Comte, M.K. Nazeeruddin, F.P. Rotzinger, A.J. Frank, M. Grätzel, J. Mol. Catal. 52 (1989) 63-84. [55] P.K. Ghosh, B.S. Brunschwig, M. Chou, C. Creutz, N. Sutin, J. Am. Chem. Soc. 106 (1984) 4772-4783. [56] A. Lewandowska-Andralojc, D.E. Polyansky, R. Zong, R.P. Thummel, E. Fujita, Phys. Chem. Chem. Phys. 15 (2013) 14058-14068. [57] M.S. Baptista, J. Cadet, P.D. Mascio, A.A. Ghogare, A. Greer, M.R. Hamblin, C. Lorente, S.C. Nunez, M.S. Ribeiro, A.H. Thomas, M. Vignoni, T.M. Yoshimura, Photochem. Photobiol. 93 (2017) 912-919. [58] L. Francàs, C. Richmond, P. Garrido-Barros, N. Planas, S. Roeser, J. Benet-Buchholz, L. Escriche, X. Sala, A. Llobet, Chem. Eur. J. 22 (2016) 5261-5268. [59] I. Bazzan, A. Volpe, A. Dolbecq, M. Natali, A. Sartorel, P. Mialane, M. Bonchio, Catal. Today 290 (2017) 39-50. [60] A.L. Kaledin, Z. Huang, Y.V. Geletii, T. Lian, C.L. Hill, D.G. Musaev, J. Phys. Chem. A 114 (2010) 73-80. [61] H.S. White, W.G. Becker, A.J. Bard, J. Phys. Chem. 88 (1984) 1840-1846. [62] D. Rehm, A. Weller, Bunsenges Ber. Phys. Chem. 73 (1969) 834. [63] D. Rehm, A. Weller, Isr. J. Chem. 8 (1970) 259. [64] F. Bolletta, M. Maestri, L. Moggi, J. Phys. Chem. 77 (1973) 861-862. [65] A.J. Bard, M.A. Fox, Acc. Chem. Res. 28 (1995) 141-145. 17